You are on page 1of 31

Thermal evolution of the impact-induced cryomagma

chamber beneath Occator Crater on Ceres


1,2 3
M. A. Hesse and J. C. Castillo-Rogez

Marc A. Hesse, mhesse@jsg.utexas.edu

1
The University of Texas at Austin,

Department of Geological Sciences, 2305

Speedway Stop C1160, Austin, TX

78712-1692, United States

2
The University of Texas at Austin,

Institute for Computational Engineering

and Sciences, 201 E 24th Street, Stop 0200,

Austin, TX 78712-1229, United States

3
Jet Propulsion Laboratory, California

Institute of Technology, Pasadena, CA,

United States

This article has been accepted for publication and undergone full peer review but has not been through
the copyediting, typesetting, pagination and proofreading process, which may lead to differences be-
tween this version and the Version of Record. Please cite this article as doi: 10.1029/2018GL080327

2018
c American Geophysical Union. All Rights Reserved.
The faculae in Occator Crater on dwarf planet Ceres are an accumulation

of salts that have been interpreted as cryovolcanic products. Current age es-

timates from crater counting suggest a maximum 18 Ma difference between

the crater forming impact and the formation of Cerealia Facula, the central

and most recent region in the crater. Here we model the thermal evolution

of the potential impact-induced cryomagma chamber beneath Occator Crater

and show that it cools in less than 12 Ma. To reach cooling times of 18 Ma

requires initial melt volumes exceeding 11000 km3 . However, simulations sug-

gest that smaller initial cryomagma chambers may lead to partial melting

of the lower crust. This may allow recharge of the magma chamber by deep

brines located in the porous upper mantle of Ceres and may extend the longevity

of cryovolcanic activity.

Keypoints:

• Likely cooling times for the impact-induced cryomagma beneath Occa-

tor Crater are less than 12 Ma.

• Cooling times approaching the 18 Ma age difference between Occator

Crater and Cerealia Facula require melt volumes exceeding 11000 km3 .

• Large impact-induced cryomagma chambers may partially melt lower crust,

leading to recharge from a deep mantle brine reservoir.

2018
c American Geophysical Union. All Rights Reserved.
1. Introduction

Dwarf planet Ceres is the largest body in the asteroid belt with a mean radius of 470 km

and has partially differentiated into a rocky mantle and a volatile rich crust [Park et al.,

2016; Russell et al., 2016]. Analysis of gravity and shape data suggest that Ceres crust is

40 to 60 km thick and largely isostatically compensated [Ermakov et al., 2017; Fu et al.,

2017]. It is strong despite its low density and hence assumed to contain significant amounts

of hydrated salts and clathrate hydrates [Fu et al., 2017; Bland and Travis, 2017; Castillo-

Rogez , 2018]. The top of the silicate mantle below this crust is several orders of magnitude

weaker, which has been attributed to the presence of small amounts of pore fluid [Fu et al.,

2017]. This is consistent with thermal models suggesting temperatures of 240 and 250 K at

the base of the crust that allow the presence of eutectic brines [Castillo-Rogez and McCord ,

2010; Neveu and Desch, 2015; Castillo-Rogez et al., 2018]. The surface composition of

Ceres is nearly uniform [De Sanctis et al., 2015; Ammannito et al., 2016] and suggests

extensive global aqueous alteration [McCord and Castillo-Rogez , 2018]. However, there

are significant local deviations in surface composition that show abundance of carbonates

and likely other salts [Carrozzo et al., 2018]. These are almost entirely associated with

impact craters and landslides, which excavated the material from shallow depth. A few

carbonate and salt-rich regions have been attributed to cryovolcanism [De Sanctis et al.,

2016; Ruesch et al., 2016; Krohn et al., 2016]. Cryovolcanism has also been invoked in the

origin of tholi, such as Ahuna Mons [Sori et al., 2017].

Here we are interested in the origin of the bright salt-rich surface deposits called faculae,

which are observed widely on Ceres’ surface [Li et al., 2016; De Sanctis et al., 2016; Stein

2018
c American Geophysical Union. All Rights Reserved.
et al., 2017; Longobardo et al., 2018]. The faculae associated with large impact craters have

been preferentially related to the brine reservoir located at the base of the crust [Quick

et al., 2018; Buczkowski et al., 2018; Nathues et al., 2018]. However, the abundance of

sodium carbonates observed across the faculae [De Sanctis et al., 2016; Raponi et al., 2018]

and the significant volume of material involved in the dome are not consistent with the

temperatures expected at the base of the crust [Castillo-Rogez et al., 2018]. The source

of the faculae needs to be above 245 K for the sodium carbonate to get in solution and

greater than 255 K for a significant amount of melt to be present.

Occator is a complex crater with a diameter of 92 km located in Ceres’ northern hemi-

sphere (20 ◦ N, 239 ◦ E). The crater lies within Hanami Planum, a topographically high

region of thickened crust [Buczkowski et al., 2017; Ermakov et al., 2017]. Estimates of

the formation age of Occator crater are derived from the crater size-frequency distribu-

tion exposed on Occator’s ejecta and lunar-derived chronology and production functions

of Hiesinger et al. [2016]. Initial estimates suggested that Occator formed ∼34 Ma ago

[Nathues et al., 2015, 2017], but a more recent detailed study based on the highest reso-

lution images has reduced the age to ∼22 Ma [Neesemann et al., 2018]. This latter age

has been adopted in a synthesis paper of Occator’s evolution by Scully et al. [2018a] and

will form the basis for discussion below.

Cerealia Facula is located within an approximately 9 km diameter, 800 m deep pit in

the center of Occator Crater [Scully et al., 2018b]. They contain mostly sodium carbonate

with a small fraction of ammonium chloride [De Sanctis et al., 2016; Raponi et al., 2018],

suggesting a crustal origin [Zolotov , 2017; Castillo-Rogez , 2018]. Cerealia Facula shows

2018
c American Geophysical Union. All Rights Reserved.
a decrease in grain size and an increase in the abundance of ammonium chloride from

the outside to the center of the dome [Raponi et al., 2018], suggesting an evolution of

the source chamber over time. The few craters found on the dome suggest that Cerealia

Facula formed as recently as 4 Ma [Nathues et al., 2017] and might still be in the process

of being emplaced. Hence, Cerealia Facula was still forming at least 18 Ma after the

formation of Occator crater.

The Occator faculae are generally interpreted as cryovolcanic deposits formed by brine

fountaining [Quick et al., 2018; Ruesch et al., 2018; Scully et al., 2018b]. The source of

the brine is either thought to be impact heating of the crust [Bowling et al., 2018], a pre-

existing crustal brine pocket [Stein et al., 2017], or a deeper brine reservoir in the mantle

[Quick et al., 2018]. Given Cerealia Facula’s location in the center of Occator Crater,

impact-induced processes (partial melting and pervasive fracturing) must have played a

major role in its formation. Bowling et al. [2018] have shown that the impact that pro-

duced Occator generated enough heat to melt the crust and create a cryomagma reservoir

with a temperature in excess of 350 K. Under these conditions, sodium carbonates and

other salts present in the crust would go in solution and can be extruded onto the surface

via fractures [Buczkowski et al., 2018; Ruesch et al., 2018]. Bowling et al. [2018] suggest

cooling timescales of no more than 4 Ma, significantly shorter than the age difference

between the Cerealia Facula and the impact that formed Occator Crater. However, they

assumed a geothermal gradient in the crust that is only one fourth of gradient predicted

by combined geophysical and petrological modeling [Castillo-Rogez et al., 2018], i.e., the

2018
c American Geophysical Union. All Rights Reserved.
background temperature field in which the magma chamber was simulated is much colder

than expected for Ceres.

In an attempt to resolve this apparent discrepancy, we study the cooling timescales of

an impact-induced cryomagma chamber in the light of the recent inference that a large

fraction of Ceres’ crust is composed of clathrate hydrates and hydrated salts [Fu et al.,

2017; Castillo-Rogez , 2018]. The presence of these phases significantly reduces the thermal

conductivity of the crust. This affects both the initial thermal state of Ceres’ crust and

the cooling timescale of the impact-induced cryomagma chamber. We also investigate the

possibility that the shallow impact-induced cryomagma communicates with a deeper brine

reservoir below the base of the crust. Such a deep reservoir has been suggested as the

source of brines for Occator by Quick et al. [2018], inferred from topography analysis by

Fu et al. [2017], and that might be consistent with the remnant muddy ocean suggested by

Travis et al. [2018]. Connection to a deep overpressured reservoir may recharge the shallow

chamber and extend the lifetime of impact-induced cryovolcanism in Occator Crater.

2. Thermal evolution model for impact-induced cryomagma chamber

To study the effect of the new estimates of crustal composition on the longevity of the

potential impact-induced cryomagma chamber beneath Occator Crater, we have developed

a 2D thermal evolution model in cylindrical coordinates. Given the large uncertainties

in the parameters, we choose a simple purely conductive thermal model. This allows

us to obtain upper bounds on the possible cooling timescales and how they vary with

parameters.

2018
c American Geophysical Union. All Rights Reserved.
We assume the crust comprises five phases: ice, water/brine, silicate rock, hydrated salts

and clathrate hydrates. The volume fractions of the phases are denoted, φp , and the phases

are referred to by three letter subcripts, p ∈ [ice, wat, sil, sal, hyd]. Values for the specific

heat capacity, cp , and thermal conductivity, κ, of each phase are given in supplementary

material (Table S1 and Figure S1). The abundance of the phases depends on temperature

as shown in Figure 1a. Melting and freezing of ice and clathrate hydrates is assumed

to occur linearly over the temperature interval, Ts ≤ T ≤ Tl , which is bounded by the

solidus and liquidus temperatures Ts and Tl , respectively. Castillo-Rogez [2018] shows that

the eutectic temperature for likely compositions of Ceres’ crust is approximately 245 K,

therefore we choose Ts = 245 K and set Tl = 273 K. Based on estimates of crustal density,

all models presented here assume that the sub-solidus crust comprises 20% silicate rock,

15% hydrated salts and variable amounts of ice and clathrate hydrate, consistent with the

mean crustal density of 1300 kg/m3 derived by Ermakov et al. [2017]. Hydrated salt is

assumed to dissolve/precipitate linearly during the melting/freezing interval (Figure 1a).

To derive an evolution equation for the multi-phase medium that includes the latent

heat of the phase transformation, we use an enthalpy method [Katz et al., 2006; Jordan

and Hesse, 2015]. With the assumption outlined above, this approach leads to an effective

heat equation, similar to those commonly used in thermal modeling of asteroids [Levin,

1962; Merk et al., 2002; Neumann et al., 2012], see supplementary materials for derivation.

Hence we have the following governing equation

dT
ρc
fp (T ) − ∇ · (κ̄(T )∇T ) = Q, (1)
dt

2018
c American Geophysical Union. All Rights Reserved.
where ρc
fp is the effective heat capacity shown in Figure 1b, κ̄ is the mean thermal con-

ductivity shown in Figure 1c, and Q is radiogenic source term. In the melting/freezing

interval the release of latent heat increases ρc


fp by almost an order of magnitude. The

mean thermal conductivity of the crust decreases with increasing abundance of clathrate

hydrate, in particular at low temperatures. Given the high volatile content of Ceres’ crust

we assume that decay heating is negligible, Q = 0.

The crustal thickness beneath Occator Crater is 46 km, the basal temperature is as-

sumed to be Ts = 245 K and the temperature at the surface is 150 K. For the full range

in crustal compositions this thermal gradient results in geothermal heat flows between 2

and 5 mW/m2 (Figure S3) and a mean crustal geotherm of approximately 2 K/km. Heat

flows consistent with Castillo-Rogez et al. [2018] are less than 2.8 mW/m2 and require

significant amounts of clathrate hydrate in the crust, φhyd ≥ 0.45. Even with the low con-

ductivity crust this is more than twice the heat flow of 1.35 mW/m2 implicit in the initial

geotherm used by Bowling et al. [2018] and the 1 mW/m2 surface heat flow resulting from

the hydrothermal models of Travis et al. [2018].

Our numerical solutions impose a constant heat flow at the base of the domain, such

that the steady geotherm reaches Ts at 46 km depth. The origin at r = 0 km is a

symmetry boundary and the vertical far-field boundary at r = 60 km is assumed to be

insulating. The initial temperature field is a superposition of the steady geotherm with

a heated region patterned after impact simulations of Bowling et al. [2018], see Figure

S5a. The impact-induced cryomagma chamber comprises a fully molten core, with initial

temperature Tini ≥ Tl . The radius of this core is 4 to 5 km, approximately equal to the

2018
c American Geophysical Union. All Rights Reserved.
pit in the center of Occator Crater that contains Cerealia Facula. This central chamber

is surrounded by a partially molten halo that extends to a radius of 10 to 15 km, see

supplementary materials. This simple initial condition allows us to explore the large

uncertainty in the size of the impact-induced cryomagma chamber.

3. Simulation results

Typical simulation results for an intermediate case with an initial melt volume of ∼4000

km3 , comparable to the study by Bowling et al. [2018], are shown in Figure 2a-e. The

interior temperature remains above the solidus for 2 Ma and the chamber fully solidifies

after 3.8 Ma (Figure 2g). However, the total melt volume declines much more rapidly than

the central temperature (Figure 2h), due to the rapid cooling of the partially molten outer

halo. Similar rapid declines of melt volume have been predicted by Quick et al. [2018],

using an analytic model. Significant cryovolcanic activity associated with an impact-

induced cryomagma chamber therefore likely ceases long before it is fully solidified.

Figures 2g and 2h show that the release of latent heat by the crystallizing cryomagma

more than doubles the cooling time of the chamber from 1.6 Ma to 3.8 Ma. The effect of

low conductivity phases in the ambient crust is explored in Figures 2i and 2j by increasing

φhyd from 0 to 0.65. This lowers the maximum thermal conductivity of the crust from up

to 3 W/m/K to less than 1 W/m/K (Figure 1c) and increases the cooling time from 2.6

Ma for pure ice to 6.2 Ma for a clathrate hydrate-rich crust. Both latent heat and the

low-conductivity crust extend the longevity of the impact-induced cryomagma, but they

fall short of the 18 Ma required to explain the age difference between Occator Crater and

the Cerealia Facula.

2018
c American Geophysical Union. All Rights Reserved.
This suggests that a larger initial cryomagma chamber is required. Here we increase

the initial volume by simply scaling the initial condition (Figure 3a). This increases the

initial depth up to 30 km and triples the initial melt volume. We justify this increase in

initial melt volume with the uncertainties inherent in the calculation of impact melting

and several restrictive assumptions made in Bowling et al. [2018] that are discussed below.

Figure 3b shows the cooling time of a cryomagma chamber as function of its initial volume,

V , its initial temperature, Tini , and the volume fraction of hydrates in the crust, φhyd .

The cooling time increases with all three parameters, but is most sensitive to the initial

melt volume. The envelope of cooling times for different crustal compositions and initial

temperatures is shown in Figure 3c. For initial melt volumes that are consistent with

kinetic energy of impactors considered by Bowling et al. [2016] the cooling ages are less

than ∼12 Ma. To reach the cooling times suggested by current estimates of the duration

of cryovolcanism requires significantly large melt volumes of at least 11000 km3 .

However, large initial volumes of cryomagma can lead to heating and partial melting of

the lower crust beneath the chamber. The base of the crust in our simulations is not a

compositional boundary, but simply the depth where the temperature exceeds Ts so that

deep brine is present. This lower crustal brine would be in contact with the deep mantle

brine reservoir, inferred from topography data [Fu et al., 2017]. For a small chamber such

as that shown in Figure 2a-f the lower crust is not heated enough to perturb the base

of the crust. Figure 4a shows the evolution of the Ts contours for a larger chamber that

heats the lower crust enough to deflect the base of the crust upward.

2018
c American Geophysical Union. All Rights Reserved.
For an even larger initial melt volume, shown in Figure 4b, the contours defining the

chamber and the base of the crust merge, generating a connection between the impact-

induced melt and the deep brine. The regime diagram in Figure 4c shows which parameter

combinations allow such a connection to occur. Again, the size of the impact-induced cry-

omagma chamber is the dominant parameter and all chambers deeper than approximately

28 km connect. This is likely an overestimate because our initial condition ignores that

increase on the lower crustal temperature due to the uplift of hot material beneath the

crater [Bowling et al., 2018].

4. Discussion

The cooling times of up to 18 Ma for the potential impact-induced cryomagma chamber

beneath Occator Crater computed here are five times longer than previous estimates by

Bowling et al. [2018], which are based on a thermal evolution model developed by Davison

et al. [2012]. For the same initial melt volume, our cooling timescales are longer because

our computations include temperature dependent thermophysical properties, the presence

of low conductivity phases in the crust, which also result in a higher geothermal gradient,

the latent heat of crystallization, and a constant basal heat flux (for direct comparison see

SI and Figure S6). All of these factors increase the longevity of the cryomagma - except

for the increased geotherm, which increases the mean thermal conductivity (Figure 1c).

However, the single biggest factor in determining the cooling timescale is the initial

volume of melt generated by the impact. These volumes are highly uncertain, because

of the complex behavior of volatile-rich crusts as discussed in detail by Bowling et al.

[2018] and the lack of proper equations of state for mixtures including large amounts

2018
c American Geophysical Union. All Rights Reserved.
of hydrates. Here we consider significantly larger volumes of impact-induced melt than

previously considered (Figure 3b). We believe that the amount of melt produced by the

Occator-forming impact could be larger for two main reasons.

First, Ceres’ crust is now considered significantly more volatile-rich than assumed by

Bowling et al. [2018]. They assumed ice/rock mass ratio in the crust is less than 0.3, which

corresponds to a total volatile volume fraction, φice + φhyd , of less than approximately 0.4.

In this study we assume φice +φhyd = 0.65, based on the recent estimates crustal density of

approximately 1300 kg/m3 [Hiesinger et al., 2016; Ermakov et al., 2017; Fu et al., 2017].

Bowling et al. [2018] show that increasing the volatile content of the crust increases both

the melt volumes and the depth of the cyromagma chamber.

Second, in Castillo-Rogez et al. [2018] we argue that the crustal geothermal gradient

beneath Occator Crater is higher than the 0.5 K/km assumed by Bowling et al. [2018].

This is based on the recent recognition that the uppermost mantle contains brines [Fu

et al., 2017], which requires that temperatures exceed the eutectic at ¿245 K [Castillo-

Rogez , 2018]. Given the surface temperature of 150 K and the current estimate of the

crustal thickness of 46 km beneath Occator Crater, the geothermal gradient should be

closer to 2 K/km. This leads to a significantly warmer crust that is closer to the eutectic

temperature (Figure S4).

Our simulations show that initial melt volumes 3 times larger than those considered

by Bowling et al. [2018] are needed to span the 18 Ma age difference between Occator

Crater and the Cerealia Facula. The largest melt volume considered here is approximately

13000 km3 and requires an energy of 1.1·109 TJ, relative to the steady-state geotherm.

2018
c American Geophysical Union. All Rights Reserved.
However, these energies exceed the kinetic energy of the range of impactors considered

by Bowling et al. [2016] and are hence unrealistic (Figure S5b). If the ages assigned to

the formation of Occator Crater and the Cerealia Facula are robust, this argues against

the simple scenario of the progressive freezing of an isolated impact-induced upper-crustal

brine reservoir.

However, age estimates derived from crater counting have large uncertainties. The main

sources of uncertainty are: the choice of the the chronology model, the target material

properties, the possibility of secondary impacts, the possibility that not all craters found

in the faculae are of impact origin, and the retention of craters on the slopes. The 22

Ma formation age of Occator Crater assumed here is based on the improved lunar-derived

chronology and production functions of Hiesinger et al. [2016]. We note that an indepen-

dent approach, based on an asteroid-flux derived chronology model, gives formation ages

of Occator Crater as young as 1.6 Ma [Neesemann et al., 2018]. The uncertainty in the 4

Ma model age of Cerealia Facula is particularly large due to small surface area, the large

slopes which could affect crater retention, and the unknown material properties for salt-

rich materials. The uncertainty in the formation crater based ages should be kept in mind

when evaluating the hypothesis that Cerealia Facula is produced by an impact-induced

near surface brine reservoir.

We believe therefore that it is possible that the current estimate for the duration of

cryovolcanic activity of 18 Ma will be reduced in the future. This would potentially allow

a resolution of model ages with our significantly extended cooling ages. Such revised

2018
c American Geophysical Union. All Rights Reserved.
model ages will likely come from the higher resolution images obtained during the final

orbit of the Dawn spacecraft at an altitude as low as 35 km.

Our models also show the possibility that the crustal melt develops connection to the

deep brine reservoir in Ceres’ mantle. This is possible even at small degrees of partial

melting, because brine wets the grain boundaries [Nye, 1989; De La Chapelle et al., 1999;

McCarthy et al., 2013] and hence percolates at very small melt fractions [von Bargen and

Waff , 1986; Ghanbarzadeh et al., 2014, 2017; Rudge, 2018]. Brine percolation in the lower

crust would establish a hydraulic connection with the deep brine reservoir below Ceres’

crust. If the brine reservoir beneath the crust is overpressured, due to continued freezing

of the interior [Quick et al., 2018], a permeable lower crust would allow recharge of the

crustal magma chamber and extend the longevity of the crustal cryomagma chamber.

Partial melting of the lower crust is not the only process that can establish a connection

between the impact-induced brine in the upper crust and the deep brine reservoir in the

mantle. Pervasive impact-induced fracturing/shattering and porosity creation would have

a similar effect and are commonly observed in terrestrial impacts into rocky crust [Collins,

2014; Morgan et al., 2016; Christeson et al., 2018]. The geophysical signal of the reduction

in crustal density is a large circular gravity low [Pilkington and Grieve, 1992]. However,

the resolution of the current gravity model is too coarse to confidently detect anomalies

associated with Occator Crater [Ermakov et al., 2017], although additional information is

expected from the high-resolution gravity data acquired during the last phase of the Dawn

mission. The creation of new void space could initially drain brine produced by impact-

induced melting, before brines from the mantle reservoir rise to replenish the crustal

2018
c American Geophysical Union. All Rights Reserved.
reservoir. It is also not clear if fracturing would affect the soft ductile lower crust and

how long any impact-induced fractures remain open. Another potential mechanism is the

gravity driven downward migration of the dense brines through viscous ice as a porosity

wave [Stevenson, 1988; Scott and Stevenson, 1986; Kalousová et al., 2014; Jordan et al.,

2018]. But again this would initially drain the brine from the upper crust. In either case,

overpressured brines from the mantle will eventually reverse the downward drainage and

recharge the impact-induced upper crustal brine chamber.

The model presented here is relatively simple and ignores advective mass and energy

transport. This likely leads to an overestimate of the cooling times, because convective

circulation in the crustal chamber would increase the cooling rates. Our ages should

therefore be seen as upper bounds. We also ignore the compositional evolution of the

brine which increases in salinity as freezing proceeds [Travis et al., 2012, 2013; Quick et al.,

2018] and may increase cooling time [Buffo et al., 2018]. Another important question is

the evolution of the gas content in the cryomagma. If the gas is vented, the frozen lid

will be ice-rich and have a high thermal conductivity and lead to faster cooling. The

conductivity of the lid will be greatly reduced, if gases are trapped in a crystal mush on

top of the convecting brine and freezing leads to the preferential formation of clathrate

hydrates with lower conductivity. The Cerealia Facula lacks the diffuse features seen in

other Occator faculae that have been attributed to venting [Quick et al., 2018]. This

may indicate the entrapment of gases at depth. Finally, our simulations do not consider

the large porosities of impact breccias that can isolate the underlying cyromagma system

2018
c American Geophysical Union. All Rights Reserved.
[Davison et al., 2012]. We believe the large heat flux and cryomagmatic activity would

lead to rapid compaction of a volatile-rich crust.

Conclusion

It has recently been recognized that Ceres’ crust contains low-conductivity phases that

significantly increase the cooling times of impact-induced cryomagmas beneath Occator

Crater. Cooling ages of an isolated impact-induced cryomagma chamber are likely less

than 12 Ma. To reach the 18 Ma cooling times currently suggested for the age difference

between the Occator Crater and the Cerealia Facula the initial melt volumes need to

exceed 11000 km3 . Such large melt volumes are not consistent with the likely size of

the impactor based on the current state of knowledge of material equations of state. A

potential way to extend the faculae exposure timescale is to account for the possible

recharge of the impact-induced upper-crustal melt reservoir with mantle brines sourced

from >45 km depth. Partial melting of the lower crust beneath the impact-induced brine

reservoir is one mechanism that provides such a hydraulic connection and whose existence

is more likely than a deep fracture network, considering the high temperatures expected

in the lower crust.

Acknowledgments. M.A.H. was supported by National Science Foundation (NSF)

Grant DMS-1720349. Part of this work has been carried out at the Jet Propulsion Labo-

ratory, California Institute of Technology, under contract to NASA. M.A.H also acknowl-

edges fruitful discussions with the students of GEO 325J in spring 2018 at UT Austin, who

worked on this problem as a class project. The authors would like to thank Nathaniel

2018
c American Geophysical Union. All Rights Reserved.
Stein and Tom Davison for constructive reviews. All MATLAB scripts and inputs are

available at this GitHub repository: https://github.com/mhesse/OccatorCoolingGRL.

References

Ammannito, E., M. Desanctis, M. Ciarniello, A. Frigeri, F. Carrozzo, J. Combe,

B. Ehlmann, S. Marchi, H. McSween, A. Raponi, M. Toplis, F. Tosi, J. Castillo-Rogez,

F. Capaccioni, M. Capria, S. Fonte, M. Giardino, R. Jaumann, A. Longobardo, S. Joy,

G. Magni, T. McCord, L. McFadden, E. Palomba, C. Pieters, C. Polanskey, M. Rayman,

C. Raymond, P. Schenk, F. Zambon, and C. Russell (2016), Distribution of phyllosili-

cates on the surface of Ceres, Science, 353 (6303), doi:10.1126/science.aaf4279.

Bland, P., and B. Travis (2017), Giant convecting mud balls of the early solar system,

Science Advances, 3 (7), e1602,514, doi:10.1126/sciadv.1602514.

Bowling, T., F. Ciesla, S. Marchi, B. Johnson, T. Davidson, J. Castillo-Rogez, M. De Sanc-

tis, C. Raymond, and C. Russell (2016), Impact induced heating of Occator Crater on

Asteroid 1 Ceres, in 47th Lunar and Planetary Science Conference, 2787, pp. 2015–2016,

doi:10.1038/ngeo2474.

Bowling, T., F. Ciesla, T. Davidson, J. Scully, J. Castillo-Rogez, and S. Marchi (2018),

Post-Impact Thermal Structure and Cooling Timescales of Occator Crater on Asteroid

1 Ceres, Icarus, pp. 1–22, doi:10.1016/j.icarus.2018.08.028.

Buczkowski, D., D. Williams, J. Scully, S. Mest, D. Crown, P. Schenk, R. Jaumann,

T. Roatsch, F. Preusker, T. Platz, A. Nathues, M. Hoffmann, M. Schaefer, S. Marchi,

M. De Sanctis, C. Raymond, and C. Russell (2017), The geology of the occator quad-

rangle of dwarf planet Ceres: Floor-fractured craters and other geomorphic evidence of

2018
c American Geophysical Union. All Rights Reserved.
cryomagmatism, Icarus, 0, 1–12, doi:10.1016/j.icarus.2017.05.025.

Buczkowski, D., J. Scully, L. Quick, J. Castillo-Rogez, P. Schenk, R. Park, F. Preusker,

R. Jaumann, and C. Raymond (2018), Tectonic analysis of fracturing associated with

occator crater Debra, Icarus, 000, 1–11, doi:10.1016/j.icarus.2018.05.012.

Buffo, J., B. Schmidt, and C. Walker (2018), Cold case: Fractional crystallization in

cryomagmatic systems, in Cryovolcanism in the Solar System Workshop 2018, p. 2045.

Carrozzo, F., M. De Sanctis, A. Raponi, E. Ammannito, J. Castillo-Rogez, B. Ehlmann,

S. Marchi, N. Stein, M. Ciarniello, F. Tosi, F. Capaccioni, M. Capria, S. Fonte,

M. Formisano, A. Frigeri, M. Giardino, A. Longobardo, G. Magni, E. Palomba, F. Zam-

bon, C. Raymond, and C. Russell (2018), Nature, formation, and distribution of car-

bonates on Ceres, Science Advances, 4 (3), e1701,645, doi:10.1126/sciadv.1701645.

Castillo-Rogez, J. (2018), Insights into Ceres’ evolution from surface composition, Mete-

oritics & Planetary Science, pp. 1–46.

Castillo-Rogez, J., and T. McCord (2010), Ceres’ evolution and present state constrained

by shape data, Icarus, 205 (2), 443–459, doi:10.1016/j.icarus.2009.04.008.

Castillo-Rogez, J., R. Fu, A. Ermakov, and M. Hesse (2018), Conditions for the Long-Term

Preservation of a Brine Reservoir in Ceres, Geophysical Research Letters, submitted.

Christeson, G. L., S. P. Gulick, J. V. Morgan, C. Gebhardt, D. A. Kring, E. Le Ber,

J. Lofi, C. Nixon, M. Poelchau, A. S. Rae, M. Rebolledo-Vieyra, U. Riller, D. R.

Schmitt, A. Wittmann, T. J. Bralower, E. Chenot, P. Claeys, C. S. Cockell, M. J.

Coolen, L. Ferrière, S. Green, K. Goto, H. Jones, C. M. Lowery, C. Mellett, R. Ocampo-

Torres, L. Perez-Cruz, A. E. Pickersgill, C. Rasmussen, H. Sato, J. Smit, S. M. Tikoo,

2018
c American Geophysical Union. All Rights Reserved.
N. Tomioka, J. Urrutia-Fucugauchi, M. T. Whalen, L. Xiao, and K. E. Yamaguchi

(2018), Extraordinary rocks from the peak ring of the Chicxulub impact crater: P-wave

velocity, density, and porosity measurements from IODP/ICDP Expedition 364, Earth

and Planetary Science Letters, 495, 1–11, doi:10.1016/j.epsl.2018.05.013.

Collins, G. S. (2014), Numerical simulations of impact crater formation with dila-

tancy, Journal of Geophysical Research: Planets, 119 (12), 2600–2619, doi:10.1002/

2014JE004708.

Davison, T. M., F. J. Ciesla, and G. S. Collins (2012), Post-impact thermal evolution of

porous planetesimals, Geochimica et Cosmochimica Acta, 95, 252–269, doi:10.1016/j.

gca.2012.08.001.

De La Chapelle, S., H. Milsch, O. Castelnau, and P. Duval (1999), Compressive creep

of ice containing a liquid intergranular phase: Rate-controlling processes in the dis-

location creep regime, Geophysical Research Letters, 26 (2), 251–254, doi:10.1029/

1998GL900289.

De Sanctis, M., A. Raponi, E. Ammannito, M. Ciarniello, M. Toplis, H. McSween,

J. Castillo-Rogez, B. Ehlmann, F. Carrozzo, S. Marchi, F. Tosi, F. Zambon, F. Ca-

paccioni, M. Capria, S. Fonte, M. Formisano, A. Frigeri, M. Giardino, A. Longobardo,

G. Magni, E. Palomba, L. McFadden, C. Pieters, R. Jaumann, P. Schenk, R. Mugnuolo,

C. Raymond, and C. Russell (2016), Bright carbonate deposits as evidence of aqueous

alteration on (1) Ceres, Nature, 536 (7614), 1–4, doi:10.1038/nature18290.

De Sanctis, M. C., E. Ammannito, A. Raponi, S. Marchi, T. B. McCord, H. Y. McSween,

F. Capaccioni, M. T. Capria, F. G. Carrozzo, M. Ciarniello, A. Longobardo, F. Tosi,

2018
c American Geophysical Union. All Rights Reserved.
S. Fonte, M. Formisano, A. Frigeri, M. Giardino, G. Magni, E. Palomba, D. Turrini,

F. Zambon, J. P. Combe, W. Feldman, R. Jaumann, L. A. McFadden, C. M. Pieters,

T. Prettyman, M. Toplis, C. A. Raymond, and C. T. Russell (2015), Ammoniated

phyllosilicates with a likely outer Solar System origin on (1) Ceres, Nature, 528 (7581),

241–244, doi:10.1038/nature16172.

Ermakov, A. I., R. R. Fu, J. C. Castillo-Rogez, C. A. Raymond, R. S. Park, F. Preusker,

C. T. Russell, D. E. Smith, and M. T. Zuber (2017), Constraints on Ceres’ Internal

Structure and Evolution From Its Shape and Gravity Measured by the Dawn Spacecraft,

Journal of Geophysical Research: Planets, 122, 2267–2293, doi:10.1002/2017JE005302.

Fu, R., A. Ermakov, S. Marchi, J. Castillo-Rogez, C. Raymond, B. H. Hager, M. T. Zuber,

S. D. King, M. T. Bland, M. Cristina De Sanctis, F. Preusker, R. S. Park, and C. T.

Russell (2017), The interior structure of Ceres as revealed by surface topography, Earth

and Planetary Science Letters, 476, 153–164, doi:10.1016/j.epsl.2017.07.053.

Ghanbarzadeh, S., M. Prodanovic, and M. Hesse (2014), Percolation and grain boundary

wetting in anisotropic texturally equilibrated pore networks, Physical Review Letters,

113 (048001), 1–5.

Ghanbarzadeh, S., M. Hesse, and M. Prodanović (2017), Percolative core formation in

planetesimals enabled by hysteresis in metal connectivity, Proceedings of the National

Academy of Sciences, 114 (51), 13,406–13,411, doi:10.1073/pnas.1707580114.

Hiesinger, H., S. Marchi, N. Schmedemann, P. Schenk, J. Pasckert, A. Neesemann,

D. OBrien, T. Kneissl, A. Ermakov, R. Fu, M. Bland, A. Nathues, T. Platz,

D. Williams, R. Jaumann, J. Castillo-Rogez, O. Ruesch, B. Schmidt, R. Park,

2018
c American Geophysical Union. All Rights Reserved.
F. Preusker, D. Buczkowski, C. Russell, and C. Raymond (2016), Cratering on Ceres:

Implications for its crust and evolution, Science, 353 (6303), aaf4759–aaf4759, doi:

10.1126/science.aaf4759.

Jordan, J., and M. Hesse (2015), Reactive transport in a partially molten system with

binary solid solution, Geochemistry, Geophysics, Geosystems, 16 (12), 4153–4177, doi:

10.1002/2015GC005956.

Jordan, J., M. Hesse, and J. Rudge (2018), On mass transport in porosity waves, Earth

and Planetary Science Letters, 485, 65–78, doi:10.1016/j.epsl.2017.12.024.

Kalousová, K., O. Souček, G. Tobie, G. Choblet, and O. Čadek (2014), Ice melting and

downward transport of meltwater by two-phase flow in Europa’s ice shell, Journal of

Geophysical Research E: Planets, 119 (3), 532–549, doi:10.1002/2013JE004563.

Katz, R., M. Spiegelman, and B. Holtzman (2006), The dynamics of melt and shear

localization in partially molten aggregates., Nature, 442 (7103), 676–679, doi:10.1038/

nature05039.

Krohn, K., R. Jaumann, K. Stephan, K. A. Otto, N. Schmedemann, R. J. Wagner, K. D.

Matz, F. Tosi, F. Zambon, I. von der Gathen, F. Schulzeck, S. E. Schrı̈der, D. L.

Buczkowski, H. Hiesinger, H. Y. McSween, C. M. Pieters, F. Preusker, T. Roatsch,

C. A. Raymond, C. T. Russell, and D. A. Williams (2016), Cryogenic flow features

on Ceres: Implications for crater-related cryovolcanism, Geophysical Research Letters,

43 (23), 994–12, doi:10.1002/2016GL070370.

Levin, B. (1962), Thermal history of the moon, in The Moon, edited by Z. Kopal and

Z. Mikhailov, pp. 152–167, IAU Symposium.

2018
c American Geophysical Union. All Rights Reserved.
Li, J.-Y., V. Reddy, A. Nathues, L. Corre, M. R. M. Izawa, E. A. Cloutis, M. V. Sykes,

U. Carsenty, J. Castillo-Rogez, M. Hoffmann, R. Jaumann, K. Krohn, S. Mottola,

T. Prettyman, M. Schaefer, P. Schenk, S. Schröder, D. Williams, D. Smith, M. Zuber,

A. Konopliv, R. Park, C. Raymond, and C. Russell (2016), Surface Albedo and Spectral

Variability of Ceres, The Astrophysical Journal, 817 (2), L22, doi:10.3847/2041-8205/

817/2/L22.

Longobardo, A., E. Palomba, A. Galiano, M. C. De Sanctis, M. Ciarniello, A. Raponi,

F. Tosi, S. E. Schröder, F. G. Carrozzo, E. Ammannito, F. Zambon, K. Stephan, M. T.

Capria, E. Rognini, C. A. Raymond, and C. T. Russell (2018), Photometry of Ceres

and Occator faculae as inferred from VIR/Dawn data, Icarus, 0, 1–13, doi:10.1016/j.

icarus.2018.02.022.

McCarthy, C., J. Blackford, and C. Jeffree (2013), Low-temperature-SEM study of dihe-

dral angles in the ice-I/sulfuric acid partially molten system., Journal of Microscopy,

249 (2), 150–7, doi:10.1111/jmi.12003.

McCord, T., and J. Castillo-Rogez (2018), Ceress internal evolution: The view after Dawn,

Meteoritics & Planetary Science, pp. 1–15, doi:10.1111/maps.13135.

Merk, R., D. Breuer, and T. Spohn (2002), Numerical Modeling of 26Al-Induced Ra-

dioactive Melting of Asteroids Considering Accretion, Icarus, 159 (1), 183–191, doi:

10.1006/icar.2002.6872.

Morgan, J. V., S. P. S. Gulick, T. Bralower, E. Chenot, G. Christeson, P. Claeys, C. Cock-

ell, G. S. Collins, M. J. L. Coolen, L. Ferrière, C. Gebhardt, K. Goto, H. Jones, D. A.

Kring, E. Le Ber, J. Lofi, X. Long, C. Lowery, C. Mellett, R. Ocampo-Torres, G. R. Os-

2018
c American Geophysical Union. All Rights Reserved.
inski, L. Perez-Cruz, A. Pickersgill, M. Poelchau, A. Rae, C. Rasmussen, M. Rebolledo-

Vieyra, U. Riller, H. Sato, D. R. Schmitt, J. Smit, S. Tikoo, N. Tomioka, J. Urrutia-

Fucugauchi, M. Whalen, A. Wittmann, K. E. Yamaguchi, and W. Zylberman (2016),

The formation of peak rings in large impact craters, Science, 354 (6314), 878–882, doi:

10.1126/science.aah6561.

Nathues, A., M. Hoffmann, M. Schaefer, L. Le Corre, V. Reddy, T. Platz, E. A. Cloutis,

U. Christensen, T. Kneissl, J.-Y. Li, K. Mengel, N. Schmedemann, T. Schaefer, C. T.

Russell, D. M. Applin, D. L. Buczkowski, M. R. M. Izawa, H. U. Keller, D. P. OBrien,

C. M. Pieters, C. A. Raymond, J. Ripken, P. M. Schenk, B. E. Schmidt, H. Sierks,

M. V. Sykes, G. S. Thangjam, and J.-B. Vincent (2015), Sublimation in bright spots

on (1) Ceres, Nature, 528 (7581), 237–240, doi:10.1038/nature15754.

Nathues, A., T. Platz, G. Thangjam, M. Hoffmann, K. Mengel, E. A. Cloutis, L. L.

Corre, V. Reddy, J. Kallisch, and D. A. Crown (2017), Evolution of Occator Crater on

(1) Ceres, The Astronomical Journal, 153 (3), 112, doi:10.3847/1538-3881/153/3/112.

Nathues, A., T. Platz, G. Thangjam, M. Hoffmann, J. E. Scully, N. Stein, O. Ruesch, and

K. Mengel (2018), Occator crater in color at highest spatial resolution, Icarus, 0, 1–15,

doi:10.1016/j.icarus.2017.12.021.

Neesemann, A., S. Van Gasselt, N. Schmedemann, S. Marchi, S. Walter, F. Preusker,

G. Michael, T. Kneissl, H. Hiesinger, R. Jaumann, T. Roatsch, C. Raymond, and

C. Russell (2018), The various ages of Occator crater, Ceres: Results of a compre-

hensive synthesis approach, Icarus, (Ldm), doi:10.1016/j.icarus.2018.09.006.

2018
c American Geophysical Union. All Rights Reserved.
Neumann, W., D. Breuer, and T. Spohn (2012), Differentiation and core formation in

accreting planetesimals, Astronomy & Astrophysics, 543, A141, doi:10.1051/0004-6361/

201219157.

Neveu, M., and S. Desch (2015), Geochemistry, thermal evolution, and cryovolcanism on

Ceres with a muddy ice mantle, Geophysical Research Letters, 42 (23), 10,197–10,206,

doi:10.1002/2015GL066375.

Nye, J. (1989), The geometry of water veins and nodes, Journal of Glaciology, 35 (119),

17–22.

Park, R., A. Konopliv, B. Bills, N. Rambaux, J. Castillo-Rogez, C. Raymond, A. Vaughan,

A. Ermakov, M. Zuber, R. Fu, M. Toplis, C. Russell, A. Nathues, and F. Preusker

(2016), A partially differentiated interior for (1) Ceres deduced from its gravity field

and shape, Nature, 537 (7621), 515–517, doi:10.1038/nature18955.

Pilkington, M., and R. Grieve (1992), The geophysical signature of terretrial impact

craters, Reviews of Geophysics, 30 (92), 161–181.

Quick, L., D. Buczkowski, O. Ruesch, J. Scully, J. Castillo-Rogez, C. Raymond, P. Schenk,

H. Sizemore, and M. Sykes (2018), A Possible Brine Reservoir Beneath Occator Crater:

Thermal and Compositional Evolution and Formation of the Cerealia dome and Vinalia

Faculae, Icarus, pp. 1–62, doi:10.1016/j.icarus.2018.07.016.

Raponi, A., M. De Sanctis, F. Carrozzo, M. Ciarniello, J. Castillo-Rogez, E. Ammannito,

A. Frigeri, A. Longobardo, E. Palomba, F. Tosi, F. Zambon, C. Raymond, and C. Russell

(2018), Mineralogy of Occator crater on Ceres and insight into its evolution from the

properties of carbonates, phyllosilicates, and chlorides, Icarus, 0, 1–14, doi:10.1016/j.

2018
c American Geophysical Union. All Rights Reserved.
icarus.2018.02.001.

Rudge, J. F. (2018), Textural equilibrium melt geometries around tetrakaidecahedral

grains, Proceedings of the Royal Society A: Mathematical, Physical and Engineering

Sciences, 474 (2212), doi:10.1098/rspa.2017.0639.

Ruesch, O., T. Platz, P. Schenk, L. McFadden, J. Castillo-Rogez, L. Quick, S. Byrne,

F. Preusker, D. OBrien, N. Schmedemann, D. Williams, J.-Y. Li, M. Bland,

H. Hiesinger, T. Kneissl, A. Neesemann, M. Schaefer, J. Pasckert, B. Schmidt,

D. Buczkowski, M. Sykes, A. Nathues, T. Roatsch, M. Hoffmann, C. Raymond, and

C. Russell (2016), Cryovolcanism on Ceres, Science, 353 (6303), aaf4286–aaf4286, doi:

10.1126/science.aaf4286.

Ruesch, O., L. Quick, M. Landis, M. Sori, O. Čadek, P. Brož, K. Otto, M. Bland, S. Byrne,

J. Castillo-Rogez, H. Hiesinger, R. Jaumann, K. Krohn, L. McFadden, A. Nathues,

A. Neesemann, F. Preusker, T. Roatsch, P. Schenk, J. Scully, M. Sykes, D. Williams,

C. Raymond, and C. Russell (2018), Bright carbonate surfaces on Ceres as remnants of

salt-rich water fountains, Icarus, doi:10.1016/j.icarus.2018.01.022.

Russell, C., C. Raymond, E. Ammannito, D. Buczkowski, M. De Sanctis, H. Hiesinger,

R. Jaumann, A. Konopliv, H. McSween, A. Nathues, R. Park, C. Pieters, T. Prettyman,

T. McCord, L. McFadden, S. Mottola, M. Zuber, S. Joy, C. Polanskey, M. D. Rayman,

J. Castillo-Rogez, P. Chi, J. Combe, A. Ermakov, R. R. Fu, M. Hoffmann, Y. Jia,

S. King, D. Lawrence, J. Li, S. Marchi, F. Preusker, T. Roatsch, O. Ruesch, P. Schenk,

M. Villarreal, and N. Yamashita (2016), Dawn arrives at ceres: Exploration of a small,

volatile-rich world, Science, 353 (6303), 1008–1010, doi:10.1126/science.aaf4219.

2018
c American Geophysical Union. All Rights Reserved.
Scott, D., and D. Stevenson (1986), Magma ascent by porous flow, Journal of Geophysical

Research B, 91 (B9), 9283, doi:10.1029/JB091iB09p09283.

Scully, J., T. Bowling, C. Bu, D. Buczkowski, A. Longobardo, A. Nathues, A. Neese-

mann, E. Palomba, L. Quick, A. Raponi, O. Ruesch, P. Schenk, N. Stein, E. Thomas,

C. Russell, J. Castillo-Rogez, C. Raymond, and R. Jaumann (2018a), Synthesis of

the special issue: The formation and evolution of Ceres’ Occator crater, Icarus, doi:

10.1016/j.icarus.2018.08.029.

Scully, J., D. Buczkowski, C. Raymond, T. Bowling, D. Williams, A. Neesemann,

P. Schenk, J. Castillo-Rogez, and C. Russell (2018b), Ceres Occator crater and its

faculae explored through geologic mapping, Icarus, 0, 1–17, doi:10.1016/j.icarus.2018.

04.014.

Sori, M., S. Byrne, M. Bland, A. Bramson, A. Ermakov, C. Hamilton, K. Otto, O. Ruesch,

and C. Russell (2017), The vanishing cryovolcanoes of Ceres, Geophysical Research

Letters, 44 (3), 1243–1250, doi:10.1002/2016GL072319.

Stein, N., B. Ehlmann, E. Palomba, M. De Sanctis, A. Nathues, H. Hiesinger, E. Amman-

nito, C. Raymond, R. Jaumann, A. Longobardo, and C. Russell (2017), The formation

and evolution of bright spots on Ceres, Icarus, 0, 1–14, doi:10.1016/j.icarus.2017.10.014.

Stevenson, D. (1988), Fluid Dynamics of Core Formation, in Conference Origin of the

Earth, pp. 87–88.

Travis, B., P. Bland, W. Feldman, and M. Sykes (2018), Hydrothermal dynamics in a

CM-based model of Ceres, Meteoritics & Planetary Science, 25, 1–25, doi:10.1111/

maps.13138.

2018
c American Geophysical Union. All Rights Reserved.
Travis, B. J., J. Palguta, and G. Schubert (2012), A whole-moon thermal history model of

Europa: Impact of hydrothermal circulation and salt transport, Icarus, 218 (2), 1006–

1019, doi:10.1016/j.icarus.2012.02.008.

Travis, B. J., W. C. Feldman, and S. Maurice (2013), A mechanism for bringing ice and

brines to the near surface of Mars, Journal of Geophysical Research E: Planets, 118 (5),

877–890, doi:10.1002/jgre.20074.

von Bargen, N., and H. Waff (1986), Permeabilities, interfacial areas and curva-

tures of partially molten systems: Results of numerical computations of equilibrium

microstructures, Journal of Geophysical Research, 91 (B9), 9261–9276, doi:10.1029/

JB091iB09p09261.

Zolotov, M. Y. (2017), Aqueous origins of bright salt deposits on Ceres, Icarus, 296,

289–304, doi:10.1016/j.icarus.2017.06.018.

2018
c American Geophysical Union. All Rights Reserved.
Figure 1. Thermal model properties: a) Example of the variation of the phase volume

fractions, φp , as function of temperature. b) The effective heat capacity of the multi-phase

medium, ρc
fp = dH/dT , as function of temperature. c) The mean thermal conductivity, κ̄, of the

multi-phase medium, for different amounts of ice and hydrate. Solidus temperature, Ts = 245

K, is the eutectic temperature of relevant salt-ice mixtures [Castillo-Rogez , 2018]. The liquidus

temperature, Tl = 273 K, is the melting point of pure ice.

2018
c American Geophysical Union. All Rights Reserved.
Figure 2. Thermal evolution: Evolution of a cryomagma chamber with initial temperature of 320
K and initial depth of 20 km (volume of 3787 km3 ) in a crust containing 35% hydrates. a-e) Thermal
fields at different times around the cooling, impact-induced cryomagma. f ) Solidus contours, T = Ts ,
from the simulation shown in panels a to e. g) Evolution of the maximum temperature in the center of
the cryomagma, Tmax , with and without latent heat. The open symbols show the times corresponding
to panels a to e. h) Evolution of the melt volume in the cryomagma chamber with and without latent
heat. i) Evolution of the maximum temperature in the center of the cryomagma for increasing volume
fractions of hydrate in the crust, φhyd . j) Evolution of the melt volume in the cryomagma chamber with
increasing φhyd . The gray bands in panels g and i indicate the partially molten temperature interval.
The dots in panels g to j indicate the complete solidification of the cryomagma.

2018
c American Geophysical Union. All Rights Reserved.
Figure 3. Cooling time cryomagma: a) Outlines of the initial cryomagma chamber for

depths of 10, 15, 20, 25 and 30 km. Solid line corresponds to Tl and dashed line corresponds to

Ts . The dot indicates the location chosen to denote the depth. b) Longevity, tm , of the impact-

induced cryomagma chamber as function of initial temperature, Tini , initial melt volume, V , and

the volume fraction of low conductivity hydrates in Ceres’ crust, φhyd . c) The most optimistic

and most conservative cooling ages as function of initial melt volume. The red and blue lines

correspond to the far right and far left vertical corners of panel 3b, respectively. The area shaded

in grey corresponds to initial volumes consistent with the kinetic energy of impactors considered

by Bowling et al. [2016], see SI.

2018
c American Geophysical Union. All Rights Reserved.
Figure 4. Connection to deep reservoir: Simulations shown in panels a and b have

Tini = 320 K and a crust with φhyd = 0.5 and Ts contours outlining the cryomagma chamber are

shown at times 0.1 Ma and then at 1 to 9 Ma. a) Initial chamber of depth 27 km and volume 9,317

km3 . b) Initial chamber of depth 28 km and volume 10,391 km3 . c) Regime diagram showing

the combinations of initial temperature, Tini , chamber depth, D, and crustal composition, φhyd ,

that lead to connection with the deep reservoir.

2018
c American Geophysical Union. All Rights Reserved.

You might also like