A&A 667, A95 (2022)
https://doi.org/10.1051/0004-6361/202244092
© N. Oberg et al. 2022
Astronomy
&
Astrophysics
Circumplanetary disk ices
I. Ice formation vs. viscous evolution and grain drift
N. Oberg1,2 , I. Kamp1 , S. Cazaux2,3 , P. Woitke4,5 , and W. F. Thi6
1
2
3
4
5
6
Kapteyn Astronomical Institute, University of Groningen, PO Box 800, 9700 AV Groningen, The Netherlands
e-mail: oberg@astro.rug.nl
Faculty of Aerospace Engineering, Delft University of Technology, Delft, The Netherlands
University of Leiden, PO Box 9513, 2300 RA, Leiden, The Netherlands
Space Research Institute, Austrian Academy of Sciences, Schmiedlstrasse 6, 8042 Graz, Austria
Centre for Exoplanet Science, University of St. Andrews, North Haugh, St. Andrews, KY16 9SS, UK
Max-Planck-Institut für extraterrestrische Physik, Giessenbachstrasse 1, 85748 Garching, Germany
Received 23 May 2022 / Accepted 13 August 2022
ABSTRACT
Context. The large icy moons of Jupiter formed in a circumplanetary disk (CPD). CPDs are fed by vertically infalling circumstellar
gas and dust which may be shock-heated upon accretion. Accreted material is then either incorporated into moons, falls into the planet,
or is lost beyond the disk edge on relatively short timescales. If ices are sublimated during accretion onto the CPD we know there must
be sufficient time for them to recondense or moons such as Ganymede or Callisto could not form. The chemical timescale to form
sufficiently icy solids places a novel constraint on the dynamical behaviour and properties of CPDs.
Aims. We aim to explore the process of ice formation in CPDs to constrain which disk properties (such as the mass, viscosity, and
dust-to-gas ratio) are consistent with the formation of an icy moon system.
Methods. We use the radiation thermochemical code P RO D I M O (Protoplanetary Disk Model) to analyze how the radial ice abundance
evolves in CPDs. We consider different initial chemical conditions of the disk to explore the consequences of infalling material being
inherited from the circumstellar disk or being reset to atomic conditions by shock-heating. We contrast the timescales of ice formation
with disk viscous timescales and radial dust drift.
Results. We have derived the radial ice abundance and rate of ice formation in a small grid of model CPDs. Water ice can form very
efficiently in the CPD from initially atomic conditions, as a significant fraction is efficiently re-deposited on dust grains within <1 yr.
Radial grain drift timescales are in general longer than those of ice formation on grains. Icy grains of size a < 3 mm retain their icy
mantles while crossing an optically thin circumstellar disk gap at 5 au for L∗ < 10 L⊙ .
Conclusions. Three-body reactions play an important role in water formation in the dense midplane condition of CPDs. The CPD
midplane must be depleted in dust relative to the circumstellar disk by a factor 10–50 to produce solids with the ice to rock ratio of
the icy Galilean satellites. The CPD snowline is not erased by radial grain drift, which is consistent with the compositional gradient of
the Galilean satellites being primordial.
Key words. planets and satellites: formation – planets and satellites: composition – accretion, accretion disks –
protoplanetary disks – planets and satellites: individual: Jupiter – methods: numerical
1. Introduction
A general feature of regular satellite formation theory is that the
circumplanetary disk (CPD) consists of circumstellar material
accreted from within the vicinity of the planet (Lubow et al.
1999; Canup & Ward 2002; Alibert et al. 2005; Shibaike et al.
2019; Ronnet & Johansen 2020). If the planet is massive enough
to open a gap in the circumstellar disk, material continues to
flow into the gap (Kley 1999; Teague et al. 2019) and falls nearly
vertically onto the CPD (Tanigawa et al. 2012; Morbidelli et al.
2014). The CPD achieves a steady-state mass when this inflow is
balanced by outflow where gas either spirals into the planet or is
decreted beyond the Hill sphere (Canup & Ward 2002; Batygin
& Morbidelli 2020). Independently of disk viscosity, stellar tides
induce spiral waves in the CPD which transport angular momentum and promote accretion onto the planet at a rate on the order
10−7 MJ yr−1 (Rivier et al. 2012), suggesting that for a CPD
of mass ∼10−4 MJ (D’Angelo et al. 2002; Gressel et al. 2013;
Szulágyi et al. 2014) infalling gas spends only a limited time
inside the CPD before being lost (Canup & Ward 2002).
The timescale of radial dust drift in small disks is also predicted to be short (Pinilla et al. 2013; Shibaike et al. 2017; Rab
et al. 2019). A CPD could lose mm-size grains within 102 –103 yr
to aerodynamic drag against highly sub-Keplerian gas due to a
very steep radial pressure gradient (Zhu et al. 2018). The CPD
is thus a very dynamical system, potentially with both inwards
and outwards radial transport of both gas and dust on very short
timescales.
The amount of time which gas and dust spend within the
CPD becomes highly relevant if a chemical “reset” occurs. The
infalling circumstellar material may pass through one or more
accretion shocks (Lubow et al. 1999; Tanigawa et al. 2012; Zhu
2015; Szulágyi et al. 2016; Schulik et al. 2020) and can be
heated ≥1000 K during accretion onto the CPD (Szulágyi 2017;
Szulágyi & Mordasini 2017; Aoyama et al. 2018). For pre-shock
velocities in excess of 90 km s−1 the shock can be sufficiently hot
A95, page 1 of 19
Open Access article, published by EDP Sciences, under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0),
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
This article is published in open access under the Subscribe-to-Open model. Subscribe to A&A to support open access publication.
A&A 667, A95 (2022)
to leave most of the infalling gas atomic or ionized (Aoyama et al.
2018). The shock may also desorb the icy mantles of grains via
sputtering and thermal desorption, which for pre-shock velocities in excess of 8–10 km s−1 can effectively strip a grain of H2 O
ice (Woitke et al. 1993; Aota et al. 2015; Tielens 2021). Small
icy grains passing through the gap may also lose their volatile
contents to photodesorption prior to shock-heating (Turner et al.
2012). We refer to a “reset” scenario if shock-heating or photodesorption effectively destroys molecules in the accretion flow
to the CPD. In a reset scenario, the re-formation of ices in the
CPD must compete with viscous accretion and decretion of gas
and radial drift of dust. Alternatively, if circumstellar disk ices
survive incorporation into the CPD, we refer to an “inheritance”
scenario.
The Galilean satellites characteristically exhibit a radial compositional gradient of decreasing density with increasing distance from Jupiter. The inner moon Io is ice-free, while Europa
has an ice mass fraction of ∼6–9% (Schubert et al. 2004; Kuskov
& Kronrod 2005) and Ganymede and Callisto have ice mass fractions of 40–55% (McKinnon 1997; Schubert et al. 2004; Sohl
et al. 2002). Theoretically it appears challenging to reproduce the
compositional gradient by tidal heating (Bierson & Steinbrügge
2021) or impact-driven escape of volatiles (Dwyer et al. 2013).
Previously it has been proposed that the gradient was imprinted
during the formation of the moons by a radial temperature gradient in the CPD (Lunine & Stevenson 1982), but the relevant
chemical timescales have rarely been taken into account (Mousis
& Alibert 2006). It is an open question whether the gradient can
be produced primordially if the chemistry of infalling gas and
dust is reset. By analyzing the composition and abundance of
ices that are able to form within the relevant timescales we can
place a lower bound on how efficiently angular momentum is
transported within the CPD.
In this work, we investigated the balance of the competing
timescales of ice formation, dust grain drift, and gas viscous
flow to seek constraints on properties of the CPD such as viscosity, ice mass fraction, and dust-to-gas ratio. We considered
the two opposing extreme cases of full chemical inheritance and
reset in chemically evolving disk models utilizing a rate-based
modeling approach. In Sect. 2, we describe our modeling setup and the assumptions that we make. In Sect. 3, we analyze
the CPD time-dependent ice abundances for both the reset and
inheritance cases, and place novel constraints on the properties
of CPDs. In Sect. 4, we discuss the implications of these constraints and place them in the context of solar system formation.
We consider also the role that radial grain drift has in competing
with ice adsorption and desorption. In Sect. 5, we summarize our
key findings.
2. Methods
We considered two opposing scenarios; one in which the molecular gas-phase and ice chemistry of the circumstellar disk is
preserved during accretion onto the CPD (inherit) and one in
which it is lost (reset). In the former case, the CPD is initially
populated with gas and ices extracted from the circumstellar
disk. In the latter case the disk is initially populated by a fully
atomic gas and dust is free of ice. We followed the build-up
of ices in the CPD over time in a thermochemical disk model
(see Sect. 2.1) and extracted the molecular ice abundance and
composition as a function of time.
The net inflow and outflow rate of gas and solids from and
to the CPD is assumed to be zero, and that the disk mass is in
steady-state. The rate of gas outflow then sets an upper limit on
A95, page 2 of 19
the applicable chemical evolutionary timescales. Hereafter we
refer to this as the viscous timescale, defined as
MCPD
tvisc =
,
(1)
Ṁ
where MCPD is the mass of the CPD and Ṁ is the infall rate
of circumstellar material onto the CPD. The relatively short
tvisc considered in this work (see Sect. 2.1.2) may cause reactions with high activation energies to be kinetically inhibited,
although it has been noted that the relatively high densities
characteristic of CPDs may allow these reactions to proceed to
equilibrium (de Pater & Lissauer 2010). Nevertheless we contrast
the time-dependent results with the assumption of steady-state
chemistry. In steady-state chemistry the disk is allowed to evolve
for an indefinite time period until the rate of formation for every
gas and ice species is balanced by the corresponding rate of
destruction.
2.1. Disk modeling code
We used the radiation thermochemical disk modeling code
P RO D I M O1 (Protoplanetary Disk Model; Woitke et al. 2009,
2016; Kamp et al. 2010, 2017; Thi et al. 2011, 2020a) to explore
the formation rates and resulting abundances of various ices
in CPDs. P RO D I M O uses a rate equation based approach to
compute the gas chemistry using either a time-dependent or
steady-state solver. The model represents a 2D slice through
an axisymmetric disk, extending radially in distance from the
planet r and vertically in distance from the disk midplane z.
Our chemical network contains 13 elements and 235 atomic
and molecular species. Where not explicitly specified we used
the “large DIANA chemical standard” network as described in
Kamp et al. (2017). Reaction rates are mainly selected from the
UMIST2012 database (McElroy et al. 2013). Important threebody collider reactions are adopted from the UMIST2006 rate
file (Woodall et al. 2007). Gas-phase reactions within the CPD
produce molecular species which can then freeze-out on grain
surfaces. The rate of ice formation is determined by the available
grain surface area, dust temperature, and the rates of thermal,
photo-, and cosmic-ray desorption (see Sect. 2.1.1 for a detailed
description of ice formation).
We made the simplifying assumption that the chemical composition of the infalling material is distributed instantaneously
and homogeneously throughout the disk (see Appendix E for a
discussion of the potential impact of the rate of vertical gas mixing). We assumed that the CPD inherits micrometer to mm-sized
dust grains directly from the circumstellar disk. The vertical dust
stratification is calculated according to the method of Dubrulle
et al. (1995) and kept fixed prior to the calculation of chemical
evolution. The timescales of radial dust drift are calculated in
a post-processing step described in Sect. 2.2. The temperature
and radiation structure of the CPD is solved in steady-state and
then kept fixed during chemical evolution of inherited or reset
infalling material.
In Sect. 3.2, we considered the implications of including grain-surface chemistry reactions. With surface chemistry
P RO D I M O models explicitly the formation of H2 in the CPD
for which the inclusion of additional chemical species such
as hydrogenated PAH is necessitated (Thi et al. 2020b,a). The
selection of the additional species and reactions in the surface
chemistry network and their role in the eventual composition of
the ices will be discussed in an accompanying work focused on
the composition of the ices (Oberg et al., in prep.).
1
https://www.astro.rug.nl/∼prodimo/
N. Oberg et al.: Circumplanetary disk ices. I.
2.1.1. Ice formation
2.1.2. Properties of the disk models
Where conditions in the disk are appropriate, ices can condense
onto the grains in successive layers by physical van der Waals
bonding (physisorption). The adsorption rate of species i onto a
physisorption sites is
The circumstellar disk. To generate the chemical abundances for our “inheritance” scenario we considered the properties of the surrounding circumstellar disk from which material
is accreted onto the CPD. We used a two-step approach to model
the chemistry in our circumstellar disk. As a first step, the initial conditions are derived from a zero-dimensional “molecular
cloud” model, the parameters of which are listed in Table G.1.
This stage represents 1.7 × 105 yr (the estimated lifetime of the
Taurus Molecular Cloud TMC-1) of chemical evolution in a
pre-collapse molecular cloud state (McElroy et al. 2013). The
resulting chemical abundances of the majority of the most
common species agree within a factor 10 with the observed abundances in TMC-1(see Fig. G.1). These abundances are then used
as initial conditions for the 2D grid of cells in the circumstellar
disk model in the second step.
In the second step, the circumstellar disk model is evolved
for an additional 4 Myr to be consistent with the formation
timeline of Jupiter proposed to account for the distinct isotopic population of meteorites found in the solar system wherein
Jupiter undergoes runaway accretion >3.46 Myr after the formation of calcium-aluminium rich refractory inclusions (Kruijer
et al. 2017; Weiss & Bottke 2021). The surface density power law
and physical extent of the circumstellar disk is based on a modified “Minimum Mass Solar Nebula” (MMSN; Hayashi 1981).
A parameteric gap has been introduced which reduces the dust
and gas density at 5.2 au centered on the location of Jupiter. The
gap dimensions are parameterized by an analytical gap scaling
relation derived from hydrodynamical simulations and are consistent with a circumstellar disk viscosity of α ∼ 10−4 and disk
age of 4 Myr (Kanagawa et al. 2016). A detailed description of
the circumstellar disk model and gap structure methodology can
be found in Oberg et al. (2020). The relevant circumstellar disk
model parameters can be found in Table 1. The dust temperature,
surface density profile, and UV field strength in and around the
circumstellar disk gap can be seen in Fig. 1.
Finally we extract the chemical abundances from the circumstellar disk model at the outer edge of the gap at a radius of
8.2 au. The gap edge is defined as the radius at which the perturbed surface density profile reaches 50% of the unperturbed
profile. As material flows into the gap from above one pressure scale height (Morbidelli et al. 2014), we extract our initial
conditions for the CPD model at z = H (z = 0.5 au at r = 8.2 au).
The ambient conditions at this point are listed in Table 2.
The extracted abundances are then used as the initial chemical
composition for our “inheritance” scenario CPD.
Throughout this work we quantify the iciness of solids with
the ice mass fraction,
−1
Rads
= 4πa2 vth
i
i nd S i s ,
(2)
where 4πa2 is a dust grain surface area, a is the grain radius, vth
i is
the thermal speed (kB T g /2πmi )1/2 , kB is the Boltzmann constant,
T g is the gas temperature, mi is the mass of the gas-phase species,
nd is the dust number density, and S i is the sticking coefficient.
The ice adsorption rate coefficients are then
dn#,i
−3 −1
= Rads
i ni cm s ,
dt
(3)
where ni is the number density of the gas-phase species.
Physisorbed species can desorb thermally, photodesorb, or
desorb after a cosmic ray impact deposits energy in the grain.
For physisorption there is no desorption barrier and the desorption energy is equal to the binding energy Eib . The Arrhenius
formulation for the rate of thermal desorption is then
b
Rides,th = v0,i e−Ei /kB Td in s −1 ,
where the pre-exponential (frequency) factor v0,i is
s
2Nsurf Eib
v0,i =
.
π2 mi
(4)
(5)
Nsurf is the density of surface binding sites 1.5 × 1015 cm−2
(Hasegawa et al. 1992), and T d is the dust temperature.
Adsorption energies of major volatile species are listed in
Appendix B.
The photodesorption rate is computed using the UV field
strength calculated by the radiative transfer and photodissociation cross-sections. The photodesorption rate for species i is
des,ph
Ri
= πa2
nd
Yi χFDraine s−1 ,
nact
i
(6)
where Yi is the photodesorption yield, nact
i is the concentration of
the species in the active layers,
nact
= ni
i
if n#, tot 6 4πNsurf a2 nd
= ni Nact /Nlayer
if n#, tot > 4πNsurf a2 nd .
(7)
The number ofPphysisorbed layers Nlayer = n#, tot /(4πNsurf a2 nd )
and n#, tot = i n#,i is the total number of the physisorbed
species, 4πNsurf a2 is the number of binding sites per layer, and
Nact is the number of chemically active layers. χFDraine is a measure of the local UV energy density from the 2D continuum
radiative transfer (Woitke et al. 2009). The rate of cosmic-ray
induced desorption Rides,CR is calculated according to the method
of Hasegawa & Herbst (1993). The total desorption rate Rdes
i
is the sum of the thermal, photo- and cosmic ray induced desdes,ph
orption rates Rides,th + Ri
+ Rides,CR . The desorption rate of the
physisorbed species i is then
dni
act
−3 −1
= Rdes
i ni cm s ,
dt
(8)
fice =
mice
,
mrock + mice
(9)
where mice is the ice mass and mrock is the total rock (in this
case, dust) mass. The fice of solids in the inherited circumstellar
material is 0.48. At a single pressure scale height (z = H) there
is a factor ∼20 reduction of the initial, global dust-to-gas ratio
due to settling which is calculated according to the method of
Dubrulle et al. (1995) with αsettle = 10−2 such that the dust-togas ratio d/g at one scale height d/gz=H = 10−3.2 . Nevertheless
we test a range of different d/g values for the CPDs both above
and below this value (10−4 –10−2 ).
Survival of icy grains passing through the gap. Icy grains
must orbit within the optically thin gap for an unknown amount
A95, page 3 of 19
A&A 667, A95 (2022)
Table 1. Parameters for the solar circumstellar disk.
Parameter
Symbol
Value
Unit
Stellar mass
Stellar Luminosity
Effective temperature
UV luminosity
X-ray luminosity
M∗
L∗
T eff
LUV,∗
LX
1.0
0.84
4395
0.01
1030
M⊙
L⊙
K
L⊙
erg s−1
Disk mass
Disk inner radius
Disk outer radius
Column density power ind.
Reference scale height
Flaring index
Mdisk
Rin
Rout
ǫ
H10au
β
0.001
0.1
100
1.5
1
1.15
M⊙
au
au
–
au
–
Minimum dust size
Maximum dust size
Dust size power law index
Dust-to-gas ratio
amin
amax
apow
d/g
0.05
3000
3.5
10−2
➭m
➭m
Dust composition:
Mg0.7 Fe0.3 SiO3
Amorphous carbon
Vacuum
Table 2. Conditions at the gap edge “inheritance” point of the circumstellar disk at heliocentric distance 8.2 au, altitude 0.5 au (1 pressure
scale height) above the midplane.
Parameter
Hydrogen density
Optical extinction
Dust-to-gas ratio
Dust temperature
Gas temperature
Ice mass fraction
Symbol
Value
Unit
nH,in
AV,in
d/g in
T d,in
T g,in
fice,in
1010
1.01
10−3.23
57.0
57.3
0.48
cm−3
–
–
K
K
–
–
–
60%
15%
25%
Notes. Stellar temperature and luminosity are selected from the premain sequence stellar evolutionary tracks of Siess et al. (2000) for t =
4 Myr. Stellar UV and X-ray luminosities for a representative Class II
T Tauri star are adopted from Woitke et al. (2016).
Fig. 2. Ice abundance in the circumstellar disk gap normalized to the
initial abundance of ice for various stellar luminosities as function of
time after the onset of UV exposure. The width of each track represents
the range of ice sublimation rates corresponding to variable conditions
across the gap region (z = 0–0.5 au, r = 5.2–8.2 au).
Fig. 1. 2D dust temperature distribution around the circumstellar disk
gap in the range 40–120 K (top). Solid, dashed, and dotted black contours indicate the relative UV field strength χ (see Eq. (12)). The white
circle with black border represents the position of Jupiter and the horizontal bar indicates the physical extent of the CPD. The white cross with
black border represents the location from which the inherited chemistry
is extracted at one scale height z = H. Hydrogen nuclei column density (blue line) with unperturbed circumstellar disk density profile for
comparison (light blue dashed line; bottom).
of time prior to being accreted onto the CPD. We considered
whether ices on grains can survive against thermal- or photodesorption while crossing through the circumstellar disk gap. The
dust and gas temperatures in the gap are not closely coupled
as a consequence of the low densities. While the gas temperature extracted from the model at the midplane is ∼200 K, the
corresponding dust temperature is 48 ± 2 K. Given that pressures
A95, page 4 of 19
in the gap range from 10−12 –10−10 bar, water ice is stable on
the relevant timescales in the absence of irradiation (Lodders
2003). However, the actual ice abundance at the gap midplane
is negligible in steady-state. Despite the presence of a shadowing inner disk, ice within the gap are sublimated as a result of the
significant stellar background radiation scattered into the gap.
To assess the longevity of ices crossing through the gap we
populated the gap region with the “inheritance” chemical abundances found exterior to the gap and produce snapshots at regular
intervals. The resulting decline in ice abundance as a function
of time is shown in Fig. 2 for various stellar luminosities. The
differing luminosities correspond to expected properties of the
sun for a stellar age of 0.1, 0.5, 1 and 4 Myr (Siess et al. 2000).
Turner et al. (2012) suggests that grains entering the gap are
accreted generally within a single orbital period or 10 yr. We
find that for a moon formation time >1 Myr after CAI formation (L∗ ≤ 2.34 L⊙ ), grains retain >99% of their volatile content
during gap-crossing if they reach the vicinity of the planet within
10 yr. A “full” inheritance scenario is thus not excluded by conditions within the gap, but would instead rely on shock-heating
at the CPD surface.
The circumplanetary disks. The CPD is an actively-fed
accretion disk with a steady-state mass proportional to its viscosity and mass infall rate. We considered an optically thick
CPD of mass 10−7 M⊙ as well as a lower mass CPD (10−8 M⊙ )
which is optically thin everywhere outside the orbit of Callisto
(r > 0.03 RH where RH = 0.34 au is the Hill radius). For a
N. Oberg et al.: Circumplanetary disk ices. I.
Table 3. Variation of parameters for the circumplanetary disk model
grid.
Model id
(7–10)
(7–11)
(8–11)
(8–12)
M (M⊙ )
Ṁ (M⊙ yr−1 )
tvisc (yr)
10
10−7
10−8
10−8
10
10−11
10−11
10−12
10
104
103
104
−7
−10
3
α
10−2.7
10−3.6
10−2.7
10−3.6
Notes. Model id’s are of the format (A–B) where the CPD mass is
10−A M⊙ and where the accretion rate onto the CPD is 10−B M⊙ yr−1 .
Table 4. Parameters common to the CPD models.
Parameter
Symbol
Value
Unit
Mp
Lp
T eff,p
LUV,p
χ
T back
1.0
10−5
1000
0.01
3 × 103
50
MJ
L⊙
K
Lp
–
K
Disk inner radius
Taper radius
Disk outer radius
Column density power ind.
Flaring index
Reference scale height
Rin,CPD
Rtap,CPD
Rout,CPD
ǫCPD
βCPD
H0.1 au
0.0015
0.12
0.35
1.0
1.15
0.01
au
au
au
–
–
au
Maximum dust size
Dust-to-gas ratio
amax,CPD
d/g
3000
10−3.3
µm
–
Planet mass
Planetary luminosity
Effective temperature
UV luminosity
Background UV field
Background temperature
Notes. Parameters which are not common to the CPD models are listed
in Table 3. Where not specified the CPD parameters are identical to the
circumstellar disk parameters listed in Table 1.
Jovian-mass planet these represent planet-disk mass ratios of
∼10−4 and 10−5 , respectively. The CPDs are thus of the “gasstarved” type, and do not instantaneously contain the mass
required to form a moon system as massive as the Galilean one.
The outer radius of the CPD is set as the planetary Hill radius
RH , however an exponential decline in the surface density profile
is parameterized to begin at RH /3 corresponding to the outermost stable orbit set by tidal interaction and angular momentum
considerations (tidal truncation radius; Quillen & Trilling 1998;
Ayliffe & Bate 2009; Martin & Lubow 2011). An empty inner
magnetospheric cavity is assumed to exist as the result of magnetic interaction with the planet (Takata & Stevenson 1996;
Shibaike et al. 2019; Batygin 2018). The parameterized radial
surface density profiles of the high- and low-mass CPDs can
be found in Fig. 3 together with the resulting optical extinction
profiles derived from the continuum radiative transfer.
The small parameter variation grid of models exploring possible CPD mass and viscosity can be found in Table 3 along with
the model id’s. The format of the id is (x–y) where x is related
to the CPD mass by MCPD = 10−x M⊙ and y to the mass infall
rate (and thus viscosity) by ṀCPD = 10−y M⊙ yr−1 . The list of
parameters which are common between the reference CPDs can
be found in Table 4.
While the mechanism that produces angular momentum transport in accretion disks is not well
understood, it is known that molecular viscosity alone is far too
CPD viscosity.
weak to explain observations (Shakura & Sunyaev 1973; Pringle
1981). The efficiency of angular momentum transport is parameterized by the dimensionless α-viscosity (Shakura & Sunyaev
1973) which for a circumstellar disk may have a broad distribution ranging from ∼10−5 –10−1 (Rafikov 2017; Ansdell et al.
2018; Villenave et al. 2022).
Disk gas with a sufficiently high ionization fraction couples
to the stellar or planetary magnetic field such that the magnetorotational instability (MRI) induced turbulence may provide the
source of the effective viscosity and produce α ≥ 10−3 (Balbus
& Hawley 1991; Hawley et al. 1995; Balbus 2003). In the case of
a CPD, MRI induced-turbulence may be inhibited by the short
orbital periods and presence of magnetic dead-zones, limiting
gas transport to a thin surface layer (Fujii et al. 2014). Even if
the CPD were effectively inviscid, tidal interaction with the star
may promote a minimum rate of angular momentum transport
through the excitation of spiral waves (Rivier et al. 2012). In
our P RO D I M O model, the α-viscosity of the disk is not explicitly specified. Instead, a mass accretion rate is specified which
controls the heating rate of the disk through viscous dissipation.
The disk mass, accretion rate, and viscosity are highly
degenerate properties. 3D hydrodynamical modeling of gas
delivery into the vicinity of a Jupiter mass planet at 5 au
suggest Ṁ = 10−9.3 M⊙ yr−1 of gas (Szulágyi et al. 2022). Stellar tidal perturbation may produce a minimum accretion rate
10−9.7 M⊙ yr−1 (Rivier et al. 2012). In the PDS 70 system, two
massive planets are observed to be accreting gas in a large dust
cavity (Wagner et al. 2018; Keppler et al. 2018; Haffert et al.
2019). K-band observations of PDS 70 b with the VLT are
consistent with Ṁ = 10−10.8 −10−10.3 M⊙ yr−1 (Christiaens et al.
2019) with similar values estimated for PDS 70 c (Thanathibodee
et al. 2019). HST UV and Hα imaging of the protoplanet
PDS 70 b suggest Ṁ = 10−10.9 −10−10.8 M⊙ yr−1 (Zhou et al.
2021). Based on these observational and theoretical constraints
we adopt Ṁ = 10−10 M⊙ yr−1 (with a heating rate corresponding to α ≈ 10−2.7 ) and Ṁ = 10−11 M⊙ yr−1 (α ≈ 10−3.6 ) for the
high-mass CPD, representing viscous timescales of 103 and
104 yr, respectively, over which the majority of the CPD mass is
replaced by freshly accreted material (see Eq. (1)). For the lowmass CPD we adopt Ṁ = 10−11 M⊙ yr−1 and Ṁ = 10−12 M⊙ yr−1
to represent the same α-viscosities and tvisc .
The viscous heating is determined according to the method
of D’Alessio et al. (1998). The half-column heating rate is
Fvis (r) =
q
3GMp Ṁ
(1
−
Rp /r) erg cm−2 s−1 ,
8πr3
(10)
where G is the gravitational constant, Mp is the planet mass, r
is the distance to the planet, and Rp is the planetary radius. We
convert the surface-heating to a heating rate per volume as
Γvis (r, z) = Fvis (r) R
ρd (r, z)
ρd
(r, z′ )dz′
erg cm−3 s−1 ,
(11)
where ρd is the mass density of the dust at radius r and height z.
The heating is applied directly to the dust. The resulting midplane dust temperature profile of the CPDs can be found in
Fig. 4. The temperature profiles have been calculated using a
new diffusion-approximation radiative transfer solver, which is
described in Appendix A.
Sources of CPD external irradiation. The rate of photodesorption plays an important role in determining the ice
abundance in the outer optically-thin region of the CPDs. Potential sources of radiation include the planet, the star, and nearby
A95, page 5 of 19
A&A 667, A95 (2022)
Fig. 3. Surface density (red) and midplane FUV field strength in units
of the Draine field χ (blue) of the high (10−7 M⊙ ) and low (10−8 M⊙ )
mass CPDs. The four circles indicate the present-day radial position
of the Galilean satellites and their position on the ordinate is arbitrary.
The striped gray region on the left indicates an empty inner magnetic
cavity and the light gray region on the right indicates the gravitationally
unstable zone outside of RH /3 (where RH = 0.35 au).
Fig. 4. Radial midplane dust temperature profiles of the four reference
CPDs (d/g = 10−3.3 ), labelled by the model id’s in Table 3. The four
circles indicate only the present-day radial position of the Galilean satellites. The striped gray area on the left indicates the inner cavity. The
light gray area on the right indicates the gravitationally unstable zone
outside of RH /3 (where RH = 0.35 au).
massive cluster stars. For the planet we assume the runaway gas
accretion phase has ended and that the luminosity has correspondingly declined to 10−5 L⊙ with a surface temperature of
1000 K where it remains relatively constant for 10 Myr (Hubickyj
et al. 2005; Marley et al. 2007; Spiegel & Burrows 2012).
We parameterize the isotropic background radiation intensity
with the dimensionless parameter χ. The background intensity
is then the sum of a diluted 20 000 K blackbody field and the
cosmic microwave background,
bg
Iν = χ · 1.71 · Wdil Bν (20 000 K) + Bν (2.7 K),
(12)
where the dilution factor Wdil = 9.85357 × 10−17 such that a
value of χ = 1 corresponds to the quiescent interstellar background, or “unit Draine field” (Draine & Bertoldi 1996; Röllig
et al. 2007; Woitke et al. 2009). Irradiation by the young sun
A95, page 6 of 19
provides a minimum χ ∼ 3000 at 5 au (see Fig. 1) despite partial shadowing by an inner disk (Oberg et al. 2020). We adopted
the same value for the strength of the FUV irradiation in the
midplane at Jupiter’s location although it is contingent on our
assumptions regarding the stellar UV luminosity and geometry
of the inner circumstellar disk. Independently of these factors,
Oberg et al. (2020) find that interstellar radiation results in a
mean χ of O(3) in the gap, as the Sun is believed to have formed
in a relatively dense stellar cluster (Adams 2010; Portegies Zwart
2019) containing massive OB stars (Pfalzner 2013). External
irradiation heats dust and gas on the surface and in the outer
regions of the optically-thin CPD midplane. 3D dust radiative
transfer modelling of gap-embedded CPDs suggests scattered
stellar radiation can be the dominant heating source in the outer
regions of a CPD (Portilla-Revelo et al. 2022). We assumed that
the outer edge of the CPD is in thermal equilibrium with the
surroundings and set a CPD background temperature lower limit
of 50 K equal to that of dust in the circumstellar disk gap (see
Fig. 1).
CPD dust mass and grain size population. The dust-togas ratio is a key parameter that regulates the eventual ratio of
ice to rock in the CPD. Planet-induced pressure bumps at the
gap edges of the circumstellar disk may filter out dust particles
above ∼100 ➭m in size (Rice et al. 2006; Zhu et al. 2012). For a
dust grain size population power law exponent of –3.5, minimum
grain size amin 0.05 µ m, maximum grain size amax 3000 µm, a
filtering of grains larger than 100 µm would be reduced the mass
of dust entering the gap by a factor ∼30. As an opposing view
Szulágyi et al. (2022) find that a planet within the gap can stir
dust at the circumstellar disk midplane and produce a very high
rate of dust delivery to the CPD, such that the dust-to-gas ratio
can even be enhanced in the CPD to ∼0.1 for a Jupiter mass
planet at 5 au.
The dust population within the CPD may also rapidly evolve.
It is possible that mm-sized grains are quickly lost to radial
drift in 102 –103 yr (Zhu et al. 2018). Similarly Rab et al. (2019)
used the dust evolution code two-pop-py (Birnstiel et al. 2012) to
show that for an isolated CPD onto which new material does not
accrete, an initial dust-to-gas ratio of 10−2 can in only 104 yr be
reduced to <10−3 and in 105 yr to <10−4 . However, larger dust
grains may become trapped in CPDs (Drażkowska
˛
& Szulágyi
2018; Batygin & Morbidelli 2020). Additionally, we expected
that in an embedded, actively-accreting CPD the dust will continually be replenished and that a higher steady-state dust-to-gas
ratio will be achieved. Given these considerations we tested
various possible dust to gas ratios ranging from 10−4 –10−2 in
Appendix H.
2.2. Dust grain drift within the CPD
In P RO D I M O, the chemistry is solved for each grid cell without accounting for radial dust or gas transport. Instead we
used properties of the P RO D I M O model output to inform the
radial dust drift calculations in a post-processing step to compare timescales of chemical evolution vs. dust drift. Disk gas
which is pressure-supported orbits at sub-Keplerian velocities
such that larger grains feel a headwind and rapidly drift inwards
(Weidenschilling 1977). In a circumstellar disk the degree of subKeplerianity can be <1% while in a CPD it could be as high
as 20–80% due to significant gas pressure support (Armitage
2007; Drażkowska
˛
& Szulágyi 2018). We considered whether
the timescale of grain drift can compete with the processes that
shape grain composition In the Epstein regime, the radial grain
N. Oberg et al.: Circumplanetary disk ices. I.
drift velocity vr,d can be approximated as
vr,d =
vr,g T s−1 − ηvK
T s + T s−1
,
(13)
where vr,g is the radial velocity of the gas, T s is the dimensionless
stopping time of a grain,
v
k
T s = ts
,
(14)
r
where ts is the stopping time, vk is the Keplerian orbital velocity,
and r is the radius in the CPD. The stopping time is
!
!
ρgrain a
ts =
,
(15)
ρgas vth
where a is the grain size, ρgas is the gas density, ρgrain is the material density of the dust grains, and vth is the thermal velocity of
the gas: cs (8/π)0.5 where cs is the speed of sound. The parameter
η is
2
c
(16)
η = n 2s ,
vk
and n is the power law exponent of the radial pressure gradient
(Armitage 2010). We extract the gas density ρgas , sound speed
cs , and pressure gradient from our disk models to consistently
determine the grain drift velocities for a grain material density
ρgrain = 3 g cm−3 . The Epstein regime is valid where the dust particle size a ≤ 9λ/4 where λ is the mean free path of the gas
particles. In the inner CPD the gas density is sufficiently high
that this condition can be violated. For the high-mass CPD this
occurs inside of the orbit of Callisto and for the low-mass CPD
inside of the orbit of Europa for grains less than 1 cm in size, in
which case we transition to the Stokes regime and calculate the
drift velocities accordingly (Weidenschilling 1977).
We adopted several simplifying assumptions regarding the
radial velocity structure of the gas. The centrifugal radius rc
of the CPD is the orbital radius at which the specific angular
momentum is equal to the average of the infalling material and
where solid material is accreted onto the CPD (Canup & Ward
2002; Sasaki et al. 2010). In the case of a Jupiter-mass planet this
lies near the orbit of Callisto. Rather than accreting at precisely
this radius infalling matter have some intrinsic spread in angular
momentum (Machida et al. 2008). We adopted the prescription
of Mitchell & Stewart (2011) that material will accrete between
16 and 28 RJ . The gas falling onto the CPD at rc will viscously
spread radially away from where it is accreted (Pringle 1981).
Hence interior to rc gas flows towards the planet and exterior to
rc it will flow outwards. Recently it has been proposed that to
effectively trap dust and allow for planetesimal growth within
the CPD, gas may indeed need to flow predominantly away from
the planet and thus form a decretion disk (Batygin & Morbidelli
2020).
In our high viscosity models, a parcel of gas that accretes
onto the CPD near rc and flows outwards must travel ∼0.3 au
within 103 yr to be consistent with tvisc and thus has a mean
outwards radial velocity on the order of 1 m s−1 . For our low
viscosity case vr,gas ∼ 0.1 m s−1 .
Fig. 5. Rate of water ice formation as a function of time at different
radii beyond the snowline in the high-mass, high-viscosity CPD (7–10),
to illustrate the two distinct stages of water ice formation. All values are
normalized to the maximum formation rate at 10−5 yr, r = 0.08 RH . The
times where the decline in the abundance of free O and free H end are
indicated by the gray vertical lines. The starting time t0 is defined as
when the infalling gas is shock-heated to an atomic and ionized state.
optically thin CPD (10−8 M⊙ ) both with d/g = 10−3.3 . For each
unique mass we considered viscous timescales of 103 and 104 yr.
For each of the four resulting CPDs we tested two initial compositions: of chemical inheritance and reset, for a total of eight
models.
3.1. Timescales of ice formation
In the following section we discuss by which pathways and at
what rate water (ice) formation occurs in a chemically reset CPD.
Thereafter we contrast these results with that of the CPDs which
inherit an initial chemical composition from the circumstellar
disk.
3.1.1. Reset
In our “reset” scenario the species initially present are exclusively atomic (with the exception of PAHs) and singly or doubly
ionized. All hydrogen is initially present in the form of H+ .
The reset case is characterized by the formation of ice where
it is stable against desorption. Ice formation is suppressed in
the innermost region of the CPD due to dust temperatures in
excess of 160–180 K. In the outer region of the CPD ice formation is suppressed by the low optical depths and correspondingly
high intensity of background radiation which causes ices to
photodesorb. Between these two boundaries ices begin to form.
The freeze-out occurs step-wise, with two characteristic
timescales on which ice formation proceeds. The rate of water
ice formation at different disk radii as a function of time is shown
in Fig. 5. Within 0.01−1 yr the first step is completed. This initial, rapid water formation occurs primarily via a path that begins
with a three-body recombination reaction important at temperatures below 2800 K (Hidaka et al. 1982; Tsang & Hampson
1986),
3. Results
H + O + M → OH + M,
(17)
In our reference model grid (Table 3), we considered the case of
a higher mass, optically thick CPD (10−7 M⊙ ), and lower mass,
where the third body M = H, H2 , or He, and to a lesser extent
by H + O → OH + photon. The water then forms by radiative
A95, page 7 of 19
A&A 667, A95 (2022)
association of the OH with free hydrogen;
OH + H → H2 O + photon,
3.2. The midplane ice mass fraction
(18)
after which it adsorbs to a grain. Water formation via the
reactions (17) and (18) remains proportional to the declining
abundance of free H and ends when it is depleted around 10−1 yr.
Typically half of the maximum possible amount of water ice that
could form is produced during this stage.
The first stage of water ice formation proceeds inside-out
as a result of the strong density-dependence of the collider
reaction. The second stage proceeds outside-in due to the pathway’s dependence on intermediate species produced by photoreactions. This can be seen in Fig. 5 where the inner disk
formation rate (red lines) is initially higher while later the outer
disk rate (blue lines) is relatively higher. At this lower rate of formation the total mass of water ice doubles at the midplane within
∼1 Myr, approximately half of which forms by
NH2 + NO → N2 + H2 O,
(19)
near the snowline. The second stage of ice formation also
involves the freezing-out of NH3 and other more volatiles
species. We find that the relatively unstable NH2 exists in abundance at such high densities (nH ∼ 1015 cm−3 ) due to the threebody reaction N + H2 + M → NH2 + M. Although the reaction
rate has been experimentally determined k0 = 10−26 cm6 s−1
(Avramenko & Krasnen’kov 1966), it has been noted as being
in need of revisiting due to its importance in water formation via
NH2 + O → NH + OH (Kamp et al. 2017). We have adopted a
rate more typical of collider reactions (10−30 cm6 s−1 ), which still
produces enough NH2 for this path to play an important role. The
other half of water ice is formed via the more “classical” route
H3 O+ + e− → H2 O + H,
(20)
and in the outermost part of the disk where water ice is stable via
H3 O+ + H2 CO → H3 CO+ + H2 O.
(21)
In the high-mass CPDs water ice formation is typically still
ongoing by the end of the viscous timescale, and so the midplane ice abundance is not able to converge to the level seen in
steady-state within the viscous timescale.
3.1.2. Inheritance
The evolution of the inheritance case is characterized by the
desorption of ices in regions where they are not stable against
thermal or photo-desorption. Where water ice is stable very little
additional water formation occurs within the viscous timescale.
Icy grains interior to the snowline sublimate typically within
10−5 yr, and a “snowline” (water iceline) is clearly established.
In some cases the snowline can take longer to stabilize, shifting
outwards over time, and sometimes continues to evolve radially
for up to 105 yr. This is notable in the (8–11) CPD where the
snowline moves from 0.01 RH at 10−5 yr to 0.03 RH after 104 yr.
In practice, this implies that there exists a radial span in which
the composition of radially drifting icy grains does not immediately begin to reflect the ambient conditions. This is discussed in
Sect. 4.2.
In the outer optically thin region of the CPDs, ices are also
lost to photodesorption although the process is slower than the
thermal desorption occurring in the inner disk. Desorption in this
area is typically complete within 103 –105 yr and has not always
reached steady-state by the end of the viscous timescale.
A95, page 8 of 19
While water ice can form efficiently in chemically reset CPDs,
the final fice of the solids depends strongly on the total dust mass
in the midplane. We explore the role of the global d/g ratio in the
reset and inheritance cases in Fig. H.1. A canonical dust-to-gas
ratio of 10−2 produces at most grains with an ice mass fraction of
<0.1 and is nowhere consistent with the composition of the icy
Galilean satellites. In contrast a dust-to-gas ratio of 10−3.3 results
in solids with fice both above and below the maximum Galilean
fice for the high and low-mass CPDs, respectively. Hereafter the
global dust-to-gas ratio of 10−3.3 is considered in discussions of
the four reference CPDs.
We show the state of the radial ice mass fraction for the CPDs
with d/g = 10−3.3 on their respective viscous timescales to allow
for direct comparison between the inheritance and reset cases
in Fig. 6. For completeness we include the results where grainsurface chemical reactions are utilized.
The midplane fice is also dependent on the degree of dust
sedimentation (settling). A general feature of the fice profiles is
an inner maximal peak at the snowline followed by a decline
towards the outer edge of the disk. As dust settling is more
efficient at larger radii, fice reduces accordingly with increasing radius. Settling of dust to the midplane is counteracted by
stochastic advection by turbulent eddies in the gas. We assume
that the value of turbulent-α used to determine the degree of
settling is equal to the global viscous α used to determine the
heating rate by viscous dissipation. In the low viscosity CPDs
(7–11) and (8–12) dust settling towards the midplane is thus proportionally enhanced. In the low-viscosity cases the dust density
is enhanced in the midplane minimally by a factor ∼3 over the
global d/g, increasing to a factor ∼20 at RH /3. As the degree of
settling is also dependent on the adopted surface density power
law exponent ǫ we explore the impact of deviation from the
assumed ǫ = 1 in Appendix C. Given that we have no reason
to believe this value will depart significantly from the range
1.0–1.3, the resulting fice in the inner disk should differ from our
reference result by no more than 25–30% interior to the ammonia
iceline at ∼70 K.
In general the high-mass chemically reset CPDs (7–10) and
(7–11) are not able to converge entirely towards a steady-state
ice abundance in either the “fast” (103 yr) or “slow” (104 yr) viscous timescales as gas-phase CO is more stable and contains a
relatively larger fraction of the total oxygen budget for longer.
As a result chemically reset CPDs contain on average less water
ice than those which inherit their ices from the circumstellar
disk. In contrast, the low-mass chemically reset CPDs (8–11) and
(8–12) are able to converge towards the ice abundances seen in
the inheritance cases within 100 yr.
The role of surface chemistry. The duration of the initial stage in which water formation is rapid is dependent on
the availability of atomic H. When H2 formation is complete
this stage ends. The formation of H2 is treated differently with
the inclusion of grain surface chemistry. When the chemistry is
limited to gas-phase reactions and direct adsorption/desorption
only, H2 formation proceeds via the analytic rate determined by
Cazaux & Tielens (2002). When surface chemistry is included
H2 formation is instead modeled explicitly via reactions involving hydrogenated PAHs (H-PAH; Boschman et al. 2012; Thi et al.
2020a).
A chemical reset poses a scenario in which H2 and H2 O
formation occur simultaneously. The analytic rate of Cazaux &
Tielens (2002) presupposes that chemisorbed H plays a role in
N. Oberg et al.: Circumplanetary disk ices. I.
(a) High-Mass High-Viscosity (7-10)
(b) High-Mass Low-Viscosity (7-11)
(c) Low-Mass High-Viscosity (8-11)
(d) Low-Mass Low-Viscosity (8-12)
Fig. 6. Midplane ice fraction at t = tvisc for the CPDs with d/g = 10−3.3 . The four circles indicate the present day radial location of the Galilean
satellites and the uncertainty bars represent their possible range of ice fractions. The thin dotted line indicates the initial radial ice fraction in the
inheritance cases. The filled gray region on the right indicates the gravitationally unstable zone outside of RH /3.
H2 formation on silicate or carbonaceous surfaces, in which H
goes through an intermediate stage of being chemically, rather
than physically, bound to the grain surface. We find that prior to
atomic H depletion several (>3) monolayers of H2 O have formed
on the average sized grain. The formation of H2 via chemisorbed
H should thus be suppressed in these regions as the water layers prevent H atoms from chemisorbing to the grain surface
(Wakelam et al. 2017). In the absence of chemisorbed H, H2 formation on dust is strongly reduced and H2 formation proceeds
primarily via H-PAH. The H2 formation rate under these conditions is less efficient than the analytic rate from Cazaux &
Tielens (2002) near the inner edge of the snowline at 150–170 K.
The relatively slower formation of H2 and the resulting prolonged availability of atomic H results in a ∼30–100% increase
in water ice abundance interior to the NH3 iceline prior to tvisc
and hence narrows the gap between the inheritance and reset
cases in this region. Water ice formation in the inner disk is
further enhanced by the inclusion of O2 H in the surface chemistry network. Gas-phase three-body reactions with O2 produce
O2 H which in turn lead to early OH formation. The gas-phase O2
reservoir is thus depleted and efficiently converted via OH into
H2 O via the three-body reaction
O2 + H + M → O2 H + M,
(22)
with rates adopted from UMIST2006 (Atkinson et al. 2004;
Woodall et al. 2007), which is highly efficient at the densities
in the CPD, and thereafter
O2 H + H → OH + OH,
(23)
OH + H → H2 O + photon.
(24)
In the outer disk, the ice mass fraction can be enhanced
relative to the gas-phase chemical network as the freezingout of more volatile species is facilitated by grain surface
reactions. CO2 ice is readily formed on the surface via
OH + CO → CO2 + H (Oba et al. 2010; Liu & Sander 2015) for
which we adopt an effective barrier of 150 K (Fulle et al. 1996;
Ruaud et al. 2016). The formation of carbon bearing ices begins
A95, page 9 of 19
A&A 667, A95 (2022)
timescales of thermal desorption and the timescale of rapid ice
formation. The composition of grains will thus correspond to the
fice profiles derived in Sect. 3 except in the case of the low-mass,
high-viscosity CPD (8–11). In this CPD icy grains can cross into
the “thermal desorption region” but only begin desorbing once
they approach the position of Europa.
Fig. 7. Radial drift velocity of particles (upper panel) and radial velocity of the gas (lower panel) in the high-mass, high-viscosity CPD
(7–10). The position of the Galilean satellites on the ordinate is arbitrary. The gray region on the right indicates the tidally unstable region
beyond RH /3. The striped region on the left indicates the inner cavity. Solid lines indicate the case of a decretion disk where gas falls
onto the CPD at the centrifugal radius (0.03 RH ). Dashed lines indicate
the grain drift and gas radial velocity in the case of a pure accretion
disk.
to significantly influence the fice of the chemically reset CPDs
only after 103 yr and hence the effect on the high-viscosity CPDs
with tvisc = 103 yr is less pronounced. The formation and composition of these ice impurities will be discussed in detail in an
accompanying work (Oberg et al., in prep.).
3.3. Grain drift vs. adsorption and desorption
We calculated the velocity of radially drifting grains in the four
reference CPDs and showcase the results for the high-mass highviscosity CPD (7–10) in Fig. 7. We solved for the total time it
takes grains of several sizes to reach the inner disk edge. The
resulting timescale of grain drift can be seen in Fig. 8 which
shows the time for a grain deposited at radius r to reach the CPD
inner edge. Grains which become trapped are not included in
Fig. 8, as they do not reach the disk inner edge. The regions
where thermal desorption, ice formation, and photodesorption
predominantly shape grain mantle composition, as well as their
respective timescales, are indicated in the figure. These are the
timescales with which grain drift competes.
The regions have been defined as follows: the “thermal desorption region” is found interior to the snowline where inherited,
initially icy grains lose their icy mantles within the viscous
timescale. The “snow border” is the region where the reset and
inheritance cases are not able to converge towards a common
snowline within 106 yr. Grains here are able to retain their icy
mantles but no significant additional ice adsorption occurs. The
“ice formation region” is where water begins to adsorb to grains
after the CPD has been chemically reset. The “photodesorption
region” is the optically thin region exterior to the snowline where
inherited grains eventually lose their icy mantles within ∼106 yr
at most.
We focus on grains of size <10 mm as the adsorption and
desorption timescales derived in Sect. 3.1 have only been derived
with the thermochemical disk model for grains up to this size.
In most cases the grain drift timescale tdrift is longer than the
A95, page 10 of 19
Dust traps. It is clear from Fig. 7 that in a decreting
CPD dust traps are present. Grains deposited near the centrifugal radius drift outwards with the gas until the force of radial
advection is balanced by the loss of orbital energy from drag
against gas orbiting at sub-Keplerian velocity. Trapped grains
thus become radially segregated by size, with smaller grains
drifting to the outer edge of the trap and larger grains remaining near the inner edge. In the high-mass high-viscosity CPD
(7–10) grains 0.1–1 mm in size become trapped near 0.1–0.2 RH .
Grains smaller than 0.05 mm advect with the gas and are able
to reach the outermost stable orbit at RH /3 where they would be
eventually lost to e.g. tidal stripping.
In general, the dust traps are spatially coincident with the ice
formation region and in the lower-mass CPDs also partially with
the photodesorption region. Hence continued ice deposition on
trapped grains could facilitate grain growth. This phenomena is
discussed further in Appendix F.
4. Discussion
We set out to explore the process of ice formation in CPDs
with relatively short viscous timescales to constrain their physical, chemical, and dynamical properties. We find that even if
infalling gas and ice is fully atomized, re-freezing proceeds
quickly in the CPD and solids reach an fice ∼ 0.5 by tvisc for
an appropriate midplane dust-to-gas ratio. The midplane ice
abundance at tvisc is generally insensitive to the initial chemical conditions. Only in the inner disk (r < 0.05–0.1 RH ) of the
high-mass CPDs is ice formation too slow for the reset and
inheritance cases to converge. With our standard chemical network the efficiency of water production in this region is closely
tied to the availability of atomic H and thus to the H2 formation rate, which is not well constrained in these conditions. The
three-body reaction H + O + M → OH + M is also critical to the
process. For this reaction we have adopted the rate coefficients
listed in UMIST2006 (Woodall et al. 2007), the value of which
originates from a recommendation of Tsang & Hampson (1986)
who noted that literature values represented only rough estimates
and suggested a factor 10 uncertainty (Baulch 1972; Day et al.
1973). More recent estimates of this rate at high temperatures
(>3000 K) suggest this value is accurate to within ∼40% (Javoy
et al. 2003), but modern low temperature measurements remain
desirable. In our expanded grain-surface chemical network, gasphase O2 H plays an important role in accelerating OH formation
in the inner disk, diverting atomic oxygen into water rather
than O2 .
4.1. Constraints on CPD properties
For reasonable assumptions regarding the properties of the CPD
the snowline can fall very near the present-day position of
Europa. The CPD with mass 10−7 M⊙ (10−4 MJ ), α = 10−3.6 and
d/g = 10−3.3 matches best the compositional gradient of the
Galilean moons at their present-day orbits. While this seems
like a promising outcome, we emphasize that both inwards gasdriven migration (Canup & Ward 2002; Madeira et al. 2021)
and long-term outwards tidally-driven migration (Yoder 1979;
N. Oberg et al.: Circumplanetary disk ices. I.
(a) High-Mass High-Viscosity (7-10)
(b) High-Mass Low-Viscosity (7-11)
(c) Low-Mass High-Viscosity (8-11)
(d) Low-Mass Low-Viscosity (8-12)
Fig. 8. Inwards drift timescale for dust grains (colored lines) relative to the timescales of desorption and ice formation (black dashed lines) in
the reference CPDs with d/g = 10−3.3 . From left to right the colored regions correspond to areas where all ices are eventually thermally desorbed
(pink), the iceline (light yellow), where grains become icier (light blue) and where all ices are eventually photodesorbed (dark gray). The red line
corresponds to the minimum grain size which is not trapped in the CPD (if there is a sign change in the gas radial velocity) and the blue line to
the maximum for which the timescales of desorption and ice formation have been verified. The vertical dotted line indicates the approximate outer
region of gravitational stability in the CPD.
Tittemore 1990; Showman & Malhotra 1997; Lari et al. 2020)
have potentially played a role in repositioning the satellites
both during and post-formation. Other regions of the CPD
parameter space are equally capable of producing solids with
0.4 < fice < 0.55, but vary in the position of their snowline. In
any case, we believe that it is more meaningful to determine
whether a particular CPD can form enough ice on the relevant viscous timescale, rather than at precisely which radii a
particular abundance of ice can be found.
To produce solids with a minimum fice ∼ 0.5 the global dustto-gas ratio of the CPD must be ≤10−3 and does not need to be
<10−3.7 . We suggest that this does not represent a minimum d/g
limit as imperfect accretion (Dwyer et al. 2013), (hydrodynamic)
proto-atmospheric escape (Kuramoto & Matsui 1994; Bierson
& Nimmo 2020), impact-vaporization (Nimmo & Korycansky
2012), or tidal heating (Canup & Ward 2009) may all have played
a role in reducing the volatile mass of the satellites either during
or post-accretion. A minimum d/g is instead implied by the minimum time required to accrete the solid total mass of the Galilean
satellites (∼10−7 M⊙ ) into the CPD. We consider the lifetime of
the gaseous circumstellar disk (∼10 Myr) to be the maximum
time over which the CPD can continue to accrete gas. Assuming
Jupiter’s runaway accretion ended <3.94 Myr after CAI formation (Weiss & Bottke 2021) and that moon formation only began
at this stage, this leaves ∼6 Myr to accrete the refractory material
for the moons. We emphasize that this limit is very approximate,
as it is possible that circumstellar gas disk lifetimes regularly
exceed 10 Myr (Pfalzner et al. 2014). Conversely an even shorter
lifetime (<4.89 Myr after CAI formation) has been proposed for
the solar nebula on the basis of paleomagnetic measurements
(Weiss & Bottke 2021).
The midplane dust-to-gas ratio and thus the fice in the CPD
will differ from what has been derived from our models if the
grain size distribution is significantly altered in some way. The
circumstellar disk gap edge may filter out larger grains from
the accretion flow (Rice et al. 2006; Zhu et al. 2012). Grains
larger than 100 µm, which settle efficiently, are those primarily responsible for enhancement of the dust mass at the CPD
A95, page 11 of 19
A&A 667, A95 (2022)
midplane. Massive planets may however vertically stir circumstellar disk midplane dust outside the gap (Binkert et al. 2021).
Szulágyi et al. (2022) found that significant vertical stirring of
large grains occurs at the gap outer wall in the presence of planets with mass above that of Saturn, resulting in a substantial
delivery of mm-sized grains to the CPD. The high-mass tail
of the dust distribution could alternatively be depleted by the
more rapid inwards drift of these grains in the CPD. We have
shown that grains larger than ∼1 mm no longer advect with the
gas in the CPD. The steady-state dust grain size distribution will
thus likely be truncated. In the absence of these grains the limits we have derived on the CPD mass are revised upwards by a
factor ∼5.
4.2. Does grain drift erase the radial distribution of ices?
We have tested only the “gas-starved” disk paradigm in which
a CPD must over time accrete the solids to form large moons.
The sequential formation, episodic growth, and potential loss
of migrating moons is a characteristic of this theory (Canup &
Ward 2002). In such a dynamical environment the relevancy of
the instantaneous radial distribution of icy grains remains to be
established. The simplest way in which the chemical properties
of dust in the CPD could be imprinted on the final satellite system would be through in-situ formation: the satellites accrete the
bulk of their material at fixed radial positions in the CPD (Ronnet
et al. 2017). This might occur if the innermost proto-moon were
prevented from migrating into Jupiter by the presence of a gasfree magnetically-truncated inner cavity (Takata & Stevenson
1996; Batygin 2018; Shibaike et al. 2019). Additional protosatellites could then pile-up in a resonant chain and be stabilized
against further migration by the proto-moon anchored at the disk
inner edge (Ogihara & Ida 2009; Sasaki et al. 2010; Izidoro et al.
2017; Madeira et al. 2021). Drifting grains would still be free to
cross the orbit of proto-moons as accretion efficiency remains
low when the proto-moons are only a fraction of their final mass
(Shibaike et al. 2019). In this paradigm proto-moons at relatively
fixed positions, continually accreting grains that drift into their
feeding-zone (Ronnet & Johansen 2020). We find that the ice
fraction of small (<1 cm) drifting grains in the inner disk will
almost always reflect ambient conditions (Sect. 3.3) independently of whether the gas in the CPD flows radially inwards or
outwards (the fate of trapped grains is discussed in Appendix F).
If the proto-moons (resonantly anchored at fixed radii within the
CPD) accrete the majority of their total mass from these drifting grains, their bulk ice fraction would reflect the temperature
gradient of the CPD.
5. Conclusions
Circumplanetary disks represent a unique chemical environment
characterized by high-densities and a relatively short timescale
on which gas and dust are viscously or aerodynamically lost. We
aimed to explore the process of ice formation in this environment
from sharply contrasting initial chemical conditions, knowing
that solids with ice/rock ∼1 must be able to form within the
viscous timescale of a Jovian CPD. We tested the paradigm in
which solids are delivered directly from the circumstellar disk in
the form of small grains (<1 cm) to a “gas-starved”, relatively
low-mass CPD. We highlight our key conclusions as follows:
If infalling material is chemically reset:
1. High densities in the CPD facilitate three-body “collider”
reactions that lead to rapid water ice formation. Roughly
half of the water ice is formed within a single year by the
hydrogenation of OH.
A95, page 12 of 19
2. Solids with the ice fraction of Ganymede or Callisto are
produced within tvisc for α ≈ 10−3 −10−4 if the midplane is
depleted in dust by a factor 20–50 relative to the canonical
d/g = 10−2 .
If chemical inheritance occurs:
1. Ices near the planet efficiently sublimate and establish a
snowline at a similar location to that of the reset case within
tvisc . Additional ice formation is minimal.
In either case:
1. Icy circumstellar dust grains preserve the majority of their
volatile content during gap-crossing if accreted onto the
CPD within 100 yr unless the stellar luminosity is >10 L⊙ .
2. The compositional imprint of the CPD temperature profile is
not erased by radial dust drift for grains of size a < 1 cm.
3. Only the “high-mass” CPDs (MCPD = 10−4 MJ ) are sensitive to the initial chemical conditions: water ice formation
in the inner disk is less efficient if a chemical reset occurs as
oxygen tends to remain locked in the gas-phase CO.
In our solar system icy moons are common. No matter whether
or not ices sublimate upon incorporation into the CPD, we have
demonstrated that ices can be efficiently re-deposited onto dust
grains and enable the general ubiquity of icy moons.
Acknowledgements. The research of N.O. and I.K. is supported by grants from
the Netherlands Organization for Scientific Research (NWO, grant number
614.001.552) and the Netherlands Research School for Astronomy (NOVA).
“This research has made use of NASA’s Astrophysics Data System Bibliographic
Services. This research has made extensive use of Numpy (Harris et al. 2020),
Matplotlib (Hunter 2007), scipy (Virtanen et al. 2020), and prodimopy https:
//gitlab.astro.rug.nl/prodimo/prodimopy. The authors would like to
thank the anonymous referee for comments that contributed to the accuracy,
clarity, and focus of this work.
References
Adams, F. C. 2010, ARA&A, 48, 47
Alibert, Y., Mousis, O., & Benz, W. 2005, A&A, 439, 1205
Ansdell, M., Williams, J. P., Trapman, L., et al. 2018, ApJ, 859, 21
Aota, T., Inoue, T., & Aikawa, Y. 2015, ApJ, 799, 141
Aoyama, Y., Ikoma, M., & Tanigawa, T. 2018, ApJ, 866, 84
Armitage, P. J. 2007, ArXiv e-prints [arXiv:astro-ph/0701485]
Armitage, P. J. 2010, Astrophysics of Planet Formation (Cambridge University
Press)
Atkinson, R., Baulch, D. L., Cox, R. A., et al. 2004, Atmos. Chem. Phys., 4, 1461
Auer, L. 1984, in Numerical Radiative Transfer, ed. W. Kalkhofen (Cambridge:
Cambridge University Press), 101
Avramenko, L. I., & Krasnen’kov, V. M. 1966, Bull. Acad. Sci. USSR, Div.
Chem. Sci., 15, 394
Ayliffe, B. A., & Bate, M. R. 2009, MNRAS, 397, 657
Balbus, S. A. 2003, ARA&A, 41, 555
Balbus, S. A., & Hawley, J. F. 1991, ApJ, 376, 214
Batygin, K. 2018, AJ, 155, 178
Batygin, K., & Morbidelli, A. 2020, ApJ, 894, 143
Baulch, D. L. 1972, Evaluated kinetic data for high temperature reactions
(Butterworths)
Bierson, C. J., & Nimmo, F. 2020, ApJ, 897, L43
Bierson, C. J., & Steinbrügge, G. 2021, Planet. Sci. J., 2, 89
Binkert, F., Szulágyi, J., & Birnstiel, T. 2021, MNRAS, 506, 5969
Birnstiel, T., Klahr, H., & Ercolano, B. 2012, A&A, 539, A148
Boschman, L., Reitsma, G., Cazaux, S., et al. 2012, ApJ, 761, L33
Bosman, A. D., Cridland, A. J., & Miguel, Y. 2019, A&A, 632, A11
Brown, W. A., & Bolina, A. S. 2006, MNRAS, 374, 1006
Canup, R. M., & Ward, W. R. 2002, AJ, 124, 3404
Canup, R. M., & Ward, W. R. 2009, Origin of Europa and the Galilean Satellites,
eds. R. T. Pappalardo, W. B. McKinnon, & K. K. Khurana, 59
Cazaux, S., & Tielens, A. G. G. M. 2002, ApJ, 575, L29
Christiaens, V., Cantalloube, F., Casassus, S., et al. 2019, ApJ, 877, L33
Collings, M. P., Anderson, M. A., Chen, R., et al. 2004, MNRAS, 354, 1133
D’Alessio, P., Cantö, J., Calvet, N., & Lizano, S. 1998, ApJ, 500, 411
D’Angelo, G., Henning, T., & Kley, W. 2002, A&A, 385, 647
Day, M., Thompson, K., & Dixon-Lewis, G. 1973, Symp. (Int.) Combust., 14, 47
de Pater, I., & Lissauer, J. J. 2010, Planet. Sci. (Cambridge, UK: Cambridge
University Press)
N. Oberg et al.: Circumplanetary disk ices. I.
Draine, B. T., & Bertoldi, F. 1996, ApJ, 468, 269
Drażkowska,
˛
J., & Szulágyi, J. 2018, ApJ, 866, 142
Dubrulle, B., Morfill, G., & Sterzik, M. 1995, Icarus, 114, 237
Dwyer, C. A., Nimmo, F., Ogihara, M., & Ida, S. 2013, Icarus, 225, 390
Fujii, Y. I., Okuzumi, S., Tanigawa, T., & Inutsuka, S.-i. 2014, ApJ, 785, 101
Fulle, D., Hamann, H. F., Hippler, H., & Troe, J. 1996, J. Chem. Phys., 105,
983
Garrod, R. T., & Herbst, E. 2006, A&A, 457, 927
Gressel, O., Nelson, R. P., Turner, N. J., & Ziegler, U. 2013, ApJ, 779, 59
Haffert, S. Y., Bohn, A. J., de Boer, J., et al. 2019, Nat. Astron., 3, 749
Harris, C. R., Millman, K. J., van der Walt, S. J., et al. 2020, Nature, 585,
357
Hasegawa, T. I., & Herbst, E. 1993, MNRAS, 261, 83
Hasegawa, T. I., Herbst, E., & Leung, C. M. 1992, ApJS, 82, 167
Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1995, ApJ, 440, 742
Hayashi, C. 1981, Progr. Theor. Phys. Suppl., 70, 35
Heinzeller, D., Nomura, H., Walsh, C., & Millar, T. J. 2011, ApJ, 731, 115
Helling, C., Woitke, P., Rimmer, P. B., et al. 2014, Life, 4, 142
Hidaka, Y., Takahashi, S., Kawano, H., Suga, M., & Gardiner, W. C. 1982,
J. Phys. Chem., 86, 1429
Hubickyj, O., Bodenheimer, P., & Lissauer, J. J. 2005, Icarus, 179, 415
Hunter, J. D. 2007, Comput. Sci. Eng., 9, 90
Izidoro, A., Ogihara, M., Raymond, S. N., et al. 2017, MNRAS, 470, 1750
Javoy, S., Naudet, V., Abid, S., & Paillard, C. 2003, Exp. Thermal Fluid Sci., 27,
371
Kamp, I., Tilling, I., Woitke, P., Thi, W. F., & Hogerheijde, M. 2010, A&A, 510,
A18
Kamp, I., Thi, W. F., Woitke, P., et al. 2017, A&A, 607, A41
Kanagawa, K. D., Muto, T., Tanaka, H., et al. 2016, PASJ, 68, 43
Keppler, M., Benisty, M., Müller, A., et al. 2018, A&A, 617, A44
Kley, W. 1999, MNRAS, 303, 696
Kruijer, T. S., Burkhardt, C., Budde, G., & Kleine, T. 2017, Proc. Natl. Acad. Sci.
U.S.A., 114, 6712
Kuramoto, K., & Matsui, T. 1994, J. Geophys. Res.: Planets, 99, 21183
Kuskov, O. L., & Kronrod, V. A. 2005, Icarus, 177, 550
Lari, G., Saillenfest, M., & Fenucci, M. 2020, A&A, 639, A40
Liu, Y., & Sander, S. 2015, J. Phys. Chem. A, 119, 1
Lodders, K. 2003, ApJ, 591, 1220
Lubow, S. H., Seibert, M., & Artymowicz, P. 1999, ApJ, 526, 1001
Lunine, J. I., & Stevenson, D. J. 1982, Icarus, 52, 14
Machida, M. N., Kokubo, E., Inutsuka, S.-i., & Matsumoto, T. 2008, ApJ, 685,
1220
Madeira, G., Izidoro, A., & Giuliatti Winter, S. M. 2021, MNRAS, 504, 1854
Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer, P., & Lissauer, J. J.
2007, ApJ, 655, 541
Martin, R. G., & Lubow, S. H. 2011, MNRAS, 413, 1447
McElroy, D., Walsh, C., Markwick, A. J., et al. 2013, A&A, 550, A36
McKinnon, W. B. 1997, Icarus, 130, 540
Min, M., Dullemond, C. P., Kama, M., & Dominik, C. 2011, Icarus, 212, 416
Mitchell, T. R., & Stewart, G. R. 2011, AJ, 142, 168
Morbidelli, A., Szulágyi, J., Crida, A., et al. 2014, Icarus, 232, 266
Mousis, O., & Alibert, Y. 2006, A&A, 448, 771
Nimmo, F., & Korycansky, D. 2012, Icarus, 219, 508
Oba, Y., Watanabe, N., Kouchi, A., Hama, T., & Pirronello, V. 2010, ApJ, 712,
L174
Öberg, K. I., & Wordsworth, R. 2019, AJ, 158, 194
Öberg, K. I., Garrod, R. T., van Dishoeck, E. F., & Linnartz, H. 2009, A&A, 504,
891
Oberg, N., Kamp, I., Cazaux, S., & Rab, C. 2020, A&A, 638, A135
Ogihara, M., & Ida, S. 2009, ApJ, 699, 824
Pfalzner, S. 2013, A&A, 549, A82
Pfalzner, S., Steinhausen, M., & Menten, K. 2014, ApJ, 793, L34
Pinilla, P., Birnstiel, T., Benisty, M., et al. 2013, A&A, 554, A95
Pinte, C., Harries, T. J., Min, M., et al. 2009, A&A, 498, 967
Portegies Zwart, S. 2019, A&A, 622, A69
Portilla-Revelo, B., Kamp, I., Rab, C., et al. 2022, A&A, 658, A89
Pringle, J. E. 1981, ARA&A, 19, 137
Quillen, A. C., & Trilling, D. E. 1998, ApJ, 508, 707
Rab, C., Kamp, I., Ginski, C., et al. 2019, A&A, 624, A16
Rafikov, R. R. 2017, ApJ, 837, 163
Rice, W. K. M., Armitage, P. J., Wood, K., & Lodato, G. 2006, MNRAS, 373,
1619
Rivier, G., Crida, A., Morbidelli, A., & Brouet, Y. 2012, A&A, 548, A116
Röllig, M., Abel, N. P., Bell, T., et al. 2007, A&A, 467, 187
Ronnet, T., & Johansen, A. 2020, A&A, 633, A93
Ronnet, T., Mousis, O., & Vernazza, P. 2017, ApJ, 845, 92
Ruaud, M., Wakelam, V., & Hersant, F. 2016, MNRAS, 459, 3756
Ruffle, D. P., Hartquist, T. W., Caselli, P., & Williams, D. A. 1999, MNRAS, 306,
691
Sakai, N., Saruwatari, O., Sakai, T., Takano, S., & Yamamoto, S. 2010, A&A,
512, A31
Sasaki, T., Stewart, G. R., & Ida, S. 2010, ApJ, 714, 1052
Schubert, G., Anderson, J. D., Spohn, T., & McKinnon, W. B. 2004,
Interior composition, structure and dynamics of the Galilean satellites, eds.
F. Bagenal, T. E. Dowling, & W. B. McKinnon, 281
Schulik, M., Johansen, A., Bitsch, B., Lega, E., & Lambrechts, M. 2020, A&A,
642, A187
Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337
Shibaike, Y., Okuzumi, S., Sasaki, T., & Ida, S. 2017, ApJ, 846, 81
Shibaike, Y., Ormel, C. W., Ida, S., Okuzumi, S., & Sasaki, T. 2019, ApJ, 885,
79
Showman, A. P., & Malhotra, R. 1997, Icarus, 127, 93
Siess, L., Dufour, E., & Forestini, M. 2000, A&A, 358, 593
Smith, I. W. M., Herbst, E., & Chang, Q. 2004, MNRAS, 350, 323
Sohl, F., Spohn, T., Breuer, D., & Nagel, K. 2002, Icarus, 157, 104
Spiegel, D. S., & Burrows, A. 2012, ApJ, 745, 174
Suutarinen, A., Geppert, W. D., Harju, J., et al. 2011, A&A, 531, A121
Szulágyi, J. 2017, ApJ, 842, 103
Szulágyi, J., & Mordasini, C. 2017, MNRAS, 465, L64
Szulágyi, J., Morbidelli, A., Crida, A., & Masset, F. 2014, ApJ, 782, 65
Szulágyi, J., Masset, F., Lega, E., et al. 2016, MNRAS, 460, 2853
Szulágyi, J., Binkert, F., & Surville, C. 2022, ApJ, 924, 1
Takata, T., & Stevenson, D. J. 1996, Icarus, 123, 404
Tanigawa, T., Ohtsuki, K., & Machida, M. N. 2012, ApJ, 747, 47
Teague, R., Bae, J., & Bergin, E. A. 2019, Nature, 574, 378
Thanathibodee, T., Calvet, N., Bae, J., Muzerolle, J., & Hernández, R. F. 2019,
ApJ, 885, 94
Thi, W. F., Woitke, P., & Kamp, I. 2011, MNRAS, 412, 711
Thi, W. F., Hocuk, S., Kamp, I., et al. 2020a, A&A, 634, A42
Thi, W. F., Hocuk, S., Kamp, I., et al. 2020b, A&A, 635, A16
Tielens, A. 2021, Molecular Astrophysics (Cambridge University Press)
Tittemore, W. C. 1990, Science, 250, 263
Tsang, W., & Hampson, R. F. 1986, J. Phys. Chem. Ref. Data, 15, 1087
Turner, N. J., Choukroun, M., Castillo-Rogez, J., & Bryden, G. 2012, ApJ, 748,
92
Villenave, M., Stapelfeldt, K. R., Duchêne, G., et al. 2022, ApJ, 930, 11
Virtanen, P., Gommers, R., Oliphant, T. E., et al. 2020, Nat. Methods, 17, 261
Wagner, K., Follete, K. B., Close, L. M., et al. 2018, ApJ, 863, L8
Wakelam, V., Bron, E., Cazaux, S., et al. 2017, Mol. Astrophys., 9, 1
Walsh, C., Harada, N., Herbst, E., & Millar, T. J. 2009, ApJ, 700, 752
Weidenschilling, S. J. 1977, MNRAS, 180, 57
Weiss, B. P., & Bottke, W. F. 2021, AGU Adv., 2, e00376
Woitke, P., Dominik, C., & Sedlmayr, E. 1993, A&A, 274, 451
Woitke, P., Kamp, I., & Thi, W.-F. 2009, A&A, 501, 383
Woitke, P., Min, M., Pinte, C., et al. 2016, A&A, 586, A103
Woodall, J., Agúndez, M., Markwick-Kemper, A. J., & Millar, T. J. 2007, A&A,
466, 1197
Yoder, C. F. 1979, Nature, 279, 767
Zangi, R., & Mark, A. E. 2003, Phys. Rev. Lett., 91, 025502
Zhou, Y., Bowler, B. P., Wagner, K. R., et al. 2021, AJ, 161, 244
Zhu, Z. 2015, ApJ, 799, 16
Zhu, Z., Nelson, R. P., Dong, R., Espaillat, C., & Hartmann, L. 2012, ApJ, 755,
6
Zhu, Z., Andrews, S. M., & Isella, A. 2018, MNRAS, 479, 1850
A95, page 13 of 19
A&A 667, A95 (2022)
Appendix A: The RT diffusion solver
P RO D I M O solves the radiative transfer equation (see Eq. (12)
in Woitke et al. 2009) together with the condition of radiative
energy conservation, which in general can be written as
divF = Γ ,
(A.1)
R
where F = Fν dν [erg/cm2 /s] is the bolometric radiation flux
vector and Γ [erg/cm3 /s] is the non-radiative heating rate per
volume. In the viscous case, we use Γ = Γvis , see Eq. (10) with
stellar mass M⋆ and stellar radius R⋆ instead of Mp and Rp for
circumstellar discs. The additional non-radiative heating leads to
a surplus emission of photon energy according to
Z
4π κνabs Bν (T ) − Jν dν = Γ ,
(A.2)
where κνabs is the dust absorption coefficient at frequency ν, Bν (T )
the Planck function, and Jν the mean intensity. For passive discs,
we have Γ = 0, in which case Eq. (A.2) simplifies to the ordinary
condition of radiative equilibrium.
The numerical solution method for the radiative transfer (RT)
in P RO D I M O involves iterations where formal solutions with
isotropic scattering are computed along multiple rays to cover the
full 4π solid angle as seen from every point in the disc, see Sect.
4 in Woitke et al. (2009). A formal solution results in new Jν (r, z)
which are used to update the dust temperatures T dust (r, z) and
source functions. The convergence of this Λ-iteration is accelerated by the Ng-method (see Auer 1984). However, despite this
acceleration, the convergence is still slow in the midplane, which
is a serious problem for all radiative transfer codes for discs,
including the Monte-Carlo codes, see Pinte et al. (2009).
Here we describe a method how to avoid this problem. In the
diffusion approximation, the bolometric radiation flux
F(r, z) = −
4π
grad J(r, z)
3κR (r, z)
(A.3)
is given by the gradient of the bolometric mean intensity J(r, z).
The Rosseland-mean and Planck-mean opacities are defined as
Z
. Z dBν (T )
1
1
dBν (T )
=
dν
dν
(A.4)
κR (r, z)
dT
κext (r, z) dT
Z ν
Z
.
κP (r, z) =
κνabs (r, z)Bν (T ) dν
Bν (T ) dν ,
(A.5)
where κνext (r, z) is the extinction coefficient and T = T dust (r, z) the
dust temperature at position (r, z) in the disk.
At the beginning of a new RT iteration, the Rosseland
and Planck opacities are calculated based on the frequencydependent disk opacity structure and the current T dust (r, z). Next,
we
R compute radial and vertical Rosseland optical depths τRoss =
κR ds from every point. When the radial inward, radial outward
and vertical upward Rosseland optical depths from that point are
all larger than a threshold value (we use a value of 10 here), the
point is flagged as being optically thick, and added to the subset
of optically thick points
M = { (i, j) | (ri , zi, j ) is optically thick} .
(A.6)
where i and j are the 2D-indices of a grid point at radius ri and
height zi, j . The following method only updates the mean intensities Ji, j and dust temperatures T i, j on the optically thick grid
points (i, j) ∈ M, whereas all other points are regarded as fixed
boundary conditions for this problem. To pick up the bolometric
A95, page 14 of 19
Zi+1/2,j+1/2
Zi,j+1/2
Zi-1/2,j+1/2
Zi+1/2,j-1/2
Zi,j
Zi,j-1/2
Zi-1/2,j-1/2
ri-1
ri-1/2
ri
ri+1/2
ri+1
Fig. A.1: Volume and areas for the spatial cell around point (i, j).
The mean intensities J and Rosseland opacities κR are available
on the grid points (i, j) marked with red dots.
mean intensities on the boundary points, we integrate Eq. (A.2)
assuming that Jν is close to a Planckian, hence
J(r, z) = B(T ) −
Γ(r, z)
.
4π κP (r, z)
(A.7)
where B(T ) = σ T dust (r, z)4 /π is the frequency-integrated Planck
function. Integration of Eq. (A.1) over the volume associated
with grid point (i, j) as sketched in Fig. A.1 results in the
following numerical equation
Ji, j − Ji−1, j
Ji, j − Ji+1, j
+ Aver
(A.8)
1/2, j Di+ 1/2, j
i+
∆ri−1/2
∆ri+1/2
Ji, j − Ji, j−1
Ji, j − Ji, j+1
+ Ahor
+ Ahor
= Vi, j Γi, j ,
i Di, j−1/2
i Di, j+1/2
∆zi, j−1/2
∆zi, j+1/2
Aver
i−1/2, j Di−1/2, j
where we note that the vertical fluxes through the cell boundaries
involve a scalar product with the slanted normal vector of the surface area, hence Ahor
i is the cell’s horizontal area after projection
onto the vertical direction. The following abbreviations are used
for the distances, vertical and horizontal areas, and the volume.
They are given by the geometry of the P RO D I M O grid points,
which are aligned on radial rays on which z/r is constant
√
(A.9)
ri−1/2 = ri ri−1
√
1
ri+ /2 = ri ri+1
(A.10)
∆ri−1/2 = ri − ri−1
(A.11)
∆ri+1/2 = ri+1 − ri
(A.12)
1
zi, j−1/2 = (zi, j + zi, j−1 )
(A.13)
2
1
(A.14)
zi, j+1/2 = (zi, j + zi, j+1 )
2
ri−1/2
zi−1/2, j−1/2 = zi, j−1/2
(A.15)
ri
ri−1/2
zi−1/2, j+1/2 = zi, j+1/2
(A.16)
ri
ri+1/2
zi+1/2, j−1/2 = zi, j−1/2
(A.17)
ri
ri+1/2
zi+1/2, j+1/2 = zi, j+1/2
(A.18)
ri
Aver
(A.19)
i−1/2, j = 2π ri−1/2 (zi−1/2, j+1/2 − zi−1/2, j−1/2 )
Aver
i+1/2, j = 2π ri+1/2 (zi+1/2, j+1/2 − zi+1/2, j−1/2 )
Ahor
i
=
Vi, j =
2
2
π(ri+
1/2 − ri− 1/2 )
Ahor
i (zi, j+1/2 − zi, j−1/2 )
(A.20)
(A.21)
(A.22)
N. Oberg et al.: Circumplanetary disk ices. I.
The radiative diffusion coefficients are defined as
4π
Di, j =
3κR (ri , zi, j )
p
Di−1/2, j = Di, j Di−1, j
p
Di+1/2, j = Di, j Di+1, j
p
Di, j−1/2 = Di, j Di, j−1
p
Di, j+1/2 = Di, j Di, j+1
(A.23)
(A.24)
(A.25)
(A.26)
(A.27)
Equation (A.8) states a system of linear equations for the
unknown bolometric mean intensities Ji, j on the optically thick
points (i, j) ∈ M of the form
A·X = B,
(A.28)
where all quantities in Eqs. (A.9) to (A.27) are constants forming the matrix A, and the volumes Vi, j and heating rates Γi, j
are constants forming the rest vector B. The unknowns {Ji, j } at
the optically thick points (i, j) ∈ M constitute the solution vector X. All terms in Eq. (A.8) that involve the other Ji, j on the
adjacent points are also included into B. The matrix equation
to solve (Eq. A.28) has a typical dimension of a few hundred
to a few thousand, depending on disk mass, geometry, and dust
parameters.
This way we can solve the 2D radiative diffusion problem
for the unknown mean intensities in the optically thick region as
a linear boundary value problem in one go, where there is one
layer of points surrounding the optically thick regions which sets
the boundary values. Our method calculates how the disk transports the photon energy through the optically thick core inside of
the boundary layer. It is applicable to both cases, passive discs
without viscous heating and active discs with Γ > 0.
Once the {Ji, j } on (i, j) ∈ M have been determined, we revert
the process described by Eq (A.7)
Γ(r, z)
4π κP (r, z)
π
1/4
T dust (r, z) =
B(T )
σ
B(T ) = J(r, z) +
Fig. A.2: Benchmark test for a viscous circumstellar disk with
a disk mass of 0.01 M⊙ and a mass accertion rate of Ṁacc =
10−8 M⊙ /yr, see text. The green lines are temperature cuts at
selected radii calculated by MCMax, the small wiggles are due
to the Monte-Carlo noise. The black dots show the temperatures
calculated by P RO D I M O.
Figure A.2 reveals a number of interesting features in the disk
temperature structure:
– The dust temperature at the inner rim is not much affected
by viscous heating.
(A.29)
(A.30)
and multiply the frequency-dependent mean intensities Jν (r, z),
as they were determined prior to the application of the diffusion solver, by a constant factor to make Eq. (A.2) valid again,
thereby keeping the previously calculated frequency distribution
of Jν (r, z).
After having modified T dust (r, z) and Jν (r, z) this way on all
grid points (i, j) ∈ M, the normal RT solution method resumes,
which begins by calculating the source functions on all grid
points and continues by performing a formal solution.
Figure A.2 shows a benchmark test against the Monte Carlo
radiative transfer program MCMax (Min et al. 2011). We consider the disk model that is described in detail by Woitke et al.
(2016), see their table 3. The central star is a 2 Myr old T Tauri
star with a mass of 0.7 M⊙ and a luminosity of 1 L⊙ , the disk has
a mass of 0.01 M⊙ , with a gas/dust ratio of 100. The dust is composed of 60% silicate, 15% amorphous carbon and 25% porosity.
The dust grains have sizes between 0.05 µm and 3 mm, with an
unsettled powerlaw size-distribution of index -3.5. The dust is
settled according to the prescription of Dubrulle et al. (1995)
with αsettle = 0.01. In contrast to this standard passive T Tauri
model, we use here a mass accretion rate of Ṁacc = 10−8 M⊙ /yr
to set the viscous heating of the disk according to Eq. (10). We
use 140 radial ×100 vertical grid points, and 40 wavelength bins.
Figure A.2 shows good agreement.
Fig. A.3: Midplane temperature T dust (r, 0) as function of the
log-distance from the inner rim. Again, the green line is the
temperature profile calculated by MCMax, and the small black
diamonds show the temperature values calculated by P RO D I M O.
A95, page 15 of 19
A&A 667, A95 (2022)
– From top to midplane, the temperature first decreases in the
disk shadow, but then the trend is reversed and the temperature re-increases towards the midplane as the viscous heat
pumped into the disk needs to flow outward, that is mostly
upward, which according to the diffusion approximation
requires a negative temperature gradient.
– There is little effect of viscous heating on T dust outside of the
optically thick region which extends outward to about 10 au
and upward to about z/r ≈ 0.1 − 0.15 in this model.
– The temperature profile across the midplane beyond the
tapering radius (Rtap = 100 au, see Woitke et al. 2016) shows
a deep minimum around the midplane z = 0. This is because
of the extreme dust settling that occurs in these diluted
outer disk regions, creating more optical thickness along the
midplane, which brings down the midplane temperature to
only 6 K in this model.
Figure A.3 compares the calculated midplane temperatures
between P RO D I M O and MCMax, which reveals dust temperatures as high as 2800 K, which is of course questionable because
at such temperatures, the dust grains are expected to sublimate.
100
Appendix B: Adsorption energies
Table B.1: Adsorption energies of the most prevalent molecular
ices found in our model CPDs.
Ice
H2 O
NH3
NH2
C2 H6
C2 H4
C2 H2
CO2
CH3 OH
OH
Eads [K]
4800
5534
3956
2300
3487
2587
2990
5534
2850
ref.
a
b
b
c
b
d
e
a
b
(a) (Brown & Bolina 2006)
(b) (Garrod & Herbst 2006)
(c) (Öberg et al. 2009)
(d) (Collings et al. 2004)
(e) (McElroy et al. 2013)
The adsorption energies of our most common ices and their
respective references are listed in Table B.1.
T-change
10-1
Appendix C: Surface density slope
10-2
10-3
10-4
10-5
0
50
100
RT iterations
150
Our reference CPDs have a surface density powerlaw exponent
ǫ = 1. The steady-state solution for a constant- Ṁ decretion αdisk is ǫ ≈ 1.25 (Batygin & Morbidelli 2020). The midplane ice
mass fraction for a variety of possible values of ǫ is shown in Fig.
C.1. For a higher ǫ the NH3 iceline responsible for the “bump"
in the fice profile at 0.07 RH moves outwards only negligibly.
In the inner disk however the ice mass fraction increases due
to a combination of the lower midplane dust-to-gas and more
efficient H2 O formation.
Fig. A.4: Maximum relative temperature change between iterations as function of RT iteration number.
Figure A.4 shows the convergence of the RT method, achieving residual relative temperature changes smaller than 10−4
after about 150 RT iterations. The maximums occuring each 5th
iteration are due to the Ng-acceleration algorithm.
Fig. C.1: Midplane ice mass fraction for a variety of surface density powerlaw exponents ǫ for the (7-11) reset CPD with global
d/g = 10−3.3 .
A95, page 16 of 19
N. Oberg et al.: Circumplanetary disk ices. I.
Appendix D: Background temperature
Throughout this work we have assumed that the CPD is embedded in a radiation field in which the equilibrium dust temperature
is 50 K. The midplane dust temperature at the gap center within
the circumstellar disk is 50 ± 2 K, for a solar luminosity 0.83
L⊙ , gap AV = 0.008, and heliocentric distance 5.2 au. For earlier
formation times with correspondingly higher solar luminosities
(2.34-13.6 L⊙ ), we find gap midplane dust temperatures ranging
from 70-120 K at 5.2 au.
We have assumed that the final stage of Jupiter’s accretion
and moon formation occured at a radial distance from the sun of
5.2 au. Volatile enrichment in Jupiter’s atmosphere indicates it
may have formed further out at circumstellar disk temperatures <
25 K or at radii > 30 au (Öberg & Wordsworth 2019). The Nitrogen abundance of Jupiter, approximately 4× solar, may suggest
additional N2 was accreted from the solar nebula near the N2 iceline (Bosman et al. 2019). In light of this possibility we consider
also lower background temperatures down to 20 K. The midplane
fice for the reference (7-11) CPD can be seen in Fig.D.1. Inside
the optically thick region of the CPD the influence of the background temperature T back is marginal for temperatures ≤ 70 K.
Above 70 K the more volatile NH3 and CO2 are unstable as
ices and only water ice remains. Below 40 K the outer disk is
able to retain ices at radii where AV < 1 as the photodesorption
timescale is in excess of the viscous timescale.
Fig. D.1: Midplane ice mass fraction for a variety of background temperatures T back for the (7-11) reset CPD and global
d/g = 10−3.3 . The four empty circles with error bars represent
the Galilean satellites at their present day locations and composition. The gray striped region on the left represents the inner
cavity and the light gray shaded region on the right represents
the gravitationally unstable zone.
Appendix E: Vertical mixing
Fig. E.1: Evolution of the water ice abundance relative to the
total number of hydrogen nuclei at the midplane (blue) and at
5 scale heights above the CPD midplane (red), and in the case
of the gas parcel which drifts from z = 5H downwards to the
midplane over a period of 100 yr (grey). The period in which
the altitude of the gas parcel was iteratively lowered towards the
midplane is highlighted in light green.
We have made the simplifying assumption that material which
accretes onto the CPD is instantaneously distributed vertically
throughout the disk. The shock front may be found at a few (∼ 5)
scale heights above the CPD midplane (Tanigawa et al. 2012). At
5 scale heights above the centrifugal radius the dust temperature
T dust = 123 K (relative to 89.5 K at the midplane), and optical
extinction AV =0.004 (16.4 at the midplane). The velocity of vertical mixing by turbulent diffusion can be estimated as vz = αcs
where cs is the local speed of sound (Heinzeller et al. 2011). We
find vz ∼ 0.5 − 1 m s−1 in this region for the high-viscosity CPDs,
assuming that the magnitude of the turbulence is constant from
the midplane up to z = 5H. The resulting vertical diffusion mixing timescale is 0.01 tvisc (10-100 yr). We perform a test in the
(7-11) CPD in which a parcel of gas is accreted at z = 5H and
iteratively evolve its chemistry in steps as it diffuses towards the
midplane over 10 yr to understand the impact of more tenuous
and high temperature conditions in the initial stages of ice formation (see Fig. E.1). By the end of the stage of rapid formation
of water (1-2 yr), the ice abundance has equalized to the fiducial
case at the midplane.
A95, page 17 of 19
A&A 667, A95 (2022)
Appendix F: Continued ice deposition on trapped
grains
In each of the reference CPDs grains of a certain size range
remain trapped within the CPD if we assume that gas actively
decretes via the outer edge of the disk. A trapped grain will
increase physically in size as ice adsorbs onto its surface and thus
alter its aerodynamic properties. Batygin & Morbidelli (2020)
proposes that grain growth and sublimation could play a role in
trapping and radial cycling of grains with size 0.1-10 mm. The
size range of trapped grains is ∼ 0.01 − 1 mm in our high-mass
CPDs and ∼ 10−3 − 0.01 mm in the low-mass CPDs, representing
2.5% and 0.1% of the infalling dust mass that reaches the midplane, respectively. A modal icy grain is typically coated in no
more than 4000 monolayers of water ice. Assuming a monolayer
thickness ∼ 0.5 nm (Zangi & Mark 2003) and compact morphology, an icy mantle no more than 1 ➭m thick will form. A grain of
size 0.05 ➭m can thus increase in size by a factor 20 and have a
density corresponding more closely to water ice rather than silicate. For a 1 ➭m thick mantle the aerodynamic effect of increased
cross-section is negated by the corresponding reduction in grain
density. If the trapped grain icy mantles continue to grow mantles beyond several 1000 monolayers, the new equilibrium trap
radius drifts inwards. Realistically only a fraction (0.01-0.1%) of
the total CPD gas mass is accreted per year. Mantle growth for a
trapped grain will thus not exceed ∼ 2 nm yr−1 on average. The
time for a grain to grow an icy mantle that allows it to drift to
the trap inner edge is then ∼ 106 yr (high-mass CPD) or ∼ 105 yr
(low-mass CPD) assuming a compact grain structure. Ice deposition is thus unlikely to allow grains to escape traps. This estimate
does not take into consideration grain growth by coagulation or
fragmentation by collision.
Appendix G: Chemical abundances of the 0D
"molecular cloud" model
The input parameters of the 0D molecular cloud model can
be found in Table G.1. A comparison between the model column densities of several common species with observations of
TMC-1 can be found in Fig. G.1. While most of the common
species column density agrees relatively well with observations,
the abundance of S-bearing species hinge on the uncertainties
regarding the S elemental abundance (Ruffle et al. 1999).
Table G.1: Molecular cloud parameters
Parameter
Value
Unit
Hydrogen density 104
cm−3
Gas temperature 10.0
K
Dust temperature 10.0
K
Optical extinction 10.0
Mean grain radius 0.1
➭m
Cloud Lifetime 1.7 × 105 yr
Note: Parameter values of the molecular cloud model and integration
time are chosen according to the method of Helling et al. (2014) recommended for TMC-1 (Taurus Molecular Cloud) by McElroy et al.
(2013). Initial atomic abundances intended to represent typical diffuse
interstellar medium conditions are also adopted from (McElroy et al.
2013).
A95, page 18 of 19
Fig. G.1: Ratio of the column density of several common species
relative to H2 in the 0D molecular cloud model (circles with error
bars) to observational values for TMC-1 (squares with black
border). Error bars on model values represent a factor 10x uncertainty for illustrative purposes. TMC-1 abundances are taken
from (Suutarinen et al. 2011) (OH,CH), (Sakai et al. 2010) (C2 H),
(Smith et al. 2004; Walsh et al. 2009) otherwise.
Appendix H: CPD dust-to-gas ratio
In Fig. H.1 we explore the resulting midplane ice mass fraction
fice for possible values of the global dust-to-gas ratio from the
canonical 10−2 down to 10−4 . The maximum grain size and dust
population size distribution is kept constant. A global dust-to-gas
ratio of 10−3.3 results in maximum midplane fice values most consistent with the Galilean satellite bulk compositions and hence is
adopted as the reference value throughout this work.
N. Oberg et al.: Circumplanetary disk ices. I.
(a) High-Mass High-Viscosity (7-10) reset
(b) High-Mass High-Viscosity (7-10) inherit
(c) High-Mass Low-Viscosity (7-11) reset
(d) High-Mass Low-Viscosity (7-11) inherit
(e) Low-Mass High-Viscosity (8-11) reset
(f) Low-Mass High-Viscosity (8-11) inherit
(g) Low-Mass Low-Viscosity (8-12) reset
(h) Low-Mass Low-Viscosity (8-12) inherit
Fig. H.1: Midplane radial ice mass fraction fice of the dust-to-gas ratio parameter exploration for the chemically reset (left column)
and chemically inherited (right column) reference CPDs. The black dotted, dashed, and solid lines indicate the radial position of
Europa, Ganymede, and Callisto. The white dotted, dashed, and solid lines indicate where fice is equivalent to the estimated ice mass
fraction of Europa (0.05), Ganymede (0.48), and Callisto (0.52).
A95, page 19 of 19