+ All Categories
Home > Documents > Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

Date post: 03-Feb-2022
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
19
Impact-induced hydrothermal activity on early Mars Oleg Abramov and David A. Kring Lunar and Planetary Laboratory, University of Arizona, Tucson, Arizona, USA Received 15 April 2005; revised 5 August 2005; accepted 15 August 2005; published 4 November 2005. [1] We report on numerical modeling results of postimpact cooling of craters with diameters of 30, 100, and 180 km in an early Martian environment, with and without the presence of water. The effects of several variables, such as ground permeability and the presence of a crater lake, were tested. Host rock permeability is the main factor affecting fluid circulation and lifetimes of hydrothermal systems, and several permeability cases were examined for each crater. The absence of a crater lake decreases the amount of circulating water and increases the system lifetime; however, it does not dramatically change the character of the system as long as the ground remains saturated. It was noted that vertical heat transport by water increases the temperature of localized near-surface regions and can prolong system lifetime, which is defined by maximum near-surface temperature. However, for very high permeabilities this effect is negated by the overall rapid cooling of the system. System lifetimes, which are defined by near-surface temperatures and averaged for all permeability cases examined, were 67,000 years for the 30-km crater, 290,000 years for the 100-km crater, and 380,000 for the 180-km crater. Also, an approximation of the thermal evolution of a Hellas-sized basin suggests potential for hydrothermal activity for 10 Myr after the impact. These lifetimes provide ample time for colonization of impact-induced hydrothermal systems by thermophilic organisms, provided they existed on early Mars. The habitable volume reaches a maximum of 6,000 km 3 8,500 years after the impact in the 180-km crater model. Citation: Abramov, O., and D. A. Kring (2005), Impact-induced hydrothermal activity on early Mars, J. Geophys. Res., 110, E12S09, doi:10.1029/2005JE002453. 1. Introduction 1.1. Impact-Induced Hydrothermal Systems [2] Impact-induced hydrothermal systems are initiated by an impact event into a water-rich or ice-rich target material. A fraction of the kinetic energy of the impactor raises the temperature of the planetary crust, providing a thermal driver for the circulation of water and the release of steam and other volatiles such as CO 2 . In particular, several long- term heat sources are generated during an impact event: shock-heated solid material, impact melt, and the central uplift. In every hypervelocity impact, a shock wave is generated in the target, compressing the material and depos- iting a large amount of internal energy. Since a shock is not thermodynamically reversible, waste heat is produced upon decompression of the shocked material by the rarefaction waves that follow [e.g., Ahrens and O’Keefe, 1972; Kieffer and Simonds, 1980]. A phase change can occur if enough heat is deposited, resulting in the melting or vaporization of the target rocks. If sufficient melt is produced, a melt sheet is formed in the crater basin. Additionally, material below the central region of the crater is uplifted during the formation process, delivering warm material closer to the surface. The relative importance of these heat sources depends on the diameter of the crater. The dominant heat source for small, simple craters (less than 7 km in diameter on Mars) is shock-emplaced heat, due to negligible amounts of melt and uplift. For larger, complex craters, melt sheets and central uplifts become important, with melt sheets contrib- uting significantly more energy than central uplifts [Daubar and Kring, 2001; Thorsos et al., 2001]. [3] Evidence of impact-induced hydrothermal activity is present at several terrestrial impact craters in the form of mineral assemblages indicative of high-temperature hydro- thermal alteration. Examples of known alteration sites include, in order of increasing diameter, the 4-km Ka ¨rdla crater [Versh et al., 2003], the 24-km Haughton crater [Osinski et al., 2001], the 35-km Manson crater [e.g., McCarville and Crossey , 1996], the 80-km Puchezh- Katunki crater [e.g., Naumov , 1993, 2002], the 180-km Chicxulub crater [Kring and Boynton, 1992; Ames et al., 2004; Hecht et al., 2004; Zu ¨rcher and Kring, 2004] and the 150- to 250-km Sudbury crater [e.g., Farrow and Watkinson, 1992; Ames et al., 1998]. Impact-induced hydrothermal activity has been suggested for Martian craters as well [Newsom, 1980; Allen et al., 1982; Newsom et al., 1996]. [4] While there are no known active impact-induced hydrothermal systems today, their presence may have been dramatically greater at 3.9 Ga. Several lines of evidence indicate that the inner solar system was subjected to a sharp increase in the number of impacts at that time. Analysis of lunar crust samples [e.g., Turner et al., 1973; Tera et al., 1974] and impact melts [e.g., Dalrymple and Ryder , 1993, 1996; Cohen et al., 2000] returned by the Apollo and Luna JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, E12S09, doi:10.1029/2005JE002453, 2005 Copyright 2005 by the American Geophysical Union. 0148-0227/05/2005JE002453$09.00 E12S09 1 of 19
Transcript
Page 1: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

Impact-induced hydrothermal activity on early Mars

Oleg Abramov and David A. KringLunar and Planetary Laboratory, University of Arizona, Tucson, Arizona, USA

Received 15 April 2005; revised 5 August 2005; accepted 15 August 2005; published 4 November 2005.

[1] We report on numerical modeling results of postimpact cooling of craters withdiameters of 30, 100, and 180 km in an early Martian environment, with and without thepresence of water. The effects of several variables, such as ground permeability and thepresence of a crater lake, were tested. Host rock permeability is the main factor affectingfluid circulation and lifetimes of hydrothermal systems, and several permeability caseswere examined for each crater. The absence of a crater lake decreases the amount ofcirculating water and increases the system lifetime; however, it does not dramaticallychange the character of the system as long as the ground remains saturated. It was notedthat vertical heat transport by water increases the temperature of localized near-surfaceregions and can prolong system lifetime, which is defined by maximum near-surfacetemperature. However, for very high permeabilities this effect is negated by the overallrapid cooling of the system. System lifetimes, which are defined by near-surfacetemperatures and averaged for all permeability cases examined, were 67,000 years for the30-km crater, 290,000 years for the 100-km crater, and 380,000 for the 180-km crater.Also, an approximation of the thermal evolution of a Hellas-sized basin suggests potentialfor hydrothermal activity for �10 Myr after the impact. These lifetimes provide ampletime for colonization of impact-induced hydrothermal systems by thermophilic organisms,provided they existed on earlyMars. The habitable volume reaches amaximumof 6,000 km3

8,500 years after the impact in the 180-km crater model.

Citation: Abramov, O., and D. A. Kring (2005), Impact-induced hydrothermal activity on early Mars, J. Geophys. Res., 110, E12S09,

doi:10.1029/2005JE002453.

1. Introduction

1.1. Impact-Induced Hydrothermal Systems

[2] Impact-induced hydrothermal systems are initiated byan impact event into a water-rich or ice-rich target material.A fraction of the kinetic energy of the impactor raises thetemperature of the planetary crust, providing a thermaldriver for the circulation of water and the release of steamand other volatiles such as CO2. In particular, several long-term heat sources are generated during an impact event:shock-heated solid material, impact melt, and the centraluplift. In every hypervelocity impact, a shock wave isgenerated in the target, compressing the material and depos-iting a large amount of internal energy. Since a shock is notthermodynamically reversible, waste heat is produced upondecompression of the shocked material by the rarefactionwaves that follow [e.g., Ahrens and O’Keefe, 1972; Kiefferand Simonds, 1980]. A phase change can occur if enoughheat is deposited, resulting in the melting or vaporization ofthe target rocks. If sufficient melt is produced, a melt sheet isformed in the crater basin. Additionally, material below thecentral region of the crater is uplifted during the formationprocess, delivering warm material closer to the surface. Therelative importance of these heat sources depends on thediameter of the crater. The dominant heat source forsmall, simple craters (less than 7 km in diameter on Mars)

is shock-emplaced heat, due to negligible amounts of meltand uplift. For larger, complex craters, melt sheets andcentral uplifts become important, with melt sheets contrib-uting significantly more energy than central uplifts [Daubarand Kring, 2001; Thorsos et al., 2001].[3] Evidence of impact-induced hydrothermal activity is

present at several terrestrial impact craters in the form ofmineral assemblages indicative of high-temperature hydro-thermal alteration. Examples of known alteration sitesinclude, in order of increasing diameter, the 4-km Kardlacrater [Versh et al., 2003], the �24-km Haughton crater[Osinski et al., 2001], the �35-km Manson crater [e.g.,McCarville and Crossey, 1996], the �80-km Puchezh-Katunki crater [e.g., Naumov, 1993, 2002], the �180-kmChicxulub crater [Kring and Boynton, 1992; Ames et al.,2004; Hecht et al., 2004; Zurcher and Kring, 2004] and the150- to 250-km Sudbury crater [e.g., Farrow and Watkinson,1992; Ames et al., 1998]. Impact-induced hydrothermalactivity has been suggested for Martian craters as well[Newsom, 1980; Allen et al., 1982; Newsom et al., 1996].[4] While there are no known active impact-induced

hydrothermal systems today, their presence may have beendramatically greater at �3.9 Ga. Several lines of evidenceindicate that the inner solar system was subjected to a sharpincrease in the number of impacts at that time. Analysis oflunar crust samples [e.g., Turner et al., 1973; Tera et al.,1974] and impact melts [e.g., Dalrymple and Ryder, 1993,1996; Cohen et al., 2000] returned by the Apollo and Luna

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, E12S09, doi:10.1029/2005JE002453, 2005

Copyright 2005 by the American Geophysical Union.0148-0227/05/2005JE002453$09.00

E12S09 1 of 19

Page 2: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

missions, as well as meteorites, indicates that the rocks inthe lunar crust were either thermally metamorphosed ormelted at �3.9 Ga. This is interpreted as a consequence of adramatic increase in the number of impacts in a relativelybrief time span of 20 to 200 Myr [e.g., Tera et al., 1974;Ryder, 2000]. This cataclysm was not limited to the Earth-Moon system as meteorites from multiple bodies in theasteroid belt, as well as the only sample of the ancientMartian crust (meteorite ALH 84001), show effects ofimpact-induced metamorphism at �3.9 Ga [Bogard, 1995;Kring and Cohen, 2002].

1.2. Importance of Impact-Induced HydrothermalSystems on Early Mars

[5] The sites of ancient hydrothermal systems are verypromising places to search for biomarkers indicative of pastbiological activity onMars. Of the two types of hydrothermalsystems, volcanogenic and impact-induced, the latter mayhave been more important early in Mars’s history, becauseold (Noachian-age) terrains are dominated by impact crater-ing processes and show little evidence of volcanism [Carr,1996]. This implies that impact cratering removed volcaniclandforms that may have been produced during this period ofhigh heat flow, and was therefore the dominant process forsome time. Also, while volcanic process on Mars haveproduced positive topographic features, such as the Tharsisshield volcanoes, impact events formed basins that couldhave filled with water to form crater lakes [e.g., Newsom etal., 1996; Cabrol et al., 2001], enhancing their biologicalpotential. Finally, during the impact cataclysm at �3.9 Ga,the heat delivered by impact events may have exceeded thatgenerated by volcanic activity [Kring, 2000].[6] Ancient valley networks, some of which are inter-

preted as surface runoff [e.g., Williams and Phillips, 2001;Craddock and Howard, 2002], as well as chemical, miner-alogical, and structural data from the Opportunity rover[Squyres et al., 2004], indicate that liquid water was presentand stable on the surface (at least episodically) in theNoachian epoch, during which the cataclysm took place.The cataclysm may have resurfaced Mars, forming most ofthe craters observed in the Martian highlands [Kring andCohen, 2002], and likely would have resulted in cycles ofvaporization of any surface water or ice following very largeimpacts (forming craters at least �600 km in diameter)[Segura et al., 2002], eliminating any life that may haveexisted at the surface. At the same time, new subsurfacehabitats were created [Zahnle and Sleep, 1997] in the formof impact-induced hydrothermal systems, which may haveprovided sanctuary to existing life or were a site of itsorigin. There are several lines of evidence suggesting thatmay have been the case on Earth. For example, phylogeniesof terrestrial organisms constructed from rRNA sequencesimply a thermophilic or hyperthermophilic common ances-tor [e.g., Pace, 1997], which, along with the earliest isotopicevidence of life at �3.85 Ga [Mojzsis and Harrison, 2000]coinciding with the cataclysm, suggest that impact-inducedhydrothermal systems played an important role in the originand evolution of early life on Earth. The same may be truefor Mars. It is also important to note, however, that impact-generated hydrothermal systems were not limited to earlyMars. Present-day subsurface ice has been inferred at highlatitudes (poleward of 60�) on the basis of the detection of

hydrogen by the Gamma Ray Spectrometer (GRS) on boardMars Odyssey [Boynton et al., 2002], and indirectly by thepresence of fresh craters with fluidized ejecta blankets [e.g.,Mouginis-Mark, 1987; Squyres et al., 1992; Barlow andPerez, 2003] and rootless cones [Lanagan et al., 2001] atlower latitudes. Thus a present-day impact may still gener-ate hydrothermal activity.[7] Unlike Earth, many ancient Martian craters are well-

preserved and presumably contain a mineralogical record ofhydrothermal activity. Future remote sensing missions, suchas the upcoming Mars Reconnaissance Orbiter (MRO), willhave spectrometers with high spatial resolution, potentiallycapable of detecting localized outcrops of hydrothermallyaltered lithologies. In anticipation of these missions, it isvital to model the spatial distribution of the alteration inMartian craters, predict mineral assemblages indicative ifhydrothermal activity, and outline a search strategy.

2. Goals

[8] One of the goals of this work is to constrain thelifetimes of impact-induced hydrothermal systems onearly Mars. The duration of hydrothermal activity can affectthe magnitude of chemical and mineralogical alteration of theMartian crust and is a strong factor in determining thepotential biological importance of these systems. As dis-cussed qualitatively by Newsom et al. [2001], the systemlifetime depends on the size and initial temperature of the heatsource, physical and thermal parameters of host rocks (such aspermeability, heat capacity, and thermal conductivity), avail-ability of liquid water, and other factors, such as self-sealingdue to mineral precipitation. Crater cooling models suggestthat the lifetimes of hydrothermal systems in craters 20 to200 km in diameter are�103 to 106 years if purely conductivecooling is assumed [e.g., Daubar and Kring, 2001; Turtle etal., 2003; Ivanov, 2004]. In order to better constrain theexpected lifetimes of these systems we are using a finitedifference computer simulation to evaluate the additionaleffects of heat transport by water and steam.[9] Another goal is to further understand the mechanics of

postimpact hydrothermal circulation, with a focus on thelocations of near-surface activity. This in turn can aid inspectroscopic and visual identification of hydrothermalvents and hydrothermally altered minerals at Martian craters.[10] Finally, we are seeking to understand the biological

potential of these systems in terms of their habitablevolume, or the rock volume within which there are temper-atures and fluid flow suitable for thermophilic microorgan-isms. The evolution of habitable volume as a function oftime can place important constraints on the long-termhabitability of these systems.[11] The present work aims to achieve these goals by

numerically modeling hydrothermal activity in craters of 30,100, and 180 km diameter in an early Martian environment.In addition, a conductive cooling timescale for a Hellas-sizeimpact basin is estimated.

3. Model

3.1. Computer Code HYDROTHERM

[12] A modified version of the publicly available pro-gram HYDROTHERM (source code available from

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

2 of 19

E12S09

Page 3: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

authors) is used for the simulations in this work.HYDROTHERM is a three-dimensional finite differencecomputer code developed by the U.S. Geological Surveyto simulate water and heat transport in a porous medium[Hayba and Ingebritsen, 1994]. Its operating range is 0 to1200� C and 0.5 to 1,000 bars; however, the uppertemperature limit has been extended by the authors to1700 �C for the modeling of impact melt sheets. For amore detailed theoretical discussion of the thermodynam-ics and fluid mechanics of hydrothermal modeling andtheir implementation in HYDROTHERM, please refer toAbramov and Kring [2004] and references therein.[13] HYDROTHERM has been previously used for

scientific applications; in particular, it was applied tohydrothermal systems of volcanic origin on Earth [e.g.,Hayba and Ingebritsen, 1997] and Mars [Gulick, 2001], aswell as hydrothermal systems at terrestrial [Abramov andKring, 2004] and Martian [Rathbun and Squyres, 2002]impact craters. This work presents several specificimprovements on Rathbun and Squyres [2002]. Whilethe earlier simulationswere limited to 50,000–100,000 years,in this work hydrothermal activity is simulated for up toseveral million years, allowing estimates of system life-times for larger craters. Crater lakes and the latent heat offusion are now explicitly included in the model. Thecrater topography has been improved on the basis ofobservations of lunar craters, and is preserved throughoutthe simulation, rather than being removed shortly aftercrater formation. In addition, the postimpact temperaturedistributions that serve as starting conditions have beengenerated by hydrocode simulations either specifically forMars or were adapted for Mars with consideration of thedifferent kinetic energy requirements for the formation ofMartian craters.[14] In general, HYDROTHERM is well suited for

modeling impact-induced hydrothermal systems, althoughthe code relies on several assumptions. Perhaps its mostsignificant shortcoming for the purposes of this work is itsinability to model brines. The program makes the assump-tion that the fluid is pure water, while hydrothermal fluidsgenerally contain some dissolved solids. In particular,several studies point to the existence and composition ofMartian brines. These include laboratory simulations ofbrine formation based on the mineralogy of the SNCmeteorites [Bullock et al., 2004] and a compositionalanalysis of the Nakhla meteorite [Sawyer et al., 2000],which suggests that it has been in contact with a seawater-like brine or a hydrothermal fluid. System parameters, likethe boiling point, depend on the concentration of thesesolutes, which in turn depends on a variety of factors suchas the amount of water available, surface and subsurfacetemperatures, erosion rates, atmospheric pressure and com-position, etc., which are not well quantified for early Mars.If we assume the bulk composition of water on early Marsis similar to terrestrial seawater, then its thermodynamicproperties at subcritical temperatures are sufficiently closeto that of pure water for the purposes of this model. Atsupercritical temperatures, the effect of solutes in H2O isminimized by extremely low permeabilities at those tem-peratures [Hayba and Ingebritsen, 1997]. Another impor-tant assumption made by HYDROTHERM is that the rockand water are in a local thermal equilibrium. This assump-

tion is valid if the fluid flow is relatively slow and steady,and breaks down in cases of rapid transients. For thisbreakdown to occur, water would have to pass through�300 to 500 meters of rock (vertical resolution of ourmodels) without reaching equilibrium. Such transients areunlikely except perhaps in the very early stages of thesystem. HYDROTHERM also assumes that the groundremains fully saturated throughout the simulation, meaningthat all pore spaces remain filled by water or steam. Withthe exception of the elevated rim, which drains rapidlyafter crater formation, this is a fine assumption because thesurface of a crater lake represents the water table and theground below that datum is expected to be fully saturated.Meanwhile the permeability and porosity of the crater rimare set to near-zero to simulate an unsaturated elevatedsurface. Water fluxes through the rim due to atmosphericprecipitation are highly uncertain and are not included inthe model, but, for an annual precipitation of less than 10 cm,would be significantly lower than the hydrothermalfluxes.

3.2. Model Conditions

[15] Taking advantage of an impact crater’s radial sym-metry, we examine a vertical cross section from the centerof the crater to beyond the outer rim. All models arerepresented on a 75 � 33 grid, with a total of 2,475 blocks.The upper boundary of the model represents a layer of cooledbreccia, with pressure and temperature held constant at0.5 bars and 1�C. The thickness of the breccia layer isequal to the vertical resolution of the model, which is 333 mfor the 30-km crater and 500 m for the 100- and 180-kmcraters. It is effectively an infinite source or sink of the fluid,donating or accepting water depending on underlyinghydrologic conditions. It also functions as a heat sink;so when the thermal energy reaches the upper boundary,it is permanently removed from the system. This con-struct is reasonable in a rapidly convecting crater lakesituation, where heat is rapidly removed from the uppersurface layer, and water is freely exchanged. The bottomboundary is impermeable with a constant basal heat fluxof 32.2 mW/m2 to match an average geothermal gradientof 13�C km�1 [Babeyko and Zharkov, 2000]. The left-hand boundary of the model is the axis of symmetry andis thus impermeable and insulating. The right-handboundary is permeable for both fluid and heat and islocated sufficiently far away from the center of the craterthat the temperatures are close to an average geothermalgradient. The depth of the models varies, with 10 km forthe 30-km crater, 15 km for the 100-km crater, and 14.5 kmfor the 180 km crater, and it was found by trial and errorthat further extending the depth had no appreciable effecton hydrothermal activity. This can also be shown analyt-ically using an expression for the thermal conductivetime:

t ¼ z2rCp

k; ð1Þ

where z is depth and r, Cp and k are density, heat capacity,and thermal conductivity, respectively, given in Table 2.This expression indicates that the time required for theheat to conductively propagate to the surface from a depth

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

3 of 19

E12S09

Page 4: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

of 10 km is close to 3 million years, well over the durationof hydrothermal activity.

4. Input Parameters

4.1. Topography

4.1.1. The 30-km Crater[16] The topography for the 30-km model crater was

obtained directly from the Mars Orbiter Laser Altimeter(MOLA) data for a 29-km fresh Martian crater located at23�N, 207�E (Figure 1). This crater has been characterizedas very young by Mouginis-Mark et al. [2003], on the basisof the presence of ejecta rays, extensive secondary craters,high thermal inertia, lack of superimposed small impactcraters, large depth/diameter ratio, and other geomorphicindicators. In addition, the rim height of this crater is�680 mand the central peak height is�820 m, which is significantlyhigher than 340 m and 220 m, respectively, predicted byGarvin et al. [2003] on the basis of topography analysis of alarge sample of Martian craters. In fact, topography derivedfrom lunar craters (Table 1),which predicts the rim and centralpeak heights of 920 m and 490 m, respectively, provides abetter fit. This is likely an indication of an advanced degree ofdegradation of most Martian craters.4.1.2. The 100-km and 180-km Craters[17] While the recent MOLA data on the topography of

large Martian craters is of very high quality, it is likely

not representative of the state of the craters immediatelyafter their formation, since the analyzed craters beendegraded in various degrees by erosional and depositionalprocesses. It is also noteworthy that the topography of the29-km fresh crater is better approximated by lunar mor-phometry than by MOLA-derived predictions. Thereforeour topography models for the 100- and 180-km cratersare based on the morphometry of lunar craters, summa-rized in Table 1.

4.2. Temperature Distribution

4.2.1. The 30-km Crater[18] The temperature distribution for the 30-km crater

model was generated by a hydrocode simulation of craterformation on Mars [Pierazzo et al., 2004]. The simulationmodeled a 90� impact of an asteroid 2 km in diameter ata velocity of 8 km/s. A hydrocode simulation for acometary impact, which produced higher temperatures,was also available, but was not used because asteroidimpacts are far more prevalent. Since the geothermalgradient chosen for our models (13�C/km) matched thegeothermal gradient in the Pierazzo et al. [2004] simula-tion, no further changes to the temperature distributionwere made.4.2.2. The 100-km and 180-km Craters[19] The temperature distribution for the 100-km crater

model was obtained from a hydrocode simulation of the

Figure 1. (a) A 29-km fresh Martian crater located at 23�N, 207�E, just west of Tharsis (VikingMDIM). (b) An elevation profile of the crater (running north to south), which was used to providetopography for the 30-km crater model. Derived from a global MOLA digital elevation model gridded at128 pixels per degree.

Table 1. Parameters Used for Reconstruction of the Original Topography for Craters With D = 30, D = 100, and D =

180 kma

Parameter Dependence on Rim-to-Rim Diameter (D, km) Source

Crater depth 1.044 D0.301 Pike [1977]Crater floor diameter 0.19 D1.25 Pike [1977]Peak ring diameter 0.5 D Wood and Head [1976]Peak ring height 3 Hale and Grieve [1982]Peak ring thickness 0.11 D Pike [1985]Rim height (0.236 D0.399)((0.5D)3/r3) Pike [1977], Melosh [1989]

aThe variable r is the distance from the center of the crater. All parameters were obtained from morphometric studies of lunar craters.After Melosh [1989].

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

4 of 19

E12S09

Page 5: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

impact that formed Popigai crater on Earth [Ivanov, 2004],and adapted for Mars. An important factor that wasconsidered was that the kinetic energy of the projectileneeded to form a 100-km crater on Mars is lowercompared to Earth (Figure 2). The main reasons for thisare the lower gravity on Mars, which allows more materialto be ejected, and the lower average impact velocity forMars, which requires a larger projectile to generate thesame kinetic energy if the density stays constant. Figure 2indicates that the kinetic energy needed to form a crater ofa given diameter is �50% less for Mars than for Earth.Since the amount of shock heating and uplift is approx-imately proportional to the kinetic energy of the projectile,this implies that immediately after an impact a 100-kmcrater on Mars would be �50% cooler than a 100-kmcrater on Earth. Consequently, the net temperature increase(DT) in the Popigai crater temperature distribution wasreduced by 50%.[20] Similarly, the temperature distribution for our 180-km

crater model was obtained from a hydrocode simulation ofSudbury crater formation [Ivanov and Deutsch, 1999] andadapted for early Mars using the methods outlined in theprevious paragraph.4.2.3. Hellas Basin (2,000 km)[21] Hydrocode simulations for the formation of the

Hellas Basin have not yet been conducted. Consequently,the temperature distribution for a Hellas-sized impact basin(�2,000 km in diameter) was computed analytically using

an expression for specific waste heat (DEw) derived fromthe Murnaghan equation of state by Kieffer and Simonds[1980]:

DEw ¼ 1

2PV0 �

2K0V0

n

� �1� Pn

K0

þ 1

� ��1=n" #

þ K0V0

n 1� nð Þ

� 1� Pn

K0

þ 1

� �1� 1=nð Þ" #

; ð2Þ

where P is the peak shock pressure, K0 is the adiabatic bulkmodulus at zero pressure, n is the pressure derivative of thebulk modulus, and V0 is the specific uncompressed volume(1/r0). For basalt, the uncompressed density r0 is 2600 kg/m

3,K0 is 19.3GPa, and n is 5.5 [Gault andHeitowit, 1963]. Shockpressure P drops off with distance r from the impact pointaccording to the power law

P ¼ Ar

Rpr

� ��n

; ð3Þ

where Rpr is the radius of the projectile, n is the decayexponent, and A is pressure at r = Rpr [e.g., Pierazzo andMelosh, 2000]. Rprwas estimated at 91,500 m using Pi-groupscaling laws, assuming a vertical asteroid impact with avelocity typical for Mars (�7,000 m/s) into competent rock.

Figure 2. Final crater diameter as a function of projectile kinetic energy for Mars and Earth. Calculatedfrom the Pi-group scaling laws, assuming a stony impactor with a density of 3,000 kg/m3, a target densityof 2,600 kg/m3, and a competent rock or saturated soil target type. Impact velocities typical for asteroidimpacts are used: 7 km/s for Mars and 17 km/s for Earth.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

5 of 19

E12S09

Page 6: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

For a 90� impact, the decay exponent n was estimated at2.016 ± 0.215 by Pierazzo and Melosh [2000] usinghydrocode simulations. A can be estimated by

A ¼ r0 C þ Su0½ u0; ð4Þ

where C and S are constants, with C = 2600 m/s and S =1.62 [Melosh, 1989, p. 232]. Assuming that the density ofthe projectile is roughly equivalent to that of the target,the initial particle velocity u0 can be calculated by

u0 ¼ffiffiffiffiffiffiffiffi1

2v2

r; ð5Þ

where v is the impactor velocity [Melosh, 1989, p. 65], inthis case �7,000 m/s. To obtain the final temperatureincrease, specific waste heat DEw is divided by the heatcapacity, which is �800 J kg�1 K�1 for basalt [Mellon,2001].[22] The temperature at the base of the lithosphere is

estimated at �1300�C [Schubert et al., 2001], which gives alithospheric thickness of 100 km with a 13�C/km temper-ature gradient. The temperature gradient in the mantle istaken to be 0.1�C/km [Schubert et al., 2001]. The finaltemperature distribution obtained using this method agreeswell with a hydrocode-generated temperature distributionfor a hypothetical impact of a 200-km body on the Earth[Ivanov, 2004].

4.3. Melt Sheet Properties

[23] The volume of melt sheets in the 100- and 180-kmcraters was estimated using an analytical expression derivedby S. M. Wong et al. (Differential melt scaling for obliqueimpacts on the Earth, Moon, and Mars, submitted toMeteoritics and Planetary Science, 2005):

Vmelt ¼ 1:23� 10�8rprtg0:18D0:83

tc D2:35pr v1:63 sin1:63 q; ð6Þ

where rp is the density of the projectile (�3,000 kg/m3), rt isthe density of the target (2600 kg/m3), g is the accelerationdue to gravity (3.72 m/s2), Dtc is the transient craterdiameter as measured at the pre-impact surface, Dpr is theprojectile diameter, v is the impact velocity (7,000 m/s), andq is the impact angle (90�).[24] Several expressions for relating the final crater di-

ameter and the transient crater diameter have been putforward. Croft [1985] suggested the following relationshipbased on the terrace widths and central peak diameters oflunar craters:

D ¼ D1:18tr

D0:18Q

; ð7Þ

where D is the rim-to-rim diameter of the final crater, Dtr isthe rim-to-rim diameter of the transient crater, and DQ is thesimple-to-complex transition diameter. DQ is estimated to be�8,400 m for early Mars assuming inverse scaling withgravity and density [Holsapple, 1993]:

DQ ¼ gMoonrMoonDQMoon

grt: ð8Þ

Kring [1995] derived a different expression for the transient/final crater relation based on the morphology of ejectablankets (in SI units):

D ¼ 0:82D1:07tr : ð9Þ

Croft [1985] and Kring [1995] expressions eventuallyintersect at large diameters but strongly disagree at smalldiameters (Dtr � DQ) at which the relationship should beclose to the simple crater relation D = 1.19Dtr [Melosh,1989]. Thus, for small complex craters, the Croft [1985]expression underestimates the final crater diameter D whilethe Kring [1995] expression overestimates it. For thatreason, we use a geometric mean of the two expressions:

D ¼ 0:91D1:125

tr

D0:09Q

; ð10Þ

which also agrees with the proportionality D / Dtr1.13

proposed by McKinnon and Schenk [1985], who estimatedthe degree of collapse on the basis of the ratio of ejectablanket diameter to the crater diameter. Equation (9) yieldstransient diameters of 62,000 m and 105,000 m for the100-km crater, and the 180-km crater, respectively.[25] However, equation (5) requires the transient crater

diameter measured at the original surface (Dtc, also calledthe apparent diameter), while equation (9) yields the rim-to-rim diameter Dtr of the transient crater. Pike [1977] deter-mined the ratio of the rim-to-rim diameter to the apparentdiameter for 164 lunar craters, finding that it decreases withincreasing crater diameter, ranging from 1.20 for craters lessthan 0.4 km to 1.11 for craters greater than 100 km. Sincethere is little collapse in very small craters, we assume thatthe ratio Dtr/Dtc is close to 1.20, which agrees with hydro-code simulations of transient crater formation by Shuvalovet al. [2002]. It is also close to the average of the ratiossuggested byHolsapple [1993] (1.30) andGrieve andGarvin[1984] (1.13). The apparent transient craters diameters arethen 52,000 m and 88,000 m for the 100- and 180-km craters,respectively.[26] The projectile diameter Dpr is calculated using the Pi-

group scaling laws, resulting in diameters of 8,500 m for the100-km crater and 16,500 m for the 180-km crater. Thisthen yields total melt volumes of 460 km3 and 3,400 km3

for the 100- and 180-km craters, respectively. However, asubstantial fraction of the melt is ejected from the crater. Forlunar craters in the 100 to 180 diameter range, the fractionof melt ejected is estimated at �0.5 [Cintala and Grieve,1998]. The fraction of melt ejected scales with gravity and isestimated at �0.24 for the terrestrial crater Chicxulub[Kring, 1995]. However, this value is not well definedand strongly depends on the assumed value of z in thez-model of Maxwell [1977]. Therefore, using a conserva-tive approach, we assume a lunar value of �1/2 meltejected. The remaining melt volumes of 230 km3 and1700 km3 are distributed in central melt sheets and smallmelt sheets, the latter of which are located between thepeak rings and rims in the annular trough of the craters.[27] The initial temperature of the melt was conservatively

estimated at 1700�C on the basis of temperature estimates inexcess of 1700�C for melt sheets of terrestrial impact craters

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

6 of 19

E12S09

Page 7: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

[e.g., Grieve et al., 1977; Ostermann et al., 1996]. Themelt/clast ratio in the melt sheet increases with craterdiameter and is expected to be high for the 180-kmcrater. However, for the 100-km crater, the volumeproportion of cold clasts becomes an important factor indetermining the initial volume and temperature of themelt sheet. The proportion of clasts in the melt sheet of asimilarly sized Popigai crater on Earth is �50% [Masaitiset al., 1998]. Thus the corresponding volume of clastswas added to the melt sheet of the 100-km Martian crater,doubling its volume to 460 km3 and lowering its tem-perature to 1125�C after thermal equilibration.[28] The latent heat of fusion is included in the model

using the approximation of Jaeger [1968], after Onorato etal. [1978], replacing the heat capacity Cp in the temperaturerange between the liquidus (TL) and the solidus (TS) with

C0p ¼ Cp þ L= TL � TSð Þ: ð11Þ

[29] Here, L is the latent heat of fusion of diopside,431 kJ kg�1. The liquidus and solidus temperatures of1280�C and 1070�C, respectively, were chosen on the basisof those of gabbro [Ernst, 1976], a coarse-grained equivalentof basalt, because the surface ofMars appears to be dominatedby basaltic lithologies [Bandfield et al., 2000].[30] An important question pertaining to the melt sheets

is how much convection occurs during the cooling pro-cess. There is no heat source below the melt sheet, sincethe underlying rocks have a cooler temperature and act asa heat sink. Thus cooling occurs at both the top andbottom of the melt sheet, and the former may generatesome convective overturn in the form of cold downwellingplumes below the stagnant lid. This mechanism has beenobserved at the Makaopuhi lava lake in Hawaii but forvarious reasons is not active at other lava lakes [Davailleand Jaupart, 1993]. Convection is not expected to occur inthe melt sheet of the 100-km crater, since its initialtemperature is well below the liquidus, a temperature atwhich convection in top-cooled magma reservoirs typicallyceases [e.g., Brandeis and Marsh, 1989]. While someconvection may have occurred in the melt sheet of the180-km crater, potentially cooling it faster, it was notincluded in the model due to its uncertain extent andduration and the limitations of the code.

4.4. Rock Parameters

[31] The porosity in our model decreases exponentiallywith depth, accounting for the closing of pore spaces by

lithostatic pressure, following the approach suggested byBinder and Lange [1980] for the lunar crust:

F zð Þ ¼ F0 exp �z=Kð Þ; ð12Þ

where F0 is surface porosity (20%) and K is the decayconstant, which scales with gravity and is 1.07 km for Earthand 2.80 km for Mars [Clifford, 1993]. The depth z ismeasured with respect to local topography, not the pre-impact surface level.[32] It is reasonable to assume that the number of fractures

in an impact crater decreases with depth [e.g., Nordyke,1964], and thus permeability in our model decays exponen-tially with depth similarly to porosity. Due to lower gravity onMars, the depth to the base of the fractures is�2.5 times whatit would be on Earth and permeability decreases moregradually with depth. Permeability is also a function oftemperature, approximating the effect of the brittle/ductiletransition at about 360�C [Fournier, 1991] by log linearlydecreasing permeability with increasing temperature between360 and 500�C:

k zð Þ ¼ k0 exp �z=Kð Þ T < 360�C

log k z;Tð Þ ¼ log k zð Þ þ 11

500� 360500� Tð Þ � 11 360 � T � 500�C

k ¼ 10�11 darcies T > 500�C:

ð13Þ

[33] Since permeability has important effects on thedynamics and duration of a hydrothermal system, and isnot well constrained, several values of k0 were investigatedin this study. It was also recognized that lower permeabil-ities may become more appropriate due to mineralization asthe system ages. Rock density, thermal conductivity, andheat capacity were assumed to be that of basalt [Mellon,2001] and are summarized in Table 2. Additionally, thephysical properties of the mantle for the Hellas basinsimulation were taken from Schubert et al. [2001].

4.5. Crater Lakes

[34] Lakes may have commonly formed in impact craterson early Mars: one survey, conducted by Cabrol and Grin[1999] suggested paleolakes occurred in at least 179 craters.Whether these crater lakes can form shortly after craterformation on a timescale significantly less than the lifetimeof the hydrothermal system depends on hydrologic condi-tions and the amount of atmospheric precipitation. A craterlake would certainly form rapidly if the impact occurredinto an area that had abundant surface water, such as thatinferred for an ancestral stage of the Meridiani Planumregion on the basis of observations by MER Opportunity[Squyres et al., 2004]. A crater lake can also form throughgroundwater drainage from an underground aquifer, asobserved at Meteor Crater on Earth [Shoemaker and Kieffer,1974] and suggested for the 150-km Gusev crater on Mars[Grin and Cabrol, 1997]. In addition, Newsom et al. [1996]suggested that crater lakes may form in large craters (>65 kmin diameter) shortly after an impact and persist even undercurrent climatic conditions. Consequently, crater lakes maybe a common phenomena, particularly on early Mars, andwere incorporated into the models presented in this paper.

Table 2. Rock Parameters Used in the Model

Parameter Value Units

Porosity f(z), 20% at the surface unitlessPermeability f(z,T), 10�2 at the surfacea darciesThermal conductivity (crust) 2.5 W m�1 K�1

Heat capacity (crust) 800 J kg�1 K�1

Density (crust) 2600 kg m�3

Thermal conductivity (mantle) 3.3 W m�1 K�1

Heat capacity (mantle) 1250 J kg�1 K�1

Density (mantle) 3300 kg m�3

aUnless otherwise indicated.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

7 of 19

E12S09

Page 8: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

However, a model was also run without a crater lake to defineits role in hydrothermal activity.

5. Results

5.1. Hydrothermal System Mechanics and Lifetimes

5.1.1. The 30-km Crater[35] Results of the numerical simulation of a hydrother-

mal system in a 30-km crater are shown in Figure 3.Overall, this system is driven by a central hot spot, whichheats up water and causes it to rise, generating a single largeupwelling that persists throughout the system’s lifetime. At25 years the system is primarily characterized by a largeregion within the crater’s central peak where temperaturesand pressures are compatible with water’s gaseous phase.

This results in emission of large quantities of steam up tothree kilometers (horizontally) from the center of the crater.Water is drawn toward the crater’s center to replenish theescaping steam. While not modeled explicitly due to thelimitations of the code, this phase transition would have leftbehind minerals and caused a degree of clogging in thisregion. However, subsequent flow of hot water through thecentral peak may have redissolved these minerals. Thesource and sink of the water is the crater lake; virtuallyno water is drawn from the permeable right boundary afterthe formation of the crater lake.[36] By 1,000 years, the steam emission from the central

peak has essentially ceased. While there are still smallquantities of steam being generated within the central peak,it condenses before reaching the surface. The temperatures

Figure 3. Results of a numerical simulation of the hydrothermal system at a 30-km crater on earlyMars. The central peak of the crater is on the left side of each figure. Surface permeability k0 is 10

�2 darcies.Solid arrows and dotted arrows indicate the water and steam fluxes, respectively. The lack of arrows in someregions indicates that fluxes are less than 2 orders of magnitude smaller than the maximum flux. Solid linesare isotherms, labeled in degrees Celsius. The length of the arrows scales logarithmically with the fluxmagnitude, and the maximum value changes with each plot. (a) 25 years, max. water flux = 6.66 �10�5 kg s�1 m�2, max. steam flux = 1.84 � 10�5 kg s�1 m�2; (b) 1,000 years, max. water flux =1.50 � 10�5 kg s�1 m�2, max. steam flux = 2.33 � 10�7 kg s�1 m�2; (c) 10,000 years, max. waterflux = 6.01 � 10�6 kg s�1 m�2; (d) 100,000 years, max. water flux = 8.49 � 10�7 kg s�1 m�2.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

8 of 19

E12S09

Page 9: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

within the central peak are now noticeably cooler, havingdecreased by more than 100�C. Due to this smaller thermaldriver, the water fluxes are about 5 times smaller than theywere at 25 years. However, the system on the whole stillremains quite hot, with temperatures of over 100�C ob-served in the near-surface up to 6 km away from the crater’scenter.[37] At 10,000 years, steam generation has completely

stopped, and temperatures and water fluxes have furtherdecreased. The center of the single giant convection cell thatmakes up the system has started moving downward, a trendthat continues in subsequent time steps. This allows water torecirculate without entering the lake. Finally, at 100,000years, a small amount of circulation continues, but the water

fluxes are 2 orders of magnitude lower than they were earlyin the system.5.1.2. The 100-km Crater[38] Evolution of an impact-induced hydrothermal sys-

tem in a 100-km crater is shown in Figure 4. Unlike the30-km crater model, this crater is large enough to possessa high-temperature and impermeable melt sheet andtemperatures high enough to render parts of the subsur-face impermeable. One similarity to the 30-km crater,which had steam emission from the central peak, is thatthe steam is being generated and emitted from the peakring early in the system, and is being replenished bywater flowing in on both sides of the peak ring (notshown). However, this activity ceases by 500 years due to

Figure 4. Results of a numerical simulation of the hydrothermal system at a 100-km crater on earlyMars. The center of the crater is on the left side of each figure. Surface permeability k0 is 10

�2 darcies.Solid arrows and dotted arrows indicate the water and steam fluxes, respectively. The lack of arrowsindicates that fluxes are less than 2 orders of magnitude smaller than the maximum flux. Solid lines areisotherms, labeled in degrees Celsius. The length of the arrows scales logarithmically with the fluxmagnitude, and the maximum value changes with each plot. (a) 500 years, max. water flux = 1.01 �10�5 kg s�1 m�2, max. steam flux = 1.35 � 10�6 kg s�1 m�2; (b) 4,000 years, max. water flux =1.88 � 10�5 kg s�1 m�2, max. steam flux = 3.46 � 10�6 kg s�1 m�2; (c) 20,000 years, max. waterflux = 1.63 � 10�5 kg s�1 m�2, max. steam flux = 1.00 � 10�5 kg s�1 m�2; (d) 200,000 years,max. water flux = 1.70 � 10�5 kg s�1 m�2.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

9 of 19

E12S09

Page 10: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

the rapid cooling of the peak ring, and is replaced by astrong upwelling of water within the peak ring, wherecold water enters near the base and is subsequentlyheated and transported upward. The highest fluxes ofthe 500-year time step are observed in this region. At thistime in the simulation, most of the melt in the annulartrough has crystallized, although it is not yet permeableto fluid flow. Some steam is being generated in thecrater’s modification zone on its outer wall, and thereare some strong but somewhat chaotic water fluxes in thisarea as well, that are probably venting through the faultsthat separate blocks of crust in the modification zone.The flow of water through these faults likely reorganizedinto conduits by the precipitation of silica, phyllosilicates,and other minerals and subsequent self-sealing, a phe-nomena observed at mid-ocean ridge black smokers [e.g.,Sleep, 1991] and most subaerial hydrothermal systems. Ifthe temperature drops below the freezing point, formationof ice can have the same effect. Unlike the 30-km cratermodel, a small fraction of the circulating water issupplied by the permeable right boundary and not thecrater lake. However, these fluxes about 2 orders ofmagnitude smaller than those seen elsewhere in thesystem.[39] At 4,000 years, the small melt sheet has completely

crystallized and cooled. Several convection cells havedeveloped in the annular trough and the modificationzone, and there are several obvious deflections of thetemperature contours due to the upward flow of warmerwater and downward flow of colder water. The centralmelt sheet has fully crystallized, but remains completelyimpermeable due to high temperatures. The strong upwell-ing in the peak ring continues to be very active. Also,relatively large steam fluxes are seen within the peak ring.Surprisingly, the overall flux magnitudes are now higherthan they were at 500 years, probably due to a smallerimpermeable rock volume that was previously impedingflow.[40] At 20,000 years, most of the central melt sheet has

cooled below 360�C and is now permeable to water. Thereis now vigorous activity near the right edge of the formercentral melt sheet with a couple of prominent convectioncells. The upwelling in the peak ring continues as before,and a small upward flow in the modification zone on thefar wall of the crater becomes noticeable as well. By200,000 years, the character of the system has changed.The convection cells at the site of the central melt sheethave merged into two strong upwellings, which are thedominant feature of this time step. Far smaller upwellingscontinue to trickle inside the peak ring and along the farwall. However, they are essentially insignificant, and theoverall character of the system is now close to that of the30-km crater: a strong central upward flow driven by acentral hot spot. The central upwelling continues until theend of the system when temperatures become too low tosustain it.5.1.3. The 180-km Crater[41] Figure 5 shows the evolution of a hydrothermal

system in 180-km crater. The main structural features hereare similar to those of the 100-km crater, with largercentral and annular melt sheets, higher topography, andhigher temperatures in the central region. At 4,000 the

central melt sheet has reached its liquidus temperature andis undergoing crystallization, while the small melt sheethas fully solidified. The peak ring again plays host to aprominent upwelling that continues throughout the lifetimeof the system. Another long-lived but relatively weakupwelling develops at the outer wall in the crater’smodification zone, and is being resupplied by water bothfrom the lake and the permeable right boundary. There issome steam generation deep within the crater below theright edge of the central melt sheet, with steam originatingnear the critical point of water at 374�C.[42] At 20,000 years, the small melt sheet has completely

crystallized, but a remnant hot spot remains and drives anupwelling in this area. The central melt sheet has com-pletely crystallized and is partly permeable, allowingwater to start circulating above it. There is also somesupercritical steam being produced below the central meltsheet.[43] At 200,000 years, the central melt sheet has cooled

and is fully permeable, allowing several large convectioncells to develop in this region. This is a major differencefrom the 100-km crater model, where these convection cellsat this time step were replaced by a single central upwelling.The convection cells continue to operate until the end of thesystem. The upwellings within the peak ring and the outerwall continue as before, but with reduced fluxes. Theoverall magnitude of the fluxes continues to decrease, withthe largest flux observed here being �2 times smaller thanthose seen early in the system.[44] At 2 million years the system has long ceased

operating and temperatures have returned close to ageothermal gradient. Only a hint of higher temperaturesremains in the center of the crater. There are a couple of‘‘fossil’’ flows still active, most notably in the modifica-tion zone, but the fluxes are �20 times smaller than thoseseen earlier in the system. This weak circulation, drivenmainly by surface relief, may persist for a long time. Itshould also be noted, however, that these fluxes may beeven smaller or entirely nonexistent due to fracture closingby hydrothermal mineralization and clay deposition, pro-cesses that generally operate on timescales smaller thanthose required for the cooling of this large crater. Giventhe low volumes of circulating water and the volume ofrocks they must traverse, water would be expected to cooloff completely before reaching the surface.

5.2. Effects of Crater Lake

[45] For the purposes of comparison, a simulation wasrun without a crater lake present in the crater basin, but withsaturated ground below the floor of the crater. Figure 6shows the results of this simulation for a 100-km crater. Themajor differences seen in this run are due to the lack ofpressure exerted by the crater lake, which results in waterbeing drawn from the permeable right boundary rather thanbeing resupplied from the lake. This causes the fluxes earlyin the system to be lower than before, and thus less heat isremoved from the system. Unlike the previous model, thereis no flow through the central peak and by 20,000 yearsmost of the activity is concentrated in the center of thecrater, which eventually develops into a single vigorousupwelling in the center of the crater by 200,000 years. Thislong-lived upwelling, coupled with the overall lower fluxes

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

10 of 19

E12S09

Page 11: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

seen in this run, results in a longer system lifetime (700,000versus 430,000 years) for this simulation compared to a casewith the crater lake present.

5.3. Heat Transport in the Absence of Fluid Flow

5.3.1. The 30-, 100-, and 180-km Craters[46] To understand the effects of water on crater cooling

it is helpful to compare and contrast the simulationsdescribed above to identical scenarios but without water.HYDROTHERM can be used to model crater coolingusing purely conductive heat transport in the rock matrixby setting permeability and porosity to near zero.Figures 7–9 show cooling of the 30-, 100-, and 180-kmcraters, respectively, in the absence of water. In the 30-kmcrater model, the temperatures in the 1,000 and 10,000

time steps are significantly higher due to the lack of heatremoval by water. However, in the 100,000 year time stepof the ‘‘wet’’ model, temperatures are actually higher inthe crater’s central region due to the heat deposited by along-lived upwelling of warm water. These vertical upwel-lings also result in the time required to return to geother-mal gradient being longer for the ‘‘wet’’ model comparedto the ‘‘dry’’ model.[47] In the 100-km and 180-km crater models, a similar

trend can be observed. While the ‘‘dry’’ model is on thewhole hotter, there are regions in the ‘‘wet’’ model, such asthe peak ring, where long-lived hot water upwellings havesignificantly increased the temperature. Conversely, thereare numerous downwellings of cold water that led to alocalized temperature decrease. These deflections in the

Figure 5. Results of a numerical simulation of the hydrothermal system at a 180-km crater on earlyMars. The center of the crater is on the left side of each figure. Surface permeability k0 is 10

�2 darcies.Solid arrows and dotted arrows indicate the water and steam fluxes, respectively. The lack of arrowsindicates that fluxes are less than 2 orders of magnitude smaller than the maximum flux. Solid lines areisotherms, labeled in degrees Celsius. The length of the arrows scales logarithmically with the fluxmagnitude, and the maximum value changes with each plot. (a) 4,000 years, max. water flux = 1.66 �10�5 kg s�1 m�2, max. steam flux = 1.50 � 10�6 kg s�1 m�2; (b) 20,000 years, max. water flux = 1.34 �10�5 kg s�1 m�2, max. steam flux = 9.38 � 10�8 kg s�1 m�2; (c) 200,000 years, max. water flux =8.66 � 10�6 kg s�1 m�2; (d) 2,000,000 years, max. water flux = 1.01 � 10�6 kg s�1 m�2.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

11 of 19

E12S09

Page 12: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

temperature contours are entirely absent in the ‘‘dry’’ modeldue to a lack of flowing water.5.3.2. Hellas Basin (2,000 km)[48] The thermal evolution of a Hellas-sized impact basin

is shown in Figure 10. While the vertical resolution of themodel (30 km) for this immense structure is too coarse toshow hydrothermal activity in the upper 10 km of thesurface, this activity is not expected to have a significanteffect on the cooling of the system of this magnitude. Thusonly cooling by conduction is modeled. However, most ofthe hot spot is located in the mantle, but neither mantleconvection nor convection in the melt produced by theimpact is included in this model, which should be treated asa rough approximation. The model results indicate that the

lifetime of the hydrothermal system produced by an impactof this magnitude would be on the order of 10 Myr, which isthe time it takes for the upper 1 km of the surface to coolbelow 90�C (see next section for an explanation).

5.4. Effects of Permeability on System Lifetime

[49] Permeability is arguably the single most importantparameter affecting the nature and duration of an impact-induced hydrothermal system. In addition to the main simu-lation set with a surface permeability of 10�2 darcies, valuesof 10�3 and 10�1 darcies were also tested, corresponding tothe average permeability of the Earth’s crust and mid-range permeability of crystalline rocks, respectively. Qual-itatively, observed fluxes, vertical extent of convection

Figure 6. Results of a numerical simulation of the hydrothermal system at a 100-km crater on earlyMars, without a crater lake. The center of the crater is on the left side of each figure. Surface permeabilityk0 is 10

�2 darcies. Solid arrows and dotted arrows indicate the water and steam fluxes, respectively. Thelack of arrows indicates that fluxes are less than 2 orders of magnitude smaller than the maximum flux.Solid lines are isotherms, labeled in degrees Celsius. The length of the arrows scales logarithmically withthe flux magnitude, and the maximum value changes with each plot. (a) 500 years, max. water flux = 7.74�10�6 kg s�1 m�2, max. steam flux = 4.12 � 10�6 kg s�1 m�2; (b) 4,000 years, max. water flux =4.15 � 10�6 kg s�1 m�2, max. steam flux = 3.14 � 10�6 kg s�1 m�2; (c) 20,000 years, max. waterflux = 6.95 � 10�6 kg s�1 m�2, max. steam flux = 3.26 � 10�6 kg s�1 m�2; (d) 200,000 years,max. water flux = 1.67 � 10�5 kg s�1 m�2, max. steam flux = 2.01 � 10�7 kg s�1 m�2.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

12 of 19

E12S09

Page 13: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

cells, and total heat removed from the system increasewith higher permeabilities, thus affecting the lifetime ofthe systems.[50] Defining the ‘‘system lifetime’’ is somewhat subjec-

tive as there is no clear transition between an active andinactive hydrothermal system. There are three general waysto define system lifetime:[51] 1. The time it takes for the temperatures within a

specified distance to the surface to drop below a specifiedlimit. This measures how long it takes for the near-surfaceto cool and for surface manifestations of hydrothermalactivity, such as venting, geysers, and hot springs, tocease.[52] 2. The time it takes for the maximum temperature

difference (between observed temperature and geothermaltemperature) to drop below a specified limit. This is howcrater cooling time is usually measured. However, since themaximum temperature difference is typically at high depthswhere there is no water flow, this is not applicable tohydrothermal systems. In other words, high temperaturesmay persist deep under the crater long after all hydrothermalactivity has ceased.

[53] 3. The time it takes for the rock volume between Xand Y degrees (�50 and 100�C for thermophiles) that haswater flow through it to reach 0.[54] Of the above, we use the first definition for ‘‘system

lifetime,’’ which we conservatively define as the time ittakes for the system to cool below 90�C within 1 km of thesurface everywhere in the model. Definition (3) is used todefine ‘‘conditions suitable for thermophilic organisms,’’which is closely related to system lifetime and discussedlater in the paper.[55] Figure 11a shows the effects of permeability on the

lifetime of the hydrothermal system in the 30-km crater.Estimated lifetimes are 40,000 years, 50,000 years, and110,000 years for surface permeabilities of 10�3, 10�2, and10�1 darcies, respectively. The trend for this is for lifetimesto increase with permeability; however, further increasingpermeability would cause the system to cool quickly andwould decrease lifetimes.[56] The effects of permeability on the lifetime of the

hydrothermal system in the 100-km crater are shown inFigure 11b, with estimated lifetimes of 320,000 years,430,000 years, and 120,000 years for surface permeabilities

Figure 7. Cooling progress of a 30-km crater on early Mars in the absence of water. The center of thecrater is on the left side of each figure. Solid lines are isotherms, labeled in degrees Celsius.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

13 of 19

E12S09

Page 14: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

of 10�3, 10�2, and 10�1 darcies, respectively. This plotshows an increase in lifetimes with increasing permeabil-ities followed by a decrease in lifetimes, illustrating apreviously made point of a drop in lifetimes at higherpermeabilities due to rapid heat removal. A similarly shapedplot is shown in Figure 11c for the 180-km crater, withestimated lifetimes of 230,000 years, 700,000 years, and210,000 years for surface permeabilities of 10�3, 10�2, and10�1 darcies, respectively.[57] Figure 11d summarizes the general pattern: at low

permeabilities, little heat is transported by flowing water toheat up the near-surface, resulting in lifetimes similar tothose for a purely conductive case. At mid-range perme-abilities, large amounts of heat are transported from thedeep interior to the near-surface, significantly increasing itstemperature and prolonging the lifetime. At high perme-abilities, the same process takes place but is negated by theoverall rapid cooling of the system, causing lifetimes todrop again. Interestingly, for the low-permeability case,system lifetime for the 180-km crater is actually lower thanthat for the 100-km crater. The relatively short lifetime ofthe 180-km crater at this permeability is due to the forma-

tion of multiple convection cells that rapidly cool the near-surface, dropping temperatures below 90�C. In the 100-kmcrater, on the other hand, water is drawn toward the hottestarea in the center of the crater in a single upwelling,concentrating the heat there.

6. Discussion

[58] Several noteworthy features seen in our simulationscorrespond well to observations at terrestrial impact craters.One such feature, seen in the 180-km crater model, and, to alesser extent, in the 100-km crater model, is the upwellingof water in the crater’s modification zone along its outerwell. In terrestrial craters this region contains numerousextensional faults caused by the slumping of crater wallsduring a crater’s modification stage. These faults can serveas conduits for the venting of water and steam, which hasbeen observed at the 24-km Haughton crater in arcticCanada by Osinski et al. [2001]. Also of note is theextensive fluid flow through the central peak of the 30-kmcrater (and through peak rings of larger craters) and thevertical extent of the circulation, with significant fluxes

Figure 8. Cooling progress of a 100-km crater on early Mars in the absence of water. The center of thecrater is on the left side of each figure. Solid lines are isotherms, labeled in degrees Celsius.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

14 of 19

E12S09

Page 15: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

extending to depths of over 6 km. Scientific drilling into thecentral peak of the Puchezh-Katunki crater in Siberia (rim-to-rim diameter �40 km) found significant hydrothermal alter-ation all the way down to the bottom of the Vorotilovskayaborehole at the depth of 5 km [Naumov, 2002].[59] The hydrothermal systems modeled in this paper

remain active for long periods of time, span a wide rangeof temperatures, and would have produced a variety ofminerals which may be detectable on the surface of Mars.Experimental studies of hydrothermal alteration in analogMartian basalts by Baker et al. [2000] suggested numerousalteration products, including opal-CT, quartz, carbonates,and hematite. However, most of these minerals were alsogenerated in lower-temperature (23 and 75�C) experimentalruns and thus may be indistinguishable from mineralsformed by fluvial and lacrustine activity. Nonetheless,several minerals (vesuvianite, sepiolite, richterite, and bio-tite) formed only during high-temperature (200 and 400�C)runs and could be diagnostic of an impact-induced hydro-thermal system. Geochemical model calculations of thehydrothermal alteration of Martian crust [Griffith andShock, 1997] also suggest numerous alteration products,including a production of magnetite in high-temperature

models. Additionally, the fluid used in the experimentalstudy was simply CO2-saturated water, while proposedMartian brines [e.g., Brass, 1980] would further chemicallyinteract with the basalts in a high-temperature environment.The chemical evolution of these systems should be exploredfurther to better predict the mineral assemblages that arepossible.[60] One of the goals of this study was to attempt to

quantify the capacity of impact-induced hydrothermal sys-tems to support long-lived ecosystems of thermophilicorganisms. Here we focus on the temperature range of50�C to 100�C, which is a habitable zone for most terrestrialthermophiles. Another important part of the equation iswater flow, which is needed to bring nutrients and removewaste products. However, the amount of flow required bythermophilic organisms to thrive varies greatly depending ofthe type of organism, its chemosynthetic pathway, and thecomposition of the hydrothermal fluid. In this study, wedefine ‘‘water flow’’ as more than 1% of the maximum fluxobserved in the system throughout its lifetime. Under thisdefinition, habitable conditions persist for roughly thelifetime of the system. However, a liquid water lake canbe maintained for an even longer period of time.

Figure 9. Cooling progress of a 180-km crater on early Mars in the absence of water. The center of thecrater is on the left side of each figure. Solid lines are isotherms, labeled in degrees Celsius.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

15 of 19

E12S09

Page 16: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

[61] Figure 12 shows the habitable rock volume in the180-km crater as a function of time. Shortly after theimpact, the environment in most of the crater is too hotfor thermophilic organisms. However, the periphery of thecrater cools rapidly, sharply increasing the habitable vol-ume. As the crater continues to cool, habitable volumereaches a maximum of 6,000 km3 at 8,500 years anddecreases from there as the crater cools. Gusev crater, whereMER Spirit landed, is only marginally smaller (150 kmversus 180 km) and would have also supported a similarhabitable volume.[62] There are several ways in which the models pre-

sented in this paper can be improved in the future. Betterunderstanding of the physical properties of the Martiancrust, in particular with regard to large-scale permeability,can help constrain the models. Improved knowledge ofhydrologic conditions on early Mars, such as the durationof ‘‘warm and wet’’ periods, amount of atmospheric pre-cipitation, and the existence and duration of stable surfacereservoirs is important for better model accuracy. Thecontinuing improvement in computational resources willallow higher resolutions and finer detail, such as explicit

characterization of fluid flow along faults in a crater’smodification zone. Finally, hydrothermal mineralization,which leads to a decrease in permeability and porositydue to the deposition of hydrothermal minerals in the rockmatrix, should be included in future models. However, thismechanism depends significantly on the mineralogy of thehost rocks and the composition of the circulating fluid,which are not well known for Mars and may vary indifferent craters.

7. Conclusions

[63] The hydrothermal systems at the three modeledcraters have several features in common. In all cases, astrong, long-lived upwelling developed in either the centralpeak (30-km crater) or the peak ring (100- and 180-kmcraters). All cases had substantial steam emission, initiallyfrom the surface, but as the near-surface cooled steamproduction ceased except for supercritical fluid at hightemperatures and pressures. Fluid fluxes in the early stagesof all models were comparable, although fluxes in themodel lacking a crater lake were significantly lower. While

Figure 10. Thermal evolution of a Hellas-size impact basin (2,000 km in diameter). Purely conductivecooling is assumed. The center of the basin is on the left side of each figure. Solid lines are isotherms,labeled in degrees Celsius.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

16 of 19

E12S09

Page 17: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

Figure 11. Lifetimes of hydrothermal systems at a range of early Martian craters plotted as a function oflog k0, where k0 is the ground permeability at the surface. The dashed line indicates the lifetime in theabsence of water (pure conductive case). (a) 30-km crater, (b) 100-km crater, (c) 180-km crater, and(d) general dependence of lifetime on permeability.

Figure 12. Rock volume in the 180-km crater with conditions suitable for thermophilic organisms:water flow and temperature between 50�C and 100�C. ‘‘Water flow’’ is defined as a flux of at least1% of the maximum flux encountered in the system throughout its lifetime. Surface permeability k0is 10�2 darcies.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

17 of 19

E12S09

Page 18: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

the absence of a crater lake decreases the amount ofcirculating water and increases the system lifetime, it doesnot dramatically change the character of the system as longas the ground remains saturated with water.[64] System lifetimes, averaged for all permeability cases

examined, were 67,000 years for the 30-km crater,290,000 years for the 100-km crater, and 380,000 forthe 180-km crater. The volume habitable by thermophilicorganisms reached amaximum of 6,000 km3 at 8,500 years inthe 180-km crater model. Also, an approximation of thethermal evolution of a Hellas-sized basin suggests potentialfor hydrothermal activity for �10 Myr after the impact.[65] The relatively long lifetimes given above are

explained by several circumstances. During the cooling ofa large impact crater (100- and 180-km models), hotter partsof the crater are dominated by conduction because they areimpermeable to fluids. Also, vertical heat transport by waterincreases the temperature of localized near-surface regionsand prolongs the system lifetimes, which is defined by near-surface temperatures. The lifetimes presented in this papershould provide ample time for colonization of impact-induced hydrothermal systems by thermophilic organisms,provided they existed on early Mars.

[66] Acknowledgments. This work was supported by NASA grantNAG512691 from the Mars Fundamental Research Program. We thankHorton Newsom and Norm Sleep for helpful and insightful reviews of thismanuscript.

ReferencesAbramov, O., and D. A. Kring (2004), Numerical modeling of an impact-induced hydrothermal system at the Sudbury crater, J. Geophys. Res.,109, E10007, doi:10.1029/2003JE002213.

Ahrens, T. J., and J. D. O’Keefe (1972), Shock melting and vaporization oflunar rocks and minerals, Moon, 4, 214–249.

Allen, C. C., J. L. Gooding, and K. Keil (1982), Hydrothermally alteredimpact melt rock and breccia: Contributions to the soil of Mars, J. Geo-phys. Res., 87, 10,083–10,101.

Ames, D. E., D. H. Watkinson, and R. R. Parrish (1998), Dating of aregional hydrothermal system induced by the 1850 Ma Sudbury impactevent, Geology, 26, 447–450.

Ames, D. E., I. M. Kjarsgaard, K. O. Pope, B. Dressler, and M. Pilkington(2004), Secondary alteration of the impactite and mineralization in thebasal Tertiary sequence, Yaxcopoil-1, Chicxulub impact crater, Mexico,Meteorit. Planet. Sci., 39, 1145–1167.

Babeyko, A. Yu., and V. N. Zharkov (2000), Martian crust: A modelingapproach, Phys. Earth Planet. Inter., 117, 421–435.

Baker, L. L., D. J. Agenbroad, and S. A. Wood (2000), Experimentalhydrothermal alteration of a Martian analog basalt: Implications for Mar-tian meteorites, Meteorit. Planet. Sci., 35, 31–38.

Bandfield, J. L., V. E. Hamilton, and P. R. Christensen (2000), A globalview of Martian surface compositions from MGS-TES, Science, 287,1626–1630.

Barlow, N. G., and C. B. Perez (2003), Martian impact crater ejectamorphologies as indicators of the distribution of subsurface volatiles,J. Geophys. Res., 108(E8), 5085, doi:10.1029/2002JE002036.

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, 3194–3208.

Bogard, D. D. (1995), Impact ages of meteorites: A synthesis, Meteoritics,30, 244–268.

Boynton, W. V., et al. (2002), Distribution of hydrogen in the near surfaceof Mars: Evidence for subsurface ice deposits, Science, 297, 81–85.

Brandeis, G., and B. D. Marsh (1989), The convective liquidus in a solid-ifying magmachamber: A fluid dynamic investigation, Nature, 339, 613–616.

Brass, G. W. (1980), Stability of brines on Mars, Icarus, 42, 20–28.Bullock, M. A., J. M. Moore, and M. T. Mellon (2004), Laboratory simula-tions of Mars aqueous geochemistry, Icarus, 170(2), 404–423.

Cabrol, N. A., and E. A. Grin (1999), Distribution, classification, and agesof Martian impact crater lakes, Icarus, 142(1), 160–172.

Cabrol, N. A., D. D. Wynn-Williams, D. A. Crawford, and E. A. Grin(2001), Recent aqueous environments in Martian impact craters: An as-trobiological perspective, Icarus, 154, 98–112.

Carr, M. H. (1996), Water on Mars, Oxford Univ. Press, New York.Cintala, M. J., and R. A. F. Grieve (1998), Scaling impact melting andcrater dimensions: Implications for the lunar cratering record, Meteorit.Planet. Sci, 33, 889–912.

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

Cohen, B. A., T. D. Swindle, and D. A. Kring (2000), Support for the lunarcataclysm hypothesis from lunar meteorite impact melt ages, Science,290, 1754–1756.

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

Croft, S. K. (1985), The scaling of impact craters, Proc. Lunar Planet. Sci.Conf. 15th, Part 2, J. Geophys. Res., 90, suppl., C828–C842.

Dalrymple, G. B., and G. Ryder (1993), 40Ar/39Ar age spectra of Apollo 15impact melt rocks by laser step-heating and their bearing on the history oflunar basin formation, J. Geophys. Res., 98(E7), 13,085–13,096.

Dalrymple, G. B., and G. Ryder (1996), Argon-40/argon-39 age spectra ofApollo 17 highlands breccia samples by laser step heating and the age ofthe Serenitatis basin, J. Geophys. Res., 101(E11), 26,069–26,084.

Daubar, I. J., and D. A. Kring (2001), Impact-induced hydrothermal sys-tems: Heat sources and lifetimes, Lunar Planet. Sci., XXXII, abstract1727.

Davaille, A., and C. Jaupart (1993), Thermal convection in lava lakes,Geophys. Res. Lett., 20, 1827–1830.

Ernst, W. G. (1976), Petrologic Phase Equilibria, 333 pp., W. H. Freeman,New York.

Farrow, C. E. G., and D. H. Watkinson (1992), Alteration and the role offluids in Ni, Cu and platinum-group element deposition, Sudbury IgneousComplex contact, Onaping-Levack area, Ontario, Mineral. Petrol., 46,611–619.

Fournier, R. O. (1991), The transition from hydrostatic to greater thanhydrostatic fluid pressure in presently active continental hydrothermalsystems in crystalline rock, Geophys. Res. Lett., 18, 955–958.

Garvin, J. B., S. E. H. Sakimoto, and J. J. Frawley (2003), Craters on Mars:Global geometric properties from gridded MOLA topography, in SixthInternational Conference on Mars [CD-ROM], Abstract 3277, LunarPlanet. Inst., Houston, Tex.

Gault, D. E., and E. D. Heitowit (1963), The partition of energy for hy-pervelocity impact craters formed in rock, Proc. Hypervelocity ImpactSymp. 6th, 11, 420–456.

Grieve, R. A. F., and J. B. Garvin (1984), A geometric model for excavationand modification at terrestrial simple impact craters, J. Geophys. Res., 89,11,561–11,572.

Grieve, R. A. F., M. R. Dence, and P. B. Robertson (1977), Crateringprocesses: As interpreted from the occurrence of impact melts, in Impactand Explosion Cratering, edited by D. J. Roddy, R. O. Pepin, and R. B.Merrill, pp. 791–814, Elsevier, New York.

Griffith, L. L., and E. L. Shock (1997), Hydrothermal hydration of Martiancrust: Illustration via geochemical model calculations, J. Geophys. Res.,102, 9135–9143.

Grin, E. A., and N. A. Cabrol (1997), Limnologic analysis of Gusev craterpaleolake, Mars, Icarus, 130, 461–474.

Gulick, V. C. (2001), Some ground water considerations regarding theformation of small Martian gullies, Lunar Planet. Sci., XXXII, abstract2193.

Hale, W., and R. A. F. Grieve (1982), Volumetric analysis of complex lunarcraters: Implications for basin ring formation, Proc. Lunar Planet. Sci.Conf. 13th, Part 1, J. Geophys. Res., 87, suppl., A65–A76.

Hayba, D. O., and S. E. Ingebritsen (1994), The computer model HYDRO-THERM, a three-dimensional finite-difference model to simulate ground-water flow and heat transport in the temperature range of 0 to 1,200�C,U.S. Geol. Surv. Water Resour. Invest. Rep., 94-4045, 85 pp.

Hayba, D. O., and S. E. Ingebritsen (1997), Multiphase groundwater flownear cooling plutons, J. Geophys. Res., 102, 12,235–12,252.

Hecht, L., A. Wittmann, R. T. Schmitt, and D. Stoffler (2004), Compositionof impact melt particles and the effects of post-impact alteration in sue-vitic rocks at the Yaxcopoil-1 drill core, Chicxulub crater, Mexico, Me-teorit. Planet. Sci., 39, 1169–1186.

Holsapple, K. A. (1993), The scaling of impact processes in planetarysciences, Annu. Rev. Earth Planet. Sci., 21, 333–373.

Ivanov, B. A. (2004), Heating of the lithosphere during meteorite cratering,Sol. Syst. Res., 38(4), 266–278.

Ivanov, B. A., and A. Deutsch (1999), Sudbury impact event: Crateringmechanics and thermal history, in Large Meteorite Impacts and PlanetaryEvolution II, edited by B. O. Dressler and V. L. Sharpton, Spec. Pap.Geol. Soc. Am., 339, 389–397.

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

18 of 19

E12S09

Page 19: Abramov, O., and D.A. Kring, Impact-induced hydrothermal activity on early Mars, J. Geophys. Res

Jaeger, J. C. (1968), Cooling and solidification of igneous rocks in basalts,in The Poldervaart Treatise on Rocks of Basaltic Composition, edited byH. H. Hess and A. Poldervaart, pp. 503–535, John Wiley, Hoboken, N. J.

Kieffer, S. W., and C. H. Simonds (1980), The role of volatiles and lithol-ogy in the impact cratering process, Rev. Geophys., 18, 143–181.

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

Kring, D. A. (1995), The dimensions of the Chicxulub impact crater andimpact melt sheet, J. Geophys. Res., 100, 16,979–16,986.

Kring, D. A., and W. V. Boynton (1992), The petrogenesis of an augite-bearing melt rock in the Chicxulub structure and its relationship to K/Timpact spherules in Haiti, Nature, 358, 141–144.

Kring, D. A., and B. A. Cohen (2002), Cataclysmic bombardment through-out the inner solar system 3.9–4.0 Ga, J. Geophys. Res., 107(E2), 5009,doi:10.1029/2001JE001529.

Lanagan, P. D., A. S. McEwen, L. P. Keszthelyi, and T. Thordarson (2001),Rootless cones on Mars indicating the presence of shallow equatorialground ice in recent times, Geophys. Res. Lett., 28, 2365–2368.

Masaitis, V. L.,M. S.Mashchak, A. I. Raiklin, T. V. Salivanovskaya, andG. I.Shafranovsky (1998), Diamond-Bearing Impactites of the Popigai. Astro-bleme, 179 pp., VSEGEI Press, St. Petersburg, Russia.

Maxwell, D. E. (1977), Simple Z model of cratering, ejection, and over-turned flap, in Impact andExplosionCratering, edited byD. J. Roddy,R.O.Pepin, and R. B. Merrill, pp. 1003–1008, Elsevier, New York.

McCarville, P., and L. J. Crossey (1996), Post-impact hydrothermal alterationof the Manson impact structure, Manson, Iowa, in The Manson ImpactStructure, Iowa: Anatomy of an Impact Crater, edited by C. Koeberl andR. R. Anderson, Spec. Pap. Geol. Soc. Am., 302, 347–376.

McKinnon, W. B., and P. M. Schenk (1985), Ejecta blanket scaling on theMoon and Mercury—and inferences for projectile populations, LunarPlanet. Sci., XVI, 544–545.

Mellon, M. (2001), Thermal inertia and rock abundance on Mars, paperpresented at TES Data Users Workshop, Ariz. State Univ., Tempe, 13–15 Nov.

Melosh, H. J. (1989), Impact Cratering: A Geologic Process, Oxford Univ.Press, New York.

Mojzsis, S. J., and T. M. Harrison (2000), Vestiges of a beginning: Clues tothe emergent biosphere recorded in the oldest known sedimentary rocks,GSA Today, 10(4), 1–6.

Mouginis-Mark, P. J. (1987), Water or ice in the Martian regolith?: Cluesfrom rampart craters seen at very high resolution, Icarus, 71, 268–286.

Mouginis-Mark, P. J., J. M. Boyce, V. E. Hamilton, and F. S. Anderson(2003), A very young, large impact crater on Mars, in Sixth InternationalConference on Mars [CD-ROM], Abstract 3004, Lunar Planet. Inst.,Houston, Tex.

Naumov, M. V. (1993), Zonation of hydrothermal alteration in the centraluplift of the Puchezh-Katunki astrobleme, Meteoritics, 28, 408–409.

Naumov, M. V. (2002), Impact-generated hydrothermal systems: Data fromPopigai, Kara, and Puchezh-Katunki impact structures, in Impacts inPrecambrian Shields, edited by J. Plado and L. J. Pesonen, pp. 117–171, Springer, New York.

Newsom, H. E. (1980), Hydrothermal alteration of impact melt sheets withimplications for Mars, Icarus, 44, 207–216.

Newsom, H. E., G. E. Britell, C. A. Hibbitts, L. J. Crossey, and A. M. Kudo(1996), Impact crater lakes on Mars, J. Geophys. Res., 101, 14,951–14,955.

Newsom, H. E., J. J. Hagerty, and I. E. Thorsos (2001), Location andsampling of aqueous and hydrothermal deposits in Martian impact cra-ters, Astrobiology, 1, 71–88.

Nordyke, M. D. (1964), Cratering experience with chemical and nuclearexplosives, in Proceedings of the Third Plowshare Symposium—Engi-neering With Nuclear Explosives, U.S. At. Energy Comm. Rep. TID-7695, pp. 51–73, Washington, D. C.

Onorato, P. I. K., D. R. Uhlmann, and C. H. Simonds (1978), The thermalhistory of the Manicouagan impact melt sheet, Quebec, J. Geophys. Res.,83, 2789–2798.

Osinski, G. R., J. G. Spray, and P. Lee (2001), Impact-induced hydrother-mal activity in the Haughton impact structure, Canada: Generation of atransient, warm, wet oasis, Meteorit. Planet. Sci., 36, 731–745.

Ostermann, M., U. Scharer, and A. Deutsch (1996), Impact melt dykes inthe Sudbury multi-ring basin (Canada): Implications from uranium-lead

geochronology on the Foy offset dike, Meteorit. Planet. Sci., 31, 494–501.

Pace, N. R. (1997), A molecular view of microbial diversity and the bio-sphere, Science, 276, 734–740.

Pierazzo, E., and H. J. Melosh (2000), Melt production in oblique impacts,Icarus, 145, 252–261.

Pierazzo, E., N. A. Artemieva, and B. A. Ivanov (2004), Starting conditionsfor hydrothermal systems underneath Martian craters: Hydrocode model-ing, Lunar Planet. Sci., XXXV, abstract 1352.

Pike, R. (1977), Size dependence in the shape of fresh impact craters on theMoon, in Impact and Explosion Cratering, edited by D. J. Roddy et al.,pp. 489–509, Elsevier, New York.

Pike, R. (1985), Some morphologic systematics of complex impact struc-tures, Meteoritics, 20, 49–68.

Rathbun, J. A., and S. W. Squyres (2002), Hydrothermal systems associatedwith Martian impact craters, Icarus, 157, 362–372.

Ryder, G. (2000), Ancient intense impacting: A positive effect on the originof life, Eos Trans. AGU, 81(19), Spring Meet. Suppl., Abstract B22B-01.

Sawyer, D. J., M. D. McGehee, J. Canepa, and C. B. Moore (2000), Watersoluble ions in the Nakhla Martian meteorite, Meteorit. Planet. Sci, 35,743–747.

Schubert, G., D. L. Turcotte, and P. Olson (2001),Mantle Convection in theEarth and Planets, 940 pp., Cambridge Univ. Press, New York.

Segura, T. L., O. B. Toon, A. Colaprete, and K. Zahnle (2002), Environ-mental effects of large impacts on Mars, Science, 298, 1977–1980.

Shoemaker, E. M., and S. E. Kieffer (1974), Synopsis of the geology ofMeteor Crater, in Guidebook to the Geology of Meteor Crater, Arizona,ASU Cent. for Meteorite Stud. Publ. 17, pp. 1–11, Tempe, Ariz.

Shuvalov, V., H. Dypvik, and F. Tsikalas (2002), Numerical simulations ofthe Mjølnir marine impact crater, J. Geophys. Res., 107(E7), 5047,doi:10.1029/2001JE001698.

Sleep, N. H. (1991), Hydrothermal circulation, anhydrite precipitation andthermal structure at ridge axes, J. Geophys. Res., 96, 2375–2387.

Squyres, S. W., S. M. Clifford, R. O. Kuzmin, J. R. Zimbelman, and F. M.Costard (1992), Ice in the Martian regolith, in Mars, edited by H. H.Kieffer et al., pp. 523–554, Univ. of Ariz. Press, Tucson.

Squyres, S. W., et al. (2004), In situ evidence for an ancient aqueousenvironment at Meridiani Planum, Mars, Science, 306, 1709–1714.

Tera, F., D. A. Papanastassiou, and G. J. Wasserburg (1974), Isotopicevidence for a terminal lunar cataclysm, Earth Planet. Sci. Lett., 22,1–21.

Thorsos, I. E., H. E. Newsom, and A. D. Davies (2001), Availability of heatto drive hydrothermal systems in large Martian impact craters, LunarPlanet. Sci., XXXII, abstract 2011.

Turner, G., P. H. Cadogan, and C. J. Yonge (1973), Argon selenochronol-ogy, Proc. Lunar Sci. Conf. 4th, 1889–1914.

Turtle, E. P., E. Pierazzo, and D. P. O’Brien (2003), Numerical modeling ofimpact heating of the Vredefort Impact Structure,Meteoritics Planet. Sci.,38(2), 293–303.

Versh, E., K. Kirsimae, A. Joeleht, and J. Plado (2003), Impact inducedhydrothermal system at Kardla crater: Development and biological con-sequences, in Third International Conference on Large Meteorite Impacts[CD-ROM], abstract 4120, Lunar and Planet. Inst., Houston, Tex.

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

Wood, C. A., and J. W. Head (1976), Comparison of impact basins onMercury, Mars, and the Moon, Proc. Lunar Sci. Conf. 7th, 3629–3651.

Zahnle, K. J., and N. H. Sleep (1997), Impacts and the early evolution oflife, in Comets and the Origin and Evolution of Life, edited by P. J.Thomas et al., pp. 175–208, Springer, New York.

Zurcher, L., and D. A. Kring (2004), Hydrothermal alteration in the Yax-copoil-1 borehole, Chicxulub impact structure, Mexico, Meteorit. Planet.Sci., 39(7), 1199–1222.

�����������������������O. Abramov and D. A. Kring, Lunar and Planetary Laboratory,

University of Arizona, 1629 East University Boulevard, Tucson, AZ85721-0092, USA. ([email protected])

E12S09 ABRAMOVAND KRING: HYDROTHERMAL ACTIVITY ON EARLY MARS

19 of 19

E12S09


Recommended