SHARAD soundings and surface roughness at past, present, and ...

14
SHARAD soundings and surface roughness at past, present, and proposed landing sites on Mars: Reections at Phoenix may be attributable to deep ground ice Nathaniel E. Putzig 1 , Roger J. Phillips 2,3 , Bruce A. Campbell 4 , Michael T. Mellon 1 , John W. Holt 5 , and T. Charles Brothers 5 1 Department of Space Studies, Space Science and Engineering Division, Southwest Research Institute, Boulder, Colorado, USA, 2 Planetary Science Directorate, Space Science and Engineering Division, Southwest Research Institute, Boulder, Colorado, USA, 3 Department of Earth and Planetary Sciences, Washington University in St. Louis, St. Louis, Missouri, USA, 4 National Air and Space Museum, Smithsonian Institution, Washington, District of Columbia, USA, 5 Institute for Geophysics and Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, Texas, USA Abstract We use the Shallow Radar (SHARAD) on the Mars Reconnaissance Orbiter to search for subsurface interfaces and characterize surface roughness at the landing sites of Viking Landers 1 and 2, Mars Pathnder, the Mars Exploration Rovers Spirit and Opportunity, the Phoenix Mars lander, the Mars Science Laboratory Curiosity rover, and three other sites proposed for Curiosity. Only at the Phoenix site do we nd clear evidence of subsurface radar returns, mapping out an interface that may be the base of ground ice at depths of ~1566 m across 2900 km 2 in the depression where the lander resides. At the Opportunity, Spirit, and candidate Curiosity sites, images and altimetry show layered materials tens to hundreds of meters thick extending tens to hundreds of kilometers laterally. These scales are well within SHARADs resolution limits, so the lack of detections is attributable either to low density contrasts in layers of similar composition and internal structure or to signal attenuation within the shallowest layers. At each site, we use the radar return power to estimate surface roughness at scales of 10100 m, a measure that is important for assessing physical properties, landing safety, and site trafcability. The strongest returns are found at the Opportunity site, indicating that Meridiani Planum is exceptionally smooth. Returns of moderate strength at the Spirit site reect roughness more typical of Mars. Gale crater, Curiositys ultimate destination, is the smoothest of the four proposed sites we examined, with Holden crater, Eberswalde crater, and Mawrth Vallis exhibiting progressively greater roughness. 1. Introduction The selection of landing sites on Mars has evolved into a highly collaborative process involving several years of effort by a broad range of scientists and engineers [Golombek et al., 2012]. Subsequent to landings, the evaluation of site characteristics with orbital observations continues throughout the landed missions and often extends many years beyond them [Golombek et al., 2008]. Since orbital data are typically acquired with spatial resolution that is coarser by an order of magnitude or more than that of ground observations, accurate interpretation of orbital data relies heavily on the observations made by landed spacecraft. At the same time, choosing new landing sites is driven largely by the interpretation of the orbital data sets, taking into consideration a variety of sometimes conicting constraints such as landing safety, site trafcability, and geologic interest. Historically, site evaluations have relied predominantly on orbital observations of the surface such as those from visible imagers, spectrometers, and a laser altimeter to assess parameters critical to landing-site selection and operations, including surface slopes, atmospheric conditions, mineralogy, grain sizes, rock abundance, stratigraphy, and geologic structures. More recently, other remote- sensing instrumentation such as gamma ray and neutron spectrometers and radars have been put into orbit at Mars and are acquiring data that include new information about surface and subsurface properties. Integration of these data more fully into site evaluations will allow tighter constraints on stratigraphy and geologic structures. In addition, radar observations can be used to evaluate surface roughness on scales intermediate to those available from visible imagery and laser altimetry. PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1936 PUBLICATION S Journal of Geophysical Research: Planets RESEARCH ARTICLE 10.1002/2014JE004646 Key Points: Possible base of ground ice is mapped with Shallow Radar at Phoenix Mars site Roughness at scales of 10 to 100 m is assessed at 10 landing sites on Mars High-resolution surface-elevation mod- els are critical for radar interpretation Correspondence to: N. E. Putzig, [email protected] Citation: Putzig, N. E., R. J. Phillips, B. A. Campbell, M. T. Mellon, J. W. Holt, and T. C. Brothers (2014), SHARAD soundings and surface roughness at past, present, and proposed landing sites on Mars: Reections at Phoenix may be attributable to deep ground ice, J. Geophys. Res. Planets, 119, 19361949, doi:10.1002/ 2014JE004646. Received 21 APR 2014 Accepted 26 JUL 2014 Accepted article online 1 AUG 2014 Published online 21 AUG 2014

Transcript of SHARAD soundings and surface roughness at past, present, and ...

SHARAD soundings and surface roughness at past,present, and proposed landing sites on Mars:Reflections at Phoenix may be attributableto deep ground iceNathaniel E. Putzig1, Roger J. Phillips2,3, Bruce A. Campbell4, Michael T. Mellon1, John W. Holt5,and T. Charles Brothers5

1Department of Space Studies, Space Science and Engineering Division, Southwest Research Institute, Boulder, Colorado,USA, 2Planetary Science Directorate, Space Science and Engineering Division, Southwest Research Institute, Boulder,Colorado, USA, 3Department of Earth and Planetary Sciences, Washington University in St. Louis, St. Louis, Missouri, USA,4National Air and Space Museum, Smithsonian Institution, Washington, District of Columbia, USA, 5Institute for Geophysicsand Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, Texas, USA

Abstract We use the Shallow Radar (SHARAD) on the Mars Reconnaissance Orbiter to search for subsurfaceinterfaces and characterize surface roughness at the landing sites of Viking Landers 1 and 2, Mars Pathfinder,the Mars Exploration Rovers Spirit and Opportunity, the Phoenix Mars lander, the Mars Science LaboratoryCuriosity rover, and three other sites proposed for Curiosity. Only at the Phoenix site do we find clearevidence of subsurface radar returns, mapping out an interface that may be the base of ground ice atdepths of ~15–66 m across 2900 km2 in the depression where the lander resides. At the Opportunity, Spirit,and candidate Curiosity sites, images and altimetry show layered materials tens to hundreds of metersthick extending tens to hundreds of kilometers laterally. These scales are well within SHARAD’s resolutionlimits, so the lack of detections is attributable either to low density contrasts in layers of similar compositionand internal structure or to signal attenuation within the shallowest layers. At each site, we use the radarreturn power to estimate surface roughness at scales of 10–100 m, a measure that is important forassessing physical properties, landing safety, and site trafficability. The strongest returns are found atthe Opportunity site, indicating that Meridiani Planum is exceptionally smooth. Returns of moderatestrength at the Spirit site reflect roughness more typical of Mars. Gale crater, Curiosity’s ultimate destination,is the smoothest of the four proposed sites we examined, with Holden crater, Eberswalde crater, and MawrthVallis exhibiting progressively greater roughness.

1. Introduction

The selection of landing sites on Mars has evolved into a highly collaborative process involving severalyears of effort by a broad range of scientists and engineers [Golombek et al., 2012]. Subsequent to landings,the evaluation of site characteristics with orbital observations continues throughout the landed missionsand often extends many years beyond them [Golombek et al., 2008]. Since orbital data are typicallyacquired with spatial resolution that is coarser by an order of magnitude or more than that of groundobservations, accurate interpretation of orbital data relies heavily on the observations made by landedspacecraft. At the same time, choosing new landing sites is driven largely by the interpretation of theorbital data sets, taking into consideration a variety of sometimes conflicting constraints such as landingsafety, site trafficability, and geologic interest. Historically, site evaluations have relied predominantly on orbitalobservations of the surface such as those from visible imagers, spectrometers, and a laser altimeter to assessparameters critical to landing-site selection and operations, including surface slopes, atmospheric conditions,mineralogy, grain sizes, rock abundance, stratigraphy, and geologic structures. More recently, other remote-sensing instrumentation such as gamma ray and neutron spectrometers and radars have been put into orbit atMars and are acquiring data that include new information about surface and subsurface properties. Integrationof these data more fully into site evaluations will allow tighter constraints on stratigraphy and geologicstructures. In addition, radar observations can be used to evaluate surface roughness on scales intermediate tothose available from visible imagery and laser altimetry.

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1936

PUBLICATIONSJournal of Geophysical Research: Planets

RESEARCH ARTICLE10.1002/2014JE004646

Key Points:• Possible base of ground ice is mappedwith Shallow Radar at PhoenixMars site

• Roughness at scales of 10 to 100 m isassessed at 10 landing sites on Mars

• High-resolution surface-elevation mod-els are critical for radar interpretation

Correspondence to:N. E. Putzig,[email protected]

Citation:Putzig, N. E., R. J. Phillips, B. A. Campbell,M. T. Mellon, J. W. Holt, and T. C. Brothers(2014), SHARAD soundings and surfaceroughness at past, present, and proposedlanding sites on Mars: Reflections atPhoenix may be attributable to deepground ice, J. Geophys. Res. Planets, 119,1936–1949, doi:10.1002/2014JE004646.

Received 21 APR 2014Accepted 26 JUL 2014Accepted article online 1 AUG 2014Published online 21 AUG 2014

In this work, we focus on results from the Shallow Radar (SHARAD) instrument onboard the MarsReconnaissance Orbiter (MRO), which has obtained observations over each of the past landing sites andseveral other sites proposed for the Mars Science Laboratory Curiosity rover. SHARAD has proven highlycapable of detecting subsurface interfaces, most dramatically within the icy layered deposits in the polarregions [Phillips et al., 2008; Seu et al., 2007a] but also within volcanic deposits [Campbell et al., 2008; Carteret al., 2009b] and at the base of glacial ices [Holt et al., 2008; Plaut et al., 2009] in the midlatitudes. At MeridianiPlanum, the landing site of the Mars Exploration Rover Opportunity, images of the surface both from orbitand from the rover show clear evidence for layering on the order of meters to tens of meters in thicknessthat extends laterally over several hundred kilometers. Layering at similar scales is also evident at the landingsite for Curiosity (Gale crater) and three other sites that were proposed for that mission (Eberswalde crater,Holden crater, and Mawrth Vallis). At the Phoenix Mars landing site in the northern plains, images of thesurface do not show direct evidence of layering, but they do show polygonal terrain consistent with thewidespread presence of near-surface ground ice extending to depths of at least 10 m [Mellon et al., 2008b].This interpretation is strongly supported by the detection of excess hydrogen throughout the high latitudesby the Mars Odyssey Gamma Ray Spectrometer (GRS) suite and modeling of thermal observations [Mellonet al., 2008a]. With sufficient dielectric contrasts at layer interfaces, these scales should allow detections bySHARAD, so each of these sites was targeted with high priority by the SHARAD Team and extensive coverageis available. Since there is no clear evidence for layering on these scales at the landing sites of the MarsExploration Rover Spirit (in Gusev crater), Mars Pathfinder, and Viking Landers 1 and 2, targeting these othersites was not deemed a high priority by the SHARAD Team and only limited coverage is available.

In 2008, the SHARAD Team first made note of low-power apparent subsurface returns at the Phoenix site[Safaeinili et al., 2008]. Subsequently, similar, slightly shallower returns were identified across the northernplains [Putzig et al., 2009a] and the southern highlands [Putzig et al., 2009b], suggesting widespread detection ofthe base of ground ice. However, later analysis showed that most of the widespread shallow returns occur atdelay times corresponding to those expected for sidelobes of the surface return [Putzig et al., 2012]. Whilethe returns do have about twice the expected power of the sidelobes, the close correspondence in delay timecasts doubt on a subsurface interpretation. At the Phoenix site, the delay times are typically larger and thepower of returns is greater, so a subsurface source for the late returns ismuch better supported, as we discuss insubsequent sections.

Surface roughness is important both as a factor in evaluating spacecraft landing safety and the ability forrovers to move around on the surface and as an indicator of morphology to constrain geologic properties.Measures of roughness depend on the horizontal length scale at which topography is sampled, with scales oforder 1–10 m of primary concern to landing and roving safety and scales of order 10–1000 m of primaryconcern to geomorphology. Image data such as that from the MRO High Resolution Imaging ScienceExperiment (HiRISE) provide a means of evaluating roughness at the finer scales, either by interpretingfeatures in individual images (e.g., using shadows to estimate heights of rocks or other relief) or by usingstereo pairs to obtain surface-elevation models. For coarser scales, individual observations by the MarsOrbiter Laser Altimeter (MOLA), which were typically spaced at ~300 m, have been used collectively toevaluate roughness at scales of 600 m to 20 km [Kreslavsky and Head, 2000, 2002]. Because the power of radarreturns from a surface depends on the distribution of slope facets at scales on the order of the signal wavelength,SHARAD data can be used to estimate roughness at intermediate scales of 10–100 m, complementary to theestimates provided by images and MOLA data. In this work, we extract roughness for each landing site from aglobal data set of a SHARAD-derived roughness parameter [Campbell et al., 2013], mapping it at two scales toassess the regional and local distributions of roughness.

2. Radar Observations

The SHARAD sounding radar emits an 85μs chirped pulse with a 10MHz bandwidth centered at 20 MHz,yielding range resolutions of 15 m in free space, 8.5 m in water ice with real part of dielectric permittivity(“real permittivity” hereinafter) of 3.15, and 5.3 m in basalt with real permittivity of 8. Correlation of returnedsignals with the transmitted pulse produces a band-limited waveform with the first and second sidelobesexpected at delays of 0.24 μs and 0.42 μs and powers of �20 dB and �34 dB, respectively, relative to peakreturns. In ground processing, a Hann weighting function is applied to the data, providing some suppressionof the sidelobe power without greatly reducing the range resolution. The altitude of the MRO orbit varies

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1937

between about 250 km and 320 km, allowing SHARAD a lateral resolution at the surface of ~3–6 km (1–2 Fresnelzones), which can be reduced in the azimuth (along-track) direction to 0.3–1.0 km with synthetic-aperturefocusing. See Seu et al. [2007b] for a detailed discussion of the SHARAD experiment.

For this work, we applied the radar processing technique of Campbell et al. [2011] to the SHARAD data prior tosearching for subsurface returns. Roughness data for each site were taken from the global data set of Campbell et al.[2013]. We refer to each collection of contiguous SHARAD returns along an orbit segment as an “observation”,presenting them as radargrams, which are image-format displays of returned power with the horizontal axisrepresenting distance along track and the vertical axis representing the signal delay time. Along-track resolution ismatched to that of a MOLA elevationmap [Smith et al., 2001] at 128 pixels per degree (ppd) (~460m per pixel), andthe power is sampled at 37.5 ns in delay time. From top to bottom, a radargram usually shows (1) lower-powerbackground noise associated with transmission through the atmosphere, (2) higher-power returns from points onthe surface closest to the spacecraft (often from the nadir points directly below the spacecraft, but sometimes fromhigh-standing off-nadir structures), and (3) later returns that may be from either subsurface interfaces or off-nadirstructures that are farther away from the spacecraft (termed “surface clutter”). Because of the latter ambiguity,common practice in evaluating subsurface radar data is to generate corresponding clutter simulations usingsurface-elevation models of the observed region and the spacecraft trajectory. We used an incoherent,geometrical optics model that takes into account the effects of along-track focusing [Holt et al., 2006; P.Choudhary et al., Surface clutter and echo location analysis for the interpretation of SHARAD data from Mars,IEEE Geoscience and Remote Sensing Letters, manuscript in revision, 2014] and griddedMOLA topography [Smithet al., 2001] or Mars Express High Resolution Stereo Camera (HRSC) DT4 products [Neukum et al., 2004]. Thismethodmaximizes the possible surface clutter, rather than attempting to precisely simulate the radar data. Tothe extent that actual surface structures are well represented in the models and simulations, late returnsappearing in the data but not in the simulations may be interpreted as subsurface returns. Where returns areidentified as subsurface returns, delay times can be converted to depths on a sample-by-sample basis as

Δdi¼ ½Δt c εi�½ (1)

where Δdi is the depth interval for the i-th sample in delay time, Δt is the delay-time sampling interval, c is thespeed of light in free space, and εi is the real permittivity of the traversed medium, with the factor of 1/2accounting for two-way travel (i.e., radar transmit and receive paths).

The SHARAD signal is subject to distortion and attenuation by the ionosphere of Mars, effects largely confinedto dayside observations that can be corrected in processing [Campbell et al., 2011, 2014]. In addition toproviding additional coverage, the inclusion of dayside observations is important to our analysis because theyform a data grid with the crossing nightside observations, thereby allowing confident mapping of specificsubsurface returns. During SHARAD operations, it was established that an increase in the signal-to-noise ratioup to 6 dB can be achieved by acquiring observations with the spacecraft in a rolled configuration and (on thenightside) positioning the solar array in a specific orientation [Campbell et al., 2014]. We included these

Table 1. SHARAD Observation Counts and Peak Roughness

Landing SiteArea of Interest (AOI) for

Subsurface Returns

Observation Count

Peak Roughness

Subsurface Return AOI

Roughness AOIs

3 × 3° 0.7 × 0.7° 3 × 3° 0.7 × 0.7°

Viking 1 21.2–24.2°N, 310.3–313.3°E 16 16 3 2.75 2.75Viking 2 46.8–49.8°N, 132.5–135.5°E 27 15 2 3.50 3.75Pathfinder 17.6–20.6°N, 325.3–328.3°E 12 6 1 3.25 3.25Spirit 13.2–15.7°S, 174.3–176.8°E 10 13 5 3.50 3.00Opportunity 4.0°S–9.0°N, �9.5–9.0°E 68 17 3 2.50 2.50Phoenix 67.0–69.5°N, 229.5–238.5°E 125 103 32 3.25 3.25Curiosity 4.0–6.6°S, 136.7–138.9°E 29 13 4 3.25 4.25Holden

23.5–27.2°S, 324.5–327.2°E 2838 9 3.50 3.50

Eberswalde 68 26 3.75 3.75Mawrth 22.2–25.8°N, 339.6–342.2°E 52 69 21 4.00 4.00

AOI for Roughness Observation Count Peak RoughnessMidlatitudes 70°S–70°N, 0–360°E 8000 3.25

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1938

rolled observations in our search for subsurfacereturns, but they are omitted in the derivationof roughness to avoid complicating comparisons ofchanges in relative surface reflectivity [Campbellet al., 2013].

SHARAD operates in a targeted mode, with largetargets established for the polar regions (spanning20° of latitude on either side of the pole, ~2400 km)and smaller targets at other sites of interest,including some of the past, present, and proposedlanding sites. Early in the MRO mission (2007),targets were established surrounding the landingsites for Opportunity, Phoenix, and the four finalistsites for Curiosity. Extensive coverage by SHARADhas been acquired at each of these sites. Whiletargets were not created specifically for the landingsites of Viking Landers 1 and 2, Pathfinder, andSpirit, regional targets do cover these sites andsome observations have been acquired near eachof them. For the purpose of searching forsubsurface returns, we used a variety of criteria toestablish areas of interest (AOIs) at each site. For thesites of Viking Landers 1 and 2 and Pathfinder, weused the same 3×3° areas for both the subsurface-return and regional-roughness investigations. AtSpirit, we used the bounds of Gusev crater toestablish the AOI. At Opportunity, we chose an AOIbounding the extents of layered materials thathad previously been mapped in the surroundingregion [Hynek, 2004]. At the Phoenix site, weadopted as AOI the prelanding study area known asRegion D, box 1 [Arvidson et al., 2008]. For theactual and proposed Curiosity sites, we used theprelanding study areas covering the proposedlanding ellipses, roving areas, and their surroundings[Golombek et al., 2012]. In assessing subsurfacereturns, we treated the Holden and Eberswalde sitestogether due to their adjacent locations. Table 1 liststhe AOIs for our subsurface-return investigations ateach site together with the number of SHARADobservations examined.

Our roughness analysis employs the results ofCampbell et al. [2013], who established a roughnessparameter ρ as

ρ¼ Pi =P0 (2)

where Pi is the radar power integrated over a rangeof incidence angles and P0 is the peak radarpower (i.e., that at zero incidence angle). In thisformulation, ρ is independent of reflectivity andmodulated only by surface roughness on the orderof the radar wavelength (15 m for SHARAD). Incalculating the roughness parameter, Campbell et al.

Figure 1

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1939

[2013] restricted incidence angles to0–1.5°, and the roughness measureincludes any subsurface returns todepths of ~35–60 m. For this work, weextracted roughness parameter valuesfrom the database of 8000 SHARADobservations on MRO orbits 2000through 25,999. We then selectedresults at each landing site, mappingthem at 50 ppd to assess regional(3 × 3°) roughness and at 200 ppd toassess local (0.7 × 0.7°, 40 × 40 km)roughness. The numbers of SHARADobservations within the roughnessAOIs for each site at these two scalesare listed in Table 1.

3. Results

In this section, we describe ourfindings at the landing sites for theidentification of SHARAD subsurfacereturns and for the analysis ofSHARAD-derived roughness. We foundclear evidence for subsurface returnsonly at the Phoenix site, but we includea discussion of the nondetections atthe other sites, emphasizing the resultsat Gale crater where early worksuggestive of subsurface returns waslater discounted using cluttersimulations created from a high-resolution surface-elevation model.

3.1. Subsurface Detections at thePhoenix Mars Landing Site

We examined SHARAD observationstaken on 125 orbits through July 2011 that crossed the region 67.0–69.5°N, 229–238°E surrounding thePhoenix site (corresponding to Region D, box 1 of Arvidson et al. [2008]). Figure 1 presents a subset of thisregion where we have identified and mapped subsurface returns in 81 radargrams using SeisWare™

seismic-data analysis software. The ground-track map (Figure 1a) shows the grid of nightside and daysidecoverage that we used in our search for subsurface returns. We highlight tracks for four rolled observations inred and for two observations nearest the landing site in yellow. The base map, consisting of a daytime Mars

Figure 1. Maps of the area around the Phoenix Mars landing site in the northern plains of Mars. Base map in each case is aTHEMIS daytime IR mosaic [Christensen et al., 2013]. (a) SHARAD ground tracks through July 2011, with nightside anddayside acquisition trajectories indicated by arrows. Red ground tracks are for observations acquired with the spacecraft ina rolled configuration to improve the signal-to-noise ratio. Solid lines indicate ground tracks of labeled observationsnearest the lander that are shown in Figures 2–4. (b) Delay time to subsurface returns mapped on 81 SHARAD observationsof the Green Valley where the Phoenix lander resides, interpolated to infill between ground tracks. While the modal valueis near that expected for the second sidelobe of the surface return, delay times range broadly from 0.18 to 0.78 μs. Thematerial—likely dominated by ground ice—between the surface and the subsurface return over the mapped area of2912 km2 occupies 64–99 km3 for real permittivities of 3.15–8.0. (c) Power in decibels of the subsurface return relative tothat of the surface return on 81 SHARAD observations. All power values are well above that expected for the secondsidelobe of the surface return, and the modal power of �9 dB is well above that expected for the first sidelobe.

Figure 2. Portion of SHARAD nightside rolled observation 16378–02, invertedto extend from south to north across the Phoenix landing site and theGreen Valley. (a) Uninterpreted radargram. (b) Interpreted radargram,showing delay times for the MOLA modeled surface in blue and for thesubsurface return in yellow, with the latter delayed ~0.4 μs from thesurface return. “×” symbols show corresponding delay times fromcrossing dayside observations. Interpretation of subsurface returns isdiscontinued where it is obfuscated by off-nadir surface returns. Insetshows map view of ground track. (c) Clutter simulation produced from aswath of the MOLA modeled surface along the ground track of theobservation. Returns arriving late or early relative to the nadir surface(yellow line) are off-nadir surface returns that may also appear in theSHARAD radargram (white arrows). No late returns occur where thesubsurface return is identified in the SHARAD radargram (yellow arrows).

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1940

Odyssey Thermal Emission ImagingSpectrometer (THEMIS) mosaic, showsthe relatively smooth, flat-lyingdepression (informally known asGreen Valley) that extends north andwest of the landing site as well asnearby Heimdal crater. The subsurfacereturns mapped are relatively lowpower, even for nightside observationsacquired with the spacecraft in arolled orientation that provides thehighest possible signal-to-noise ratio(e.g., Figure 2a). As discussed insection 2, having a grid of data iscritical to correlating subsurfacereturns between adjacent tracks(see yellow × symbols in Figure 2b),and here it is also important toconfident interpretation of the low-power signals within each track. Inaddition, we use clutter simulations toidentify returns from off-nadir surfacefeatures (white arrows in Figures 2band 2c) that might otherwise be

mistaken for subsurface returns and to ensure that returns identified as subsurface returns are not in factoff-nadir features present in the modeled surface (cf. yellow line in Figure 2b and yellow arrows in Figure 2c).Given the relatively sparse coverage of rolled observations (red ground tracks in Figure 1), mapping of thelaterally extensive subsurface returns found in the vicinity of the Phoenix site requires the inclusion of

nonrolled tracks. Nightside nonrolledtracks (e.g., Figure 3) exhibit lowersignal-to-noise ratio than rolled tracks(cf. Figures 2a and 3a), but thesubsurface returns have sufficientpower to identify them. Daysidenonrolled tracks (e.g., Figure 4)provide still lower signal-to-noiseratio, but as already mentioned, theyprovide a critical link betweenadjacent nightside tracks.

We found that the subsurface returnsshown in Figures 2–4 extendcontiguously across 81 of the 125radargrams included in our Phoenixstudy area. We discontinued ourtracing of the subsurface returnseither where the power fell to thenoise level or the feature isinterrupted or comingled withoff-nadir surface returns within eachradargram (e.g., Figure 2b). To the eastand west of the mapped area(see Figure 1b), a shallow, late returnpersists in the radargrams but its delaytime approaches that expected for

Figure 3. Portion of SHARAD nightside nonrolled observation 5724–01,inverted to extend from south to north across the Phoenix landing site andthe Green Valley. (a) Uninterpreted radargram. (b) Interpreted radargram,showing delay times for the MOLA modeled surface in blue and for thesubsurface return in yellow, with the latter delayed ~0.4 μs from the surfacereturn. “×” symbols show corresponding delay times from crossing daysideobservations. Inset shows map view of ground track. Nonrolled observationstypically have lower signal-to-noise ratio than rolled ones, but power ofsubsurface returns remains high enough to distinguish from sidelobes.

Figure 4. Portion of SHARADdayside nonrolled observation 12903–01 extend-ing from south to north across the Phoenix landing site and the Green Valley.(a) Uninterpreted radargram. (b) Interpreted radargram, showing delay timesfor the MOLA modeled surface in blue and for the subsurface return in yellow,with the latter delayed ~0.4 μs from the surface return. “×” symbols showcorresponding delay times from crossing nightside observations. Inset showsmap view of ground track. Dayside nonrolled observations typically have lowersignal-to-noise ratio than either nightside or rolled ones, but power of subsur-face returns remains high enough to distinguish from sidelobes, and thecrossing trajectories are critical for tracking features between observations.

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1941

sidelobes of the surface return, so wediscontinued mapping of thesubsurface return in those areas. Inaddition to tracing the subsurfacereturns, we determined the delay tothe surface along each ground trackusing MRO ephemeris data and the128ppd MOLA elevation map,displayed as a blue line in Figures 2b,3b, and 4b. Once the surface andsubsurface-return boundaries weredelineated on the radargrams, we usedthe analysis software to calculate thedelay time from the surface to thesubsurface returns, which we binnedalong track and interpolated betweentracks in an regular grid of500m×500m bins to obtain the mapin Figure 1b. Across the mapped area,the delay times vary between 0.18 μsand 0.78 μs, with a modal value of 0.41μs. Delay time of the subsurface returnscan be translated to depth via equation(1) using an appropriate value for thereal permittivity of the subsurfacematerials. In some studies, it has beenpossible to constrain the dielectricusing structural considerations [e.g.,Campbell et al., 2008; Carter et al.,2009a; Phillips et al., 2011]. Thatmethodcannot be applied to SHARAD data atthe Phoenix site, since there are noindications of structures exposingsubsurface layering that may bepresent. However, the presence of asubstantial quantity of ground iceextending to depths up to a few tens ofmeters is well supported byobservations from the Phoenix lander,GRS data, and thermal modeling[Mellon et al., 2009], and it is thereforereasonable to assume a realpermittivity consistent with a mixtureof water ice and basaltic regolith.Considering end-member cases, if thematerials at depth are entirely water ice(with real permittivity of 3.15), then

the depth to the detected interface ranges from 15 to 66 m, with a modal depth of 34 m. If the materials arepurely basaltic (with a real permittivity of 8), then the depth range is 9 to 41 m with a modal value of 22 m.Integrating the depths (using the end-member modal values) over the entire mapped area of 2912 km2 yields avolume of 64–99 km3 for the interval from the surface to the mapped subsurface returns.

As discussed in section 1, somewhat shallower and lower-power returns across other areas of the northernplains were found to coincide closely with the delay times of the first and second sidelobes of the surfacereturn (0.24 μs and 0.42 μs), with about twice the expected power. For SHARAD returns at the Phoenix site,

Figure 5. Portion of SHARAD nightside rolled observation 21762–01, invertedto extend from south to north across Gale crater and the Curiosity landingsite. (a) Uninterpreted radargram. (b) Interpreted radargram, showing delaytimes for the MOLA modeled surface in blue and for the putative subsurfacereturns in yellow,with the latter delayed ~2–3 μs from the surface return. Insetshows ground track with putative subsurface returns identified along adja-cent tracks over aMOLA color shaded relief basemap. Curiosity landing ellipseis shown with a black line. (c) Clutter simulation produced from a swath of theMOLA modeled surface along the ground track of the observation. Returnsarriving late or early relative to the nadir surface (yellow line) are off-nadirsurface returns, many of which also appear in the SHARAD radargram(white arrows). No late returns occur where the subsurface returns are identi-fied in the SHARAD radargram (yellow arrows). (d) Clutter simulation producedfrom a swath of the HRSC modeled surface along the ground track of theobservation. Many more off-nadir surface returns are identifiable relative tothat for the MOLA surface, including ones corresponding to the putativesubsurface returns identified in the SHARAD radargram (yellow arrows).

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1942

the delay times to the subsurface return are more variable and predominantly greater than that of the firstsidelobe, but the modal delay time is very near that of the second sidelobe, so one must include considerationof the power, which we do for the collection of radargrams without interpolation (Figure 1c). Thesubsurface-return power has a modal value of �9 dB relative to that of the surface return, with all valueswell above that expected for the second sidelobe (�34 dB) and most values above that expected forthe first sidelobe (�20 dB). In addition, we conducted processing tests using different weighting functions(uniform, Hann, Hamming, and Chebychev) to assess their relative effects, and in no case was thepower of the putative detections substantially suppressed. These results lend greater confidence to aninterpretation of the returns as having a subsurface source.

3.2. Nondetections of Subsurface Returns at Gale Crater and Other Landing Sites

Our search for subsurface returns at landing sites other than that of the Phoenix Mars lander has so far provenfruitless. In the other cases where the region surrounding the landing site is relatively flat lying (the VikingLander 1 site in Chryse Planitia, the Viking Lander 2 site in Utopia Planitia, the Mars Pathfinder site in AresVallis, the Spirit site in Gusev crater, and the Opportunity site in Meridiani Planum), the available SHARADobservations show a distinct surface return with no apparent subsurface returns within the study areas(see section 2 and Table 1). The lack of detections in these regions may be due in part to the sparse coverageat each of these sites except for Meridiani Planum, which was targeted extensively by SHARAD. There, wefind some of the highest-power radar returns from the surface of Mars, but the subsequent records showvery low power and no strong indicators of subsurface returns throughout the region (see, however,Watterset al. [2014]). In the cases where the region surrounding the landing site has substantial topographicrelief (the actual Curiosity site in Gale crater and the proposed Curiosity sites in Holden crater, Eberswaldecrater, and Mawrth Vallis), returns subsequent to the nadir surface return are common, but we find that theyare largely explained as surface returns from off-nadir topographic features. Typically, the off-nadirtopography is well characterized by the MOLA modeled surface, but this is not the case for Gale crater. InSHARAD observations of Gale crater (e.g., Figure 5a), we identified a laterally extensive feature at delays of~1–3 μs below the northern crater floor (Figure 5b). Clutter simulations produced from the MOLA modeledsurface show extensive off-nadir returns in the area, but none of them correspond to that of the SHARADfeature we identified (Figure 5c), and thus, our preliminary analysis supported the interpretation of thefeature as a subsurface return [Putzig et al., 2012]. However, we implemented clutter simulations usingsurface-elevation models produced from HRSC images [Neukum et al., 2004], and simulations produced froma Gale crater HRSC surface-elevation model at 75 m per pixel reveal that the feature we identified can in factbe attributed to an off-nadir return (Figure 5d). This example demonstrates that high-resolution topographyis essential to distinguishing possible subsurface returns from surface clutter, especially in limited areas withrough terrain and few observations.

Figure 6. Map of roughness parameter [Campbell et al., 2013] derived from SHARAD nonrolled observations on MRO orbits2000–25,999 between 70°S and 70°N and binned at 10 ppd in a Mollweide projection centered on 0° longitude. Whiteboxes indicate landing sites for Viking 1 (V1), Viking 2 (V2), Pathfinder (Pa), Opportunity (O), Spirit (S), Phoenix (Ph), Curiosity(C), and the proposed MSL sites in Holden crater (H), Eberswalde crater (E), and Mawrth Vallis (M).

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1943

3.3. SHARAD-Derived Roughness at All Landing Sites

In Figure 6, we present a map of the roughness parameter [Campbell et al., 2013] between 70°S and 70°N,produced from 8000 SHARAD nonrolled observations on MRO orbits 2000–25,999 and binned at 10 ppd.Locations of the landing sites analyzed in this work are indicated with white boxes. An example of how wespecify areas for regional and local analyses is presented in Figure 7a, using the Opportunity site inMeridiani Planum. Results from SHARAD are infilled (Figure 7c) and compared to those obtained previouslyfrom MOLA at the 0.6 km scale (Figure 7d) [Kreslavsky and Head, 2002]. While the specific values from thetwo roughness techniques are not directly comparable, the patterns of smooth and rough terrains arelargely consistent. In general, the region around the Opportunity landing site is shown to be extremelysmooth on the scales of ~10 m and ~1 km provided by SHARAD and MOLA, respectively.

To compare radar roughness of the landing sites included in this study, we created histograms of theroughness parameter in both roughness AOIs at each site and for the entire midlatitude map (Figure 8).Peak (modal) values for each site can be discerned from the histograms and are also listed in Table 1. Atboth regional and local scales, the Opportunity site is smoothest with a peak value of 2.5 and theMawrth site is roughest with a peak value of 4.0. Of the actual past landing sites, only those of Viking 2 andSpirit exhibit regional peak roughness greater than that of the entire midlatitude band. The sites showdiverse distributions of roughness, with Opportunity and Viking 1 having the narrowest distributions andMawrth, Eberswalde, and Curiosity having the broadest distributions. For the most part, local and regionalresults are consistent with each other, with the notable exception of the Curiosity site. Here, the localroughness results are limited to 4 SHARAD tracks that disproportionately sample the Gale crater rim and

Figure 7. Maps of Meridiani Planum and surrounding areas. X symbols mark the Opportunity landing site in each map.(a) SHARAD roughness parameter at 10 ppd, with white boxes showing the regional (3 × 3°) and local (0.7 × 0.7°) areas ofinterest used for histograms in Figure 8. (b) Shaded relief map of MOLA elevation [Smith et al., 2001]. (c) Same as Figure 7awith nearest-neighbor mean infilling applied. (d) MOLA 0.6 km scale roughness from Kreslavsky and Head [2002], wherevalues in map bins represent typical slopes in radians (M. Kreslavsky, personal communication, 2013).

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1944

Aeolis Mons (Mount Sharp) and are thus biasedtoward higher values. Future efforts to includeadditional coverage in roughness parameter canbe expected to reduce the local peak value forthe Curiosity site.

4. Discussion

The Phoenix Mars mission was motivated in largepart by the inference from GRS results ofabundant water ice in the near-subsurface allacross the northern plains of Mars. Selection ofpossible landing sites began by choosing a bandof latitudes where the ice was believed to beshallow enough to be accessible by the lander’sscoop but deep enough to retain a surface layerof dry soil. Initially, the Phoenix team consideredpotential landing sites over a broad range oflongitudes. When images of the sites began toarrive from HiRISE, the discovery of widespreadhigh rock abundance raised attendant concernsabout landing safety for the spacecraft.Eventually, the landing site was chosen based onthat region having relatively low rockabundance in HiRISE imagery and meeting thescientific objectives of the mission. EarlySHARAD observations were limited and did notprovide discriminating information towardlanding-site selection. The subsurface detectionspresented in this work are unique across theentire zone of latitudes that were being

considered, and in retrospect, the results may well have provided some scientific motivation towardchoosing this same location.

The simplest explanation for the SHARAD subsurface returns mapped at the Phoenix site is that theyrepresent the base of ground ice in this area, but if that is the case, then it is more difficult to explain why thebase of ground ice is not detected more broadly in the northern plains. An ongoing study does suggestthe presence of a widespread deposit of water ice in Arcadia Planitia [Bramson et al., 2014], but the SHARADdetections there are deeper and have higher power that is suggestive of relatively pure ice, perhapsdeposited as snowfall. More broadly, prior thermal modeling and observational data showing polygonalpatterns at scales of ~5–20 m have suggested that the base of ground ice at the Phoenix site and across thenorthern plains extends to depths of at least ~10 m [Mellon et al., 2008b]. A broad range of mechanisms foremplacing ground ice on Mars have been proposed, from freezing of an ocean or ancient flood waters[Carr, 1996] to snow accumulated at times of higher obliquity [Head et al., 2005]—perhaps coeval with dustand thereby forming the observed high-latitudemantle deposits [Mustard et al., 2001]–to vapor diffusion intoan initially dry regolith. At the Phoenix site, the lower power of the SHARAD detections suggests that theice is largely interstitial to the regolith, favoring the vapor-diffusion mechanism. In that case, ground icemay be emplaced by some combination of atmospheric water vapor diffusing downward along seasonalthermal waves due to insolation and ground-water vapor diffusing upward along the Martian geothermalgradient [Clifford, 1993; Mellon and Jakosky, 1993, 1995]. The level where the vapor-density gradientfrom each source converges is the maximum depth to which diffusion of atmospheric water vapor alonemay emplace ground ice. Below this point, any additional ground ice would necessarily be due tounderplating from upward diffusion of ground water. Using the thermal model of Mellon et al. [2004] andmaking reasonable assumptions of physical properties, we find a mean-annual flux-convergence depth of~15–20 m at the Phoenix latitude (upper curve in Figure 9a), somewhat shallower than the SHARAD

1000

100

10

1

2 3 4 5 6 7 8 9

roughness parameter

coun

t in

bins

of 0

.25

coun

t in

bins

of 0

.25

(b)

(a)

1000

100

10

1

Figure 8. Histograms of the SHARAD roughness parameterin bins of ρ = 0.25. Black curves show midlatitude valuesmapped at 10 ppd (Figure 6) and scaled by a factor of 50.Color curves show values mapped at (a) 50 ppd in 3 × 3°regions and (b) 200 ppd in 0.7 × 0.7° local areas surroundingeach site. Legend is sorted in ascending order of peak rough-ness in Figure 8a. See Figure 6 for site locations.

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1945

detections. Greater depths areachieved for lower geothermalgradients, an assumption that issupported by the lack of basaldeflection observed beneath PlanumBoreum [Phillips et al., 2008]. Onemight also argue that the maximumrather than the mean fluxconvergence could control the baseof ground ice (lower curve inFigure 9a), but then the lack ofregionally broader SHARADdetections remains unexplained. Onepossibility is that as ground ice hasbeen emplaced over time by vapordiffusion and freezing within theupper few m throughout thenorthern plains, it has reachedgreater depths here due to a higherrate of aggradation of surfacesediments in the low-lying GreenValley. Another possibility is thatcertain features or conditions uniqueto the Phoenix region allowed vapordiffusion and deposition of groundice at greater depths than elsewhere.One of those features might be alocalized source of ground water atdepth. In either case, one mightexpect to find a SHARAD basal-icedetection only in areas with such adeeper base of ground ice, sincesidelobes of the surface return wouldobfuscate returns from shallower(~10–15m) depths (Figure 9b). Athird possibility is that the ground iceextends to similar depths morebroadly, but its base is only detectedhere because of reduced loss of theradar signal resulting from somecombination of the uniquely low rockabundance in this area and alocalized lack of radar-attenuatingmaterials at depth.

We cannot rule out the possibility thatground ice is shallower at the Phoenixsite as well as elsewhere and that the

base of ground ice is not the source of the radar detection. Rather, it may be due to some other geologiccontact with a high dielectric contrast, such as sediments over a buried lava flow. In seeking alternativeexplanations, we found a geographic correspondence between the SHARAD subsurface returns at thePhoenix site and a geomorphic unit associated with the outer ejecta blanket of Heimdal crater [Heet et al.,2009] (Figure 10). Heimdal has been classified as a double-layered ejecta crater, for which the outer ejectlayers typically range from a few meters to tens of meters in thickness [Boyce and Mouginis-Mark, 2006].However, the idea that the subsurface returns could be the base of an ejecta blanket that extends to the

Figure 9. (a) Modeled depths of the mean-annual (solid curve) and maxi-mum-annual (dashed curve) convergence point between downwarddiffusion of atmospheric water vapor and upward diffusion of ground-water vapor at the latitude of the Phoenix lander (68.2°N). The flux-conver-gence point controls the maximum depth to which atmospheric vapor maydeposit ground ice. Although earlier studies have suggested a Martiangeothermal gradient of 30 mW m�2 K�1 (bold vertical line), SHARAD results[Phillips et al., 2008] suggest a much lower value (e.g., 10 mW m�2 K�1,dashed vertical line). The thermalmodel [Mellon et al., 2004] assumes a surfacethermal inertia of 200 tiu and an albedo of 0.2. (b) Cartoon cross sectiondepicting different scenarios for the SHARAD detection at the Phoenix land-ing site. The detected interface (dark grey line at depth of ~30 m) maycorrespond to a relatively deep base of ground ice or to some other geologiccontact with a dielectric contrast (e.g., sediments over bedrock such as a lavaflow). Overlain curve depicts the theoretical power response of the surfacealone to the SHARAD signal and indicates that later returns from a possibleshallower base of ground ice as might occur elsewhere in the northern plains(light grey line at ~10 m depth) may be obfuscated by sidelobes (S1, S2) ofthe surface return (S0) while returns from the deeper interface could arrivesufficiently later in delay time to be distinguishable.

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1946

surface can be dismissed solely on geometric considerations, as the size of Heimdal is insufficient by well overan order of magnitude to account for the volume of material overlying the mapped subsurface returns.Another consideration is that the geomorphic unit is characterized by low rock abundance inferred to be aresult of the impact process [Heet et al., 2009], which could allow for reduced scattering of the radar signaland thereby explain the geographically restricted detection of an interface at depth that may actually extendwell beyond the bounds of the unit and the detection area. Given all of these considerations, we favor ascenario of a locally deeper base of ground ice that may extend beyond the mapped area, with the radarreflections made more detectable in the mapped area by the relatively low rock abundance of Heimdal’souter ejecta blanket.

The lack of subsurface detections at the other landing sites may be due to a variety of factors. Despite thepresence of layering on scales nominally resolvable by SHARAD at many of the sites, it is possible that there isnot sufficient dielectric contrast between the layers to yield a detectable reflection. Another potentialexplanation is that the radar signal is being attenuated within the uppermost layers, due to either absorptionor scattering. Stillman and Grimm [2011] have suggested that surfaces older than Middle Amazonian havebeen subject to a larger extent than more recent ones to processes leading to the presence of adsorbedwater as well as pervasive fracturing, features which may lead to increased absorption and scattering of radarsignals and thereby account for the fact that SHARAD subsurface detections have largely been restrictedto younger terrains. While they implied that even low-temperature alteration by water (such as in fluvialand outwash deposits) will preclude detections, SHARAD returns apparently from the base of the EarlyAmazonian Vastitas Borealis Formation [Campbell et al., 2008] suggest that may not be the case, and Ehlmannet al. [2011] have suggested that hydrothermal alteration is needed to produce clay minerals (e.g., smectite)that are sufficiently attenuating to the radar. In any case, crater counting yields age estimates of 0.6 Ga forHeimdal ejecta materials, making the Phoenix site the youngest surface of any landing site to date [Heet et al.,

Unit map: NASA/JPL-Caltech/Washington Univ. St. Louis/JHU APL/Univ. of Arizona

SHARAD det

ectio

n ar

ea

Heimdal

Phoenix

Figure 10. Map of geomorphic units of Arvidson et al. [2008] in the region surrounding the Phoenix Mars landing site, overlain with an outline (red) showing theextents of the SHARAD subsurface returns mapped in Figure 1. The SHARAD detections largely overlap with the Lowland Bright geomorphic unit that extendsover much of the outer ejecta blanket of Heimdal crater (dashed black line).

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1947

2009]. While the counts for the eroded portions of the outer ejecta unit yield a significantly older age of 3.0 Ga[Heet et al., 2009], the Heimdal impact may have produced temperatures and pressures sufficient to vaporizenear-surface water ice, shatter rocks, and otherwise alter surface materials [Wrobel et al., 2006] in ways thatcould reduce their attenuation of radar signals.

SHARAD roughness at the Phoenix site is about average for the northern plains (see Figure 6), but landing-safety concerns were of a different scale, focused on meter-scale boulders found in greater abundance at theother sites being considered. The scales of roughness characterized by SHARAD (~10–100 m) come moreinto play for factors such as rover trafficability and access to outcrops. From a SHARAD roughness standpoint,the ideal site for a roving spacecraft is onewith a broad distribution of roughness that allows both a safe landinglocation with low roughness and features of scientific interest with high roughness. The Curiosity site fits thiscriterion quite well, as do the Spirit site and the other proposed MSL sites (see Figure 8).

5. Conclusions

The SHARAD instrument has mapped an interface at depths of 9–66 m extending over a 2900 km2 area thatincludes the landing site for the Phoenix Mars mission. This interface may represent the base of ground ice,although no similar detection has been found more broadly in the northern plains, and other geologiccontacts cannot be ruled out. Relatively rapid aggradation or unique near-surface conditions in the PhoenixGreen Valley area may have led to deeper ground ice and/or lower loss of the radar signal, perhaps explainingthe geographically limited nature of this detection. Radar reflections from a shallower (~10–15m) base ofground ice are likely to be obfuscated by sidelobes of the surface return.

Landing sites other than that of Phoenix have not yielded any clear evidence of radar returns from subsurfaceinterfaces, likely due to some combination of feeble dielectric contrasts at volcanic and sedimentary layerinterfaces and radar signal attenuation by absorption and/or scattering in the uppermost strata. A putativedetection from beneath the floor of Gale crater in and around the Curiosity landing site has proven to beattributable to an off-nadir return from a surface feature captured in a HRSC clutter simulation but not in oneproduced from MOLA data. This result provides an important cautionary example that should be consideredwhen interpreting radar results in any area with substantial topographic relief.

Analysis of near-surface roughness from SHARAD returns at ~10–100m scales complements both that fromHiRISE imagery and surface-elevation models at ~1m scales and that from MOLA data at ~1 km scales. Theintermediate scales characterized by SHARAD provide unique information relevant to landing-site trafficabilityand to local geologic features and processes. While landing-safety considerations typically require smoothsurfaces, features of scientific interest (i.e., rocks, outcrops, and geologic structures) are often associated withrougher surfaces, so optimal landing sites—especially for roving spacecraft—will generally exhibit a broaddistribution of surface roughness. We found such distributions atmany of the past landing sites, including thoseof the Curiosity and Spirit spacecraft. The global data set of roughness results from SHARAD could contributetoward improved assessments of sites being proposed for future landed missions.

ReferencesArvidson, R., et al. (2008), Mars Exploration Program 2007 Phoenix landing site selection and characteristics, J. Geophys. Res., 113, E00A03,

doi:10.1029/2007JE003021.Boyce, J. M. and P. J. Mouginis-Mark (2006), Martian craters viewed by the Thermal Emission Imaging System instrument: Double-layered

ejecta craters, J. Geophys. Res., 111, E10005, doi:10.1029/2005JE002638.Bramson, A. M., S. Byrne, N. E. Putzig, S. Mattson, J. J. Plaut, and J. W. Holt (2014), Thick, excess water ice in Arcadia Planitia, Mars, Lunar Planet.

Sci., XLV, Abstract 2120.Campbell, B. A., N. E. Putzig, L. M. Carter, and R. J. Phillips (2011), Autofocus correction of phase distortion effects on SHARAD echoes, IEEE

Geosci. Remote Sens. Lett., 8, 939–942, doi:10.1109/LGRS.2011.2143692.Campbell, B. A., N. E. Putzig, L. M. Carter, G. A. Morgan, R. J. Phillips, and J. J. Plaut (2013), Roughness and near-surface density of Mars from

SHARAD radar echoes, J. Geophys. Res. Planets, 118, 436–450, doi:10.1002/jgre.20050.Campbell, B. A., N. E. Putzig, F. J. IIFoss, and R. J. Phillips (2014), SHARAD signal attenuation and delay offsets due to the Martian ionosphere,

IEEE Geosci. Remote Sens. Lett., 11, 632–635, doi:10.1109/LGRS.2013.2273396.Campbell, B., L. Carter, R. Phillips, J. Plaut, N. Putzig, A. Safaeinili, R. Seu, D. Biccari, A. Egan, and R. Orosei (2008), SHARAD radar sounding of the

Vastitas Borealis Formation in Amazonis Planitia, J. Geophys. Res., 113, E12010, doi:10.1029/2008JE003177.Carr, M. H. (1996), Water on Mars, pp. 248, Oxford Univ. Press, New York.Carter, L. M., B. A. Campbell, J. W. Holt, R. J. Phillips, N. E. Putzig, S. Mattei, R. Seu, C. H. Okubo, and A. F. Egan (2009a), Dielectric properties of

lava flows west of Ascraeus Mons, Mars, Geophys. Res. Lett., 36, L23204, doi:10.1029/2009GL041234.

AcknowledgmentsThe Shallow Radar (SHARAD) instrumentwas provided to NASA’s MarsReconnaissance Orbiter (MRO) missionby the Italian Space Agency (ASI) and isoperated from the Sapienza University ofRome. The MRO mission is managed bythe Jet Propulsion Laboratory, CaliforniaInstitute of Technology, for the NASAScienceMission Directorate, Washington,DC. We are grateful to all the people atthese and other institutions who havebeen instrumental in the success of theSHARAD experiment. We also thankSeisWare International Inc. for access toand support of the SeisWare interpreta-tion software, which serves an integralrole in our mapping efforts. Data used inthis work are available from the NASAPlanetary Data System. Insightfulcomments fromMisha Kreslavsky and ananonymous reviewer greatly improvedthe manuscript.

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1948

Carter, L. M., et al. (2009b), Shallow Radar (SHARAD) sounding observations of the Medusae Fossae Formation Mars, Icarus, 199, 295–302,doi:10.1016/j.icarus.2008.10.007.

Christensen, P. R., N. S. Gorelick, G. L. Mehall, and K. C. Murray (2013), THEMIS Public Data Releases, Planetary Data System node, Arizona StateUniversity. [Available at http://themis-data.asu.edu.]

Clifford, S. M. (1993), A model for the hydrologic and climatic behavior of water on Mars, J. Geophys. Res., 98, 10,973–11,016, doi:10.1029/93JE00225.

Ehlmann, B. L., J. F. Mustard, S. L. Murchie, J.-P. Bibring, A. Meunier, A. A. Fraeman, and Y. Langevin (2011), Subsurface water and clay mineralformation during the early history of Mars, Nature, 479, 53–60.

Golombek, M. P., A. F. C. Haldemann, R. A. Simpson, R. L. Fergason, N. E. Putzig, R. E. Arvidson, J. F. Bell III, and M. T. Mellon (2008), Martiansurface properties from joint analysis of orbital, Earth-based, and surface observations, in The Martian Surface: Composition, Mineralogy,and Physical Properties, edited by J. F. Bell III, pp. 468–497, Cambridge Univ. Press, Cambridge, U. K.

Golombek, M., J. Grant, D. Kipp, A. Vasavada, R. Kirk, R. Fergason, P. Bellutta, F. Calef, K. Larsen, and Y. Katayama (2012), Selection of the MarsScience Laboratory landing site, Space Sci. Rev., 170, 641–737, doi:10.1007/s11214-012-9916-y.

Head, J. W., G. Neukum, R. Jaumann, H. Hiesinger, E. Hauber, M. Carr, P. Masson, B. Foing, H. Hoffmann, and M. Kreslavsky (2005), Tropical tomid-latitude snow and ice accumulation, flow and glaciation on Mars, Nature, 434, 346–351.

Heet, T. L., R. E. Arvidson, S. C. Cull, M. T. Mellon, and K. D. Seelos (2009), Geomorphic and geologic settings of the Phoenix Lander missionlanding site, J. Geophys. Res., 114, E00E04, doi:10.1029/2009JE003416.

Holt, J. W., M. E. Peters, S. D. Kempf, D. L. Morse, and D. D. Blankenship (2006), Echo source discrimination in single-pass airborne radarsounding data from the Dry Valleys, Antarctica: Implications for orbital sounding of Mars, J. Geophys. Res., 111, E06S24, doi:10.1029/2005JE002525.

Holt, J. W., A. Safaeinili, J. J. Plaut, J. W. Head, R. J. Phillips, R. Seu, S. D. Kempf, P. Choudhary, D. A. Young, and N. E. Putzig (2008), Radarsounding evidence for buried glaciers in the southern mid-latitudes of Mars, Science, 322, 1235–1238, doi:10.1126/science.1164246.

Hynek, B. M. (2004), Implications for hydrologic processes on Mars from extensive bedrock outcrops throughout Terra Meridiani, Nature, 431,156–159, doi:10.1038/nature02902.

Kreslavsky, M. A., and J. W. IIIHead (2000), Kilometer-scale roughness of Mars: Results from MOLA data analysis, J. Geophys. Res., 105,26,695–26,711, doi:10.1029/2000JE001259.

Kreslavsky, M. A., and J. W. Head III (2002), Mars: Nature and evolution of young latitude-dependent water-ice-rich mantle, Geophys. Res. Lett.,29(15), 1719, 10.1029/2002GL015392.

Mellon, M. T., and B. M. Jakosky (1993), Geographic variations in the thermal and diffusive stability of ground ice on Mars, J. Geophys. Res., 98,3345–3364, doi:10.1029/92JE02355.

Mellon, M. T., and B. M. Jakosky (1995), The distribution and behavior of Martian ground ice during past and present epochs, J. Geophys. Res.,100, 11,781–11,799, doi:10.1029/95JE01027.

Mellon, M. T., W. C. Feldman, and T. H. Prettyman (2004), The presence and stability of ground ice in the southern hemisphere of Mars, Icarus,169, 324–340.

Mellon, M. T., W. V. Boynton, W. C. Feldman, R. E. Arvidson, T. N. Titus, J. L. Bandfield, N. E. Putzig, and H. G. Sizemore (2008a), A prelandingassessment of the ice table depth and ground ice characteristics in Martian permafrost at the Phoenix landing site, J. Geophys. Res., 113,E00A25, doi:10.1029/2007JE003067.

Mellon, M. T., R. E. Arvidson, J. J. Marlow, R. J. Phillips, and E. Asphaug (2008b), Periglacial Landforms at the Phoenix Landing Site and theNorthern Plains of Mars, J. Geophys. Res., 113, E00A23, doi:10.1029/2007JE003039.

Mellon, M. T., et al. (2009), Ground ice at the Phoenix Landing Site: Stability state and origin, J. Geophys. Res., 114, E00E07, doi:10.1029/2009JE003417.

Mustard, J. F., C. D. Cooper, and M. K. Rifkin (2001), Evidence for recent climate change on Mars from the identification of youthfulnear-surface ground ice, Nature, 412, 411–414.

Neukum, G., R. Jaumann, and the HRSC Co-Investigator and Experiment Team (2004), HRSC: The High Resolution Stereo Camera of MarsExpress, in Mars Express: The Scientific Payload, vol. 1240, edited by A. Wilson and A. Chicarro, pp. 17–35, ESA Special Publication, Noordwijk,Netherlands.

Phillips, R. J., et al. (2008), Mars north polar deposits: Stratigraphy, age, and geodynamical response, Science, 320, 1182–1185, doi:10.1126/science.1157546.

Phillips, R. J., B. J. Davis, K. L. Tanaka, S. Byrne, M. T. Mellon, N. E. Putzig, R. M. Haberle, M. A. Kahre, B. A. Campbell, and L. M. Carter (2011),Massive CO2 ice deposits sequestered in the south polar layered deposits of Mars, Science, 332, 838–841, doi:10.1126/science.1203091.

Plaut, J. J., A. Safaeinili, J. W. Holt, R. J. Phillips, J. W. Head III, R. Seu, N. E. Putzig, and A. Frigeri (2009), Radar evidence for ice in lobate debrisaprons in the mid-northern latitudes of Mars, Geophys. Res. Lett., 36, L02203, doi:10.1029/2008GL036379.

Putzig, N. E., R. J. Phillips, J. W. Head, B. A. Campbell, A. F. Egan, J. J. Plaut, L. M. Carter, R. Seu, M. T. Mellon, and the SHARAD Team (2009a),Do shallow radar soundings reveal possible near-surface layering throughout the northern lowlands of Mars?, Lunar Planet. Sci., XL,Abstract 2477.

Putzig, N. E., R. J. Phillips, J. J. Plaut, M. T. Mellon, J. W. Head, B. A. Campbell, L. M. Carter, A. Egan, and R. Seu (2009b), Shallow Radar soundingsof the southern highlands of Mars, Eos Trans. AGU, 90(52), Fall Meet. Suppl., Abstract P13B-1280.

Putzig, N. E., B. A. Campbell, R. J. Phillips, and M. T. Mellon (2012), SHARAD sounding and surface roughness of once and future Mars landingsites, Lunar Planet. Sci., XLIII, Abstract.

Safaeinili, A., R. Phillips, N. Putzig, J. Plaut, R. Seu, D. Biccari, J. Holt, S. Smrekar, and D. Nunes (2008), Radar detection of a subsurface horizon atthe Phoenix landing site, Eos Trans. AGU, 89(53), Fall Meet. Suppl., Abstract P32B-06.

Seu, R., et al. (2007a), Accumulation and erosion of Mars’ south polar layered deposits, Science, 317, 1715–1718, doi:10.1126/science.1144120.Seu, R., et al. (2007b), SHARAD sounding radar on the Mars Reconnaissance Orbiter, J. Geophys. Res., 112, E05S05, doi:10.1029/2006JE002745.Smith, D. E., H. V. Frey, J. B. Carvin, H. J. Zwally, F. G. Lemoine, D. D. Rowlands, J. B. Abshire, R. S. Afzal, X. Sun, and M. T. Zuber (2001), Mars

Orbiter Laser Altimeter: Experiment summary after the first year of global mapping of Mars, J. Geophys. Res., 106, 23,689–23,722,doi:10.1029/2000JE001364.

Stillman, D. E. and R. E. Grimm (2011), Radar penetrates only the youngest geological units onMars, J. Geophys. Res., 116, E03001, doi:10.1029/2010JE003661.

Watters, T. R., C. J. Leuschen, B. A. Campbell, J. A. Grant, M. Ga, N. E. Putzig, and R. J. Phillips (2014), MARSIS Subsurface radar sounding ofMeridiani Planum, Mars: Implications for the properties of the plains deposits, Lunar Planet. Sci., XLV, Abstract 2521.

Wrobel, K., P. Schultz, and D. Crawford (2006), An atmospheric blast/thermal model for the formation of high-latitude pedestal craters,Meteorit. Planet. Sci., 41, 1539–1550, doi:10.1111/j.1945-5100.2006.tb00434.x.

Journal of Geophysical Research: Planets 10.1002/2014JE004646

PUTZIG ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1949