martian post-impact hydrothermal systems incorporating freezing

17
Martian post-impact hydrothermal systems incorporating freezing Charles J. Barnhart a, * , Francis Nimmo a , Bryan J. Travis b a Department of Earth and Planetary Sciences, University of California, Santa Cruz, CA 95064, USA b Earth and Environmental Sciences Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA article info Article history: Received 11 June 2009 Revised 8 December 2009 Accepted 14 January 2010 Available online 25 January 2010 Keywords: Impact processes Mars Mars, Interior Mars, Surface Geological processes abstract We simulate the evolution of post-impact hydrothermal systems within 45 km and 90 km diameter cra- ters on Mars. We focus on the effects of freezing, which alters the permeability structure and fluid flow compared with unfrozen cases. Discharge rates, total discharge and water–rock ratios increase with per- meability. Systems with permeabilities of 10 10 m 2 or higher exhibit convection in the hydrosphere, allowing them to derive heat from greater depths. Surface discharges persist for 10 3 –10 5 years under freezing surface conditions, with higher permeabilities permitting longer lifetimes. Maximum discharge rates and total discharges range from 0.1 to 10 m 3 s 1 and 10 9 to 10 12 m 3 , respectively, for systems with permeabilities between 10 14 and 10 12 m 2 . Near-surface water–rock ratios range from <1 for low per- meability, frozen cases to 10 3 for high permeabilities and/or unfrozen cases. Propagation of the freezing front radially inwards focuses flow towards the center of the crater resulting in a diagnostic increase in water–rock ratios there. This process may explain the phyllosilicate assemblages observed at some crater central peaks. Ó 2010 Elsevier Inc. All rights reserved. 1. Introduction Atmospheric pressure and temperature conditions at Mars have apparently conspired against the presence of liquid water on the surface for much of its history (Carr, 1996). The surface, however, tells a different story. It records a diverse collection of geochemical and geomorphic signatures left behind by the presence of liquid water (Craddock and Howard, 2002; Mustard et al., 2008). Explor- ing how various climatic and geological processes permit water on the surface, and tying those processes to observable data, can con- tribute to our understanding of the hydrological evolution of Mars. One likely process involves the interaction of high velocity bolides with Mars’s volatile-rich crust. A kilometer-scale bolide traveling at several kilometers per second delivers enough energy to melt ice, heat subsurface water and drive hydrothermal circulation (Newsom, 1980). This post-impact hydrothermal (PIH) circulation can lead to surface discharge of water and chemical alteration of rock—both of which are potentially detectable. Here we show how hydrothermal systems initiated by bolide impacts can deliver liquid water to the surface even under current atmospheric conditions. Our analysis focuses on how freezing can affect PIH systems in ways that can be remotely sensed. Specifi- cally we employ a numerical model to determine geophysical quantities—discharge rate, total discharge, near-surface tempera- ture evolution, and water-to-rock ratios—that govern observable geochemical and geomorphological surface features. We model the evolution of PIH systems as multi-phase, convective flow through a saturated, heterogeneous, porous medium in which both advection and conduction contribute to temporal and spatial changes in temperature and ice content. The code we employ is MAGHNUM (Mars Global Hydrology Numerical Model) (Travis and Rosenberg, 1991; Travis et al., 2003). MAGHNUM solves the governing equations for mass and energy transport in a saturated porous medium. Our models differ from previous efforts (Newsom, 1980; Newsom et al., 1996; Rathbun and Squyres, 2002; Abramov and Kring, 2005) in that we incorporate freezing aquifers. In this section we provide background on the martian cryo- sphere, previous Mars-based groundwater studies, and terrestrial PIH systems. In Section 2, we explore theoretical considerations and describe our model setup. In Sections 3 and 4 we present and discuss results, and in Section 5 we summarize our conclusions. 1.1. Are PIH systems possible on ‘cold–dry’ Mars? PIH systems require water in either solid or liquid form. Mars’s crust likely contains significant quantities (e.g. Clifford, 1993). Epi- thermal neutron counts collected by the Neutron Spectrometer aboard 2001 Mars Odyssey leave little doubt that the near-subsur- face of Mars holds water ice, especially in high latitude regions (Boynton et al., 2002; Feldman et al., 2002). Discharge estimates for large outflow channels (e.g. Kasei Vallis), argue for 7 10 7 km 3 —or up to 500 precipitable (pr) m—of water in the 0019-1035/$ - see front matter Ó 2010 Elsevier Inc. All rights reserved. doi:10.1016/j.icarus.2010.01.013 * Corresponding author. E-mail address: [email protected] (C.J. Barnhart). Icarus 208 (2010) 101–117 Contents lists available at ScienceDirect Icarus journal homepage: www.elsevier.com/locate/icarus

Upload: doanhanh

Post on 12-Feb-2017

217 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: Martian post-impact hydrothermal systems incorporating freezing

Icarus 208 (2010) 101–117

Contents lists available at ScienceDirect

Icarus

journal homepage: www.elsevier .com/locate / icarus

Martian post-impact hydrothermal systems incorporating freezing

Charles J. Barnhart a,*, Francis Nimmo a, Bryan J. Travis b

a Department of Earth and Planetary Sciences, University of California, Santa Cruz, CA 95064, USAb Earth and Environmental Sciences Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA

a r t i c l e i n f o a b s t r a c t

Article history:Received 11 June 2009Revised 8 December 2009Accepted 14 January 2010Available online 25 January 2010

Keywords:Impact processesMarsMars, InteriorMars, SurfaceGeological processes

0019-1035/$ - see front matter � 2010 Elsevier Inc. Adoi:10.1016/j.icarus.2010.01.013

* Corresponding author.E-mail address: [email protected] (C.J. Barnhar

We simulate the evolution of post-impact hydrothermal systems within 45 km and 90 km diameter cra-ters on Mars. We focus on the effects of freezing, which alters the permeability structure and fluid flowcompared with unfrozen cases. Discharge rates, total discharge and water–rock ratios increase with per-meability. Systems with permeabilities of 10�10 m2 or higher exhibit convection in the hydrosphere,allowing them to derive heat from greater depths. Surface discharges persist for �103–105 years underfreezing surface conditions, with higher permeabilities permitting longer lifetimes. Maximum dischargerates and total discharges range from 0.1 to 10 m3 s�1 and 109 to 1012 m3, respectively, for systems withpermeabilities between 10�14 and 10�12 m2. Near-surface water–rock ratios range from <1 for low per-meability, frozen cases to �103 for high permeabilities and/or unfrozen cases. Propagation of the freezingfront radially inwards focuses flow towards the center of the crater resulting in a diagnostic increase inwater–rock ratios there. This process may explain the phyllosilicate assemblages observed at some cratercentral peaks.

� 2010 Elsevier Inc. All rights reserved.

1. Introduction

Atmospheric pressure and temperature conditions at Mars haveapparently conspired against the presence of liquid water on thesurface for much of its history (Carr, 1996). The surface, however,tells a different story. It records a diverse collection of geochemicaland geomorphic signatures left behind by the presence of liquidwater (Craddock and Howard, 2002; Mustard et al., 2008). Explor-ing how various climatic and geological processes permit water onthe surface, and tying those processes to observable data, can con-tribute to our understanding of the hydrological evolution of Mars.One likely process involves the interaction of high velocity bolideswith Mars’s volatile-rich crust. A kilometer-scale bolide travelingat several kilometers per second delivers enough energy to meltice, heat subsurface water and drive hydrothermal circulation(Newsom, 1980). This post-impact hydrothermal (PIH) circulationcan lead to surface discharge of water and chemical alteration ofrock—both of which are potentially detectable.

Here we show how hydrothermal systems initiated by bolideimpacts can deliver liquid water to the surface even under currentatmospheric conditions. Our analysis focuses on how freezing canaffect PIH systems in ways that can be remotely sensed. Specifi-cally we employ a numerical model to determine geophysicalquantities—discharge rate, total discharge, near-surface tempera-ture evolution, and water-to-rock ratios—that govern observable

ll rights reserved.

t).

geochemical and geomorphological surface features. We modelthe evolution of PIH systems as multi-phase, convective flowthrough a saturated, heterogeneous, porous medium in which bothadvection and conduction contribute to temporal and spatialchanges in temperature and ice content. The code we employ isMAGHNUM (Mars Global Hydrology Numerical Model) (Travisand Rosenberg, 1991; Travis et al., 2003). MAGHNUM solves thegoverning equations for mass and energy transport in a saturatedporous medium. Our models differ from previous efforts (Newsom,1980; Newsom et al., 1996; Rathbun and Squyres, 2002; Abramovand Kring, 2005) in that we incorporate freezing aquifers.

In this section we provide background on the martian cryo-sphere, previous Mars-based groundwater studies, and terrestrialPIH systems. In Section 2, we explore theoretical considerationsand describe our model setup. In Sections 3 and 4 we presentand discuss results, and in Section 5 we summarize ourconclusions.

1.1. Are PIH systems possible on ‘cold–dry’ Mars?

PIH systems require water in either solid or liquid form. Mars’scrust likely contains significant quantities (e.g. Clifford, 1993). Epi-thermal neutron counts collected by the Neutron Spectrometeraboard 2001 Mars Odyssey leave little doubt that the near-subsur-face of Mars holds water ice, especially in high latitude regions(Boynton et al., 2002; Feldman et al., 2002). Discharge estimatesfor large outflow channels (e.g. Kasei Vallis), argue for�7 � 107 km3—or up to 500 precipitable (pr) m—of water in the

Page 2: Martian post-impact hydrothermal systems incorporating freezing

102 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

past (Carr, 1996). Crater softening studies argue for layers of ice-rich regolith up to 150 m deep in the mid-latitudes, which yieldsa total ice volume of 107 km3 (Squyres and Carr, 1986; Parsonsand Nimmo, 2009). The volume of the upper 10 km of the martiancrust is roughly 1.4 � 109 km3. Assuming a depth-averaged poros-ity of 0.01, the crust could hold �1.4 � 107 km3 (�100 pr m) at sat-uration. These volumes are comparable to estimates for the polarcaps; 1.2–1.7 � 106 km3 and 2–3 � 106 km3 of ice reside in thenorthern and southern polar caps, respectively (in total 22–33 pr m) (Smith et al., 1999).

Presently, water is not thermodynamically stable on the surfaceof Mars. Mean annual surface temperatures range from �154 K atthe poles to �218 K at the equator (±5 K) (Clifford, 1993). Themean surface pressure is 6.1 mbar with a partial pressure of waterpH2O = 10�6–10�5 bars (Carr, 2006). At such low humidities, thefrost point, as measured by the Viking Mars Atmospheric WaterDetectors (MAWD), is a mere 200 K (Farmer and Doms, 1979).

Surface temperatures and theoretically derived backgroundgeothermal fluxes imply a region of subfreezing temperatures(i.e. a cryosphere), that is at least hundreds of meters thick andlikely averages a few kilometers (Clifford and Hillel, 1983; Clifford,1993; Mellon et al., 1997; Travis et al., 2003). Kilometer-scale bo-lide impacts deliver enough energy to locally eliminate the cryo-sphere by melting and/or vaporizing the subsurface ice (Wohletzand Sheridan, 1983; Mouginis-Mark, 1987; Newsom et al., 1996).Some fraction of the bolide’s kinetic energy becomes residual heatthat can drive hydrothermal flow (Newsom et al., 1996; Rathbunand Squyres, 2002; Abramov and Kring, 2005).

Given the evidence for volatiles within the crater-riddled crustof Mars, PIH systems are likely commonplace. Their capacity to cre-ate and preserve detectable geochemical and geomorphic signa-tures under a range of climate and geological conditions remainsan open question and is the focus of this study. Mineral assem-blages (Ehlmann et al., 2008) and geomorphic features (Mooreand Howard, 2005) associated with complex craters motivate ourdecision to model PIH systems in 45 km and 90 km diametercraters.

1.2. Previous Mars-based hydrothermal studies

Several studies, including simple analytic models and 3-Dnumerical porous-flow models, have investigated the behavior ofPIH systems on Mars. Early work focused on potential associationwith crater lakes. When in communication with a large subsurfaceaquifer, energy balance calculations show that crater lakes will re-main at least partially unfrozen for millions of years due to subli-mation removal of ice from the top and release of latent heat asice plates to the base of the ice sheet (McKay et al., 1985; McKayand Davis, 1991). Newsom et al. (1996) demonstrated that theadditional heat supplied by a post-impact melt sheet extends lakelifetimes by tens of millions of years.

Rathbun and Squyres (2002) was the first study to employnumerical models. Their work showed that PIH systems associatedwith complex craters 180 km in diameter could support a lake upto 1 km thick. Abramov and Kring (2005) modeled PIH systemsassociated with 30, 100, and 180 km diameter craters exposed to5 �C surface temperatures and showed that host rock permeabilityis the main factor affecting fluid circulation, and that these systemsprovide hospitable conditions for microbial life for �100 kyr. Thecode that these studies used, HYDROTHERM, does not simulatefreezing. Freezing has important effects on the thermohydrologicproperties of the aquifer including latent heat release and drasticchanges in permeability.

Other sources of heat can drive hydrothermal circulation. Underspecialized circumstances, magmatic intrusions may drive enoughflow to generate fluvial valleys (Gulick, 1998). Harrison and Grimm

(2002) showed that magma chamber-induced hydrothermal circu-lation and surface discharge increase linearly with permeabilityand that total discharge volumes span a range comparable to theinferred discharges of valley networks in volcanic provinces onMars.

Travis et al. (2003) explored the effect of a background geother-mal flux on putative sub-cryospheric aquifers, and found that forsufficiently high permeabilities, background geothermal heat fluxalone could drive hydrothermal convection. These buoyant convec-tive plumes are capable of thinning the martian permafrost to athickness of 1 km or less.

1.3. Terrestrial impact structures

On Earth, water saturates the upper crust, and any hyperveloc-ity impact capable of generating a crater will disrupt the localhydrosphere (Osinski and Spray, 2001). Hydrothermal systemshave been inferred at many terrestrial impact structures including:Haughton (Osinski and Spray, 2001), Vredefort (Grieve and Garvin,1984), Sudbury (Ames et al., 1998), Chicxulub (Kring, 2000), Ries(Newsom et al., 1986), Chesapeake Bay (Sanford, 2005), Manson(McCarville and Crossey, 1994), and Lockne (Sturkell et al., 1998).

Studies of terrestrial impact structures add to our generalunderstanding of their internal structure. For example, Haughtoncrater—a 23 km diameter structure located on Devon Island, Nun-avut, Canada—demonstrates how permeability controls PIH systemplumbing, cooling, and surface mineralogical expression. Pipestructures, characterized by pronounced Fe-hydroxide alterationof the carbonate country rocks, are concentrated throughout thefaulted annulus on crater wall slumps, but have not been foundin the impact breccias or in the central uplifted area (Osinski andSpray, 2001). This observation indicates that relatively high perme-abilities are concentrated in the pipes and that the brecciated floorhas relatively low permeability (Osinski and Spray, 2001). Concen-tric listric faults and interconnecting radial faults provide path-ways for hot fluid and steam.

1.4. Mars mineralogical observations

The TES, CRISM and OMEGA spectrometers orbiting Mars havedetected a variety of clay alteration minerals (Christensen and Ruff,2004; Bibring et al., 2006; Mustard et al., 2008). Clay alterationmineralogy depends on both the peak fluid temperature experi-enced by the rock and the water/rock (W/R) ratio (Griffith andShock, 1997; Schwenzer and Kring, 2008). Recent work that simu-lates equilibrium chemical reactions shows that higher W/R ratios(�1000) occurring at modest temperatures (<250 �C) will producenontronite and hematite with lower percentages of chlorite, andthat low W/R (�1–10) will produce chlorites, smectites, mica,amphibole and garnets (Schwenzer and Kring, 2008, 2009). More-over, depending on host-rock conditions, serpentine or feldsparmay also form (Schwenzer and Kring, 2009). The type and spatialdistribution of these phyllosilicates associated with PIH systemswill thus be affected by thermal and mass transport characteristicscentral to our models. We discuss the implications of our resultswith respect to these detections and geochemical models inSection 4.2.

2. Theory and model setup

Here, we discuss the thermal and hydrological properties ofMars’s crust. Next we use a simple theoretical analysis to illumi-nate PIH system behavior and to justify model assumptions. Then,we establish the model’s initial conditions and temperature field asprescribed by an analytical calculation of impact shock-heating.

Page 3: Martian post-impact hydrothermal systems incorporating freezing

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 103

Finally, we present the numerical code we employ and the param-eter space explored by this study.

2.1. Thermohydrological properties

The following section details various thermal and hydrologicalparameters important to PIH system evolution. Table 1 lists param-eters used in this study.

2.1.1. PorosityPorosity, /, is given by:

/ ¼ Vvoid

Vtotalð1Þ

where Vvoid and Vtotal are the volume of void space and total rock,respectively. The porosity structure of the martian surface is poorlyconstrained. Measurements are limited to soil samples, yieldinghigher porosities than expected at depth (Clark et al., 1976; Toonet al., 1980; Clifford and Hillel, 1983). Of course, the bulk porosityof the regolith is likely to display considerable heterogeneity. Marsimages show complex stratigraphy and planet-wide layering (Edg-ett and Malin, 2004). If the crust is composed of a megaregolith, thehydraulic properties will depend primarily upon the abundance ofbreccia and the compressional state of fractures (Hanna and Phil-lips, 2005). To complicate porosity estimates further, impact shockwaves may be affected by interactions at material boundaries and,in turn, change the porosity by crushing pores (Pierazzo et al.,2005). For the Moon, seismic properties suggest that there is an

Table 1A list of hydrologic, thermal, and material parameters that govern PIH systemsincluding parameters used in the governing equation for energy conservation, Eq.(11). The base of the table lists the four parameters varied in our numericalsimulations of PIH systems.

Hydrological, thermal and material parameters

Symbol Parameter Value Units

/0 Surface porosity 0.2 Dimensionlessqr Density of rock 2600 kg m�3

q Density of water Variable � (103) kg m�3

qice Density of ice 917 kg m�3

cr Specific heat of rock 1000 J kg�1 K�1

cice Specific heat of ice E.O.S. dependent�2050

J kg�1 K�1

r Liquid fraction Node waterfraction

Dimensionless

Qice Latent heat of fusion of ice 3.33 � 105 J kg�1

b Melt temperature range 25 K�2

Tm Melt temperature of ice 273.15 pure H2O KT Temperature Node temperature KI Internal energy of water E.O.S. dependent J kg�1

E Enthalpy of liquid water E.O.S. dependent J kg�1

~u Darcy velocity of liquid Node dependent m s�1

kT Thermal conductivity of ice–rock matrix

Rock: 2.5, ice:temp. dep.

W m�1 K�1

g Gravitational acceleration(Mars)

3.71 m s�2

a Thermal expansivity of water E.O.S. dependent k�1

l Dynamic viscosity of water E.O.S.dependent � 10�3

Pa s

Hs Hydrological scale height 2800 mqgeo Background geothermal flux 32.5 mW m�2

E.O.S. temperature limits 0–900 �CE.O.S. pressure limits 105–2 � 108 Pa

D Final crater diameter 45, 90 kmRp Impactor radius 1900, 4700 mT0 Ambient surface temperature 5, �53.15 �Ck0 Surface permeability 10�16, 10�14, 10�12,

10�10m2

exponential decline in porosity with depth with a best fit surfaceporosity, /0, of 0.2 and a decay constant, Hs, of 6.5 km (Toksöz,1979):

/ðzÞ ¼ /0e�z

Hsð Þ ð2Þ

When scaled to martian gravity, porosity becomes <0.01 at �8.5 kmdepth and the decay constant, Hs, is 2.8 km (Clifford, 1981). This isthe formalism used in several previous studies (e.g. Clifford, 1993;Clifford and Parker, 2001), and we adopt it for our work as well.

2.1.2. PermeabilityThe permeability, k, characterizes a porous medium’s resistance

to flow through it (Turcotte and Schubert, 2002, §9-2). Previousstudies (Rathbun and Squyres, 2002; Abramov and Kring, 2005)show that permeability governs fluid flow in PIH systems. Likeporosity, the bulk permeability of the martian regolith is unknownand is likely to exhibit strong spatial variability following an im-pact. We test a wide range of surface permeabilities, 10�16,10�14, 10�12 and 10�10 m2, in an attempt to explore the conse-quences of reasonable geologic materials. We assume that porosityis primarily in the form of interconnected fractures, allowing us toexpress permeability through a porosity-dependent cubic flow law(Travis et al., 2003), which we assume can be modified to accountfor ice content to yield:

kðzÞ ¼ k0ð1� rÞ3e�3zHs ð3Þ

where r is the fraction of pore or fracture space filled with ice, Hs isthe same decay constant used in the porosity function and k0 is thesurface permeability (m2)—a free parameter in this study.

The permeability of geologic materials spans many orders ofmagnitude. The permeability of breccias in known terrestrial im-pact craters depends on the geologic setting of the impact crater.For example, the Chesapeake Bay impact crater formed in Cenozoicsediments that are generally characterized by sands, silts and claysdeposited by marine transgressions and regressions (Sanford,2005). Based on these materials and expected time ranges requiredto flush out the original water in the breccia, permeability likelyranges from 10�16 to 10�14 m2, but was likely higher at the onsetof PIH system development before subsequent compaction (San-ford, 2005). Higher permeabilities, 10�12–10�10 m2, might be rea-sonable for plains made up of lava flows on Mars (Berman andHartmann, 2002; Manga, 2004).

2.1.3. Background heatAssuming conductive equilibrium, prior to an impact event the

crust has a background temperature gradient and water/ice distri-bution established by the crustal geothermal heat flux and theambient surface temperature. The basic relation for conductiveheat transport is given by:

qgeo ¼ kTdTdz

� �ð4Þ

where qgeo is the geothermal heat flux (W m�2), kT is the thermalconductivity (W m�1 K�1), and dT

dz is the temperature gradient withrespect to height (K m�1).

The thermal properties of the martian crust have been approx-imated using both frozen soil and basalt as analogs (Clifford, 1993).Thermal conductivities will cover a wide range on Mars but thecolumn-averaged thermal conductivity over the cryosphere, alength-scale comparable to the modeling domain of this study, islikely to vary between �1.0 and 3.0 W m�1 K�1 (Travis et al.,2003). Thermal conductivity increases with ice content and de-creases with increasing temperature. Here we calculate the effec-tive thermal conductivity as a volume-weighted mean of thebulk rock thermal conductivity (2.5 W m�1 K�1) and ice and/or

Page 4: Martian post-impact hydrothermal systems incorporating freezing

104 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

water content for each model node throughout the domain at a gi-ven time. We obtained expressions for the thermal conductivitiesof ice and liquid water from Grimm and McSween (1989).

Although no direct measurements of qgeo have been made forMars, estimates based on compositional (Fanale, 1976) and geo-physical arguments (Davies and Arvidson, 1981; Schubert et al.,1992; Nimmo and Tanaka, 2005) generally cluster around�30 mW m�2(±15 mW m�2). During the early to mid-NoachianPeriod, higher concentrations of radiogenic elements would havefueled a higher geothermal flux of �60 mW m�2 (Nimmo and Ta-naka, 2005). This study will assume a constant background geo-thermal flux of 32.5 mW m�2 that enters through the base of thedomain. A heat flux of 32.5 mW m�2 is equivalent to the tempera-ture gradient of 13 �C km�1 used by Abramov and Kring (2005),assuming a thermal conductivity of 2.5 W m�1 K�1. The crustalheat flux of Mars, like Earth, is likely spatially nonuniform (Cliffordand Hillel, 1983; Squyres et al., 1992). However, given the consid-erable uncertainty in the global average value, the addition of high-er order spatial variations is unwarranted.

2.2. Analytical expressions

Following an impact event, the distribution of subsurface iceand the surface discharge of liquid water depend on heat sourcesand permeability. The impact event itself is the most important ini-tial condition; the energy delivered by the impact and the redistri-bution of shock-heated material and rock at depth sets the stagefor any subsequent hydrothermal activity. Before exploring resultsfrom simulated evolution of PIH systems, it is informative to devel-op simple analytic expressions that approximate the governingphysics.

Domain Average Permeability (m2

Surface Permeability (m )2

Tim

e (y

r)

ΔT=10K

ΔT=30K

ΔT=300K

ΔT=100K

)

Fig. 1. The characteristic timescales required for a 1-D advective plume to travel1 km depend strongly on permeability: equation (7). The time required for anadvective plume to equilibrate with the background geothermal profile from aninitial DT of 10, 30, 100, or 300 K varies from �10 to 107 years for reasonable surfacepermeabilities (top x-axis). Timescales longer than �100 kyr are longer than theconductive cooling timescale and thus cool conductively. In our numericalsimulations (Section 2.1.2) permeability is assumed to decay with depth; the lowerx-axis contains the depth-averaged (10 km) permeability. In general, lowerpermeabilities yield longer lived but less vigorous PIH systems. Plumes travelingthrough highly permeable substrates advect heat rapidly and cool quickly. Weassumed the following parameter values for this calculation: surface permeability,l = 10�3 Pa s, l = 103 m, a = 1.5 � 10�4 K�1, and g = 3.71 m s�2.

2.2.1. Cooling timescalesAn order of magnitude estimate of the thermal conductive cool-

ing time of an impact structure is:

tcond �l2

jð5Þ

where l is a characteristic length scale, and j is the thermal diffusiv-ity. An l of 1 km and a j of �10�6 m2 s�1 (i.e. basalt) yields a con-ductive cooling time of �30,000 years.

The advection of fluid reduces cooling timescales. For advection,we use the Dupuit approximation (Turcotte and Schubert, 2002), toobtain a time scale for the flow as a function of permeability, k, inthe absence of heat:

t � l2lkqgh

� tcondjl

kqghð6Þ

Here, we have introduced a distance along the direction of flow, l,over which the head changes by an amount h. Thus, h is technicallya head difference, and the ratio h/l is the head gradient driving flow.Other variables are defined as follows: density, q, gravity, g, and dy-namic viscosity, l.

To first order, the shock-heated region of an impact structureoccupies a hemisphere that will be desiccated as a result of theshock. Assuming a saturated region with a permeability of10�12 m2 exists outside the zone of desiccation and thatl � h � 1 km due to the shocked zone’s spherical geometry, Eq.(6) shows that for a fluid viscosity of �10�3 Pa s, the time requiredfor fluid to inundate the shocked region is �10 years—over four or-ders of magnitude shorter than the conductive cooling time (Eq.(5)). This means that the hottest and driest portion of the PIH sys-tem experiences negligible cooling before water and steam beginto interact with it.

Permeability exerts the strongest control on cooling time. High-er permeabilities allow greater advection of heat and cool hydro-thermal systems more rapidly than low permeabilities. Assuming1-D steady state-flow driven by the buoyancy of hot water, it ispossible to derive the Darcy velocity, u, of a rising plume and hencea thermally driven advective timescale, tadv (cf. Turcotte and Schu-bert, 2002):

tadv ¼ll

hkiaqg1

DT

� �ð7Þ

where a is the thermal expansivity of water, hki is the depth-aver-aged permeability, and DT is the temperature contrast across acharacteristic length scale, l. Fig. 1 plots tadv against the surface per-meability and shows that bigger DT results in quicker cooling, be-cause velocity scales with DT and heat flow ð~q ¼~uqCPDTÞ scalesas DT2. For likely surface permeabilities, 10�14–10�10 m2, systemswill advect heat to the surface in �104–108 years. The timescalesgenerated by this analytical calculation are similar to results ob-tained by more sophisticated numerical calculations (see Section 3).

2.2.2. Geothermal fluxThe geothermal heat flux, qgeo, plays two important roles in the

behavior of a hypothetical hydrosphere residing below the cryo-sphere (Travis et al., 2003). Together with permeability, qgeo deter-mines whether or not the hydrosphere transfers heat conductivelyor by convection. The Rayleigh number (Ra) characterizes the vigorof convection:

Ra ¼qghkiaqgeok

2

ljkTð8Þ

where hki is the domain averaged permeability and k is the domainheight. For Ra values greater than 20–40 meta-stability gives way to

Page 5: Martian post-impact hydrothermal systems incorporating freezing

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 105

fluid convection in an open porous medium with a porosity that de-cays exponentially with depth (Nield and Bejan, 1992). Fig. 2 showsthat conductive or convective driven heat transfer depends primar-ily on permeability—k0 P 10�11 m2 assures background convectionin absence of impacts. The geothermal flux also determines thedepth of the cryosphere. PIH systems that have enough initial heat

Crit

ical

Ra

Valu

e~30

Domain Average Permeability (m )2

Geo

ther

mal

Flu

x (W

m -2

Surface Permeability (m 2Silt, Loess

Permeable BasaltFractured Igneous and Metamorphic Rock

Sand and Gravela

bd

efc

Terrestrial Continents

Terrestrial Oceans

Mars Present (±0.015)

Mars 4.1 Ga

g

10 3

10 210

)

)

Fig. 2. The Rayleigh number (Eq. 8), an indication of conduction or convectiondominated heat transport is contoured here as a function of permeability (x-axis,surface at top, depth-averaged at base) and background geothermal flux (y-axis). Inporous media, convection dominates (gray region) for Rayleigh numbers above 20–40 (Nield and Bejan, 1992). Estimates for Mars’s past and present geothermal fluxvaries by about a factor of two: from �30 to �60 mW m�2. Conversely, thepermeability of reasonable geological materials spans many orders of magnitude.Consequently, permeability plays the principle role in determining heat transport inthe martian hydrosphere. The ‘X’ symbols indicate parameter values—surfacepermeability, k0, and background geothermal flux, qgeo—modeled by this study.Finally, the letters a–g indicate permeabilities explored by previous analytical andnumerical hydrology studies of Mars’s subsurface as specified: (a) Harrison andGrimm (2002); (b) Abramov and Kring (2005); (c) Rathbun and Squyres (2002); (d)Newsom et al. (1996); (e) Gulick (1998); (f) Manga (2004); (g) Travis et al. (2003).

-54 -15 23 62 101

Hot Mat Fro

BackgroundGeothermal

Shoc

Tempe

Dep

th B

elow

Sur

face

(m)

radial d

Fig. 3. The initial temperature distribution used in simulations exposed to ambient scharacteristics: a radially decaying shock-heated region, a shallow region far from thecollapses from transient to final form, and a background geothermal profile of 32.5 mWcalculations with pi-scaling arguments that define material redistribution (cf. Appendix).1900 m impacting at 7 km s�1.

to penetrate through the cryosphere to an aquifer below will havemuch more water available for hydrothermal flow.

2.3. Energy delivered by bolide impact

The bolide impact delivers heat and redistributes material. Weestimate the post-impact subsurface temperature field by combin-ing an analytic calculation for the change in temperature as a func-tion of shock-emplaced heat with the effects of excavation andgeothermal uplift (Fig. 3). A short description of our method forgenerating the initial temperature fields is provided below. TheAppendix contains calculation details. Our temperature fields com-bine aspects of previous analytical calculations for shock-heating(Kieffer and Simonds, 1980; Melosh, 1989; Rathbun and Squyres,2002; Abramov and Kring, 2005) with analytical calculations forcrater formation (Melosh, 1989; Cintala and Grieve, 1998). Initialtemperature fields agree well with temperature fields generatedby hydrocode calculations (Pierazzo et al., 2005; Senft and Stewart,2008).

Initial temperature distributions are established for two finalcrater diameters (45 km and 90 km), two surface temperatures(5 �C and �53.15 �C) and a background geothermal flux of32.5 mW m�2. The 5 �C surface temperature was chosen to matchprevious PIH studies (Rathbun and Squyres, 2002; Abramov andKring, 2005); the �53.15 �C or 220 K temperature (hereafterrounded to �53 �C) was chosen to approximate present-day meanannual surface temperatures at the equator of Mars. If the surfacepermeability we specify yields a convective hydrosphere (seeFig. 2), we then obtain an initial average geothermal temperatureprofile from numerical simulations run to a dynamic equilibriumstate in the absence of impact-induced heating (cf. Appendix). Inall cases we assume an initially static aquifer. For the 45 km craterwe assume a bolide with velocity, radius (Rp) and density of7 km s�1, 1900 m, and 2600 kg/m3, respectively. The same velocityand density are assumed for the 90 km diameter crater; we simplyassume a larger bolide diameter of 4700 m. Using an analyticalexpression for the amount of waste heat derived from the Murna-ghan equation of state (Kieffer and Simonds, 1980), the amount ofshock energy initially imparted by the bolide is 66% of the kineticenergy (cf. Appendix). Excavation and vaporization removes a por-tion of the shock energy bringing final waste heat budgets intoagreement with hydrocode studies (Gisler et al., 2006).

139 178 217 255 294 332

Frozen Region (Cryosphere)

erial Upliftedm Depth

k Heating

rature (°C)

istance (m)

Transient Radius

Crater Rim

urface temperatures that are well below freezing (�53 �C) exhibits four notablecenter that remains frozen, hot material that is brought from depth as the crater

m�2. We generate the initial temperature analytically by combining shock-heatHere, we constructed a 45 km diameter crater generated by a bolide with a radius of

Page 6: Martian post-impact hydrothermal systems incorporating freezing

106 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

Fig. 3 is a plot of our initial temperature distribution for a PIHsystem occupying a 45-km crater exposed to subfreezing surfacetemperature. It includes contributions from shock-heating, thebackground geothermal temperature profile, and advected heatfrom the centrally uplifted region. The shock-heating is stronglyconcentrated at the top center of the crater and falls off rapidlywith distance. Fig. 3 shows how this steep falloff leaves most ofthe crater’s floor and its rim frozen. Hotter material brought fromdepth as the crater collapses from its transient to its final form sup-plies additional heat to the PIH system as indicated by the concavedown temperature contour lines in Figs. 3 and 11.

2.4. The numerical model: MAGHNUM

We simulate the evolution of PIH systems using the Mars GlobalHydrology Numerical Model (MAGHNUM) (cf. Travis et al., 2003).MAGHNUM iteratively provides a finite difference numerical solu-tion of the governing partial differential equations for 1-, 2-, or 3-D,time-dependent transport of non-isothermal water and heatthrough a porous, permeable medium. MAGHNUM addresses thedynamics of vapor, water and ice in extraterrestrial environments(Travis et al., 2003; Travis and Schubert, 2005; Schubert et al.,2007). In this application we do not treat vapor dynamics. Fluidflow is governed by the thermohydrologic properties of the mar-tian subsurface and modified by the melting or freezing of ice clog-ging the pores.

2.4.1. Governing equationsThe mass balance equation for the water component is:

r � ðq~uÞ ¼ 0 ð9Þ

Darcy’s law determines the fluid velocity:

~u ¼ � klðrP þ qgzÞ ð10Þ

where additional variables are pressure (P), and the height (z) abovesome reference (z0). The use of Darcy’s law assumes relatively lowvelocities so that flow is entirely laminar. During the earliest stagesof the PIH circulation, flow is likely to be turbulent in the near sur-face, in the hottest regions, where our results may slightly overesti-mate flow velocities.

Incorporating freezing complicates the energy balance equation(Travis et al., 2003), yielding the following equation—variables de-fined in Table 1:

@

@tð1� /ÞqrcrT þ /rqI þ ð1� rÞqice �

ciceT

"

þffiffiffiffibp

rQ ice

Z T

�1e�bðT�TmÞ2 dT

!#þr � ð~uqEÞ ¼ r � ðkTrTÞ ð11Þ

The left-hand side of Eq. (11) quantifies the time rate of change ofenergy for five processes: the rate of change of energy in the threematerials (rock, water, and ice); the melting/freezing process; andthe advection of enthalpy. These are balanced on the right-handside by the diffusion of heat. The integral term represents cumula-tive release and absorption of latent heat around temperature Tm.The ice fraction’s dependence on the capillary pressure is omitted;its effect will be minor compared to the temperature dependencefor many permeable materials, and it can be mimicked to some de-gree by choice of b (Travis et al., 2003). b controls how smoothly thephase change occurs and is chosen here to be 25 K�2.

The equation of state tables of Harr et al. (1984) determine thepressure and temperature dependencies of density, viscosity, heatcapacity, thermal expansivity and enthalpy for liquid water. Theequation of state limits for pressure and temperature in these ta-

bles are between 105 and 2 � 108 Pa (1–2000 bars) and 0 �C and900 �C. Although the average surface pressure on Mars, �103 Pa,is below the table’s limits, a hydrostatic pressure of 105 Pa isreached at a depth of �30 m under Mars’s gravity. This depth iswell within the first grid row. The change of state between liquidwater and water ice occurs over a sharp temperature interval�0.1–0.2 �C (Travis et al., 2003). Ice fraction within pores controlspermeability (Eq. (3)) which, in turn, controls the Darcy velocity(Eq. (10)) and affects the energy balance (Eq. (11)).

2.4.2. NumericsMAGHNUM discretizes the governing equations in space using

an integrated finite difference algorithm on a grid, and discretizesin time using the fully implicit backward Euler method. MAGH-NUM has 1-D, 2-D, or 3-D capability, on irregular (but orthogonal)grids. In this application, we use 2-D cylindrical geometry and auniform grid. The resulting non-linear system of flow equationsis solved at each time-step using the Newton–Raphson iterativemethod. At each time-step, a sequential process first solves forpressure; this allows calculation of fluid velocities. Then transportof heat updates energy and mass in each grid cell. Finally, temper-ature and ice fraction are determined using the equation of state.

2.5. Domain setup and boundary conditions

Taking advantage of the crater’s radial symmetry, we model PIHsystems in a 2-D axisymmetric domain with 100 radial nodes and50 vertical nodes. We place the center of the crater’s surface at thetop left-hand corner of the domain. The right half of Fig. 3 repre-sents the dimensions of the 2-D domain; the left half is simplyits reflection. Depending on crater diameter, 45 km or 90 km, thedomain spans 30 km or 60 km radially with respective node widthsof 333 m or 666 m. We set the domain height to the depth of as-sumed pore closure, 10 km, yielding a node thickness of 200 m.Our work differs from Rathbun and Squyres (2002) and Abramovand Kring (2005) in that we ignore topographic relief—an assump-tion we discuss further in Section 5.3.

Although energy and mass are conserved across all cell inter-faces, the domain’s boundary is open, allowing heat and fluid toexit or enter. We assume that the domain is saturated at all times;any mass exiting the domain must be balanced by mass influx else-where. The following boundary conditions control the four sides ofthe domain. Along the base, we prescribe a constant geothermalflux and allow free fluid flow (although permeability is very low).At the axis of symmetry (left-hand side) the temperature gradientis zero, temperature is variable, and no horizontal fluid flow is per-mitted. We assume the right-hand side to be far from the PIH sys-tem, fix the temperature at values assigned by the geothermalprofile, and permit fluid flow.

The top boundary condition allows fluid to exit and enter thedomain and is designed to simulate the physical effects imposedby subfreezing surface temperatures. At each time step, the bound-ary condition of each surface cell is determined by whether or notpositive fluid flow out of the cell is occurring. This binary conditioncontrols the behavior of a row of dummy cells that, conceptually,reside above the surface cells. If a particular surface cell has posi-tive upwards flow then its corresponding dummy cell is set tothe surface cell’s temperature, neither cooling nor heating that sur-face cell. Otherwise, the dummy cell is set to the ambient temper-ature value and conductively cools the surface cell. The row ofdummy cells is essential, because no surface melting would occurin their absence. Our approach attempts to recreate the physics offluid entering a frozen environment—while flow is occurring thecell is well out of equilibrium with its environment and when flowceases the cell cools conductively toward thermal equilibrium withthe atmosphere.

Page 7: Martian post-impact hydrothermal systems incorporating freezing

5°C

-53°C

perm. (m2)

equiv. crater depth (m)

disc

harg

e (m

3 )

10-12 m2

10-16 m2

10-10 m2

10-14 m2

10-12 m2

10-16 m2

10-10 m2

10-14 m2

Surfa

ce D

isch

arge

Rat

e (m

3 /s)

Time (yr)

A

B

Fig. 4. Fluvially related geomorphologies depend on discharge rates and totalsurface discharges. Here 45 km diameter crater discharge rates for four permeabil-ities are plotted as a function of time for PIH systems exposed to a surfacetemperature of 5 �C (A) or �53 �C (B). The inset plots show total discharge and theequivalent depth that the total discharge would fill a 45-km crater ignoringatmospheric loss for frozen cases (D) and clement cases (+). Convective systemswith especially high permeabilities (10�10 m2) exhibit prolonged lifetimes andsignificant surface discharge. In the 5 �C simulations the heat from the central upliftbegins to contribute to surface discharge rates at �60 kyr for the 10�10 m2 case andat �80 kyr for the 10�12 m2 case. The diamonds show that peak fluxes (kg s�1 m�2)measured by Abramov and Kring (2005) compare well with our dischargemeasurements for a simulation of similar size, surface permeability, and surfacetemperature.

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 107

2.6. Parameter space

In this study we focus on how surface temperatures and freez-ing affect geochemical and geomorphological characteristics of PIHsystems. Given this aim, we vary three parameters: crater diame-ter, ambient surface temperature and surface permeability. Craterdiameter, set to 45 or 90 km, determines the amount of heat avail-able to drive hydrothermal flow. The surface temperature affectsdischarge and W/R ratios. By setting the temperature to 5 �C wetest hypothetical warm periods of Mars’s past and compare our re-sults with the work of Abramov and Kring (2005). A surface valueof �53 �C simulates PIH behavior throughout most of Mars’s pastincluding the present epoch. The permeability of Mars’s subsurfaceis unknown and values for reasonable geological materials spanmany orders of magnitude. We prescribe a wide range of surfacepermeability values (10�16, 10�14, 10�12, and 10�10 m2) in orderto explore the effects of both conductive and convective hydro-spheres (Fig. 2). These variable parameters are listed in Table 1.

3. Results

3.1. Discharge rate and total discharge

We record the Darcy velocity throughout the domain at eachtime step and assume surface discharge to be the positive verticalcomponent of the velocity vector for each surface node. We calcu-late the surface discharge rate as the product of Darcy velocity andarea for the entire PIH system by summing the discharge rates forall surface nodes as follows:

q ¼Xnodes

i¼1

maxð0; v iÞp r2i � r2

i�1

� �ð12Þ

where ri, the radius at each node, and vi, the magnitude of the posi-tive vertical component of velocity determine q, the discharge rateat a particular time. The total discharge is an integration of the en-tire system’s instantaneous discharge rate, q, over time; we assumethat no steam or fluid is lost to the atmosphere.

Fig. 4A shows how discharge rate varies with permeability for5 �C surface temperature runs. Fig. 4A shows that discharge ratesfor a system with an especially low surface permeability,10�16 m2, are negligible. Higher surface permeabilities, 10�12 and10�10 m2, produce higher initial discharge rates of �3 and�60 m3 s�1, respectively. Although discharge still scales with per-meability, these systems continue to produce low discharge ratesof about 1 m3 s�1 for millions of years as heat from the central up-lift reaches the surface beginning around 100 kyr after the impactevent.

Systems with a permeability of 10�14 m2 produce initial dis-charge rates on order of 0.1 m3 s�1 that decay to negligible valuesover 106 years. These discharge rates are comparable to dischargerates we infer for a smaller, 30 km diameter PIH system simulatedby Abramov and Kring (2005). To make this comparison we con-verted three reported peak flux values (kg s�1 m�2) simulated byAbramov and Kring (2005) to discharge rate (m3 s�1) by assumingthat the peak flux occurred uniformly at the surface from the cen-ter of the crater to a radius of 1 km, an area of A = p(103)2 m2, andapplying the following conversion: qdischarge ¼ A qflux

q . Inferred dis-charge rates are 0.05, 0.02, and 0.002 m3 s�1 occurring at 103,104, and 105 years, respectively. Fig. 4A shows that these inferreddischarge rates (diamonds) are in reasonable agreement with ourresults.

Fig. 4B shows that for PIH systems exposed to frozen surfacetemperatures (�53 �C), surface discharge rates again scale withpermeability, but cease when the entire surface refreezes. FrozenPIH systems with low surface permeabilities, 10�16 m2, freeze in

only �1700 years and produce negligible discharge rates and totaldischarge. Greater surface permeabilities, 10�14 and 10�12 m2,yield greater initial discharge rates and increase the length of timethat the surface remains unfrozen. Systems with high permeability,10�10 m2, behave in a fundamentally different way. They harbor apre-existing vigorously convecting hydrosphere capable of miningheat from the central uplift and delivering it to the surface beforethe surface can freeze. This additional heat yields discharge ratesand total discharges that are an order of magnitude greater and al-low the system to maintain contact with the surface for >50 kyr.

A bolide capable of generating a 90 km diameter crater containsroughly 10 times as much kinetic energy as a bolide that forms a45 km diameter crater: �1022 J versus �1021 J. Fig. 5 shows thatthis additional shock-heating and a more extensive central upliftproduce surface discharge rates and total surface discharges thatare roughly 10 times as great as the 45 km case for systems ex-posed to frozen temperatures. Low permeability systems(k0 = 10�16, 10�14 m2) freeze within 3000 years and produce lowtotal discharges of 107 and 109 m3, respectively. The larger shockedregion of the 90 km crater and greater central uplift allows heatmined from the central uplift to advect to the surface before itfreezes for systems with surface permeabilities of 10�12 and10�10 m2. Fig. 5 shows that, ignoring loss to the atmosphere, the to-tal discharge from these systems would fill their 90 km craters to adepth of 1–8 km.

Page 8: Martian post-impact hydrothermal systems incorporating freezing

10-12 m2

10-16 m2

10-10 m2

10-14 m2

perm. (m2)

disc

harg

e (m

3 )

Time (yrs)

Surfa

ce D

isch

arge

Rat

e (m

3 /s)

totalsurfacedischarge

equiv. crater depth (m)

Fig. 5. Similar to the 45 km simulations, the permeability of 90 km PIH systemscontrols discharge rates and total surface discharge. Low permeability systems,10�16–10�14 m2 produce geomorphologically insignificant discharge rates and areincapable of advecting enough heat to remain unfrozen at the surface much longerthan a few 1000 years. The 10�12 m2 system exhibits much larger discharge rates,total discharge and system lifetime when compared to the 45 km system. It behaveslike the 45 km, 10�10 m2 case—the additional, more deeply deposited, heat presentin the 90 km crater allows the surface of the 10�12 m2 system to remain unfrozenlong enough for heat from the central uplift to advect to the surface. In general, the90 km systems produces greater discharge rates and total discharges whencompared to the 45 km systems but permeability plays the primary role indetermining how quickly the surface freezes.

108 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

3.2. W/R ratios and temperature evolution

The chemical alteration of rock by hydrothermal circulation de-pends on host-rock chemistry, temperature, peak pressure, expo-sure time, and the amount of water that flowed through the rock(Griffith and Shock, 1997). W/R ratios strongly influence the chem-istry and mineralogy of hydrothermal alteration products in basalt(Fisher and Narasimhan, 1991). Although Osinski (2005) showsthat host-rock chemistry may play an important role in porosity-

sryk 33 tA

Rim

w:r

- ra

tiode

pth

(m)

radius (m)

Surface : 5°C

Depth

Fig. 6. Profile plots of W/R ratios for the upper 200 m of rock (top panels) and contour psurface permeability of 10�10 m2 exposed to ambient surface temperatures of 5 �C (left) oenable convection in the hydrosphere fueled by geothermal heat alone. These two effectview, convective plumes may produce a diagnostic ‘bulls-eye’ pattern of geochemical altunder frozen conditions, the cryosphere prevents convective plumes from reaching the susystems actually produce higher W/R ratios at the center of the crater than clement systeincreasingly constricted toward the center of the crater, focusing flow there.

permeability evolution with time, we did not include chemical ef-fects in our model. We recorded the cumulative W/R ratio by timeintegrating the incremental W/R ratio measured at each nodethroughout PIH system evolution. The incremental W/R ratio is adimensionless measure of the water mass that flowed through agiven rock mass (Taylor, 1977; Fisher and Narasimhan, 1991).We calculated the incremental W/R ratio as follows:

WR¼ Dt

qH2O

qrockð1� /Þ

� �uz

Dzþ ur

Dr

h ið13Þ

where Dt, Dz, and Dr, are the time step, vertical node size and radialnode size, respectively. The radial and vertical components of theDarcy velocity vector, ~u, are defined as ur and uz. Alternatively, Eq.(13) can be written as:

WR¼ Dt

qH2O

qrockð1� /Þ

� �~u� ~w

A

� �ð14Þ

In this formulation it is clear that W/R ratio depends on the w/Aratio (Fisher and Narasimhan, 1991), where A is node area, and~w ¼ ðDr;�DzÞ. As ~w goes to zero so will A and the W/R ratio willgo to an asymptotic value. Like temperature or pressure, W/R is apoint function and we approximate its value on the domain grid.

Node area is uniform throughout the domain and for all simula-tions of a particular crater size but the node size in the radial direc-tion is twice as great in the 90 km versus 45 km cases (666 mversus 333 m). This reduces the accuracy of the contribution fromthe horizontal component of velocity in the 90 km cases. However,PIH systems and bottom heated convective domains principallyproduce vertical flow so we assume that a direct comparison ofcumulative W/R ratios between runs with different crater sizes ispermissible in this study. The ‘absolute’ W/R ratios calculated hereshould exceed geochemical estimates because (1) residence timesmay be too short to allow geochemical equilibrium, and (2) wateroften passes through previously altered rock (Fisher and Narasim-han, 1991). In any case, relative differences between W/R ratios atdifferent locations or between different runs are more important

sryk 33 tA

Rim

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000w/r ratio by mass

Surface : -53°C

lots at depth (bottom panels) as a function of radius for 45 km PIH systems with ar�53 �C (right). Surface permeabilities J 10�11 m2 yield rapid advection of fluid ands produce high W/R ratios, >100 at depth, especially in convective plumes. In planareration due to alternating annuli of high and low W/R ratios (left). For PIH systemsrface near the rim of the crater (right); discharge and W/R ratios are zero. Frozen PIHms. As the PIH system freezes, the remaining unfrozen area of the surface becomes

Page 9: Martian post-impact hydrothermal systems incorporating freezing

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 109

than the absolute values. They indicate at which locations geo-chemical alteration is more likely to have occurred.

W/R ratios at all depths increase with permeability. Figs. 6–8show profiles for the W/R ratios for the uppermost node (upper200 m of rock) and contour plots of the W/R ratios for rocks inthe subsurface to a depth of 10 km. Fig. 9 tracks the evolution oftemperature and cumulative W/R ratios for the upper 200 m ofthe centermost node of the crater. Fig. 6 shows that systems withhigher permeabilities (10�10 m2) allow convection, rendering themcapable of mining heat from the central uplift before the surface

sryk 33 tA

Rim

w:r

- ra

tiode

pth

(m)

radius (m)

Surface : 5°C

Depth

Fig. 7. Profile plots of W/R ratios for the upper 200 m of rock (top panels) and contour psurface permeability of 10�12 m2 exposed to ambient surface temperatures of 5 �C (leftplumes in the shock-heated region and above the central uplift. For these systems, ambitemperatures continue to advect heat to the surface from the shock-heated region and theof �100 in areas of the surface above convective plumes. When exposed to subfreezing aW/R ratios are zero for much of the surface. Flow never occurs at the rim or in much osystems that produce modest W/R ratios of �10 may explain spectral detection of alter

sryk 001 tA

Rim

w:r

- ra

tiode

pth

(m)

radius (m)

Depth

Surface : -53°C

Fig. 8. Profile plots of W/R ratios for the upper 200 m of rock (top panels) and contour ploan ambient surface temperatures of �53 �C with a surface permeability of 10�10 m2 (leftratios that are �10 times greater than respective 45 km systems. Systems exposed to susimilar in character but larger in magnitude than the 45 km system. W/R ratios are concethe system exposed to clement surface conditions produce an alternating pattern of higplumes fail to carve through the cryosphere and do not produce discharge at the rim.

freezes. Convective systems subjected to surface temperatures be-low freezing are particularly interesting because an impermeablefront—the surface ice/water interface that marches inwards frommid-floor to the central peak—forces flow towards the center ofthe crater. This effect prolongs modest near-surface temperatures(0–100 �C for 50 kyr) at the crater center and yields W/R ratios>1000 (Fig. 9D). Vigorous convective upwelling plumes enhanceW/R ratios far from the crater’s center in the non-frozen case butare unable to melt through the cryosphere in the frozen case(Fig. 6). The profile of W/R ratios of surface-exposed rock in Fig. 6

sryk 33 tA

Rim

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000w/r ratio by mass

Surface : -53°C

lots at depth (bottom panels) as a function of radius for 45 km PIH systems with a) or �53 �C (right). Modest permeabilities allow advection of heat and convectiveent surface temperatures increase W/R ratios. Systems exposed to clement ambientcentral uplift for�106 years producing W/R ratios of �300 at the crater’s center andmbient surface conditions, the upper 200 m of the system freezes in <10,000 years.f the floor; surface exposed W/R ratios do not exceed �10 at the central peak. PIHation minerals present on the central peak of mid-sized craters.

sryk 001 tA

Rim

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000w/r ratio by mass

Surface : -53°C

ts at depth (bottom panels) as a function of radius for 90 km PIH systems exposed to) or 10�12 m2 (right). Ninety kilometer craters harbor PIH systems that produce W/Rbfreezing ambient surface temperatures produce a surface W/R ratio profile that is

ntrated at the center because that location is the last to freeze. Convective plumes inh (>1000) and modest (�100) W/R ratios. In the subfreezing case, these convective

Page 10: Martian post-impact hydrothermal systems incorporating freezing

110 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

shows that the cryosphere prevents flow and maintains zero near-surface W/R ratios at the rim and across much of the crater floor.

For 45 km diameter craters with lower permeability, 10�12 m2,exposed to frozen surface temperatures, the upper 200 m of rockat the crater’s center experience fluid temperatures between 0and 100 �C for 20,000 years and W/R ratios of �50 (Figs. 7 and9C). All flow subsequent to surface freezing occurs in deeper partsof the domain, 2–6 km below the surface, where permeability isroughly 1% of the surface value. Fig. 7(right) shows how reducedpermeability at depth restricts flow and limits W/R ratios to values�1.

With surface temperatures above freezing, the discharge rateslowly decays as heat is advected from the system for 105–106

years (Fig. 9A). This prolonged surface contact and flow in areasof low permeability yields greater W/R ratios of �300 at the centralpeak and W/R ratios of 1 in convective plumes extending to thesurface in regions near the rim (Fig. 7, left). By the time the convec-tive plumes reach the surface far from the crater’s center, the fluidhas cooled to the surface temperature, 5 �C.

Larger, 90 km, frozen PIH systems produce W/R ratios �2–5times greater than their respective 45 km counterparts at the cen-tral peak in simulations with surface permeabilities of 10�10 and10�12 m2 (Fig. 8). In the 10�10 m2 case, convective plumes create

tem

pera

ture

(°C

)

time (y

A

C

freezing

convective plume

-53 °C, 10-12 m2

5 °C, 10-12 m2

temp

temp

W/R

W/R

Fig. 9. These four plots show the time-evolution (x-axis) of temperature (solid line, left ysurface-exposed upper 200 m of rock for four 45 km PIH system simulations. (B and C) Wthe surface remains unfrozen long enough, advective plumes originating from the centraratios (A and D). As the system cools, the surface freezes beginning with more radially dcenter of the crater which greatly increases cumulative W/R ratios there (C and D).

alternating annuli with high W/R ratios (�1000) and modest W/Rratios (�400) with discharge temperatures near the surface value.(Fig. 8, left) shows a strong W/R ratio peak at the center of the cra-ter and a second one at a radius of �13 km. The second peak is gen-erated by fluid, heated by the shocked region and the central uplift,being directed by an overlying cryosphere. The 10�12 m2 case gen-erates a surface W/R ratio profile that is similar to, but larger than,the frozen 45 km case. The most significant difference is that theinitially unfrozen surface extends radially out to 20 km versus10 km in the 45 km crater. This increases the surface area thatexperiences W/R > 0 by a factor of four to �1.25 � 109 m2. Again,as the surface freezes, discharge is forced toward the crater’s cen-ter, enhancing W/R ratios there. Similar to the 45 km case, Fig. 8shows that convective plumes present in highly permeable simula-tions (10�10 m2) fail to erode the cryosphere at radial distances be-yond about 5 km. Consequently, even the most vigorous frozen PIHsystems do not produce discharge at the crater wall or rim.

In summary, permeability, crater diameter and surface temper-ature affect surface discharge rates, surface discharge duration, W/R ratio values and W/R ratio distribution. Total surface dischargeand discharge rate increase with permeability and crater size.The background geothermal heat flux allows the hydrosphereto convect for surface permeabilities J 10�11 m2. Convection

W/R

ratio

r)

B

D

convective plume

freezing

-53 °C, 10-10 m2

-53 °C, 10-14 m2

W/R

-axis) and W/R ratios (dashed line, right y-axis) in the crater’s centermost 333 m and/R ratios cease to accumulate once temperatures fall below 0 �C (shaded region). If

l uplift reach the surface, elevate surface temperatures and increase cumulative W/Ristant regions and concluding with the crater’s center. This forces flow toward the

Page 11: Martian post-impact hydrothermal systems incorporating freezing

cum

ulat

ive

W/R

ratio

surface permeability (m2)

45 km, 5 °C

45 km, -53 °C90 km, -53 °C

Fig. 10. Cumulative W/R ratios at the central peak (centermost 333 m and upper200 m) as a function of surface permeability for a suite of crater diameters andsurface temperatures: 45 km, 5 �C (D); 90 km, �53 �C (�); and 45 km, �53 �C (+).W/R ratios are negligible for low permeabilities in all cases. For modest perme-abilities, 10�12 m2, heat mined from the central uplift reaches the surface for the90 km subfreezing case—significantly increasing W/R ratios over the subfreezing45 km case. At high permeabilities, 10�10 m2, frozen cases actually generate higherW/R ratios at the central peak than cases with clement surface temperatures.Although the central peaks in clement cases experience circulation for much longertimescales, flow is focused at the central peak in frozen cases, yielding relativelyhigh W/R ratios.

Table 2Discharge rate, periods of activity, and total discharge for martian fluvial morphol-ogies. PIH system discharge rates are capable of producing gullies but are likelyincapable of generating alluvial fans even if permeability structures favored theirformation. In all cases, a permanently frozen region prevents discharge near the rim.PIH systems associated with 45 km craters are incapable of supporting valleynetworks or larger fluvial morphologies.

Comparisons of discharge rates and durations

Feature Max. discharge Duration Total discharge

45 km (m3 s�1) (m3)PIH system 0.1–10 103–105 years 109–1012

90 kmPIH system 0.1–100 103–106 years 109–1013

Outflow 104–109[a,b] Uncertain UncertainChannels Days to years[b] 1015–1016[a,b]

Valley 102–104[c–e] 103–106 years[e] UncertainNetworks Episodically 1012–1014[e]

Alluvial fans 103–105[f]6102 years[f] 1012–1014[f]

Gullies �10–30[g] 1–100 days[g] 106[g]

Values obtained from: aRobinson and Tanaka (1990), bWilliams et al. (2000), cCarr(1987), dIrwin et al. (2005), eBarnhart et al. (2009), fKraal et al. (2008), gParsons andNimmo (2010).

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 111

significantly increases surface discharge rates and duration of PIHsystems. Frozen surface temperatures reduce surface dischargeduration and affect the distribution of surface-exposed rock W/Rratios. As PIH systems freeze, ice directs and focuses flow to unfro-zen regions. This greatly enhances surface W/R ratios in the lastplace on the surface to freeze, the central peak. Fig. 10 shows thecumulative W/R ratios at the central peak as a function of surfacepermeability for the subfreezing 45 km and 90 km cases as wellas the unfrozen, 10�12 m2, 45 km case. At low permeabilities W/Rratios are negligible at the central peak. For surface permeabilitiesof 10�12 m2 crater size becomes important. The frozen 90 km caseis capable of delivering heat from the central uplift to the surfacebefore it freezes, yielding high W/R ratios of �300. At high perme-abilities, 10�10 m2, Fig. 10 shows that frozen systems have higherW/R ratios at the central peak than unfrozen systems. This occursbecause of the focusing effect of the advancing freezing front.

4. Discussion

Here we will examine the extent to which PIH hydrothermalsystems associated with 45–90 km diameter craters on Mars arecapable of generating detectable geochemical and geomorphologi-cal features. Three principal aspects of the PIH system drive hydro-thermal behavior and determine the scale and character of thesefeatures: crater size, permeability, and surface temperature. Thecrater size determines the PIH system’s heat budget. Permeabilitycontrols convective vigor, surface discharge, and W/R ratios. Sur-face temperatures determine the presence and extent of a cryo-sphere, which focuses flow towards the center of the crater,which, in turn, modulates surface discharge and W/R ratios.

4.1. Geomorphic signatures

Several craters are associated with features suggestive of pastsurface water including gullies, deltas, meanders and lake deposits(e.g. Cabrol and Grin, 2001). Table 2 compares model PIH systemdurations and discharges with those estimated to be responsiblefor martian features. Outflow channels require massive dischargerates that dwarf rates associated with PIH systems. Valley net-works, too, require discharge rates that are somewhat too large.A few craters exhibit alluvial fans with alcoves perched on the cra-ter’s wall and debris aprons deposited on the crater’s floor (Mooreand Howard, 2005). Kraal et al. (2008) suggest that a large-scalefan present in a 120 km crater formed in only a few decades andthat it required discharge rates on order of 103–105 m3 s�1. It ispossible that smaller-scale fans might be associated with PIH sys-tems, but the location of their alcoves, high up on the crater’s wall,argue against a hydrological relationship to PIH systems unlessheterogeneous permeability structures are involved. In all simula-tions of PIH systems exposed to frozen ambient surface conditions,a permanently frozen region prevents discharge near the craterwall. Alluvial fans are therefore likely not directly caused by PIHsystems subjected to freezing surface temperatures.

Groundwater discharge has been suggested for the formation ofsome gullies (Heldmann et al., 2007; Mellon and Phillips, 2001).Gullies �100 m to 1 km in length on crater wall slopes of �20� re-quire discharge rates of �10 m3 s�1 to form (Parsons and Nimmo,2010). These discharge rates argue for PIH systems with surfacepermeabilities on order 10�12–10�10 m2. However, the total dis-charge from these PIH systems would bury low-lying gullies. Thiswould bias gully preservation to topographic highs: the centralpeak, peak ring, or crater wall. Numerous craters exhibit gullieson their walls (Heldmann et al., 2007) and a few craters, such asLyot, Mojave, and Zunil exhibit gully-like, dendritic features in-cised into their central peaks (Tornabene et al., 2007; Williamsand Malin, 2008). Permeability structures, including vertical faultswithin the central peak and listric faults extending radially frombeneath the crater’s floor to the slumps at the wall, could providefluid conduits. Like alluvial fans, gullies at the crater wall, however,are unlikely to be formed by PIH systems exposed to frozen surfacetemperatures. Again, this is simply a consequence of the perma-nently frozen region occupying nearly half the radial extent ofthe crater—directing subsurface flow to the center and preventing

Page 12: Martian post-impact hydrothermal systems incorporating freezing

112 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

flow near the wall. Although PIH systems modeled here are capableof forming gullies it seems unlikely that they are responsible formost gully formation on Mars.

Very large-scale impactors—bolides 100–300 km in diameter—produce global atmospheric effects (Segura et al., 2002). Smallerscale PIH systems may generate local-to-regional atmospheric ef-fects. Like Icelandic geysers, steam and vapor may readily con-dense out near the PIH system, contributing to ‘‘pasted-onterrain” development on the rim of the crater. Orbit-induced cli-mate variations (e.g. Laskar et al., 2002) or subsequent neighboringimpact events may potentially lead to the melting of snowpacksand ice deposits, which is a suggested mechanism for gully forma-tion (Christensen, 2003).

4.2. Geochemical alteration

The OMEGA (Bibring et al., 2006) and CRISM (Mustard et al.,2008) spectrometers indicate that phyllosilicates and hydrated sil-icates are widely distributed in the southern highlands. Impact cra-ters saturate the southern highlands and indicate that they areamong the oldest surfaces on Mars (Hartmann and Neukum,2001). The crater age of the southern highlands corresponds withthe Phyllosian era during which the deep phyllosilicates and lay-ered phyllosilicates were apparently formed (Mustard et al., 2009).

Low-grade metamorphism of basaltic rock depends on temper-ature, pressure, host-rock chemistry and W/R ratios (Griffith andShock, 1997). W/R ratios for PIH systems simulated here vary inmagnitude and location depending on crater size, ambient surfacetemperature, and surface permeability. PIH systems exposed to5 �C surface temperatures yield elevated W/R at the central peakfor all simulated surface permeabilities. Systems with high surfacepermeabilities allow the hydrosphere to convect. Figs. 6–8 showthat this convection generates spatially-varying convectiveplumes. Given our axisymmetric domain, in planar view thisbehavior would generate a bull’s-eye pattern of high (>100) andmoderate (10) W/R ratios in surface-exposed rock across the entirefloor of the crater. Such a pattern might generate a spatial footprintthat might be detectable from orbit. Reality, however, is unlikely tobe so simple: heterogeneous permeability structures would inhibitthe formation of such patterns. Systems with lower surface perme-abilities (10�14–10�12 m2) exhibit modest W/R ratios at the centralpeak (1–100) and uniformly low W/R ratios of �1 in exposed rocksacross much of the floor.

PIH systems exposed to frozen surface temperatures produceelevated near-surface W/R ratios that peak at the center of the cra-ter and decay to zero at the initial unfrozen extent of the crater’sfloor. Systems with a high surface permeability yield central peakW/R ratios >1000. Systems with a surface permeability of10�12 m2 yield lower W/R ratios. Fig. 9C shows that temperatureswithin the upper 200 m of rock at the crater’s center are above100 �C for 2000 years following an impact event and remain abovefreezing for 20,000 years. As mentioned above (Section 1.4), low W/R ratios and modest temperatures (�150 �C) are likely to produceprehnite, chlorites, smectites, mica, amphibole and garnets. Someof these clays and minerals, including prehnite, are detected inthe central peak of a 45 km crater in Nili Fossae (Ehlmann et al.,2008). Prehnite, indicative of low-grade metamorphism of basalticrock, is particularly compatible with PIH systems modeled here be-cause it can form at low pressure and requires formation temper-atures below �300 �C (Liou et al., 1985; Digel and Ghent, 1994).Our model results for cases with permeabilities in the range of10�14–10�12 m2 and frozen surface temperatures are consistentwith the formation of prehnite, particularly at the central peak.

PIH systems create an environment conducive to forming min-eral assemblages. Perhaps numerous and successive PIH systemsoccupying a wide range of crater sizes along with repeated brecci-

ation and impact excavation of the southern highlands could haveproduced most of the phyllosilicates and hydrated silicates de-tected there. Subsequent reworking of the highlands (Craddockand Howard, 2002) and enhanced late-Noachian fluvial activity(Howard et al., 2005) involved low temperature fluvial erosion thatredistributed sediments from topographic highs to low-standingbasins. This redistribution would contribute to the masking of min-eralogical signatures by later sedimentation.

A few other effects might be associated with PIH systems. Re-ported spectroscopic observations (Mumma et al., 2009) suggestthat Mars may exhibit localized abundances of methane in itsatmosphere. Similar to clay mineralogy, abiogenic methane pro-duction depends on the W/R ratio and the water temperature(Lyons et al., 2005; Mumma et al., 2009). Although PIH systemsmight be related to methane production they are unlikely to ex-plain the reported ongoing production of methane: that would re-quire a large scale (�45 km) crater impact event within the last106 years. PIH systems may foster a hypothetical abode for life(Abramov and Kring, 2005). Simple organisms may potentially findrefuge from the harsh vagaries of the martian surface deep underthe cryosphere (Sleep and Zahnle, 1998). Even under frozen surfacetemperatures, PIH systems provide a chemically dynamic andwater-rich bridge that links the surface to these refugia for100,000s of years. Craters that show evidence of geochemical alter-ation and fluvial features related to PIH systems would make astrong candidate location for micro-fossils.

4.3. Limitations, assumptions, and caveats

4.3.1. Homogeneous permeabilityThe importance of PIH system permeability is second only to

the system’s heat budget in determining system behavior and itsobservable consequences. The permeabilities and porosities inour simulations decay smoothly with depth and exhibit no hori-zontal heterogeneity. Terrestrial hydrothermal systems have com-plex permeability structures and heterogeneities that greatlyinfluence flow and heat transport (Spinelli and Fisher, 2004). Thesmooth convective rolls and plumes present in the 5 �C simulationsmight be an artifact of the axisymmetric domain and probablywould not arise naturally if the domain had a complex permeabil-ity distribution (i.e. faulting, layering; Nield and Bejan, 1992). Sub-surface structures such as lava flows, faults, and buried impactcraters (cf. Edgett and Malin, 2004) would likely be disrupted bythe impact event but may still affect permeability. Impact craters,buried and exposed, and their associated fracture systems likelydominate the basement structural fabric of the ancient highlands(Rodrıguez et al., 2005).

If anything, a homogeneous permeability and porosity distribu-tion is the one distribution we can surely assume does not exist onMars. But for the sake of systematic first order parameter spaceexploration, we incorporated such a distribution with the impor-tant complication that freezing affects permeability. We can, how-ever, make inferences on how heterogeneous permeabilitystructures would affect PIH behavior. Regions of high permeabilitymay include highly fractured brecciated zones radially distributedfrom the center of the crater beneath the melt sheet; a central peakcomposed of poorly consolidated material uplifted up and throughthe melt sheet; and listric slump faults extending to the crater walland rim. Conversely, a melt sheet pooled on the crater’s floor willlikely steer flow to the center or toward the wall. Pore-cloggingprecipitation of minerals will cause flow to migrate from initial re-gions of high flow and hot temperatures to unclogged regions. Fi-nally, fractures may extend well beyond our domain depth of10 km. These fractures are unlikely to experience much geochem-ical alteration, but heat conducted or mined via fractures from this

Page 13: Martian post-impact hydrothermal systems incorporating freezing

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 113

depth will enter the domain during the lifetime of the hydrother-mal system.

This work also ignores topographic relief. There are two reasonsfor doing so: (1) under Mars’s gravity, hydrothermally driven flowcan easily overcome a few hundred meters of relief (Harrison andGrimm, 2002), and (2) post-impact permeability structures maylimit flow where a complex crater’s natural topography wouldotherwise direct it. That is, where highly fractured and faultedzones occurring at relative topographic highs at the crater’s rimand central peak would enhance flow, low-permeability meltdeposits on the crater’s floor would reduce flow. Previous workers,(Rathbun and Squyres (2002) and Abramov and Kring (2005)) didinclude topography. However, these studies assumed saturated do-mains which necessitated crater lake levels that were at least ashigh as the central peak or peak ring. This setup is essentially thesame as ours in that it maintains a zero head-gradient across thesurface that, in turn, prevents topography from altering surfacedischarge. However, Abramov and Kring (2005) show that the heatlocated in the central peak or peak ring does affect subsurface flowrouting by wicking water up and into these areas of relatively hightopography.

4.3.2. Heat budgetWe have made many assumptions in the construction of our

temperature distribution. Our approach produces shock-heated re-gions that agree well with hydrocode simulations (Pierazzo et al.,2005) and with prior modeling efforts (Abramov and Kring,2005), but contain roughly seven times more energy than shock-heated regions generated by the Gault–Heitowit approach de-scribed in Rathbun and Squyres (2002). Consequently, our workmay be compared to Abramov and Kring (2005), but not necessar-ily to Rathbun and Squyres (2002). We ran additional simulationswith temperature distributions generated using the same methodas Rathbun and Squyres (2002) and found that discharge rateswere lower by less than an order of magnitude and that the funda-mental behavior remained unchanged: permeability still con-trolled discharge rates and the time required for the system tofreeze; and flow was concentrated to the center of the crater asthe system froze.

We have ignored the role of volatiles and porosity in determin-ing shock-induced heating. The low strength of ice and the shockimpedance contrast between icy and rock layers may further mod-ify the temperature distribution and ejection volumes (Senft andStewart, 2008). Other assumptions are detailed in the Appendixincluding the decision to ignore contributions from a melt sheet.Finally, the background geothermal flux determines the depth ofthe cryosphere. Especially low background geothermal fluxes of�10 mW/m2 would support a cryosphere thick enough to preventshock-heat penetration into the hydrosphere for craters of the sizeconsidered here. PIH systems under this scenario would be water-starved and should be a topic of future study.

4.3.3. Saturated domainPIH systems require water to influence geomorphology and

geochemistry. We assumed fully saturated model domains. Re-gions of Mars, especially the equator, are likely unsaturated (Boyn-ton et al., 2002). Drier, partially saturated substrates would, ofcourse, produce lower surface discharge rates and lower W/R ra-tios. The wide variety of crater morphologies (i.e. rampart craters,lobed ejecta) not related to subsequent modification (i.e. aeolianand periglacial effects) on Mars may be due to varying amountsof water in the crust at the time of impact.

5. Conclusion

Post-impact hydrothermal systems are a likely mechanism fordelivering hot water to Mars’s cold, dry surface. The geomorphicand geochemical signatures of PIH systems occurring in 45–90 km craters on Mars depend on rock permeability and the sur-face temperature. If the PIH system occurs during a time in whichsurface temperatures are below freezing, discharge will not occurat the rim of the crater. Instead, all geomorphic activity and miner-alogical alteration will occur near the center of the crater. Dis-charge rates, total discharge and W/R ratios increase withpermeability. Systems with higher surface permeabilities(10�10 m2) allow convection, rendering them capable of miningheat from the central uplift before the surface freezes. Convectivesystems subjected to surface temperatures below freezing are par-ticularly interesting because an impermeable freezing front, thesurface ice/water interface, marches from mid-floor to the centralpeak and focuses flow towards the center of the crater. This effectprolongs modestly high temperatures (0–150 �C for 25 kyr) andyields W/R ratios >1000 for systems with high surface permeabili-ties of 10�10 m2 (Fig. 9D). For systems with modest surface perme-abilities, 10�12 m2, the upper 200 m of rock at the crater’s centerexperience fluid temperatures between 0 and 180 �C for 3500 yearsand W/R ratios of �10 (Fig. 9C). These results imply that differentsurface permeabilities should produce spatially diagnostic mineralalteration patterns. PIH systems with surface permeabilities of10�14–10�12 m2 exposed to frozen ambient temperatures may ex-plain mineral assemblages and fluvial features observed at centralpeaks of craters. In particular, the low-grade metamorphic mineralprehnite may be particularly diagnostic of the low temperature (0–300 �C), low pressure hydrothermal activity predicted by our mod-els. Discharge rates and total discharges are capable of producinggullies, ponds and lakes, but not alluvial fans, valley networks oroutflow channels. Future work should focus on the effect of heter-ogeneous permeability structures, specific case studies capable oftying theory to observations, and potential linkages with the regio-nal hydrosphere and atmosphere.

Acknowledgments

We thank Don Korycansky and Andy Fisher for their advice andconsultation. We acknowledge Keith Harrison, Oleg Abramov andRobert Lowell for their constructive reviews and valuable insight.The NASA Mars Fundamental Research Program GrantNNX07AU77G and the NASA Ames Graduate Student ResearcherProgram Grant 8254 funded this research.

Appendix A. Input temperature field generation

A.1. Heat sources

Three heat sources contribute to the initial temperature profilethat drives PIH hydrothermal activity: impact-induced shock-heat-ing, a melt sheet pooled on the crater floor, and an uplifted geo-thermal gradient. Unfortunately, there is not a goodunderstanding of how large-scale impacts deposit their heat. Cur-rent theory is highly dependent upon the equation of state ofimpactor/target material, assumptions of how shock travelsthrough the medium, and how that shock is converted into wasteheat. Porosity plays a major role in impact heat deposition, as dopore fluids. However, hydrocode simulations of impacts into a por-ous planet-scale target with, for example, an exponentially decay-ing porosity, have yet to be investigated. That said, simplifiedmodels of impact-induced shock-heating can be constructed using

Page 14: Martian post-impact hydrothermal systems incorporating freezing

-54 21 97 172 248 323 398 474 549 625 700

A

B

C

D

Temperature (°C)

Dep

th B

elow

Sur

face

(m)

radial distance (m)

Transient Radius

Crater Rim

Bolide Radius

Fig. 11. The construction of our initial temperature distribution for a 45 kmdiameter crater generated by a bolide 1900 m in radius traveling at 7 km s�1, beginswith a background geothermal profile (A). Next, the shock-heating is added to thebackground geothermal profile (B). A parabola excavates a region, indicated by thecrosshatch pattern (C), that has depth of hexc and intersects the surface at rtr (C). Weassume that the transient crater rebounds and that all material below regions ofexcavation are uplifted, vertically to the surface (D). Fig. 3 shows the finaldistribution following further temperature adjustments imposed by the limits ofthe model’s equation of state tables.

114 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

analytical formulations (e.g. Melosh, 1989) or hydrocode results(e.g. Pierazzo et al., 2005).

A.2. Shock-heating

We estimate the distribution of shock-heating produced by animpactor of radius, Rp, density, qp, and velocity, vp, by combiningan analytic expression for the change in internal energy as a func-tion of distance from the impact point with pi-scaling argumentsthat quantify the morphological change from the transient to finalcrater. The relationship between final crater diameter, D, and tran-sient crater diameter, Dt, is given by Melosh (1989): Dt = 0.6D.

Once Dt has been calculated we use the following scaling rela-tionship to determine the diameter of the bolide as a function ofthe bolide’s density and velocity (rearranged from Schmidt andHousen (1987) and Cintala and Grieve (1998)). Note: variablesdescribing projectile density, qp, target density, qt, bolide diameter,Dp, velocity, vp, transient crater diameter, Dt, and gravity, g, must beexpressed in cgs units:

Dp ¼ 0:862Dtqt

qp

!13

v�0:44p g0:22

0@

1A

1:2821

ð15Þ

In both the 45 km and 90 km case we assume an asteroid-like im-pact event and set qp equal to qt at 2600 kg m�3 and we set vp to7 km s�1. These are the same values used in both Rathbun andSquyres (2002) and Abramov and Kring (2005). There are, however,caveats concerning Eq. (15)’s derivation and application. First of all,it is scaled by gravity, g, from empirical relationships derived fromlaboratory experiments and lunar morphologies (Schmidt and Hou-sen, 1987). Furthermore, a non-unique combination of projectileand target densities, projectile diameters, and projectile velocitiescan generate the transient crater diameter required to fit the de-sired final crater diameter. Finally, hydrocode modeling indicatesthat low density comets impacting at high velocities will generatea final crater of comparable size compared with higher densityslower asteroids, but will produce �8 times as much melt (Pierazzoet al., 2005). Such craters may harbor very different PIH systemsfrom those considered here and further study is needed.

Given a bolide diameter, we determine the change in tempera-ture as a function of distance from the impact point. We follow amodified approach of the one described in Abramov and Kring(2005) which uses an expression for specific waste heat, DEw, de-rived from the Murnaghan equation of state by Kieffer and Si-monds (1980):

DEw ¼12

PV0 �2K0V0

n

� �1� Pn

K0þ 1

� ��1=n" #

þ K0V0

nð1� nÞ 1� PnK0þ 1

� �1�ð1=nÞ" #

ð16Þ

where K0 is the adiabatic bulk modulus at zero pressure, n is thepressure derivative of the bulk modulus, V0 is the specific uncom-pressed volume (1/qt) and P is the peak shock pressure (Kiefferand Simonds, 1980; Abramov and Kring, 2005). We use the samevalues for K0 = 19.3 GPa and n = 5.5 as Abramov and Kring (2005)as obtained from Gault (1963), Kieffer and Simonds (1980) and Me-losh (1989). We deviate from Abramov and Kring (2005) by choos-ing to use a steeper exponent, f, for peak pressure falloff:

P ¼ Ar

Rp

� ��f

ð17Þ

where r is the distance from the impact and A is the pressure atr = Rp and f is the pressure falloff exponent. For distances less thanRp, P = A. Melosh (1989) and Abramov and Kring (2005) estimate

A as A = qt[Ct + Stu0]u0, where Ct and St are parameters in the linearshock velocity–particle relation (Melosh, 1989; Appendix II). For ba-salt, Ct = 2600 m s�1 and St = 1.58 (dimensionless); the initial shockvelocity is u0 = 0.5vp. The change in temperature is calculated bydividing the specific waste heat by the heat capacity of basalt. Moti-vated by temperature distributions generated by a numericalhydrocode model for a 30-km martian crater by Pierazzo et al.(2005) we found that temperature contours with depth were repli-cated reasonably well analytically with a value for f of 3.0. Fig. 11Bshows the shock-heated temperature distribution added to thebackground geothermal temperature profile (Fig. 11A).

Modern hydrocode studies suggest that more than 50% of theimpactor’s kinetic energy can be converted into the internal energyof the target (e.g. Gisler et al., 2006; Abramov, personal communi-cation). One can integrate Eq. (16) over a hemisphere with a radiusset to the final crater radius to obtain an estimate for the amount ofinternal energy deposited into the target. Internal energy estimatesfor the 45 and 90 km craters are 1.3 � 1021 J and 1.9 � 1022 J,respectively. These values are roughly 67% of their respective bo-lides’ kinetic energies: 1.9 � 1021 J and 2.8 � 1022 J. The shallowerfalloff value of f = 2.016 used by Abramov and Kring (2005) depos-its roughly twice as much waste heat as the bolide’s kinetic energy.Of course some shock-heated material would be ejected or vapor-ized during the impact event.

We also generated initial post-impact temperature distribu-tions following the Gault–Heitowit approach detailed in Rathbunand Squyres (2002) and found that the shock-emplaced heat was1% of the impactor’s kinetic energy. However, simulations using

Page 15: Martian post-impact hydrothermal systems incorporating freezing

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 115

the Gault–Heitowit—Rathbun and Squyres (2002) approach in-stead of the Kieffer–Simonds—Abramov and Kring (2005) approachdid not change the fundamental behavior of PIH systems exposedto subfreezing temperatures: (1) permeability governs dischargerate, total discharge, and W/R ratios, (2) an advancing freezingfront concentrates flow to the center of the crater and (3) a perma-nently frozen subsurface region prevents discharge at or beyondthe rim of the crater. For these reasons, we chose to use the Kief-fer–Simonds analytic calculation of the initial temperature distri-bution as a tool to replicate hydrocode results, but acknowledgethat is not an exact description of the physics of shock-heatingdue to impact.

A.3. Excavation and geothermal uplift

The impact event excavates a significant portion of the shock-heated region and compresses the remaining material hydrody-namically. This forms a transient crater. We describe the volumeof material that is excavated by a parabola that extends to a depthof �1/10 the diameter of the transient crater, hexc (cf. Melosh, 1989,§5.5.4) and intersects the surface at the radius of the transient cra-ter, rtr. The crosshatch pattern in Fig. 11C shows the excavationparabola. The transient crater is defined by a second parabola cen-tered at a depth of 1

2:7 Dt (cf. Melosh, 1989, §8.2), that also intersectsthe surface profile at the radius of the transient crater. Here weintroduce three assumptions: (1) all material within the excavationparabola is removed; (2) that the final crater has no topography(i.e. the surface is flat) and that consequently; (3) the transient cra-ter rebounds all the way to the surface and that material at pre-im-pact temperatures established by the background geothermalprofile are brought closer to the surface by a height, z(r)exc, definedas a function of radius by the excavation parabola:

zðrÞexc ¼hexc

r2tr

� �r2 � hexc ð18Þ

Fig. 11D shows the resultant temperature distribution followingexcavation and rebound. This approach only moves material inthe vertical direction. In reality, material would move in from thesides too. Our domain depth, however, is only 10 km deep, allowingus to assume that material from depth is moved primarily in thevertical direction.

A.4. Further temperature adjustments

The temperatures at specific nodes are adjusted further if thevapor pressure at the node is above the hydrostatic pressure (i.e.the water is boiling), or if the temperature is outside the range de-fined by the equation of state look-up tables. We assume that va-por advects heat and escapes on a short timescale compared toother timescales. The vapor pressure for each node is calculatedas a function of temperature (Hardy, 1998). If the vapor pressureexceeds the hyrdostatic pressure, Phydro = qgz plus a backgroundsurface pressure of 105 Pa, in a particular node, then that node’stemperature is iteratively reduced until that is no longer the case.Adjustments of this kind occur in the hottest and shallowest areasof the domain—the upper shock-heated region at the center of thecrater—and are made to a depth of 800 m and 1800 m in the 45 kmand 90 km case, respectively.

MAGHNUM assigns the viscosity, density and expansivity ofwater within each node using an equation of state look-up table(see Section 2.6). The equation of state look-up table assigned tothis study does not go beyond temperatures of 900 �C. Thereforeany node with a temperature above 900 �C is initially set to900 �C. This adjustment primarily affects nodes in the uniformly

shocked hemisphere of radius Rp—a shallow region from whichheat is assumed to rapidly advect out of the system.

In reality, the melt volume produced by the impact event con-tributes additional heat to PIH systems. Melt volume has beenfound to depend on crater size and this relationship is consistentwhen scaled by gravity for terrestrial and lunar craters (Melosh,1989; Cintala and Grieve, 1998). Cintala and Grieve (1998) usescaling laws and the lunar cratering record to relate the volumeof melt to the transient-cavity diameter (Eq. (19)):

Vm ¼ cDdt ð19Þ

where Vm is the melt volume (km3), Dt, the transient crater diameter(km), and c and d are constants that for a chondritic impactor trav-eling at 10 km s�1 assume the values 1.08 � 10�4 and 3.85, respec-tively. The 45 km and 90 km craters have respective transient craterdiameters of �27 km and �54 km which yield melt volumes of�35 km3 and �504 km3.

We do not include a treatment of melt heating in our simula-tions because these volumes would fill the respective floors ofthe 45 km and 90 km craters to a depth of 22 m and 79 m—a depththat would conductively cool in less than 20 and 300 years, respec-tively, and are shallower than the node height (Dz = 200 m) used inour simulations. Abramov and Kring (2005) did not include melt intheir simulation of a 30 km diameter crater but they did include amelt sheet in simulations of 100 km craters. In the Abramov andKring (2005) simulations of 100 km craters, the melt sheet conduc-tively supplied heat to the PIH system and remained impermeablefor 20,000 years. Relative to system evolution, this is a long periodof time. However, a time of 20,000 years is likely a consequence ofan overestimation of melt production due to the vertical resolutionof their model, 500 m, which has a conductive cooling time of atleast 10,000 years. By not including the additional heat suppliedby the melt sheet our simulations may underestimate the time ittakes for the system to freeze. However, our geochemically signif-icant result, that flow is likely enhanced at the central peak, re-mains viable. Because a melt sheet pools around the centralpeaks of complex craters, an initially impermeable melt sheetmay further enhance flow at the central peak early in systemevolution.

Figs. 11 and 3 plot temperature contours which define theshock-heating and central geothermal uplift generated by animpactor with a radius of, Rp = 1900 m, a velocity of, vi = 7 km s�1,and density of, qi = qt = 2600 kg m�3. This bolide carries1.5 � 1021 J of kinetic energy and produces a transient crater diam-eter of 27 km, an excavation depth of 3333 m and a final craterdiameter of 45 km. For the 90 km case, bolide density and velocityremains the same but the impactor radius is set to 4700 m. This in-creases the kinetic energy to 2.1 � 1022 J, dtr to 54 km, and hexc to6666 m.

References

Abramov, O., Kring, D.A., 2005. Impact-induced hydrothermal activity on earlyMars. J. Geophys. Res. (Planets) 110. E12S09.

Ames, D., Watkinson, D., Parrish, R., 1998. Dating of a regional hydrothermal systeminduced by the 1850 Ma Sudbury impact event. Geology 26, 447–450.

Barnhart, C.J., Howard, A.D., Moore, J.M., 2009. Long-term precipitation and late-stage valley network formation: Landform simulations of Parana Basin, Mars. J.Geophys. Res. (Planets) 114 (E01003). doi:10.1029/2008JE003122.

Berman, D.C., Hartmann, W.K., 2002. Recent fluvial volcanic and tectonic activity onthe Cerberus Plains of Mars. Icarus 159, 1–17.

Bibring, J.-P., Langevin, Y., Mustard, J.F., Poulet, F., Arvidson, R., Gendrin, A., Gondet,B., Mangold, N., Pinet, P., Forget, F., 2006. Global mineralogical and aqueousMars history derived from OMEGA/Mars Express Data. Science 312, 400–404.

Boynton, W.V., and 24 colleagues, 2002. Distribution of hydrogen in the near surfaceof Mars: Evidence for subsurface ice deposits. Science 297, 81–85.

Cabrol, N., Grin, E., 2001. The evolution of lacustrine environments on Mars: Is Marsonly hydrologically dormant? Icarus 149, 291–328.

Carr, M.H., 1987. Water on Mars. Nature 326, 30–35.Carr, M.H., 1996. Water on Mars. Oxford University Press, New York. 248 pp.

Page 16: Martian post-impact hydrothermal systems incorporating freezing

116 C.J. Barnhart et al. / Icarus 208 (2010) 101–117

Carr, M.H., 2006. The Surface of Mars. Cambridge University Press, New York. 322pp.

Christensen, P.R., 2003. Formation of recent martian gullies through melting ofextensive water-rich snow deposits. Nature 422, 45–48.

Christensen, P.R., Ruff, S.W., 2004. The formation of the hematite-bearing unit inMeridiani Planum: Evidence for deposition in standing water. J. Geophys. Res.109 (E08003). doi:10.1029/2003JE002233.

Cintala, M.J., Grieve, R.A.F., 1998. Scaling impact-melt and crater dimensions:Implications for the lunar cratering record. Meteorit. Planet. Sci. 33, 889–912.

Clark, B.C., Castro, A.J., Rowe, C.D., Baird, A.K., Evans, P.H., Rose Jr., H.J., Toulmin III, P.,Keil, K., Kelliher, W.C., 1976. Inorganic analyses of martian surface samples atthe Viking landing sites. Science 194, 1283–1288.

Clifford, S.M., 1981. A pore volume estimate of the martian megaregolith based on alunar analog. LPI Contrib. 441, 46–48.

Clifford, S.M., 1993. A model for the hydrologic and climatic behavior of water onMars. J. Geophys. Res. 98, 10973–11016.

Clifford, S.M., Hillel, D., 1983. The stability of ground ice in the equatorial region ofMars. J. Geophys. Res. 88, 2456–2474.

Clifford, S.M., Parker, T.J., 2001. The evolution of the martian hydrosphere:Implications for the fate of a primordial ocean and the current state of thenorthern plains. Icarus 154, 40–79.

Craddock, R.A., Howard, A.D., 2002. The case for rainfall on a warm, wet early Mars.J. Geophys. Res. (Planets) 107 (E11), 5111. doi:10.1029/2001JE001505.

Davies, G.F., Arvidson, R.E., 1981. Martian thermal history, core segregation, andtectonics. Icarus 45, 339–346.

Digel, S., Ghent, E.D., 1994. Fluid–mineral equilibria in prehnite–pumpellyite togreenschist facies metabasites near Flin Flon, Manitoba, Canada: Implicationsfor petrogenetic grids. J. Metamorphic Geol. 12, 467–477.

Edgett, K.S., Malin, M.C., 2004. The geologic record of early Mars: A layered,cratered, and ‘‘Valley-ed” volume. In: Mackwell, S., Stansbery, E. (Eds.), Lunarand Planetary Institute Conference Abstracts, vol. 35, p. 1188.

Ehlmann, B.L., Mustard, J.F., Fassett, C.I., Schon, S.C., Head, J.W., Des Marais, D.J.,Grant, J.A., Murchie, S.L., 2008. Clay minerals in delta deposits and organicpreservation potential on Mars. Nature Geosci. 1 (6), 355–358.

Fanale, F.P., 1976. Martian volatiles—Their degassing history and geochemical fate.Icarus 28, 179–202.

Farmer, C.B., Doms, P.E., 1979. Global seasonal variation of water vapor on Mars andthe implications of permafrost. J. Geophys. Res. 84, 2881–2888.

Feldman, W.C., and 12 colleagues, 2002. Global distribution of neutrons from Mars:Results from Mars Odyssey. Science 297, 75–78.

Fisher, A.T., Narasimhan, T.N., 1991. Numerical simulations of hydrothermalcirculation resulting from basalt intrusions in a buried spreading center. EarthPlanet. Sci. Lett. 103, 100–115.

Gault, D.E., Heitowit, E.D., 1963. The partition of energy for hypervelocity impactcraters formed in rock. Hypervelocity Impact Symp. 11, 420–456.

Gisler, G.R., Weaver, R.P., Gittings, M.L., 2006. Energy partitions in three-dimensional simulations of the Chicxulub meteor impact. In: 37th AnnualLunar Planetary Science, p. 2095.

Grieve, R.A.F., Garvin, J.B., 1984. A geometric model for excavation and modificationat terrestrial simple impact craters. J. Geophys. Res. 89, 11561–11572.

Griffith, L.L., Shock, E.L., 1997. Hydrothermal hydration of martian crust: Illustrationvia geochemical model calculations. J. Geophys. Res. 102, 9135–9144.

Grimm, R.E., McSween Jr., H.Y., 1989. Water and the thermal evolution ofcarbonaceous chondrite parent bodies. Icarus 82, 244–280.

Gulick, V.C., 1998. Magmatic intrusions and a hydrothermal origin for fluvial valleyson Mars. J. Geophys. Res. (Planets) 103, 19365–19388.

Hanna, J.C., Phillips, R.J., 2005. Hydrological modeling of the martian crust withapplication to the pressurization of aquifers. J. Geophys. Res. (Planets) 110(E01004). doi:10.1029/2004JE002330.

Hardy, B., 1998. ITS-90 formulations for vapor pressure, frostpoint temperature,dewpoint temperature, and enhancement factors in the range �100 to +100 �C.In: Third International Symposium on Humidity and Moisture, London, England,pp. 1–8.

Harr, L., Gallagher, J., Kell, G., 1984. NBS/NRC Steam Tables, Thermodynamics, andTransport Properties and Computer Programs for Vapor and Liquid States ofWater. Taylor and Francis, Philadelphia, PA.

Harrison, K., Grimm, R., 2002. Controls on martian hydrothermal systems:Application to valley network and magnetic anomaly formation. J. Geophys.Res. (Planets), 107. E5.

Hartmann, W.K., Neukum, G., 2001. Cratering chronology and the evolution of Mars.Space Sci. Rev. 96, 165–194.

Heldmann, J.L., Carlsson, E., Johansson, H., Mellon, M.T., Toon, O.B., 2007.Observations of martian gullies and constraints on potential formationmechanisms. Icarus 188, 324–344.

Howard, A.D., Moore, J.M., Irwin III, R.P., 2005. An intense terminal epoch ofwidespread fluvial activity on early Mars: 1. Valley network incision andassociated deposits. J. Geophys. Res. (Planets), 110. E12S14.

Irwin III, R.P., Craddock, R.A., Howard, A.D., 2005. Interior channels in martian valleynetworks: Discharge and runoff production. Geology 33 (6), 489–492.

Kieffer, S.W., Simonds, C.H., 1980. The role of volatiles and lithology in the impactcratering process. Rev. Geophys. 18, 143–181.

Kraal, E.R., van Dijk, M., Postma, G., Kleinhans, M.G., 2008. Martian stepped-deltaformation by rapid water release. Nature 451, 973–976.

Kring, D.A., 2000. Impact events and their effect on the origin, evolution, anddistribution of life. GSA Today 10 (8), 1–7.

Laskar, J., Levrard, B., Mustard, J.F., 2002. Orbital forcing of the martian polar layereddeposits. Nature 419, 375–377.

Liou, J.G., Maruyama, S., Cho, M., 1985. Phase equilibria and mineral parageneses ofmetabasites in low-grade metamorphism. Mineral. Mag. 49 (321–333).

Lyons, J.R., Manning, C., Nimmo, F., 2005. Formation of methane on Mars by fluid–rock interaction in the crust. Geophys. Res. Lett. 32, L13201.

Manga, M., 2004. Martian floods at Cerberus Fossae can be produced bygroundwater discharge. Geophys. Res. Lett. 31, L02702.

McCarville, P., Crossey, L.J., 1994. Postimpact hydrothermal alteration of theManson impact structure. Meteoritics 29, 499–500.

McKay, C.P., Davis, W.L., 1991. Duration of liquid water habitats on early Mars.Icarus 90, 214–221.

McKay, C., Clow, G., Wharton Jr., R., Squyres, S.W., 1985. Thickness of ice onperennially frozen lakes. Nature 313, 561–562.

Mellon, M.T., Phillips, R.J., 2001. Recent gullies on Mars and the source of liquidwater. J. Geophys. Res. 106, 23165–23180.

Mellon, M.T., Jakosky, B.M., Postawko, S.E., 1997. The persistence of equatorialground ice on Mars. J. Geophys. Res. 102 (E8), 19357–19370.

Melosh, H.J., 1989. Impact cratering: A geologic process. Oxford Monographs onGeology and Geophysics, vol. 11. Oxford University Press, New York. 253p.

Moore, J.M., Howard, A.D., 2005. Large alluvial fans on Mars. J. Geophys. Res.(Planets) 110 (E9), E04005.

Mouginis-Mark, P.J., 1987. Water or ice in the martian regolith?—Clues fromrampart craters seen at very high resolution. Icarus 71, 268–286.

Mumma, M.J., Villanueva, G.L., Novak, R.E., Hewagama, T., Bonev, B.P., DiSanti, M.A.,Mandell, A.M., Smith, M.D., 2009. Strong release of methane on Mars innorthern summer 2003. Science 323, 1041–1045.

Mustard, J.F., Murchie, S.L., and the CRISM team, 2008. Hydrated silicate minerals onMars observed by the Mars Reconnaissance Orbiter CRISM instrument. Nature454, 305–309.

Mustard, J.F., Murchie, S.L., Bibring, J.P., Bishop, J.L., Ehlmann, B.L., McKeown, N.,Milliken, R.E., Poulet, F., Roach, L., 2009. Hydrous environments on Mars fromvisible-infrared orbital data. In: Workshop on Modeling Martian HydrousEnvironments. Lunar and Planetary Institute, Technical Report, p. 51–52.

Newsom, H.E., 1980. Post-impact hydrothermal circulation through impact meltsheets. Meteoritics 15, 339.

Newsom, H.E., Brittelle, G.E., Hibbitts, C.A., Crossey, L.J., Kudo, A.M., 1996. Impactcrater lakes on Mars. J. Geophys. Res. 101, 14951–14956.

Newsom, H.E., Sewards, T., Keil, K., Graup, G., 1986. Fluidization and hydrothermalalteration of the suevite deposit at the Ries Crater, West Germany, andimplications for Mars. J. Geophys. Res. 91, E239–E251.

Nield, D.A., Bejan, A., 1992. Convection in Porous Media, second ed. Springer-Verlag,New York.

Nimmo, F., Tanaka, K., 2005. Early crustal evolution of Mars. Annu. Rev. Earth Planet.Sci. 33, 133–161.

Osinski, G.R., 2005. Hydrothermal activity associated with Ries impact event,Germany. Geofluids 5, 202–220.

Osinski, G.R., Spray, J.G., 2001. Impact-generated carbonate melts: Evidence fromthe Haughton structure, Canada. Earth Planet. Sci. Lett. 194, 17–29.

Parsons, R.A., Nimmo, F., 2009. North–south asymmetry in martian crater slopes. J.Geophys. Res. (Planets) 114 (E13). doi:10.1029/2007JE003006.

Parsons, R. A., Nimmo, F., 2010. Numerical modeling of martian gully sedimenttransport: Testing the fluvial hypothesis. J. Geophys. Res. doi:10.1029/2009JE003517.

Pierazzo, E., Artemieva, N.A., Ivanov, B.A., 2005. Starting conditions forhydrothermal systems underneath martian craters: Hydrocode modelling.Geol. Soc. Am. 384, 443–457 (special paper).

Rathbun, J.A., Squyres, S.W., 2002. Hydrothermal systems associated with martianimpact craters. Icarus 157, 362–372.

Robinson, M.S., Tanaka, K.L., 1990. Magnitude of a catastrophic flood event at KaseiValles, Mars. Geology 18, 902–905.

Rodrguez, J.A.P., and 12 colleagues, 2005. Control of impact crater fracture systemson subsurface hydrology, ground subsidence, and collapse, Mars. J. Geophys.Res. 110 (E9), E06003.

Sanford, W.E., 2005. A simulation of the hydrothermal response to the ChesapeakeBay bolide impact. Geofluids 5, 185–201.

Schmidt, R.M., Housen, K.R., 1987. Some recent advances in the scaling of impactand explosion cratering. Int. J. Impact Eng. 5, 543–560.

Schubert, G., Solomon, S.C., Turcotte, D.L., Drake, M.J., Sleep, N.H., 1992. Origin andthermal evolution of Mars. In: Mars. The University of Arizona Press, pp. 147–183.

Schubert, G., Anderson, J., Travis, B., Palguta, J., 2007. Enceladus: Present internalstructure and differentiation by early and long-term radiogenic heating. Icarus188 (2), 345–355.

Schwenzer, S.P., Kring, D.A., 2008. Inferred impact-generated hydrothermal mineralassemblages in basaltic regions of Mars. In: Lunar and Planetary InstituteScience Conference Abstracts, vol. 39, p. 1817.

Schwenzer, S.P., Kring, D.A., 2009. Impact-generated hydrothermal alteration onMars: Clay minerals, oxides, zeolites, and more. In: Lunar and PlanetaryInstitute Science Conference Abstracts, vol. 40, Abstract 1421.

Segura, T.K., Toon, O., Colaprete, A., Zahnle, K., 2002. Environmental effects of largeimpacts on Mars. Science 298, 1977–1980.

Senft, L.E., Stewart, S.T., 2008. Impact crater formation in icy layered terrains onMars. Meteorit. Planet. Sci. 43 (12), 1993–2013.

Sleep, N.H., Zahnle, K., 1998. Refugia from asteroid impacts on early Mars and theearly Earth. J. Geophys. Res. 103 (E12), 28529–28544.

Page 17: Martian post-impact hydrothermal systems incorporating freezing

C.J. Barnhart et al. / Icarus 208 (2010) 101–117 117

Smith, D.E., and 18 colleagues, 1999. The global topography of Mars andimplications for surface evolution. Science 284, 1495.

Spinelli, G.A., Fisher, A.T., 2004. Hydrothermal circulation within topographicallyrough basaltic basement on the Juan de Fuca Ridge flank. Geochem. Geophys.Geosyst. 5, 2001.

Squyres, S.W., Carr, M.H., 1986. Geomorphic evidence for the distribution of groundice on Mars. Science 231, 249–252.

Squyres, S.W., Clifford, S.M., Kuz’min, R.O., Zimbelman, J.R., Costard, F.M.,1992. Ice in the Martian regolith. In: Kieffer, H.H., Jakosky, B.M., Snyder,C.W., Matthews, M. (Eds.), Mars. The University of Arizona Press, pp.523–554.

Sturkell, E., Broman, C., Forsberg, C., Trossander, T., 1998. Impact-relatedhydrothermal activity in the Lockne impact structure, Jämtland, Sweden. Eur.J. Mineral. 10, 589–606.

Taylor Jr., H.P., 1977. Water/rock interactions and the origin of H2O in graniticbatholiths. J. Geol. Soc. Lond. 133, 509–558.

Toksöz, M.N., 1979. Planetary seismology and interiors. Rev. Geophys. Space Phys.17, 1641–1655.

Toon, O.B., Pollack, J.B., Ward, W., Burns, J.A., Bilski, K., 1980. The astronomicaltheory of climatic change on Mars. Icarus 44, 552–607.

Tornabene, L.L., McEwen, A.S., Grant, J.A., Mouginis-Mark, P.J., Squyres, S.W., Wray,J.J., and the HiRISE team, 2007. Evidence for the role of volatiles on martianimpact craters as revealed by HiRISE. In: Lunar and Planetary InstituteConference Abstracts, vol. 38, p. 2215.

Travis, B.J., Rosenberg, N.D., 1991. Three-dimensional simulations of hydrothermalcirculation at mid-ocean ridges. Geophys. Res. Lett. 18, 1441–1444.

Travis, B.J., Schubert, G., 2005. Hydrothermal convection in carbonaceous chondriteparent bodies. Earth Planet. Sci. Lett. 240, 234–250.

Travis, B.J., Rosenberg, N.D., Cuzzi, J.N., 2003. On the role of widespread subsurfaceconvection in bringing liquid water close to Mars’ surface. J. Geophys. Res.(Planets) 108 (E4), 8040–8055.

Turcotte, D.L., Schubert, G., 2002. Geodynamics. Cambridge University Press,Cambridge, UK. p. 472. ISBN 0521661862.

Williams, R.M., Malin, M.C., 2008. Sub-kilometer fans in Mojave Crater, Mars. Icarus198 (2), 365–383.

Williams, R.M., Phillips, R.J., Malin, M.C., 2000. Flow rates and duration within KaseiValles, Mars: Implications for the formation of a martian ocean. Geophys. Res.Lett. 27 (7), 1073.

Wohletz, K.H., Sheridan, M.F., 1983. Martian rampart crater ejecta—Experimentsand analysis of melt–water interaction. Icarus 56, 15–37.