hydrological modeling of the martian crust with application to the pressurization … ·...

19
Hydrological modeling of the Martian crust with application to the pressurization of aquifers Jeffrey C. Hanna and Roger J. Phillips McDonnell Center for the Space Sciences and Department of Earth and Planetary Sciences, Washington University, St. Louis, Missouri, USA Received 26 July 2004; revised 14 October 2004; accepted 9 November 2004; published 26 January 2005. [1] We develop a hydrological model of the Martian crust, including both ancient heavily cratered terrains and younger basaltic and sedimentary terrains. The porosity, permeability, and compressibility are represented as interdependent functions of the effective stress state of the aquifer, as determined by the combination of the lithostatic pressure and the fluid pore pressure. In the megaregolith aquifer model, the crust is modeled as a 2 km thick megaregolith, composed of lithified and fractured impact ejecta, overlying the impact-fractured and partially brecciated basement rock. The hydraulic properties depend primarily upon the abundance of breccia and the compressional state of the fractures. The porosity ranges from approximately 0.16 at the surface to 0.04 at a depth of 10 km, with a sharp discontinuity at the base of the regolith. The permeability varies from approximately 10 11 m 2 at the surface to 10 15 m 2 at depths of 5 km or more and is strongly dependent upon the fluid pore pressure. The hydrologic properties of basaltic and sedimentary aquifers are also considered. These parameters are used to model the fluid pressures generated beneath a thickening cryosphere during a postulated dramatic cooling of the climate at the end of the Noachian. As a result of a negative feedback between the fluid pore pressure and the permeability, it is more difficult than previously thought to generate pore pressures in excess of the lithostatic pressure by this mechanism. The production of the outflow channels as the result of such a climatic change is deemed unlikely. Citation: Hanna, J. C., and R. J. Phillips (2005), Hydrological modeling of the Martian crust with application to the pressurization of aquifers, J. Geophys. Res., 110, E01004, doi:10.1029/2004JE002330. 1. Introduction [2] There is a substantial body of evidence pointing to the fact that groundwater plays a central role in the Martian hydrologic cycle, both past and present. The dendritic valley networks, which are the earliest recorded fluvial features on Mars, were likely formed by some combination of surface runoff and groundwater sapping [Goldspiel and Squyres, 2000; Hynek and Phillips, 2003; Williams and Phillips, 2001]. Either process is suggestive of the existence of a widespread and shallow groundwater system during the Noachian epoch. More recently in Mars history, with ages ranging from Hesparian to Amazonian [Carr and Clow, 1981], the circum-Chryse outflow channels provide the strongest evidence for the importance of groundwater. These enormous fluvial features, ranging from tens to hundreds of kilometers in width and a kilometer or more in depth, originate from groundwater sources in chaos or canyon regions [Baker and Milton, 1974; Carr, 1979]. A number of smaller outflow channels are found throughout the planet, including Athabasca and Mangala Valles. The young ages of the channel surface of Athabasca Valles indicates the persistence of outflow activity into the past 10 to 100 Ma [Berman and Hartmann, 2002]. The most recent fluvial features, with ages possibly younger than 10 Ma, are small-scale gullies, which occur on the steep slopes of crater walls, central peaks, and canyon walls in high latitudes [Malin and Edgett, 2000a]. While the origin of these features is still uncertain, one interpreta- tion is that they result from the freezing of small, confined aquifers at depths of several hundred meters [Gaidos, 2001; Mellon and Phillips, 2001]. [3] Despite this dramatic evidence for both ancient and recent hydrologic activity on Mars, many first order ques- tions regarding the distribution of groundwater and the hydrologic properties of the Martian crust remain largely unanswered. Yet any attempt to understand the nature and history of the action of water on the Martian surface must hinge upon the assumptions made regarding the hydrologic properties of the subsurface. This study presents a general- ized hydrologic model of the Martian crust that allows for the modeling of a variety of processes over the wide range of conditions thought to have existed in the Martian subsurface, and then applies that model to gain insight into the origin of the outflow channel floods. [4] Hydrologic phenomena on Mars range in spatial scale from gullies at the hundred meter to kilometer scale, to outflow channels at the scale of tens to thousands of kilo- meters. For smaller scale features, the local variability in the JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, E01004, doi:10.1029/2004JE002330, 2005 Copyright 2005 by the American Geophysical Union. 0148-0227/05/2004JE002330$09.00 E01004 1 of 19

Upload: others

Post on 27-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

Hydrological modeling of the Martian crust with application to the

pressurization of aquifers

Jeffrey C. Hanna and Roger J. PhillipsMcDonnell Center for the Space Sciences and Department of Earth and Planetary Sciences, Washington University,St. Louis, Missouri, USA

Received 26 July 2004; revised 14 October 2004; accepted 9 November 2004; published 26 January 2005.

[1] We develop a hydrological model of the Martian crust, including both ancientheavily cratered terrains and younger basaltic and sedimentary terrains. The porosity,permeability, and compressibility are represented as interdependent functions of theeffective stress state of the aquifer, as determined by the combination of the lithostaticpressure and the fluid pore pressure. In the megaregolith aquifer model, the crust ismodeled as a 2 km thick megaregolith, composed of lithified and fractured impactejecta, overlying the impact-fractured and partially brecciated basement rock. Thehydraulic properties depend primarily upon the abundance of breccia and thecompressional state of the fractures. The porosity ranges from approximately 0.16 atthe surface to 0.04 at a depth of 10 km, with a sharp discontinuity at the base of the regolith.The permeability varies from approximately 10�11 m2 at the surface to 10�15 m2 at depths of5 km or more and is strongly dependent upon the fluid pore pressure. The hydrologicproperties of basaltic and sedimentary aquifers are also considered. These parameters areused to model the fluid pressures generated beneath a thickening cryosphere during apostulated dramatic cooling of the climate at the end of the Noachian. As a result of anegative feedback between the fluid pore pressure and the permeability, it is more difficultthan previously thought to generate pore pressures in excess of the lithostatic pressure by thismechanism. The production of the outflow channels as the result of such a climatic change isdeemed unlikely.

Citation: Hanna, J. C., and R. J. Phillips (2005), Hydrological modeling of the Martian crust with application to the pressurization of

aquifers, J. Geophys. Res., 110, E01004, doi:10.1029/2004JE002330.

1. Introduction

[2] There is a substantial body of evidence pointing to thefact that groundwater plays a central role in the Martianhydrologic cycle, both past and present. The dendritic valleynetworks, which are the earliest recorded fluvial features onMars, were likely formed by some combination of surfacerunoff and groundwater sapping [Goldspiel and Squyres,2000; Hynek and Phillips, 2003; Williams and Phillips,2001]. Either process is suggestive of the existence of awidespread and shallow groundwater system during theNoachian epoch. More recently in Mars history, with agesranging from Hesparian to Amazonian [Carr and Clow,1981], the circum-Chryse outflow channels provide thestrongest evidence for the importance of groundwater.These enormous fluvial features, ranging from tens tohundreds of kilometers in width and a kilometer or morein depth, originate from groundwater sources in chaos orcanyon regions [Baker and Milton, 1974; Carr, 1979]. Anumber of smaller outflow channels are found throughoutthe planet, including Athabasca and Mangala Valles. Theyoung ages of the channel surface of Athabasca Vallesindicates the persistence of outflow activity into the past

10 to 100 Ma [Berman and Hartmann, 2002]. The mostrecent fluvial features, with ages possibly younger than10 Ma, are small-scale gullies, which occur on the steepslopes of crater walls, central peaks, and canyon walls inhigh latitudes [Malin and Edgett, 2000a]. While theorigin of these features is still uncertain, one interpreta-tion is that they result from the freezing of small,confined aquifers at depths of several hundred meters[Gaidos, 2001; Mellon and Phillips, 2001].[3] Despite this dramatic evidence for both ancient and

recent hydrologic activity on Mars, many first order ques-tions regarding the distribution of groundwater and thehydrologic properties of the Martian crust remain largelyunanswered. Yet any attempt to understand the nature andhistory of the action of water on the Martian surface musthinge upon the assumptions made regarding the hydrologicproperties of the subsurface. This study presents a general-ized hydrologic model of the Martian crust that allows forthe modeling of a variety of processes over the wide rangeof conditions thought to have existed in the Martiansubsurface, and then applies that model to gain insight intothe origin of the outflow channel floods.[4] Hydrologic phenomena on Mars range in spatial scale

from gullies at the hundred meter to kilometer scale, tooutflow channels at the scale of tens to thousands of kilo-meters. For smaller scale features, the local variability in the

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, E01004, doi:10.1029/2004JE002330, 2005

Copyright 2005 by the American Geophysical Union.0148-0227/05/2004JE002330$09.00

E01004 1 of 19

Page 2: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

crustal materials and hydrologic properties will be signifi-cant, and the range of parameter space that must be consid-ered will span several orders of magnitude. However, onlarge spatial scales, the local heterogeneity will average outand the large-scale hydraulic parameters can be estimatedwith a greater degree of confidence. The model here devel-oped is a representation of the large-scale average hydraulicproperties of the Martian crust. To a first approximation,Martian crustal materials can be divided into three categories:(1) heavily cratered, Noachian-aged crust that predominatesin the southern highlands; (2) younger basaltic material thatlikely makes up much of the crust in the Tharsis and Elysiumvolcanic provinces, and which is likely present in smallerproportions within the crust elsewhere on the planet; and(3) sedimentary deposits that appear to be most abundantwithin closed depressions [Malin and Edgett, 2000b]. Manyterrains are best represented as combinations of the threetypes, such as lava flows with interbedded sediments andimpact ejecta, and the northern plains, which are thought tobe made up of an ancient heavily cratered basement [Frey etal., 2002] overlain by volcanic plains with a thin sedimen-tary cover [Head et al., 2002]. We model the properties ofthe heavily cratered Martian crust based upon the record ofsurface impacts and the observed properties of terrestrialand lunar impact craters, drawing also on the hydrologicproperties of terrestrial and lunar analogs to Martian regolithmaterial. The younger basaltic and sedimentary regions aremodeled directly after terrestrial aquifers.[5] This model differs from previous studies in that it

includes the compressibility of the aquifer, and thus allowsfor the modeling of time-dependent flow. Furthermore, all ofthe hydraulic properties of interest are modeled interdepen-dently using a self-consistent and physically realistic model.The parameter values can be calculated for any combinationof lithostatic and fluid pore pressure. This flexibility isessential for modeling many of the hydraulic processes thatare thought to have occurred on Mars involving deeplyburied aquifers and both large and rapidly varying porepressures.[6] In the following sections we review two existing

models of the hydrologic properties of the Martian crust[Clifford, 1993; MacKinnon and Tanaka, 1989], beforepresenting the model that is developed in this study. Thishydrologic model is then used to test the theory that thelarge pore pressures necessary to form the outflow channelscould have been generated beneath a downward propagat-ing freezing front [Carr, 1979].

2. Governing Equations

[7] Before developing the hydrologic model of theMartian crust, we first review the relevant governingequations in order to introduce the physical properties ofinterest. Steady state flow within an aquifer is governed byDarcy’s Law:

q ¼ Krh; ð1Þ

where q is the volumetric flux vector of water per unit areaor Darcy velocity (m/s), K is the hydraulic conductivity(m/s), and h is the hydraulic head (m). The hydraulicconductivity describes the resistance to flow within theaquifer, and is determined by the intrinsic permeability of

the aquifer k (m2; 1 darcy = 10�12 m2), the fluid dynamicviscosity m (Pa s), the acceleration of gravity g, and thedensity of water rw:

K ¼ krwgm

: ð2Þ

The dynamic viscosity of water is temperature dependent[Demming, 2002], described by the equation:

m Tð Þ ¼ 2:4� 10�5 � 10248= T�140Kð Þ Pa � sð Þ: ð3Þ

For Martian gravity, the permeability can be converted tohydraulic conductivity values by multiplying by a factorof approximately 2 � 106 (m s)�1 at 273 K. In this studywe report values of the permeability rather than thehydraulic conductivity, as the permeability is an intrinsicproperty of the aquifer material itself and does not requirescaling to account for changes in gravity, fluid density,and viscosity. The hydraulic head is a potential term,which includes both the elevation z of a fluid parcelabove a datum, as well as the aquifer pore pressure Ppore:

h ¼ Ppore

rwgþ z: ð4Þ

Since we are only interested in the gradient of the head,it does not matter to what datum the elevation isreferenced, as long as it is done consistently throughoutthe aquifer.[8] Transient flow within an aquifer is governed by the

consolidation equation [Domenico and Schwartz, 1990]:

@h

@t¼ 1

Ssr � Krhð Þ: ð5Þ

The specific storage, Ss (m�1), describes the elastic response

of the aquifer to pressure changes:

Ss ¼ rwg nbw þ baquifer� �

; ð6Þ

where n is the porosity and bw and baquifer (Pa�1) are the

compressibilities of water and the aquifer matrix respec-tively. The aquifer matrix compressibility is defined as

baquifer ¼1

Vtotal

@Vtotal

@seff 1

Vtotal

@Vpore

@seff¼ 1

1� nð Þ@n

@seff: ð7Þ

Note that the compression of the aquifer depends uponthe effective stress state (seff), defined as the differencebetween the lithostatic and fluid pore pressures, rather thanon the lithostatic pressure alone. Since the compressibility ofthe pore space is much greater than the compressibility of theactual mineral grains, the change in the volume of the rock,Vtotal, with changing effective stress is essentially equal tothe change in volume of the pore space, Vpore. The finalequality in equation (7) can be arrived at by setting n equal toVpore/Vtotal, and @Vtotal equal to @Vpore [Domenico andSchwartz, 1990].[9] The porosity of the aquifer can be calculated as a

function of the effective stress, again assuming that the

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

2 of 19

E01004

Page 3: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

mineral grains themselves are essentially incompressibleand the bulk compression is accommodated entirely by adecrease in the pore volume:

n seff� �

¼ n0 � 1ð Þ � expZseff0

baquifer � ds

0@

1Aþ 1

¼ n0 � 1ð Þ � exp baquifer � seff� �

þ 1 for constant b; ð8Þ

where n0 is the uncompressed porosity. As will be seen inthe sections that follow, in some cases the compressibility isa function of the effective stress, and equation (8) mustremain an integral expression.[10] Thus three essential hydrologic parameters are re-

quired to represent an aquifer: the porosity, permeability,and compressibility. The porosity is important for estimatingthe total volume of water that may be present (section 4.5), formodeling the thermal conductivity of the crust (section 4.5),and for modeling pressurization by the injection of excesswater into the pore space (section 6.2). The permeability, k, isessential for modeling both steady state and transient flowwithin the aquifer (section 6.2). The compressibility, b,governs both transient flow and the pressurization of theaquifer (section 6.2). These parameters are interdependent, asboth the nature and the volume of the pore space determinesthe permeability, and the compressibility determines depen-dence of the porosity and permeability on the effective stress.Furthermore, as it is likely that extremely large pore pressureswere required to form the outflow channels, it is necessary toconsider the variation of these parameters bothwith depth andwith changing pore pressure.

3. Previous Studies

[11] MacKinnon and Tanaka [1989] modeled the hydro-logic properties of the heavily cratered Martian uppercrust, consisting of a thick megaregolith composed ofbreccia overlying the impact-fractured basement rock.They assumed that the Martian crust is made up of initiallyimpermeable igneous rock that has been fractured byimpacts. Citing fracture widths resulting from the DannyBoy nuclear test explosion [Nugent and Banks, 1966], theyrepresented the deep Martian crust as a basement rockpermeated by fractures with an average aperture of 5 mmand a spacing of 3 m, resulting in a porosity of 1.5 � 10�3

and a permeability of approximately 10�9 m2. Theysuggested that overlying this fractured basement rock isa 1–2 km thick megaregolith of impact ejecta, this basedon Monte Carlo simulations of the ejecta thickness on asurface with a random distribution of impacts [Woronow,1988] and on observations of apparent crustal discontinu-ities in this depth range. The hydraulic properties of thisejecta layer were modeled using clast size distributionsfrom the Nevada Test Site [O’Keefe and Ahrens, 1985]and from the Meteor Crater ejecta. They calculated thatthis loosely packed Martian ejecta layer has a porositybetween 0.1 and 0.2 and that the permeability is likely lessthan 10�14 m2.[12] While the MacKinnon and Tanaka [1989] model is

a good conceptual representation of the ancient Martiancrust, it does not include the variation of porosity andpermeability with increasing confining pressure or chang-

ing pore pressure. The permeability of terrestrial materialsdecreases by several orders of magnitude at depths of afew kilometers [Huenges et al., 1997; Manning andIngebritsen, 1999]. While fractures can have arbitrarilylarge apertures with little to no overburden or if they havebeen filled with particulate matter, the aperture of unfilledfractures decreases with increasing confining pressure. Thefracture apertures measured at the site of the Danny Boyexplosion are much larger than those typically found in theterrestrial crust, and are more similar to those measured inrecently activated faults. Active motion along a fault isfollowed by a period of dramatically increased permeabil-ity due to the formation of large aperture fractures withinthe fault. However, the increase in aperture and perme-ability is a temporary effect and the wide apertures are notstable over geologic time [Gudmundsson, 2001]. Thisestimate of the fracture aperture may be representative ofthe time period immediately following an impact, but it isinflated above typical values for stable fractures at shallowdepth by a factor of about 50, and neglects the variation ofaperture with depth [Snow, 1970]. Since the permeabilitydepends on the cube of fracture aperture, this overestima-tion of fracture aperture leads to overestimation of thepermeability by a factor of as much as 105. Furthermore,their model does not include the compressibility of theaquifer and thus cannot be used to model aquifer pressur-ization or transient flows.[13] The most widely cited model of the hydrologic

properties of the Martian crust is that of Clifford [1981,1993] and Clifford and Parker [2001]. This model assumedan exponential decrease in porosity with depth due to theelastic compression of the pore space and was based on thelunar model of Binder and Lange [1980] scaled to Marsgravity:

n zð Þ ¼ 0:2 � e�z=6:5 km Moon

¼ 0:2 � e�z=2:8 km Mars; ð9Þ

where z is the depth in km below the surface. The Binderand Lange [1980] model was loosely based on lunar seismicvelocities, which were originally interpreted to suggestclosure of fractures and pores at the 20-km seismicdiscontinuity [Keihm and Langseth, 1977]. The lunarequation above simply fits an exponential decrease inporosity from 0.2 at the surface to 0.01 at 20 km depth, anddoes not attempt to fit the seismic data in between thesevalues, which shows a discontinuity at the base of themegaregolith as well as the second discontinuity at 20 km[Toksoz et al., 1974; Watkins and Kovach, 1973].[14] Alternatively, Toksoz et al. [1974] suggested that the

20 km seismic discontinuity may correspond to a compo-sitional change rather than the closure of the pore space.Support for this interpretation of the seismic data came fromanalysis of the geoid to topography ratios around lunarbasins, which indicates that the discontinuity is due to adensity interface as would exist between an upper anortho-sitic crust and a lower noritic crust [Wieczorek and Phillips,1997].[15] More recent analyses of the lunar seismic data [Khan

and Mosegaard, 2002; Khan et al., 2000; Longonne et al.,2003] have called into question the existence of the 20 kmseismic discontinuity altogether. Khan et al. [2000], using

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

3 of 19

E01004

Page 4: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

an updated data set of all locatable moonquakes, found asteady increase in the P wave velocity from the base of themegaregolith at a depth of 1 km, down to the base of thecrust at a depth of 40 km. They found no evidence for anintracrustal seismic discontinuity at 20 km. Longonne et al.[2003] found a 30 km thick crust, with increasing seismicvelocity with depth. Looking at the VP/VS ratio, Longonneet al. [2003] found evidence that the crust is heavilyfractured throughout its thickness, and that the fracturesmay even extend somewhat deeper than the 30 km crustalthickness. However, it is not possible to calculate hydro-logically useful parameters such as fracture frequency andaperture based on seismic data alone, as the seismic wavevelocity is affected by a number of unknown parametersincluding the along-strike dimension and the aspect ratio ofthe fractures [O’Connell and Budiansky, 1974]. In addition,the elastic properties, and thus the effect on seismic wavevelocities, of simple spherical pores and elongated fracturesdiffer substantially [Walsh, 1965], complicating the matterfurther when both pores and fractures are present.[16] The proposed elastic decrease in porosity with depth

does not correlate with the low compressibility actuallymeasured for the lunar breccias [Warren and Trice, 1975].The demonstration of substantial crustal permeability atdepths of 9 km on the Earth [Huenges et al., 1997],corresponding by a simple pressure scaling to a depth of24 km on Mars and 56 km on the Moon, contradicts theinterpretation of the elastic closure of the lunar pore space at20 km. In section 4.5, we demonstrate that the closure ofpore space is more likely caused by either plastic deforma-tion or pressure solution.[17] The Clifford [1993] study did not predict the perme-

ability based on the model of porosity, since there is nounique relationship between porosity and permeability in afractured porous medium. Clifford and Parker [2001]modeled the permeability of the Martian crust after thestudy by Manning and Ingebritsen [1999] of the averagepermeability in the terrestrial crust as a function of depth,which they again scaled to Mars gravity:

log kð Þ ¼ �14� 3:2 log z � gMars=gEarthð Þ¼ �12:65� 3:2 log zð Þ; ð10Þ

where z is the depth in km, and k is the permeability in m2.The Manning and Ingebritsen [1999] model was based on

indirect evidence for the permeability obtained from heatflow data and metamorphic reaction rates.[18] The permeability in the Earth’s crust is influenced

by a number of different factors, including rock type,fracture frequency, fracture aperture, heat flow, degree ofmetamorphism, age, and tectonic history. A simple grav-itational scaling of the terrestrial permeability function toMars is likely an oversimplification. For the same reason,this model also does not allow for the variation ofpermeability with changing pore pressure. Note that asthe depth approaches zero, or alternatively as the porepressure approaches the lithostatic pressure, equation (10)would predict infinite permeability. This unphysical resultoccurs simply because the model was intended only toapply to greater depths within the Earth’s crust. Thus theManning and Ingebritsen [1999] model is of only limitedapplicability and does not allow for the modeling ofshallow aquifers or deep aquifers with large pore pres-sures, both of which are integral parts of the Martianhydrologic system.[19] Another drawback of these models is that the

parameters of interest were not modeled self consistently.The porosity and permeability models were based onhydrologic models of different geologic environments ondifferent planets. Furthermore, they did not consider thecompressibility of the crust, thus precluding the modelingof time-dependant flows. What is needed is a single self-consistent and physically realistic model of the porosity,permeability, and compressibility of the Martian crust forany combination of depth and pore pressure.

4. Megaregolith Aquifer Model

4.1. General Characteristics

[20] We assume, as a starting condition, that the crustof Mars at the time of its formation was composed of acrystalline basement rock [Harper et al., 1995], whichwas subsequently modified by impacts (Figure 1). On thebasis of the effect of a single impact on the porosity andfracture frequency induced in the host rock, it should bepossible to calculate the hydraulic properties of a plane-tary surface with a given crater distribution. Since thesouthern highlands of Mars are near crater-saturation, anypreferential radial or circumferential orientation of theimpact-induced damage should average out and theresulting porosity, permeability and compressibility areconsidered to be isotropic. We assume that the impactsduring and after the period of heavy bombardmentresulted in a heavily fractured and partially brecciatedbasement overlain by a megaregolith breccia layer ofconsolidated impact ejecta [Clifford, 1981; MacKinnonand Tanaka, 1989; Woronow, 1988]. For Noachian-agedsurfaces, the thickness of this upper megaregolith layerwould likely be between 1 and 3 km [Hartmann et al.,2001;MacKinnon and Tanaka, 1989;Ward, 2002;Woronow,1988]. Since cratering and megaregolith production arecontemporaneous processes, the megaregolith is subject toimpact modification as well. This megaregolith layer iscapped by a thin (tens of m up to 1 km) layer of aeolianand aqueous sediments [Malin and Edgett, 2000b], Viking-type ‘‘soil’’ [Moore et al., 1977], and unconsolidated regolithfrom the perpetual impact gardening [Hartmann et al.,

Figure 1. Conceptual model of the breccia and fracturesbeneath an isolated impact crater. The surface is coveredwith a 2 km thick megaregolith (gray). Impacts produceboth breccias (gray) and fractures (white) beneath the crater.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

4 of 19

E01004

Page 5: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

2001; Shkuratov and Bondarenko, 2001; Watkins andKovach, 1973]. However, this upper thin veneer ofmaterials will not be hydrologically active under the coldclimate conditions that have persisted for much of Mars’history.

4.2. Impact-Induced Breccias

[21] As impact-generated breccias are thought to be asignificant component of the ancient heavily crateredportions of the Martian crust, it is thus necessary toestimate the distribution, abundance, and hydraulic prop-erties of breccias in Martian aquifers. As previouslystated, the accumulation of impact ejecta on the Martiansurface likely produced a megaregolith breccia layerbetween 1 and 3 km thick. Furthermore, breccias areproduced in the bedrock beneath craters both during theinitial passage of the shockwave after impact, and duringthe subsequent decompression and modification phases[Dressler et al., 1996]. In this latter stage, large brecciabodies and anastamozing breccia dykes form as blocks ofbedrock slide past one another during the modification ofthe transient cavity, providing a mechanism for theacoustic fluidization of the target material [Ivanov,2002; Melosh and Ivanov, 1999].[22] The distribution and abundance of breccia beneath

impact craters can be inferred from a combination ofgravity studies and drill core data. Early studies ofterrestrial impact craters revealed that they are oftenassociated with negative gravity anomalies [Innes, 1961;Pilkington and Grieve, 1992]. In a study of young lunarcraters, Dvorak and Phillips [1977] found that the gravitydata was best explained by a brecciated layer extendingto a depth of one third of the crater diameter and laterallyto the crater rim, with a density contrast of 0.3 g/cm3.The deep drilling project into the 40 km diameterPuchezh-Katunki impact crater found that the crust beneaththe crater consists of blocks of bedrock, 100 to 400 macross, separated by zones of breccia [Ivanov et al., 1996;Kocharyan et al., 1996]. The average ratio of breccia tocompetent rock was 1:3 down to a depth of about 4 km.Combining these two lines of evidence, we suggest that thecrust beneath large craters is composed of a mixture ofbreccia and bedrock, in a ratio of 1:3, down to a depth of1/3 the crater diameter.[23] The hydraulic properties of breccias can be esti-

mated on the basis of observations of terrestrial and lunarbreccias. Porous fault breccias have porosities of around0.1 [Antonellini and Mollema, 2000]. MacKinnon andTanaka [1989] calculated that Martian ejecta should haveporosities in the range of 0.1 to 0.2. The porosities oflunar breccia samples lie in the range of 0 to 0.4 with amean of 0.17, and with the most representative sample ofhighlands breccia having a porosity of 0.15 [Warren andRasmussen, 1987].We adopt a value of 0.15 as representativeof the uncompressed porosity of breccias, including bothlithified impact ejecta and breccias intermixed with thebedrock at greater depth.[24] Laboratory studies of the compressibility of fault

breccia yield values of approximately 10�9 Pa�1 [Serontet al., 1982]. Similarly, the compressibility of lunarbreccias is approximately 10�9 Pa�1 at low effectivestresses, while at higher effective stresses it can be fit

by a power law with a change in slope at 10 MPa[Warren and Trice, 1975]:

bbrec ¼ c1 � sc2effc1 ¼ 7:94� 10�10 for 0:1 < seff < 10MPa

c1 ¼ 2:81� 10�9 for 10 � seff � 1000MPa

c2 ¼ �0:1 for 0:1 < seff < 10MPa ð11Þc2 ¼ �0:65 for 10 � seff � 1000MPa;

where the effective stress is in MPa and the compressibilityis in Pa�1. The breccia porosity as a function of the effectivestress can then be calculated using equations (8) and (11).[25] Breccias have low permeability relative to their po-

rosity due to the poor degree of sorting and the presence ofsmall clast sizes. Gudmundsson [2001] measured the perme-ability of fault breccias to be in the range of 10�17 to 10�20m2,much lower than the upper limit of 10�14 m2 calculated byMacKinnon and Tanaka [1989]. As will be seen in the nextsection, the permeability of intact breccia is orders of magni-tude lower than the permeability of the superimposed fracturenetwork, and thus can be neglected. Both seismic observa-tions of the lunar megaregolith [Watkins and Kovach, 1973],as well as geomorphic observations of fault scarps in the topkilometer or so of the Martian megaregolith suggest that thebreccia is likely a coherent, lithified unit. We assume thatbreccia is subject to impact fracturing to the same degree ascompetent rock, so fractured breccia is treated in the samemanner as fractured bedrock with regards to the permeability.The observation of hydraulically conducting fractures withinthe breccia filled cores of fault zones [Gudmundsson et al.,2001] supports this assumption.[26] In summary, we assume that the heavily cratered

Noachian-aged crust is composed of a 2 km thick mega-regolith overlying the partially brecciated basement rock, inwhich the ratio of breccia to bedrock is 1:3. The breccias inthe upper megaregolith and intermixed within the bedrockare assumed to have similar hydraulic properties, with anuncompressed porosity of approximately 0.15, a compress-ibility given by equation (11), and negligible permeability.The permeability within the breccia units will be determinedby the presence of the superimposed fracture network, if oneexists.

4.3. Impact-Induced Fractures

[27] The permeability resulting from the presence offractures beneath an impact crater is determined by thefracture frequency (N, expressed in terms of the number offractures per unit distance) and the fracture aperture (b), asdescribed by the cubic law [Domenico and Schwartz,1990]:

kfrac ¼Nb3

12: ð12Þ

The fracture porosity in the host rock can be calculated asthe product of the fracture frequency and the fractureaperture:

nfrac ¼ N � b: ð13Þ

Thus, to calculate the impact-induced permeability andfracture porosity, it is necessary to estimate the frequencyand aperture of fractures beneath a crater.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

5 of 19

E01004

Page 6: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

[28] Seismic data indicate that the zone of fracturesaround a crater with diameter D extends radially adistance of D/2 beyond the crater rim and down to adepth of D/2 [Ackermann et al., 1975]. This is supportedby both impact experiments [Ahrens and Rubin, 1993]and numerical modeling studies [Ivanov, 2002]. Thefracture frequency in the 218 m deep drill hole in theLappajarvi crater in Finland [Kukkonen et al., 1992] isapproximately 20 m�1. Similarly, the fracture frequencyin outcrops of the intensely fractured and shatteredbedrock in the Acraman crater in Australia is between20 and 100 m�1 [Williams, 1994]. These values aresignificantly greater than the fracture frequency of typicalterrestrial crust, which lies in the range of 0.1 to 5 m�1

[Seeburger and Zoback, 1982], but are roughly consistentwith the fracture frequency of about 10 m�1 in activefault zones [Gudmundsson et al., 2001]. The increasedfracture frequency is in agreement with the observation ofa large zone of crushed rock in the vicinity of under-ground nuclear explosions in which the density of jointsincreases up to hundreds of times above the initial value[Kocharyan et al., 1996]. For this work we adopt afracture frequency of 20 m�1.[29] A number of both theoretical and experimental

studies have been performed on the compression offractures. Experimental studies measure the change inaperture and permeability with changing confining pres-sure for small samples of actual fractures [Cook, 1992].Theoretical studies, on the other hand, produce compli-cated integral formulations for the aperture as a func-tion of pressure, based on the statistical distribution ofasperities along the fracture faces [Brown and Scholz,1986]. The theoretical studies base the geometry of thefracture face on measurements of small samples ofactual fractures, thus both the experimental and theo-retical approaches rely on laboratory-scale fracture sam-ples. However, it has been demonstrated that thepermeability of the Earth’s crust is very strongly scaledependant [Brace, 1984]. Permeability on the scale of alaboratory sample is often orders of magnitude lessthan the large-scale permeability in the same region.Furthermore, it has been demonstrated that flow withinfractures is focused along discrete channels between thefracture faces [Wang et al., 1988]. Hence the fractal-like nature of the fracture face topography and thenonuniformity of the flow within fractures make itunlikely that the hydraulic behavior of small-scalelaboratory samples of natural fractures will be represen-tative of the average behavior of large-scale in situfractures.[30] We model fracture apertures using a novel ap-

proach in which we use data on near-surface fractureapertures to calculate the fracture aperture and compress-ibility at low effective stresses. We then extrapolate thesevalues to high effective stresses using data on thepermeability and fracture frequency at greater depths.[31] Snow [1970] calculated the effective fracture aper-

tures in the near-surface on the basis of permeability andfracture frequency measurements in wells to depths inexcess of 100 m in a variety of rock types, and foundthat the apertures decrease regularly with depth and areessentially independent of host rock type. We plot these

data together on one graph (Figure 2), and fit the datawith an exponential function:

b zð ÞEarth 2� 10�4 mð Þ � e�0:0087 m�1ð Þ�z; ð14Þ

where b and z are both expressed in meters. The exponentialdecrease in aperture implies a constant compressibility ofthe fractures in the near surface. The fracture compressi-bility is then calculated as

bfrac;0 ¼1

Vfrac

� dVfrac

dseff¼ 1

rrg � b zð Þ �db zð Þdz

2:9� 10�7Pa�1

brock ¼1

Vrock

� dVrock

dseff¼ 1

Vrock

� dVfrac

dseff¼ bfrac � Nfrac � bfrac;

ð15Þ

where bfrac,0 is the compressibility of an individual fractureunder low effective stresses, brock is the whole rockcompressibility due to the presence of the fractures, rr isthe rock density, and Vfrac is the volume of the fracture.[32] We estimate fracture apertures at depths greater than

100 m on the Earth based on permeability data by assuminga typical fracture frequency. Seeburger and Zoback [1982]measured fracture frequencies to depths in excess of 1 km inthree geologically distinct areas. On the basis of their data,we consider a range of 0.1 to 1 m�1 as being representativeof the fracture frequency of the Earth’s crust at depths of akilometer or more. Permeabilities obtained in the KTB drillhole of the German Continental Deep Drilling Program areon the order of 10�16 to 10�17 m2 for depths ranging from0.5 to 5 km [Huenges et al., 1997]. Using this range inpermeabilities with the above range in fracture frequencies,equation (12) allows us to calculate the typical fractureaperture at large depths to be on the order of 10�5 to 5 �10�6 m. These minimum effective apertures are in agree-ment with laboratory experiments on the compression offractures [Cook, 1992]. While an aperture of 10 mm seemsexceptionally small to be considered an open fracture, itmust be remembered that this is an effective hydraulicaperture. The actual fracture would have a larger aperturein places and be completely closed at the contact pointsbetween the two faces.

Figure 2. Fracture aperture as a function of depth for avariety of rock types (data of Snow [1970]).

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

6 of 19

E01004

Page 7: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

[33] Fracture compressibility must decrease markedlywith depth, approaching zero so as to agree with therelatively constant permeability at large depths found byHuenges et al. [1997]. This can be justified conceptually byconsidering that on the microscopic level fracture surfaceshave a fractal-like roughness [Wang et al., 1988]. Thus theaperture is just an average value, while the fracture is heldopen by the asperities on the fracture surface. As thefracture is compressed, the opposing faces come intocontact in more places and the resistance to further com-pression is increased proportionately (Figure 3). We applythis conceptual model by assuming a simple exponentialdecrease in the fracture compressibility as a function ofconfining pressure, and fitting it to the surface compress-ibility and the inferred minimum fracture aperture at largedepths:

bfrac seff� �

¼ bfrac;0 exp �seff =sfrac;0� �

bfrac seff� �

¼ b0 � exp �Zseff0

bfrac seff� �

� ds

24

35 ð16Þ

¼ b0 � exp sfrac;0bfrac;0 e�seff =sfrac;0 � 1� h i

;

where sfrac,0 is the scale stress of the exponential decreasein fracture compressibility. Assuming a minimum fractureaperture of 10�5 m at large depth, we find that sfrac,0 isapproximately 10.3 MPa. This aperture function fits boththe near surface aperture data as well as the deeppermeability data. The general character of the plot of thefracture aperture as a function of effective stress in thisstudy (Figure 4) agrees well with both the theoretical andlaboratory studies, while avoiding the limitations of small-scale samples. This approach, while more empirical innature, is likely a better representation of the net behavior ofreal fractures within the crust.[34] In summary, we assume that heavily cratered crust

has a fracture frequency of approximately 20 m�1. Thefracture apertures decrease exponentially with increasingeffective stress at low values of the effective stress, froman uncompressed value of approximately 2 � 10�4 m.Under conditions of high effective stress, the fractureapertures plateau at a value of approximately 10�5 m,while the fracture compressibility decreases to zero. Thefracture aperture and compressibility as expressed in

equation (16) depend only upon the effective stress stateof the aquifer, making these results directly applicable toMars.

4.4. Megaregolith Aquifer Model Synthesis

[35] We use the above description of the hydrologicproperties of impact-induced breccias and fractures tosynthesize an aquifer model based on the record ofimpacts on the surface. As stated earlier, we assume thatthe top 2 km of the crust is composed of a lithified andfractured breccia. Beneath this megaregolith, the crust iscomposed of an equally fractured mixture of bedrock andbreccia. We assume that the effects of superimposedimpacts do not add, as preexisting fractures and faultswould likely be reactivated by subsequent impacts. Thuswe assume that the fracture frequency and breccia distri-bution beneath the crater-saturated Noachian-aged crust isthe same as that found beneath a single isolated crater.Since the largest fraction by area of the crater populationon Noachian aged surfaces is composed of craters withradii in excess of 15 km, there will likely be nosignificant decrease in the fracture frequency and brecciaabundance in the top 15 km of the crust. As will be seenin section 4.5, this likely constitutes the bulk of thehydrologically active crust of Mars. Crater saturation willalso lead to an averaging out of any preferred radial andcircumferential orientations of the impact-induced damagebeneath individual craters. It is thus assumed that crater-saturated terrains are underlain by a uniform and isotropicdistribution of fractures and breccia to a depth of 15 kmor more, with a breccia fraction of 0.25 and a fracturefrequency of 20 m�1. The hydraulic properties of lessdensely cratered terrains could be calculated using thecrater density, though at low crater densities the assump-tions of uniform and isotropic hydraulic parameters wouldbreak down.[36] The porosity in the crust is largely a result of the

presence of breccias, but it also includes smaller compo-nents due to the fracture porosity and the primary porosityof the crystalline rocks:

ntotal z; seff� �

¼ fbrec zð Þ �"

n0;brec � 1� �

� expZseff0

bbrec � ds

0@

1Aþ 1

#

þ Nfracbfrac seff� �

þ 1� fbrec zð Þð Þ� n0;primary � 1� ��

� exp bprimary � seff� �

þ 1�; ð17Þ

where fbrec is the breccia fraction within the crust as afunction of depth (defined as 1.0 in the megaregolith, and0.25 in the partially brecciated bedrock at greater depths),n0,primary is the uncompressed bedrock primary porosity(assumed to be 0.02 [Domenico and Schwartz, 1990]),bprimary is the compressibility of the primary pore space inthe bedrock (assumed to be 10�10 Pa�1 [Domenico andSchwartz, 1990]), and all other parameters are as previouslydefined. The permeability is determined by the fracturefrequency and aperture as described by equations (12) and(16). The net compressibility of the crust depends on thefracture compressibility, the fracture frequency, the com-pressibility of the breccia, the volumetric breccia fraction inthe crust, the primary pore space compressibility in the rock,

Figure 3. Conceptual diagram of the compression of afracture. Individual contact points are drawn as (left)asperities on the fracture surface, or (right) springs,demonstrating the increased resistance to further compressionas more asperities come into contact.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

7 of 19

E01004

Page 8: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

and the compressibility of the water within the pore space(bwater):

btotal ¼ bfrac � Nfrac � bfrac þ fbrec � bbrec þ 1� fbrecð Þ � bprimaryþ n � bwater: ð18Þ

A summary of all parameters used in the equations ispresented in Table 1.[37] The megaregolith aquifer model of this study is

presented in Figure 4 alongside the model of Clifford[1993] and Clifford and Parker [2001], as well as thesandstone model to be developed in section 6. The param-eters for the megaregolith model in Figure 4 are calculatedon the basis of the assumption of hydrostatic pore pressure,though similar plots can be produced for any amount ofexcess pore pressure.[38] In an attempt to quantify the uncertainty in the

megaregolith hydrologic model, we calculate the likely

range of parameter space using a Monte Carlo analysis.We consider the likely range of values for the key inputparameters for which the uncertainty can be estimated(Table 1). Note that for the fracture model, uncertainty isconsidered in the uncompressed fracture aperture and frac-ture frequency only. The fracture compressibility functionarrived at above is constrained by the minimum fractureaperture at great depth, and thus is not independent of theassumed uncompressed aperture. The adopted range inuncompressed fracture aperture leads to the likely range inminimum fracture aperture discussed above without varyingthe fracture compressibility function. We represent the un-certainty in the breccia compressibility function expressed inequation (11) by a scaling factor ranging from 0.3 to 3,encompassing the range in the data of Warren and Trice[1975]. The range in values for each parameter is assumedto correspond to the two-sigma variation. These parametersare then varied normally using a random number genera-tor, and the mean values and standard deviations are

Figure 4. Hydrologic parameters of the Martian crust based on the impact generation of fractures andbreccia within the megaregolith (solid black line), for the sandstone model (gray line) and for the modelof Clifford [1993] and Clifford and Parker [2001] (dashed black line). The estimated one-sigmauncertainty in the megaregolith aquifer model is represented by the shaded gray region.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

8 of 19

E01004

Page 9: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

calculated for the fracture aperture, porosity, permeabilityand compressibility as a function of depth. Values of thepermeability and compressibility are varied and averagedlogarithmically. The one-sigma variations in the hydraulicparameters are represented in Figure 4 by the shaded grayregions.

4.5. Closure of Pore Space and the Base of the Aquifer

[39] Previous studies [Clifford, 1981, 1993] emphasizedthe depth of the base of the aquifer as a means ofcalculating the total volume of fluid that could becontained within the pore space of the Martian crust,and thus constraining the total water storage capacity ofthe crust and the amount of water available for surfaceprocesses. Clifford [1993] assumes that it is the elasticclosure of the pore space with increasing pressure thatdetermines the base of the aquifer. However, using theactual compressibilities of breccias and fractures, we findthat a significant amount of pore space and permeabilityextends to depths much greater than the 8 km depthassumed in that study. Considering elastic compressionalone, the porosity decreases slowly with depth beyondthe base of the megaregolith, and will remain at values ofaround 0.04 to 0.03 until a depth is reached at which theimpacts into the surface had little effect. Similarly,fracture apertures reach a minimum value of approximately10�5 m, and thus continue to be hydraulically conductingeven under conditions of extremely large overburdenpressures.[40] Instead, we propose that an upper limit to the

thickness of aquifers on Mars is imposed by the rheologyof the rocks. The brittle to ductile transition is defined as thedepth at which the stress required to form a new fracture isequal to the frictional resistance on existing fractures[Byerlee, 1968]. The ductile flow in this case is stillcataclastic in nature, and is not relevant to the closure ofthe pore space. The brittle to plastic transition (BPT),which is more relevant to the closure of pore space andfractures within an aquifer [Brace and Kohlstedt, 1980],occurs approximately when the differential stress neces-sary for plastic flow is equal to the effective confiningpressure [Kohlstedt et al., 1995]. We calculate the max-imum differential stress that could be supported in theMartian crust at the end of the heavy bombardment usinga combined thermal and rheological model of the Martiancrust based on the megaregolith model developed above.

[41] The thermal conductivity of the Martian crust iscalculated for both dry and wet conditions. For dry breccias,the thermal conductivity is given by

k ¼ 2:0 � e�n=n0 W �m�1K�1� �

; ð19Þ

where k (W m�1K�1) is the thermal conductivity, n is theporosity, and n0 is 0.0738 [Warren and Rasmussen, 1987].For saturated breccias, the thermal conductivity is calculatedby the harmonic mean of the thermal conductivities of rockand water:

k ¼ kr � kwn � kr þ 1� nð Þ � kw

; ð20Þ

where kr and kw are the thermal conductivities of rock andwater, taken to be 3 W m�1K�1 and 0.57 W m�1K�1

respectively [Demming, 2002]. In the partially brecciatedbasement crust, the thermal conductivity is likely inter-mediate between the harmonic and arithmetic means of theconductivities of the breccia component and solid rock. Forthe saturated crust, the harmonic and arithmetic means yieldessentially the same result due to the similar thermalconductivities of rock and saturated breccia. At depthsgreater than the depth of the BPT, it is assumed that the porespace has closed and the thermal conductivity is set equal tothat of solid rock.[42] The heat flux into the base of the lithosphere at the

end of the heavy bombardment is taken to be 65 mW m�2

[Hauck and Phillips, 2002], and the surface temperature isset to be 273 K on the basis of the assumption of a warmerand wetter early Mars. If surface temperatures were cooler,as is suggested by some [e.g., Squyres and Kasting, 1994],the brittle to plastic transition would be pushed deeperwithin the crust, increasing the thickness of potentialMartian aquifers.[43] The rheology of the Martian crust is modeled after

studies of the rheology of diabase in dislocation creep:

_e ¼ A � sn exp �Ea=RTð Þ; ð21Þ

where A, n, and Ea are constants that take on differentvalues for dry [Mackwell et al., 1998] or wet [Caristan,1982] diabase. For the case of a dry crust, it is clearlyappropriate to use the rheology of dry diabase, but it is

Table 1. Summary of Parameters Used in Modelsa

Parameter Symbol Value Range

Uncompressed fracture aperture b0 2 � 10�4 m 1–3 � 10�4 mSurface fracture compressibility bfrac,0 2.9 � 10�7 Pa�1

Scale-stress of exponential decrease in fracture compressibility sfrac,0 10.3 MPaFracture frequency Nfrac 20 m�1 10–30 m�1

Regolith thickness zreg 2 km 1–3 kmBreccia fraction within crust fbrec 1.0 for z < zreg, 0.25 for z > zreg 0.15–0.35Bedrock compressibility bprimary 1 � 10�10 Pa�1 0.5–3 � 10�10 Pa�1

Water compressibility bwater 4.4 � 10�10 Pa�1

Uncompressed bedrock porosity n0,primary 0.02Uncompressed breccia porosity n0,brec 0.15 0.05–0.25Rock density rrock 2700 kg m�3

Water density rwater 1000 kg m�3

aThe likely range in values is given for the parameters varied in the Monte Carlo analysis.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

9 of 19

E01004

Page 10: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

unclear which rheology is more appropriate for the case ofwater saturated pore space. For pore pressures greater thanthe hydrostatic pressures, the water fugacity is likely to belarge and the wet rheology may be appropriate. However,experiments on wet diabase experienced partial meltingupon release of the water, which would have resulted in anoverly weak rheology [Mackwell et al., 1998]. Thus therheologies of both dry and wet diabase are plotted for thecase of the saturated crust, while only the dry rheology isplotted for the case of unsaturated crust.[44] Figure 5 shows that the BPT occurs between depths

of 6 and 10 km for the dry crust case, and between depths of17 and 26 km for the case of saturated crust. The 26 kmdepth is likely an overestimate, as the dry rheology assumedis probably stronger than the wet Martian crustal rocks.Furthermore, even for the case of a wet early Mars, an upperdry layer of regolith may have formed an insulating blanketon the surface, thus decreasing the depth to the BPT.However, hydrothermal circulation of groundwater in theMartian crust [Travis et al., 2003] would lead to a substan-tial advective heat flux, thereby decreasing the conductivetemperature gradient and increasing the depth to the BPT.

Similarly the 6 km depth is likely an underestimate, as it isunlikely that no water existed in the crustal pore space onearly Mars. The unrealistic assumption of a fully dryMartian crust results in excessively low thermal conductiv-ities and high temperatures at shallow depth, leading topartial melting at depths as shallow as 30 km underNoachian conditions. The approximate range of likelydepths for the BPT at the end of the heavy bombardment,and thus the upper limit for the thickness of ancient Martianaquifers, is between 10 and 20 km. However, in regions ofcontinued geologic activity and beneath young impactcraters it is possible that open pore space has been createdat greater depths later in Mars history when the BPT waspushed deeper by the lower surface temperatures anddecreased heat flux.[45] While the brittle to plastic transition imposes an

upper limit on the thickness of aquifers on Mars, otherprocesses can act to close off the pore space at shallowerdepths. Terrestrial observations demonstrate the importanceof pressure solution in the closure of both fractures andprimary pore space in the presence of a fluid [Renard et al.,2000]. In short, this results from preferential dissolution at

Figure 5. Models of yield stress, temperature, and thermal conductivity for the Martian crust at the endof the heavy bombardment, assuming a surface temperature of 273 K and a heat flux of 65 mW/m2 for(top) dry and (bottom) saturated pore space. Shown on the stress plots are (1) the effective stress, seff, thelithostatic pressure minus the pore pressure, and (2) and the ductile yield stress for diabase under variousconditions and a strain rate of 10�15 s�1. For the dry case, only the dry diabase rheology is used[Mackwell et al., 1998], and thermal conductivities of rock/breccia mixtures are calculated using theharmonic (h) and arithmetic (a) means. For the wet case, both dry and wet [Caristan, 1982] diabaserheologies are used, and thermal conductivities of rock/breccia mixtures are calculated with the harmonicmean. The intersection of seff with the ductile yield stress defines the brittle-plastic transition (BPT;indicated by arrows) [Kohlstedt et al., 1995], here used to define the depth of pore space closure.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

10 of 19

E01004

Page 11: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

the contact points of grains and fracture surfaces, followedby preferential precipitation in the void spaces. This is atime dependent process that is strongly sensitive to thetemperature, pressure, rock composition, and the micro-scopic geometry of both the contacts and void spaces.Furthermore, fractures and fault breccias exhibit markedlydifferent behaviors with regards to their susceptibility topressure solution. The process is poorly constrained interrestrial systems, and cannot be reliably applied to Mars.Additionally, hydrothermal circulation within the crust[Travis et al., 2003] could facilitate dissolution and mineralprecipitation, closing off the pore space and emplacingveins within the fractures.[46] Claims regarding the total water storage capacity of

the Martian crust will be limited by this uncertainty in thedepth of closure of the pore space. We use the megaregolithmodel above to estimate the total pore volume of theMartian crust for a range of aquifer thicknesses. For aquiferthicknesses of 5, 10, and 20 km we calculate total porevolumes of 6.1, 9.8, and 17 � 107 km3 respectively. Thepore space present in basaltic aquifers within the Tharsisrise and the northern plains must also be considered. Thetotal volume of Tharsis basalts has been estimated at 3 �108 km3 [Phillips et al., 2001]. Taking an average basaltporosity of 0.1 (see section 5.1), this would give a total porevolume of 3 � 107 km3. For a more conservative estimate,we also consider cases in which there is significant porosityonly in the top 2 to 5 km of the crust in this region, as couldresult from closure of the pore space at depth or if asignificant amount of low-porosity volcanic intrusions con-tributed to the Tharsis rise, leading to total pore volumes of2.5 to 6.3 � 106 km3. Representing the northern plains as a1 km thick layer of porous basalt, we find a total porevolume of 4.8 � 106 km3 above the underlying megarego-lith. Using the above ranges in values, we get low and highestimates of the total pore volume of 6.8 � 107 and 2.0 �108 km3, with an intermediate estimate of 1.1 � 108 km3.These values are in agreement with the estimates of Clifford[1981], despite the differences in the hydrologic modelused. However, we again caution the reader with respectto the great uncertainties in these estimates.

5. Alternative Aquifer Types

5.1. Basaltic Aquifer Model

[47] While the megaregolith aquifer model is likelyrepresentative of much of the Noachian-aged highlands,the younger surfaces that predominate in other areas of theplanet have not been significantly modified by impacts. TheTharsis and Elysium volcanic provinces are composed ofthick sequences of basalt flows, the upper portions of whichare young enough to have escaped the heavy bombardmentof the early Noachian era. The northern plains, as well as anumber of other areas across the planet, are likely composedof layers of basaltic lava flows and sediments overlyingmore ancient cratered terrain beneath [Frey et al., 2002;Head et al., 2002].[48] Terrestrial basaltic aquifers are characterized by both

high porosity and high permeability. While there is greatvariability in the hydraulic properties of terrestrial basalts, itis likely that the average properties of terrestrial and Martianbasalts are similar. Over small spatial scales, it is necessary to

model basaltic aquifers using a stochastic approach to repre-sent the significant heterogeneity of the aquifers [Welhan etal., 2002]. However, over the large spatial scales commonlyof interest onMars, this lateral heterogeneity averages out andit is appropriate to use the bulk properties.[49] The hydraulic properties of the Snake River Plains

aquifer summarized in studies by Gego et al. [2002] andWelhan et al. [2002] are reviewed here. A basaltic aquifer ismade up of a large number of individual lava flows,typically on the order of 1 to 10 meters thick. Each flowis composed of a thick layer of massive basalt (84% byvolume), overlain by a thinner interflow zone (12% byvolume) [Gego et al., 2002]. These interflow zones are bothvery porous and heavily fractured, as a result of thermalcooling of the outer surface of the flow and rotationalstresses at the lava flow front. The interflow zones havean average porosity of 0.22 and an average permeability ofapproximately 10�8 to 10�10 m2. The intervening layers ofmassive basalt, on the other hand, have an average porosityof 0.09 and permeability of approximately 10�11 to 10�13 m2.The average porosity of a basaltic aquifer is then calculated tobe 0.10. The average horizontal and vertical permeabilitiescan be calculated using the arithmetic and harmonic means,yielding values of approximately 10�9 to 10�11 m2 and 10�11

to 10�13 m2 respectively. If there is sufficient time betweenindividual eruptions, a layer of sediments will accumulateabove each lava flow, which on Mars could include impactejecta as well as aeolian and fluvial sediments. These sedi-ments can act to further increase the porosity and permeabil-ity, or alternatively they may fill the fractures in theunderlying interflow zone and result in a net decrease in theporosity and permeability.[50] There is not enough data to accurately constrain the

behavior of basaltic aquifers under varying conditions ofconfining stress and pore fluid pressure, making the basalticaquifer model of limited applicability. While some of thepermeability is due to simple fractures, much of it is due tothe presence of larger interconnected cavities within theinterflow zones. The permeability is likely less sensitive tothe effective stress than was found for the megaregolithmodel due to the presence of these larger cavities, though itis unlikely that the high values measured at the surfacepersist to depths of several kilometers as some compressionis inevitable.[51] In a study of basaltic aquifers in the Oregon

Cascades, Saar and Manga [2004] modeled the permeabil-ity as a function of depth based on a number of indirectapproaches, including spring discharges, thermal profiles,magmatic intrusion rates, and seasonal seismicity. Theyrepresent the permeability as following an exponentialdecrease with depth in the top 800 m, followed by a powerlaw relationship similar to that of Manning and Ingebritsen[1999] at greater depths. The horizontal permeability rangesfrom 10�9 to 10�13 m2 at the surface, to 10�15 to 10�18 m2

at a depth of 10 km. They propose that the decrease inpermeability with depth is exclusively due to mineralprecipitation, though we would suggest that the elasticcompression of the fractures plays a significant role as well.Manga [2004] suggests that lower water to rock ratios onMars would lead to negligible mineral precipitation and themaintenance of the high surface permeability values even atgreat depth. However, as discussed earlier, pressure solution

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

11 of 19

E01004

Page 12: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

and hydrothermal circulation are likely important in theclosure of pore space and the decrease in permeability withdepth on Mars. While the Saar and Manga [2004] studysheds some light on the variation of permeability with depthin basaltic aquifers, the methods of calculating the perme-ability were indirect and there are large uncertainties in thenumbers. There still remains much work to be done to fullyunderstand the dependence of the permeability, porosity,and compressibility on the depth and pore pressure inbasaltic aquifers on both Earth and Mars.

5.2. Sedimentary Aquifers

[52] A number of studies have emphasized the possibilityof extensive sedimentary deposits on Mars [Goldspiel andSquyres, 1991; Malin and Edgett, 2000b]. Sedimentaryrocks can display a wide range of hydraulic behaviors,and a comprehensive treatment of the topic is beyond thescope of the present study. We here aim only to outline thebasic properties and behavior in such a way as to allow forthe production of a generalized model.[53] The porosity of sandstone at the surface is generally

in the range of 0.05 to 0.3 [Domenico and Schwartz, 1990],and is commonly modeled as decreasing exponentially withdepth [Athy, 1930; Chapman et al., 1984]:

n zð Þ ¼ n0e�z=z0 ; ð22Þ

where n0 is the surface porosity and z0 is the scale depth.Chapman et al. [1984] used values of 0.25 for n0, and 3 kmfor z0. It is significant that this exponential decrease inporosity cannot be attributed to the simple elastic compres-sion of the pore space. Laboratory studies indicate that theporosity in sandstone decreases elastically muchmore slowlywith increasing effective stress than would be predicted onthe basis of the observed variation of the porosity with depth[Demming, 2002; Neuzil, 1986]. The rapid decrease inporosity with depth observed in sandstones is likely due tomineral precipitation within the pore space or perhaps to along-term viscoelastic response, and is more dependent uponthe thermal history than the effective stress [Neuzil, 1986].The elastic response of the pore space to changes in the fluidpressure is much less significant than the variation of theporosity with depth, amounting to an increase in porosity of0.015 with an increase in pore pressure of 10 MPa. As amodel of sandstone aquifers on Mars, we adopt the formulaof Chapman et al. [1984] and neglect the change in porositywith changing effective stress state. We do not scale the valueof z0 toMars gravity, since the decrease in porosity with depthis not due to elastic compression.While it is unknown exactlywhat processes are responsible for the decrease in porositywith depth in terrestrial sandstones and how this would scaleto Mars, the above relationship should capture the basicbehavior of Martian sandstones. On the basis of the summaryof Neuzil [1986], we represent the compressibility as anexponentially decreasing function of the effective stress:

b seff� �

¼ b0;sed � exp �seff =s0;sed� �

; ð23Þ

where we assume values for b0,sed of approximately10�9 Pa�1, and s0,sed of 25 MPa.[54] The permeability of a sedimentary rock is dependent

upon the porosity, grain size, and degree of sorting of the

particles [Nelson, 1994]. It is commonly assumed that thepermeability is proportional to the log of the porosity[Archie, 1950; Demming, 2002], but no one relationshipbetween permeability and porosity can be said to berepresentative of all sandstones or shales. On the basis ofthe data of Nelson [1994], we represent the log of thepermeability as being proportional to the porosity:

log10 kð Þ ¼ �17:3þ 25 � n; ð24Þ

where k is the permeability in m2, and n is the porosity. Theporosity, permeability, and compressibility of the sandstoneaquifer model are represented in Figure 4 alongside themegaregolith aquifer model. An important differencebetween the sandstone and megaregolith aquifers is nowapparent. In the megaregolith model the permeability isdependent on the elastic closure of the fractures and scalesdirectly with the effective stress state, whereas in thesandstone model the porosity and thus the permeability aremuch less sensitive to the effective stress state.

5.3. Mixed-Media Aquifers

[55] Single component aquifer models may be reasonablerepresentations of certain regions of the Martian crust. Forexample the upper portion of the Tharsis region is com-posed largely of basalts, while the deep crust in the heavilycratered southern highlands can be represented by themegaregolith model. However, in many cases it may benecessary to consider the effects of a mixture of twocomponents, such as sediments or basaltic lava flowsinterlayered within or overlying a megaregolith.[56] For the case of isotropic, intimate mixtures of dif-

ferent components, the permeabilities can be averaged usingthe geometric mean [Demming, 2002]:

keff ¼Y

kfii

� ; ð25Þ

where ki and fi are the permeability and volumetric fractionof the ith component, respectively. Alternatively, for thecase of horizontal layering of aquifer components, thepermeability is anisotropic. The horizontal permeability isfound by the arithmetic mean, while the vertical perme-ability is calculated using the harmonic mean:

khor ¼X

fi � kið Þ

kvert ¼X

fi=kið Þh i�1

:

ð26Þ

The effective compressibility for mixed media aquifers canbe found by a simple arithmetic mean of the compressi-bilities of the components.

6. Application: Aquifer Pressurization ThroughClimate Change

6.1. Background

[57] The outflow channels are commonly thought to haveformed when the fluid pore pressure within a confinedaquifer reached or exceeded the lithostatic pressure [Carr,1979]. Various studies ascribe these fluid pressures to be the

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

12 of 19

E01004

Page 13: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

result of (1) a perched aquifer [Carr, 1979, 1996], (2) aconfined aquifer pressurized by a downward propagatingfreezing front [Carr, 1979], (3) artesian pressures caused bythe proposed uplift of Tharsis [Clifford and Parker, 2001],(4) the intrusion of magma into ice rich crust [McKenzie andNimmo, 1999; Squyres et al., 1987], (5) the decompositionof gas hydrates [Milton, 1974], and (6) the coalescence ofgroundwater flow in the tectonic fractures surroundingTharsis [Baker et al., 1991]. While many of these ideasare conceptually valid, it has not been possible to adequatelytest them without a complete hydrologic model of theMartian crust.[58] We here apply our hydrologic model to investigate the

possibility that the fluid pressures necessary to form theoutflow channels could be generated beneath a downwardpropagating freezing front as the result of a rapid cooling ofthe climate, as proposed by Carr [1979]. As the aquiferprogressively freezes, the volumetric expansion of wateraccompanying the phase change would pressurize theremaining aquifer beneath. The premise of this hypothesisrelies on a dramatic cooling of the Martian climate to itscurrent cold state. Mars’ climate evolution is still poorlyunderstood, but there is clear evidence that the surface ofMars was both warmer and wetter during the Noachianepoch, including the high drainage densities of the valleynetworks [Baker, 1982; Hynek and Phillips, 2003] and theinferred high erosion rates [Carr, 1992]. The climatic evolu-tion was likely driven, to a large degree, by the loss of an earlythick greenhouse atmosphere to impact erosion, solar windsputtering, adsorption of CO2 onto the regolith, and thepossible precipitation of carbonates [Brain and Jakosky,1998; Melosh and Vickery, 1989]. Alternatively, other work-ers propose that theMartian climate may have been subject tomore rapid, episodic climate changes associated with redis-tribution of the volatile inventory [Baker et al., 1991].[59] For perfectly confined aquifers, arbitrarily large pres-

sures could be generated by the growth of the cryosphere,potentially leading to catastrophic breakout and the formationof the outflow channels [Carr, 1979]. However, overlarge spatial and temporal scales, all aquifers are likelyinterconnected to some degree [Clifford, 1993]. This inter-connectedness would be even more pronounced in the highlyfractured megaregolith of the Martian southern highlands, aswell as in the thick sequences of basalt in the Tharsis rise.Thus it seems likely that the entire circum-Chryse region wasunderlain by an extensive aquifer that would have beencontinuous with the high topography of both the Tharsis riseand the southern highlands. It is highly unlikely that this largeregional aquifer was fully confined. Rather, it would havebeen locally confined from above by the cryosphere, butlaterally continuous with unconfined aquifers in the Tharsisregion and the southern highlands. Such a partially confinedaquifer can be brought about by means other than topo-graphic variations as well, as illustrated in Figure 6. In theabsence of significant topographic relief (Figure 6a),locally confined aquifers that are laterally continuous withunconfined aquifers can be produced by variations in thethickness of the cryosphere (Figure 6b) as a result oflateral variations in either the thermal conductivity of thesurface materials or the local heat flux. The same outcomecould also be produced by variations in the height of thewater table (Figure 6c), which could result, for example,

from hydrothermal convection or a localized pressurizationmechanism operating within or beneath the aquifer.[60] Thus the thickening cryosphere is not simply com-

pressing a closed aquifer. Rather, as pressure is generatedbeneath the freezing front, it simultaneously diffuses awaytoward the edges of the confined portion of the aquifer. Theactual pressure generated depends upon the rate at which thefreezing front advances, the area over which the aquifer isconfined, and the hydraulic properties of the aquifer. Ifeither the confined portion of the aquifer is small or thethickening of the cryosphere is a slow process, the pressureshould diffuse away essentially as fast as it is generated.Alternatively, if the confined portion of the aquifer issufficiently large or the cryosphere thickens rapidly, signif-icant pressures should be generated, possibly approachingthose generated within a fully confined aquifer.

6.2. Pressurization of Aquifers: Model Description

[61] In order to test the feasibility of the thickening of thecryosphere as a means of producing high pressures inMartian aquifers, we consider the hydrologic response tothe cryospheric thickening for a number of climate changescenarios and aquifer geometries. The time dependantthickening of the cryosphere is modeled using a fullyexplicit finite difference code. The temperature evolutionwithin the crust is modeled using the 1-D heat equation:

@T

@t¼ 1

rc@

@zk@T

@z

� �; ð27Þ

where k is the thermal conductivity and c is the specificheat. The rate of change of the depth of the melting isotherm

Figure 6. Aquifer and cryosphere geometries that can leadto locally confined aquifers in lateral continuity withunconfined aquifers, viewed in cross section. Localconfinement of the aquifer can be brought about byvariations in (a) the topography, (b) the cryospherethickness, or (c) the elevation of the water table, asdescribed in the text.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

13 of 19

E01004

Page 14: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

(zm) is calculated using the difference between thetemperature gradient above and beneath the freezing front:

@zm@t

¼ 1

nrLk1@T

@z

����þ� k2

@T

@z

�����

� �; ð28Þ

where n is the porosity, L is the latent heat of fusion ofwater, r is the density of water, k1 is the thermalconductivity above the freezing front, and k2 is thethermal conductivity below the freezing front. Forsimplicity, we assume a fully ice-saturated cryosphere,though it is likely that the ice content of the pore spacewill decrease to zero toward the surface at a steady stateice table [Mellon et al., 1997]. The thermal conductivitiesare modeled as in Section 4.5.[62] Once the thickness of the cryosphere as a function

of time is computed, we model the pressure that would begenerated beneath the growing cryosphere as a function oftime for a given aquifer geometry. The pressure generatedat the freezing front is found by considering a finite-radiusaquifer of thickness D beneath the freezing front. As thefreezing front moves an incremental distance dz into thisaquifer, the expansion of the water as it freezes to form iceexpels an amount of water equivalent to a layer ofthickness n(av � 1) dz, where av, the coefficient ofvolumetric expansion of water upon freezing, is approxi-mately 1.09, and n is the porosity. This water is forcedinto the aquifer beneath the freezing front, resulting in apressure increase of DP:

DP ¼ 1

bln

D� dzþ n � dz av � 1ð ÞD� dz

� �

1

bn � dz av � 1ð Þ

D� dz; ð29Þ

where b is the combined compressibility of the aquifermatrix and water.[63] Since the freezing front moves slowly relative to

the vertical diffusion of the excess hydraulic head, andsince the vertical length scale of the aquifer is severalorders of magnitude less than the horizontal, the aquiferproperties are vertically averaged and diffusion is consid-ered in the horizontal direction only (i.e., @h/@z = 0). Theevolution of the hydraulic head produced by this excesspressure in the confined portions of the aquifer iscalculated using a finite difference code representing theelastic response of the aquifer as governed by theconsolidation equation (equation (5)) expressed in axisym-metric cylindrical coordinates with a pressure source term:

@h

@t¼ 1

Ss

1

r

@

@r� K � r @h

@r

� �þ 1

rwg@P

@t

� �cryos

; ð30Þ

where (@P/@t)cryo is the rate of pressurization at the freezingfront from equation (29). In the unconfined portions of theaquifer, the net flow of water into a column of aquiferresults in an increase in the hydraulic head through a directincrease in the water table height, rather than through aconfined pressurization. It is also necessary to account forthe fact that the aquifer thickness now varies with radius dueto the increase in the water table height. If the hydraulic

head is taken with respect to the base of the aquifer, suchthat it is equal to the unconfined aquifer thickness, then flowis modeled by

@h

@t¼ 1

r � ntop@

@r� K � h � r @h

@r

� �; ð31Þ

where ntop is the porosity at the water table.

6.3. Pressurization of Aquifers: Model Results

[64] We consider instantaneous climate changes from270 K to 220 K, and from 250 K to 220 K, representing atransition from the warm wet climate conditions on earlyMars to the cold conditions of today in a single rapid event.We also consider more plausible gradual climate changescenarios, with the temperature changing from 270 K to220 K over the course of 60 ka and 1 Ma. The former casecorresponds to half of the period of an obliquity oscillation,while the latter case is a somewhat more conservativetimescale for climate change. The resulting cryospherethickness as a function of time for the four scenarios isshown in Figure 7. The cryosphere thickness is sensitive tothe thickness of ice-free regolith above the permafrost, hereassumed to be negligible, and the geothermal heat flow, hereassumed to be 50 mW/m2. This heat flux is appropriate forthe Hesparian epoch on Mars [Hauck and Phillips, 2002], inwhich most of the outflow channels formed [Baker, 1982;Carr and Clow, 1981; Masursky et al., 1977].[65] We use a generic aquifer geometry in which a

circular region of the aquifer is confined from above bythe thickening cryosphere, but the adjacent aquifer isunconfined. Our nominal model consists of an aquifer witha confined region radius (R) of 1000 km and an aquiferthickness (D) of 3 km. This aquifer thickness is roughlyequal to the total change in cryosphere thickness during thesimulation. We also consider confined regions with radiiranging from 500 to 3000 km, as well as initial aquiferthickness ranging from 1 to 5 km. The megaregolith aquifermodel is representative of several of the outflow channelsource regions to the east of Tharsis, while those further upon the Tharsis rise may be better represented by the basalticaquifer model. Due to the difficulty constraining the hy-draulic behavior of basaltic aquifers under varying condi-tions of lithostatic and fluid pore pressure, we limit thefollowing investigation to the megaregolith aquifer model.Lower pore pressures would be expected for basaltic aqui-fers due to their higher permeability relative to megaregolithaquifers, thus the pore pressures generated in our models arelikely upper limits.[66] Figure 8 plots the aquifer pore pressure at the base of

the cryosphere in the center of the confined region as afunction of time for the aquifer geometries and climatechange scenarios described above. Also included in thefigure is the lithostatic pressure at the base of the cryo-sphere, which increases with time as the cryosphere thick-ens. For all scenarios, the pore pressure rises steeply withtime during the early period of rapid cryosphere thickening.As the rate of propagation of the freezing front slows, thepore pressures plateau out. After the freezing front passesthrough the base of the megaregolith, a smaller volume ofwater is expelled from the freezing front per unit increase incryosphere thickness due to the decrease in porosity. As a

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

14 of 19

E01004

Page 15: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

result, the rate of pressurization decreases dramaticallyrelative to the rate of diffusion within the aquifer, and thepore pressure drops.[67] For the instantaneous climate change from 270 K to

220 K, most scenarios result in maximum pore pressuresapproaching, but not exceeding, the lithostatic pressure(Figures 8a and 8b). As the fluid pore pressure approachesthe lithostatic pressure, the fracture apertures rapidly increase(equation (16)), resulting in a sharp rise of the permeability(equation (12)). This allows for a more rapid lateral diffusionof the pressure away from the confined portion of the aquifer,producing a negative feedback cycle that prevents fluidpressures from exceeding the lithostatic pressure. For theconfined aquifers with radii up to 2000 km, the pore pressuresapproach the lithostatic pressure during the early period ofrapid cryospheric thickening, before plateauing at lowervalues as the rate of thickening decreases. For the locallyconfined aquifer with a radius of 3000 km, pore pressuresexceed the lithostatic pressure during this early period byapproximately 0.4MPa. However, as the rate of thickening ofthe cryosphere decreases, the pressure diffuses away morerapidly than it builds up and the aquifer quickly returns tosublithostatic pore pressures.[68] For more plausible climate change scenarios, the

pore pressure at the base of the cryosphere never exceeds

the lithostatic pressure. For the change in surface tempera-ture from 270 to 220 K over the course of 60 ka, the porepressure for the 3000 km radius confined aquifer reaches,but does not exceed, the lithostatic pressure (Figures 8c and8d). Smaller confined aquifers result in lower pore pres-sures. For the change in surface temperature from 270 K to220 K over the course of 1 Ma, the pore pressure remainswell below the lithostatic pressure for all confined aquifersizes (Figure 8e). During these slower climate changes, thepore pressure has ample time to diffuse away from theslowly advancing freezing front toward the unconfinedportions of the aquifer. For an instantaneous change from250 K to 220 K, the greater initial cryosphere thicknessresults in a greater thermal diffusion timescale at the base ofthe cryosphere and a slower initial rate of cryosphericthickening, leading to the generation of significantly lowerpressures within the aquifer (Figure 8f). These more plau-sible scenarios produce pore pressures significantly lessthan the lithostatic pressure at all times.[69] The pressures generated are highly sensitive to the

original thickness of the aquifer (Figures 8g and 8h). For aninitial aquifer thickness of 5 km and confined region radiusof 1000 km, the pore pressures generated are only slightlyless than those for the 3 km thick aquifer. The water beingexpelled by the freezing front is being injected into a thickeraquifer beneath, thus significantly lower pressures might beexpected. However, since the compressibility and perme-ability decrease with depth, the ability of the aquifer toaccommodate and transport the additional water decreaseswith depth as well. On the other hand, when the aquifer isthinner, such that the freezing front passes through the baseof the aquifer, significantly higher pressures are generated.During the early stage of rapid cryosphere thickening, thepore pressures for a 2 km thick aquifer approach, but do notexceed, the lithostatic pressure, while those for a 1 km thickaquifer exceed the lithostatic pressure by nearly 1 MPa.Then, when the freezing front approaches and passesthrough the base of the aquifer, the pore pressures riseasymptotically as the volume of the pore space into whichthe excess water is being injected decreases to zero. Forboth the 1 and 2 km thick aquifers, the pore pressures don’tbegin the final asymptotic rise until the freezing front passeswithin a few tens of meters to 100 m of the base of theaquifer. Large outflow events in these cases would beimpossible, since the large volumes of water inferred forthe outflow channel floods could not be stored in such thinaquifers.[70] In order to highlight the importance of the depen-

dence of the hydraulic parameters on the fluid pore pressure,we have included one scenario in which the aquifer prop-erties are artificially held constant in time and not allowed to

Figure 8. Aquifer pore pressure and lithostatic pressure at the base of the cryosphere as a function of time for severaldifferent climate change scenarios and aquifer geometries, including (a and b) an instantaneous change from 270 to 220 Kfor 3 km thick confined aquifers with radii between 500 and 3000 km; (a and b, dotted curve) a 1000 km radius confinedarea in which the aquifer properties are modeled as being insensitive to the pore pressure; (c and d) a linear change from270 to 220 K over the course of 60 ka for a 3 km initial aquifer thickness and radii between 500 and 3000 km; (e) a linearchange from 270 to 220 K over the course of 1 Ma for a 3 km initial aquifer thickness and radii between 500 and 3000 km;(f) an instantaneous change from 250 to 220 K for a 3 km initial aquifer thickness and radii between 500 and 3000 km;(g and h) an instantaneous change from 270 K to 220 K for a 1000 km radius confined region with initial aquiferthicknesses between 1 km and 5 km.

Figure 7. Model of the depth of the base of the cryosphereas a function of time as the average surface temperaturechanges from 270 to 220 K either instantaneously (solidblack line) or linearly over a course of 1 Ma (solid gray line)or 60 ka (dashed black). Dashed gray line shows responsefor an instantaneous change from 250 to 220 K.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

15 of 19

E01004

Page 16: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

Figure 8

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

16 of 19

E01004

Page 17: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

adjust to the changing fluid pore pressure (dotted line inFigures 8a and 8b). In this case, the pore pressure risesmuch more rapidly and exceeds the lithostatic pressure by asubstantial margin. Similarly large pressures could beachieved in an unfractured sedimentary aquifer of sufficientvertical and horizontal extent, however there is no evidencethat such an aquifer exists in the region of the outflowchannel source regions.[71] Similar simulations have been run investigating the

climate change scenarios and aquifer geometries describedabove over a wider range of parameter space in thehydrologic model. Model results are most sensitive to theuncertainty in the fracture aperture, due to the cubicdependence of the permeability on this parameter. Usingan uncompressed fracture aperture of 10�4 m, at the lowestend of the plausible range, super-lithostatic pore pressurescan be generated for several of the model scenarios,including the temperature changes from 270 to 220 K bothinstantaneously and over 60 ka for confined aquifer radiigreater than 1000 km. However, the high discharges in-ferred for the outflow channel floods are suggestive ofhigher than average permeability, rather than lower.[72] The circum-Chryse region, in which most of the

outflow channels occur, is similar in size to the 2000 kmradius confined aquifer modeled above. Thus it would seemthat if the aquifer in the region were confined over an areasomewhat greater than the area of the Chryse basin (e.g.,over a 3000 km radius area), it would be possible to producesuper-lithostatic pore pressures during an instantaneouschange in surface temperature from 270 K to 220 K.However, the model pore pressures peak in the center ofthe confined region and decrease toward the periphery,while the outflow channel source regions are all along themargin of the circum-Chryse region. Thus for such ahypothetical confined aquifer in this region, the super-lithostatic pore pressures would be generated in the centerof the Chryse basin, and not where the outflow channelsource regions are located.[73] If the effects of topographic variations were to be

considered, then an extra component of hydrostatic pressurewould be added to the pore pressure in the aquifer beneathtopographic lows. However, many of the source regions ofthe circum-Chryse outflow channels are relatively high upon the Tharsis bulge and show no evidence for a preferentialoccurrence in local topographic lows. If the aquifer pres-surization mechanism acted over the entire circum-Chryseregion, the maximum hydrostatic component of the pres-sure would occur in the topographically lower center ofthe Chryse basin, enhancing the higher pressures beingproduced there. Thus the location of the outflow sourceregions at higher elevations on the margins of the Chrysebasin suggests a more localized pressurization mechanismfocused at the individual source regions.[74] A further argument against a climatic mechanism as

the driving force behind the outflow channel floods lies inthe observations that the channels have ages ranging fromHesparian to early Amazonian, a time period spanningroughly 1 billion years [Carr and Clow, 1981; Tanakaand Skinner, 2004], and that several of the outflow channelsshow evidence for multiple episodes of flow [Hanna andPhillips, 2003; Williams et al., 2000]. Thus it would benecessary to have repeated near-instantaneous climate

changes from 270 to 220 K recurring over a period ofroughly 1 Ga to explain the temporal distribution of outflowchannel events. While a number of workers have argued forthe importance of episodic climate changes on Mars [Bakeret al., 1991; Head et al., 2003], temperature swings as largeand rapid as are required to form the outflow channelsthrough this mechanism seem unlikely. Mechanisms com-monly invoked for secular climate change, such as erosionof the atmosphere by impacts and solar wind stripping[Brain and Jakosky, 1998] act over longer timescales.Gulick et al. [1997] found that the injection of 1 to 2 barsof CO2 into the Martian atmosphere resulted in a temper-ature increase to only 250 K, however the subsequentcooling took place over hundreds of Ma. Rapid climatechanges attributed to obliquity oscillations will be limitedby the CO2 content of the polar caps, which has been shownto be on the order of tens to hundreds of mbar [Mellon,1996], much less than that required for a strong greenhouseeffect.[75] In summary, the difficulty with producing super-

lithostatic pore pressures through a climate change lies bothin the slow diffusion of the thermal wave into the crust andin the negative feedback cycle between the pore pressureand the permeability. This effect would be magnified forbasaltic aquifers, which have a significantly higher perme-ability than that in the megaregolith model. These resultshave implications for any mechanism of aquifer pressuriza-tion that relies upon processes acting over timescales of tensof thousands to millions of years. The work presented heresuggests that a more rapid, repeatable, and localized mech-anism for pressure generation would be favored as a drivingforce behind the formation of the outflow channels, such asa magmatic intrusion into the aquifer or a tectonic event.[76] Due to the unconstrained nature of the problem, it is

of course not possible to fully prove or disprove the theorythat the outflow channels were generated by the pressuri-zation of an aquifer beneath a downward propagatingfreezing front. However, this study sheds some light onthe details and problems of this theory. It does not appearpossible to produce super-lithostatic pore pressures solelythrough the freezing of a megaregolith-type aquifer forplausible climate change scenarios and aquifer geometries.This does not rule out the possibility that this pressurizationmechanism could have produced the requisite pressures inconcert with another mechanism. For example, this mech-anism may have played a secondary role by causing ageneral increase in aquifer pore pressures to sublithostaticlevels over a large region, thereby allowing a more localizedand rapid pressurization mechanism to push pore pressuresto super-lithostatic levels at the outflow channel sources.

7. Conclusions

[77] We have presented here a generalized hydrologicmodel of the Martian crust based on a combination ofterrestrial and lunar analogs, in which the relevant hydraulicproperties are represented in a physically realistic and self-consistent manner. Much of the hydrologically active crustof Mars can be represented by the megaregolith aquifermodel, based on the inferred effects of impacts on thehydrologic properties of the crust. Younger terrains arebetter represented as either basalts, such as the topmost

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

17 of 19

E01004

Page 18: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

layers in the Tharsis rise, or sediments, such as the surfacematerials within craters and elsewhere [Malin and Edgett,2000b]. The hydrologic properties of all of these crustalmaterials will vary with both depth and pore pressure.[78] The sensitivity of the hydraulic properties to the pore

pressure was highlighted in the study of the pressurization ofa partially confined aquifer beneath a thickening cryosphere.We found that even for an unrealistically rapid climatechange, the diffusion of the excess pressure toward theunconfined portions of the aquifer precluded the attainmentof super-lithostatic pore pressures for almost all cases con-sidered. This lateral diffusionwas accentuated by the increasein permeability at high pore pressures, resulting in a negativefeedback cycle between pore pressure and permeability. Thisfeedback will be important for any hydrologic processinvolving large and widely varying pore pressures.[79] Hydrological modeling of groundwater processes on

Mars is essential for understanding the many observedwater-related features on the surface. While the nature andhydraulic properties of the materials in the subsurface arepoorly constrained, the model developed here provides abaseline with which we can begin to quantitatively testcurrent hypotheses regarding Martian hydrological pro-cesses, as has been done here for the origin of theoutflow channels. The parameters contained in this studyare best estimates based on our current understanding, butwill be subject to revision as more data become available.The forthcoming data from the SHARAD and MARSISradar sounders will likely shed new light on the distri-bution and abundance of Martian groundwater, therebyproviding a much-needed constraint on the aquifer model.

[80] Acknowledgments. We would like to thank Daniel Nunes andAndrew Dombard for many insightful discussions. We especially thankJules Goldspiel and an anonymous reviewer for their thorough andthoughtful reviews. This work was supported by grant NAG5-13243 fromthe NASA Planetary Geology and Geophysics Program and grant NAG5-11202 from the NASA Mars Data Analysis Program.

ReferencesAckermann, H. D., R. J. Godson, and J. S. Watkins (1975), A seismicrefraction technique used for subsurface investigations at Meteor Crater,Arizona, J. Geophys. Res., 80(5), 765–775.

Ahrens, T. J., and A. M. Rubin (1993), Impact-induced tensional failure inrock, J. Geophys. Res., 84(E1), 1185–1203.

Antonellini, M., and P. N. Mollema (2000), A natural analog for a fracturedand faulted reservoir in dolomite; Triassic Sella Group, northern Italy,AAPG Bull., 84(3), 314–344.

Archie, G. E. (1950), Introduction to petrophysics of reservoir rocks, Am.Assoc. Pet. Geol. Bull., 34, 943–961.

Athy, L. F. (1930), Density, porosity, and compaction of sedimentary rocks,Am. Assoc. Pet. Geol. Bull., 34, 943–961.

Baker, V. R. (1982), The Channels of Mars, 198 pp., Univ. of Tex. Press,Austin.

Baker, V. R., and D. J. Milton (1974), Erosion by catastrophic floods onMars and Earth, Icarus, 33, 27–41.

Baker, V. R., R. G. Strom, V. C. Gulick, J. S. Kargel, G. Komatsu, and V. S.Kale (1991), Ancient oceans, ice sheets and the hydrological cycle onMars, Nature, 352, 589–594.

Berman, D. C., and W. K. Hartmann (2002), Recent fluvial, volcanic, andtectonic activity on the Cerberus Plains of Mars, Icarus, 159, 1–17.

Binder, A. B., and M. A. Lange (1980), On the thermal history, thermalstate, and related tectonism of a Moon of fission origin, J. Geophys. Res.,85(B6), 3194–3208.

Brace, W. F. (1984), Permeability of crystalline rocks: New in situ measure-ments, J. Geophys. Res., 89(B6), 4327–4330.

Brace, W. F., and D. L. Kohlstedt (1980), Limits on lithospheric stressimposed by laboratory experiments, J. Geophys. Res., 85(B11),6248–6252.

Brain, D. A., and B. M. Jakosky (1998), Atmospheric loss since the onset ofthe Martian geologic record: Combined role of impact erosion and sput-tering, J. Geophys. Res., 103(E10), 22,689–22,694.

Brown, S. R., and C. H. Scholz (1986), Closure of rock joints, J. Geophys.Res., 91(B5), 4939–4948.

Byerlee, J. D. (1968), Brittle-ductile transition in rocks, J. Geophys. Res.,73(14), 4741–4750.

Caristan, Y. (1982), The transition from high temperature creep to fracturein Maryland diabase, J. Geophys. Res., 87(B8), 6781–6790.

Carr, M. H. (1979), Formation of Martian flood features by release of waterfrom confined aquifers, J. Geophys. Res., 84(B6), 2995–3007.

Carr, M. H. (1992), Post-Noachian erosion rates: Implications for Marsclimate change (abstract), Lunar Planet. Sci., XXIII, 205–206.

Carr, M. H. (1996), Channels and valleys on Mars: Cold climate featuresformed as a result of a thickening cryosphere, Planet. Space Sci., 44(11),1411–1423.

Carr, M. H., and G. D. Clow (1981), Martian channels and valleys: Theircharacteristics, distribution, and age, Icarus, 48, 91–117.

Chapman, D. S., T. H. Keho, M. S. Bauer, and M. D. Picard (1984), Heatflow in the Uinta Basin determined from bottom hole temperature (BHT)data, Geophysics, 49, 453–466.

Clifford, S. M. (1981), A pore volume estimate of the Martian megaregolithbased on a lunar analog, in Papers Presented to the Third InternationalColloquium on Mars, LPI Contrib. 441, pp. 46–48, Lunar and Planet.Inst., Houston, Tex.

Clifford, S. M. (1993), A model for the hydrologic and climatic behavior ofwater on Mars, J. Geophys. Res., 98(E6), 10,973–11,016.

Clifford, S. M., and T. J. Parker (2001), The evolution of the Martianhydrosphere: Implications for the fate of a primordial ocean and thecurrent state of the northern plains, Icarus, 154, 40–79.

Cook, N. G. (1992), Natural joints in rock: Mechanical, hydraulic andseismic behaviour and properties under normal stress, Int. J. Rock Mech.Min. Sci., 29(3), 198–223.

Demming, D. (2002), Introduction to Hydrology, 468 pp., McGraw-Hill,New York.

Domenico, P. A., and F. W. Schwartz (1990), Physical and ChemicalHydrogeology, 824 pp., John Wiley, Hoboken, N. J.

Dressler, B. O., V. L. Sharpton, B. Scnieders, and J. Scott (1996), Forma-tion of impact breccias at the Slate Islands structure, Northern LakeSuperior, Ontario, Canada (abstract), Lunar Planet. Sci., XXVII, 325–326.

Dvorak, J., and R. J. Phillips (1977), The nature of the gravity anomaliesassociated with large young lunar craters, Geophys. Res. Lett., 4(9), 380–382.

Frey, H. V., J. H. Roark, K. M. Shockey, E. L. Frey, and S. E. H. Sakimoto(2002), Ancient lowlands on Mars, Geophys. Res. Lett., 29(10), 1384,doi:10.1029/2001GL013832.

Gaidos, E. J. (2001), Note: Cryovolcanism and recent flow of liquid wateron Mars, Icarus, 153, 218–223.

Gego, E. L., G. S. Johnson, M. R. Hankin, A. H. Wylie, and J. A. Welhan(2002), Modeling groundwater flow and contaminant transport in theSnake River Plain aquifer: A stochastic approach, Spec. Pap. Geol.Soc. Am., 353, 249–261.

Goldspiel, J. M., and S. W. Squyres (1991), Ancient aqueous sedimentationon Mars, Icarus, 89, 392–410.

Goldspiel, J. M., and S. W. Squyres (2000), Groundwater sapping andvalley formation on Mars, Icarus, 148, 176–192.

Gudmundsson, A. (2001), Fluid overpressure and flow in fault zones: Fieldmeasurements and models, Tectonophysics, 336, 183–197.

Gudmundsson, A., S. Berg, K. Lyslo, and E. Skurtveit (2001), Fracturenetworks and fluid transport in active fault zones, J. Struct. Geol., 23,343–353.

Gulick, V. C., D. Tyler, C. P. McKay, and R. M. Haberle (1997), Episodicocean-induced CO2 greenhouse on Mars: Implications for fluvial valleyformation, Icarus, 130, 68–86.

Hanna, J. C., and R. J. Phillips (2003), Theoretical modeling of outflowchannels and chaos regions on Mars, Eos Trans. AGU, 84(46), Fall Meet.Suppl., Abstract P11B-1036.

Harper, C. L. J., L. E. Nyquist, B. Bansal, H. Wiesmann, and C. Shih(1995), Rapid accretion and early differentiation of Mars indicated by142Nd/144Nd in SND meteorites, Science, 267, 213–216.

Hartmann, W. K., J. Anguita, M. A. de la Casa, D. C. Berman, and E. V.Ryan (2001), Martian cratering 7: The role of impact cratering, Icarus,149, 37–53.

Hauck, S. A., II, and R. J. Phillips (2002), Thermal and crustal evolution ofMars, J. Geophys. Res., 107(E7), 5052, doi:10.1029/2001JE001801.

Head, J. W. I., M. A. Kreslavsky, and S. Pratt (2002), Northern lowlandsof Mars: Evidence for widespread volcanic flooding and tectonicdeformation in the Hesparian Period, J. Geophys. Res., 107(E1), 5003,doi:10.1029/2000JE001445.

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

18 of 19

E01004

Page 19: Hydrological modeling of the Martian crust with application to the pressurization … · 2016-03-01 · networks, which are the earliest recorded fluvial features on Mars, were likely

Head, J. W., J. F. Mustard, M. A. Kreslavsky, R. E. Milliken, and D. R.Marchant (2003), Recent ice ages on Mars, Nature, 426, 797–802.

Huenges, E., J. Erzinger, J. Kuck, B. Engeser, and W. Kessels (1997), Thepermeable crust: Geohydraulic properties down to 9101 m depth,J. Geophys. Res., 102(B8), 18,255–18,265.

Hynek, B. M., and R. J. Phillips (2003), New data reveal mature, integrateddrainage systems on Mars indicative of past precipitation, Geology, 31(9),757–760.

Innes, M. J. S. (1961), The use of gravity methods to study the undergroundstructure and impact energy of meteorite craters, J. Geophys. Res., 66,2225–2239.

Ivanov, B. A. (2002), Deep drilling results and numerical modeling:Puchez-Katunki impact crater, Russia, Lunar Planet. Sci. [CD-ROM],XXXIII, abstract 1286.

Ivanov, B. A., G. G. Kocharyan, V. N. Kostuchenko, A. F. Kirjakov, andL. A. Pevzner (1996), Puchez-Katunki impact crater: Preliminary dataon recovered core block structure (abstract), Lunar Planet. Sci., XXVII,589–590.

Keihm, S. J., and M. G. Langseth (1977), Lunar thermal regime to 300 km,Proc. Lunar Sci. Conf. 8th, 499–514.

Khan, A., and K. Mosegaard (2002), An inquiry into the lunar interior: Anonlinear inversion of the Apollo lunar seismic data, J. Geophys. Res.,107(E6), 5036, doi:10.1029/2001JE001658.

Khan, A., K. Mosegaard, and K. L. Rasmussen (2000), A new seismicvelocity model for the Moon from a Monte Carlo inversion of the Apollolunar seismic data, Geophys. Res. Lett., 27(11), 1591–1594.

Kocharyan, G. G., V. N. Kostuchenko, and B. A. Ivanov (1996), Mechanicsof rock massive disruption: Implementation to planetary cratering pro-cesses (abstract), Lunar Planet. Sci., XXVII, 589–590.

Kohlstedt, D. L., B. Evans, and S. J. Mackwell (1995), Strength of thelithosphere: Constraints imposed by laboratory experiments, J. Geophys.Res., 100(B9), 17,587–17,602.

Kukkonen, I. T., L. Kivekas, and M. Paananen (1992), Physical propertiesof karnaite (impact melt), suevite and impact breccia in the Lappajarvimeteorite crater, Finland, Tectonophysics, 216, 111–122.

Longonne, P., J. Gagnepain-Beyneix, and H. Chenet (2003), A new seismicmodel of the Moon: Implications for structure, thermal evolution, andformation of the Moon, Earth Planet. Sci. Lett., 211, 27–44.

MacKinnon, D. J., and K. L. Tanaka (1989), The impacted Martian crust:Structure, hydrology, and some geologic implications, J. Geophys. Res.,94(B12), 17,359–17,370.

Mackwell, S. J., M. E. Zimmerman, and D. L. Kohlstedt (1998), High-temperature deformation of dry diabase with application to tectonics onVenus, J. Geophys. Res., 103(B1), 975–984.

Malin, M. C., and K. S. Edgett (2000a), Evidence for recent groundwaterseepage and surface runoff on Mars, Science, 288, 2330–2335.

Malin, M. C., and K. S. Edgett (2000b), Sedimentary rocks of early Mars,Science, 290, 1927–1936.

Manga, M. (2004), Martian floods at Cerberus Fossae can be produced bygroundwater discharge, Geophys. Res. Lett., 31, L02702, doi:10.1029/2003GL018958.

Manning, C. E., and S. E. Ingebritsen (1999), Permeability of the conti-nental crust: Implications of geothermal data and metamorphic systems,Rev. Geophys., 37(1), 127–150.

Masursky, H., J. M. Boyce, A. L. Dial, G. G. Schaber, and M. E. Strobell(1977), Classification and time of formation of Martian channels basedon Viking data, J. Geophys. Res., 82(B28), 4016–4038.

McKenzie, D., and F. Nimmo (1999), The generation of Martian floods bythe melting of ground ice above dykes, Nature, 397, 231–233.

Mellon, M. T. (1996), Limits on the CO2 content of the Martian polardeposits, Icarus, 124, 268–279.

Mellon, M. T., and R. J. Phillips (2001), Recent gullies on Mars and thesource of liquid water, J. Geophys. Res., 106(E10), 23,165–23,179.

Mellon, M. T., B. Jakosky, and S. Postawko (1997), The persistence ofequatorial ground ice on Mars, J. Geophys. Res., 102(E8), 19,357–19,370.

Melosh, H. J., and B. A. Ivanov (1999), Impact crater collapse, Annu. Rev.Earth Planet. Sci., 27, 385–415.

Melosh, H. J., and A. M. Vickery (1989), Impact erosion of the primordialatmosphere of Mars, Nature, 338, 487–489.

Milton, D. J. (1974), Carbon dioxide hyrdrate and floods on Mars, Science,183, 654–656.

Moore, H. J., R. E. Hutton, R. F. Scott, C. R. Spitzer, and R. W. Shorthill(1977), Surface materials of the Viking landing sites, J. Geophys. Res.,82, 4497–4523.

Nelson, P. H. (1994), Permeability-porosity relationships in sedimentaryrocks, Log Anal., 35(3), 38–62.

Neuzil, C. E. (1986), Groundwater flow in low-permeability environments,Water. Resour. Res., 22(8), 1163–1195.

Nugent, R. C., and D. C. Banks (1966), Project Danny Boy: Engineering-Geologic Investigations, Rep. PNE-5005, 98 pp., U.S. Army Corps ofEng., Waterw. Exp. Stn.

O’Connell, R. J., and B. Budiansky (1974), Seismic velocities in dry andsaturated cracked solids, J. Geophys. Res., 79(35), 5412–5426.

O’Keefe, J. D., and T. J. Ahrens (1985), Impact and explosion crater ejecta,fragmentation size, and velocity, Icarus, 62, 328–338.

Phillips, R. J., et al. (2001), Ancient geodynamics and global-scale hydrol-ogy on Mars, Science, 291, 2587–2591.

Pilkington, M., and R. A. F. Grieve (1992), The geophysical signature ofterrestrial impact craters, Rev. Geophys., 30, 161–181.

Renard, F., J. Gratier, and B. Jamtveit (2000), Kinetics of crack-sealing,intergranular pressure solution, and compaction around active faults,J. Struct. Geol., 22, 1395–1407.

Saar, M. O., and M. Manga (2004), Depth dependence of permeability inthe Oregon Cascades inferred from hydrogeologic, thermal, seismic, andmagmatic modeling constraints, J. Geophys. Res., 109, B04204,doi:10.1029/2003JB002855.

Seeburger, D. A., and M. D. Zoback (1982), The distribution of naturalfractures and joints at depth in crystalline rock, J. Geophys. Res., 87(B7),5517–5534.

Seront, B., T. Wong, J. S. Caine, C. B. Forster, R. L. Bruhn, and J. T.Fredrich (1982), Laboratory characterization of hydromechanical prop-erties of a seismogenic normal fault system, J. Struct. Geol., 20, 865–881.

Shkuratov, Y. G., and N. V. Bondarenko (2001), Regolith layer thicknessmapping of the Moon by radar and optical data, Icarus, 149, 329–338.

Snow, D. T. (1970), The frequency and aperture of fractures in rock,J. Rock. Mech. Min. Sci., 7, 23–40.

Squyres, S. W., and J. F. Kasting (1994), Early Mars: How warm and howwet?, Science, 265, 744–749.

Squyres, S. W., D. E. Wilhelms, and A. C. Moosman (1987), Large-scalevolcano-ground ice interactions on Mars, Icarus, 70, 385–408.

Tanaka, K. L., and J. A. Skinner (2004), Advances in reconstructing thegeologic history of the Chryse region outflow channels on Mars, LunarPlanet. Sci. [CD-ROM], XXXV, abstract 1770.

Toksoz, M. N., A. M. Dainty, S. C. Solomon, and K. R. Anderson (1974),Structure of the Moon, Rev. Geophys., 12(4), 539–567.

Travis, B. J., N. D. Rosenberg, and J. N. Cuzzi (2003), On the role ofwidespread subsurface convection in bringing liquid water close to Mars’surface, J. Geophys. Res., 108(E4), 8040, doi:10.1029/2002JE001877.

Walsh, J. B. (1965), The effect of cracks on the compressibility of rock,J. Geophys. Res., 70(2), 381–389.

Wang, J. S., T. N. Narasimhan, and C. H. Scholz (1988), Aperture correla-tion of a fractal fracture, J. Geophys. Res., 93(B3), 2216–2224.

Ward, S. N. (2002), Planetary cratering: A probabilistic approach, J. Geo-phys. Res., 107(E4), 5023, doi:10.1029/2000JE001343.

Warren, N., and R. Trice (1975), Correlation of elastic moduli systema-tics with texture in lunar materials, Proc. Lunar Sci. Conf. 6th, 3255–3268.

Warren, P. H., and K. L. Rasmussen (1987), Megaregolith insulation, inter-nal temperatures, and bulk uranium content of the Moon, J. Geophys.Res., 92(B5), 3453–3465.

Watkins, J. S., and R. L. Kovach (1973), Seismic investigation of the lunarregolith, Proc. Lunar Sci. Conf. 4th, 2561–2574.

Welhan, J. A., T. M. Clemo, and E. L. Gego (2002), Stochastic simulationof aquifer heterogeneity in a layered basalt aquifer system, eastern SnakeRiver Plain, Idaho, Spec. Pap. Geol. Soc. Am., 353, 225–247.

Wieczorek, M. A., and R. J. Phillips (1997), The structure and compensa-tion of the lunar highland crust, J. Geophys. Res., 102(E5), 10,933–10,943.

Williams, D. E. (1994), Acraman: A major impact structure from the Neo-proterozoic of Australia, Spec. Pap. Geol. Soc. Am., 293, 209–224.

Williams, R. M. E., and R. J. Phillips (2001), Morphometric measurementsof Martian valley networks from Mars Orbiter Laser Altimeter (MOLA)data, J. Geophys. Res., 106(E10), 23,737–23,751.

Williams, R. M., R. J. Phillips, and M. C. Malin (2000), Flow rates andduration within Kasei Valles, Mars: Implications for the formation of aMartian ocean, Geophys. Res. Lett., 27(7), 1073–1076.

Woronow, A. (1988), Variation in the thickness of ejecta cover on Marswith increasing crater density, in MEVTV Workshop on the Nature andComposition of Surface Units on Mars, LPI Tech. Rep. 88–05, pp. 135–137, Lunar Planet. Inst., Houston, Tex.

�����������������������J. C. Hanna and R. J. Phillips, McDonnell Center for the Space Sciences

and Department of Earth and Planetary Sciences, Washington University,St. Louis, MO 63130, USA. ([email protected])

E01004 HANNA AND PHILLIPS: HYDROLOGICAL MODELING OF THE MARTIAN CRUST

19 of 19

E01004