Estimating erosional exhumation on Titan from drainage network...

17
Estimating erosional exhumation on Titan from drainage network morphology Benjamin A. Black, 1 J. Taylor Perron, 1 Devon M. Burr, 2 and Sarah A. Drummond 2 Received 20 March 2012; revised 15 June 2012; accepted 26 June 2012; published 15 August 2012. [1] Drainage networks on Titan, Earth, and Mars provide the only known examples of non-volcanic fluvial activity in our solar system. The drainage networks on Titan are apparently the result of a methane-ethane cycle similar to Earths water cycle. The scarcity of impact craters and the uneven distribution of fluvial dissection on Titan suggest that the surface may be relatively young. The purpose of this study is to assess the importance of erosion relative to other plausible mechanisms of resurfacing such as tectonic deformation, cryovolcanism, or deposition of aerosols. We present a new method, based on a measure of drainage network shape known as the width function, to estimate cumulative erosion into an initially rough surface. We calibrate this method with a numerical landscape evolution model, and successfully test the method by applying it to river networks on Earth with different exhumation histories. To estimate erosional exhumation on Titan, we mapped fluvial networks in all Synthetic Aperture Radar swaths obtained by the Cassini spacecraft through T71. Application of our method to the most completely imaged drainage networks indicates that for two of four regions analyzed, Titans fluvial networks have produced only minor erosional modification of the surface. For the best-constrained region in the northern high latitudes, we find that fluvial networks reflect spatially averaged erosion of more than 0.4% but less than 9% of the initial topographic relief. This result implies either a recent, non-fluvial resurfacing event or long-term fluvial incision rates that are slow relative to the rate of resurfacing. Citation: Black, B. A., J. T. Perron, D. M. Burr, and S. A. Drummond (2012), Estimating erosional exhumation on Titan from drainage network morphology, J. Geophys. Res., 117, E08006, doi:10.1029/2012JE004085. 1. Introduction [2] Prior to the arrival of the Cassini-Huygens spacecraft in orbit around Saturn in 2004, the surface of Titan was the subject of wide-ranging interest and speculation. Gerard Kuiper was the first scientist to detect methane in Titans atmosphere [Kuiper, 1944]. Methane is near its triple-point (91 K and 0.2 bars) at surface conditions on Titan of 94 K and 1.47 bars [Fulchignoni et al., 2005]. This observation led Sagan and Dermott [1982] to envision methane oceans on Titan hundreds of meters deep. Lunine et al. [1983] suggested that methane photolysis might instead result in an ocean dominated by ethane, of even greater depth. Other researchers, however, concluded that global methane oceans seemed unlikely [Eshleman et al., 1983]. [3] As a result of the thick organic-rich haze that shrouds Titan, the actual landscape on the surface of the moon remained mysterious until the first spectacular results from the Cassini spacecraft. While Titan does not host a global surface ocean, Synthetic Aperture Radar (SAR) images from the Cassini Titan Radar Mapper [Elachi et al., 2004] show 100 km-scale drainage networks with branching morpholo- gies (Figure 1). This morphology was revealed in even greater detail and at finer scales for networks imaged by the Huygens Probe Descent Imager/Spectral Radiometer [Tomasko et al., 2005]. The negative topographic relief associated with the Huygens networks (that is, they are valleys) [Tomasko et al., 2005; Soderblom et al., 2007] further suggests that they were created by fluvial erosion [Perron et al., 2006; Moore and Pappalardo, 2011; Langhans et al., 2012]. These networks were most likely carved by a mixture of liquid methane [Kouvaris and Flasar, 1991] and possibly ethane [Lunine et al., 1983; Brown et al., 2008] as they eroded bedrockcomposed of water ice [Griffith et al., 2003; McCord et al., 2006; Rodriguez et al., 2006], perhaps with a sedimentary veneer of organic atmo- spheric precipitates [Yung et al., 1984; McKay et al., 2001; Lavvas et al., 2008; Krasnopolsky, 2009]. Near the poles, river valleys feed into typically RADAR-dark hydrocarbon lakes; the RADAR-darkness of the lakes most likely corre- sponds to lower backscatter associated with a smoother 1 Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA. 2 Earth and Planetary Sciences Department, University of Tennessee, Knoxville, Tennessee, USA. Corresponding author: B. A. Black, Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Ave., Bldg. 54-810, Cambridge, MA 02139, USA. ([email protected]) ©2012. American Geophysical Union. All Rights Reserved. 0148-0227/12/2012JE004085 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, E08006, doi:10.1029/2012JE004085, 2012 E08006 1 of 17

Transcript of Estimating erosional exhumation on Titan from drainage network...

Page 1: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

Estimating erosional exhumation on Titan from drainagenetwork morphology

Benjamin A. Black,1 J. Taylor Perron,1 Devon M. Burr,2 and Sarah A. Drummond2

Received 20 March 2012; revised 15 June 2012; accepted 26 June 2012; published 15 August 2012.

[1] Drainage networks on Titan, Earth, and Mars provide the only known examples ofnon-volcanic fluvial activity in our solar system. The drainage networks on Titan areapparently the result of a methane-ethane cycle similar to Earth’s water cycle.The scarcity of impact craters and the uneven distribution of fluvial dissection on Titansuggest that the surface may be relatively young. The purpose of this study is to assessthe importance of erosion relative to other plausible mechanisms of resurfacing suchas tectonic deformation, cryovolcanism, or deposition of aerosols. We present a newmethod, based on a measure of drainage network shape known as the width function, toestimate cumulative erosion into an initially rough surface. We calibrate this method witha numerical landscape evolution model, and successfully test the method by applying it toriver networks on Earth with different exhumation histories. To estimate erosionalexhumation on Titan, we mapped fluvial networks in all Synthetic Aperture Radar swathsobtained by the Cassini spacecraft through T71. Application of our method to the mostcompletely imaged drainage networks indicates that for two of four regions analyzed,Titan’s fluvial networks have produced only minor erosional modification of the surface.For the best-constrained region in the northern high latitudes, we find that fluvial networksreflect spatially averaged erosion of more than 0.4% but less than 9% of the initialtopographic relief. This result implies either a recent, non-fluvial resurfacing event orlong-term fluvial incision rates that are slow relative to the rate of resurfacing.

Citation: Black, B. A., J. T. Perron, D. M. Burr, and S. A. Drummond (2012), Estimating erosional exhumation on Titan fromdrainage network morphology, J. Geophys. Res., 117, E08006, doi:10.1029/2012JE004085.

1. Introduction

[2] Prior to the arrival of the Cassini-Huygens spacecraftin orbit around Saturn in 2004, the surface of Titan was thesubject of wide-ranging interest and speculation. GerardKuiper was the first scientist to detect methane in Titan’satmosphere [Kuiper, 1944]. Methane is near its triple-point(91 K and 0.2 bars) at surface conditions on Titan of �94 Kand 1.47 bars [Fulchignoni et al., 2005]. This observationled Sagan and Dermott [1982] to envision methane oceanson Titan hundreds of meters deep. Lunine et al. [1983]suggested that methane photolysis might instead result inan ocean dominated by ethane, of even greater depth. Otherresearchers, however, concluded that global methane oceansseemed unlikely [Eshleman et al., 1983].

[3] As a result of the thick organic-rich haze that shroudsTitan, the actual landscape on the surface of the moonremained mysterious until the first spectacular results fromthe Cassini spacecraft. While Titan does not host a globalsurface ocean, Synthetic Aperture Radar (SAR) images fromthe Cassini Titan Radar Mapper [Elachi et al., 2004] show100 km-scale drainage networks with branching morpholo-gies (Figure 1). This morphology was revealed in evengreater detail and at finer scales for networks imaged bythe Huygens Probe Descent Imager/Spectral Radiometer[Tomasko et al., 2005]. The negative topographic reliefassociated with the Huygens networks (that is, they arevalleys) [Tomasko et al., 2005; Soderblom et al., 2007]further suggests that they were created by fluvial erosion[Perron et al., 2006; Moore and Pappalardo, 2011;Langhans et al., 2012]. These networks were most likelycarved by a mixture of liquid methane [Kouvaris and Flasar,1991] and possibly ethane [Lunine et al., 1983; Brown et al.,2008] as they eroded “bedrock” composed of water ice[Griffith et al., 2003; McCord et al., 2006; Rodriguez et al.,2006], perhaps with a sedimentary veneer of organic atmo-spheric precipitates [Yung et al., 1984; McKay et al., 2001;Lavvas et al., 2008; Krasnopolsky, 2009]. Near the poles,river valleys feed into typically RADAR-dark hydrocarbonlakes; the RADAR-darkness of the lakes most likely corre-sponds to lower backscatter associated with a smoother

1Department of Earth, Atmospheric, and Planetary Sciences,Massachusetts Institute of Technology, Cambridge, Massachusetts, USA.

2Earth and Planetary Sciences Department, University of Tennessee,Knoxville, Tennessee, USA.

Corresponding author: B. A. Black, Department of Earth, Atmospheric,and Planetary Sciences, Massachusetts Institute of Technology, 77Massachusetts Ave., Bldg. 54-810, Cambridge, MA 02139, USA.([email protected])

©2012. American Geophysical Union. All Rights Reserved.0148-0227/12/2012JE004085

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, E08006, doi:10.1029/2012JE004085, 2012

E08006 1 of 17

Page 2: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

surface [Mitri et al., 2007; Stofan et al., 2007]. Some lake-shaped features are distinctly brighter and are probably drylakebeds [Hayes et al., 2008]; possible SAR-bright paleo-lakes have also been identified in the low-latitude HoteiRegio and Tui Regio basins [Moore and Howard, 2010].[4] Titan has relatively few visible impact craters, and

these are unevenly distributed geographically. Depending onthe crater production rate, Titan’s globally averaged surfaceage may be between �0.2 to �1 billion years [Lorenz et al.,2007; Neish and Lorenz, 2012]; tropical dunefields and thenorth polar region display a particular dearth of craters[Wood et al., 2010]. SAR swaths currently cover �40% ofTitan’s surface. These images show abundant fluvial net-works at high northern latitudes (Figure 2), several largenetworks and numerous fragmentary ones near the tropicalregion of Xanadu (e.g., Figure 1) and in the southern highlatitudes, and scattered networks elsewhere. If this

apparently patchy and uneven fluvial dissection [Drummondet al., 2012] is truly representative of the surface, it furthersupports the hypothesis that Titan’s surface is locally rela-tively young, and subject to variable amounts of modifica-tion. Cryovolcanism [Sotin et al., 2005; Mitri et al., 2008],tectonic deformation [Radebaugh et al., 2007; Burr et al.,2009], deposition of organic aerosols [Khare et al., 1978],possible migration of observed dunes [Lorenz et al., 2006],and erosional exhumation all provide plausible mechanismsfor resurfacing, but determining which process has beendominant is a key question for deciphering Titan’s geologichistory.[5] The high concentration of methane in Titan’s atmo-

sphere provides some clues to the enigma of Titan’s youthfulsurface, but also underscores an additional puzzle. Titan’stotal atmospheric pressure is �1.47 bars at the surface,composed of roughly 95% nitrogen and 5% methane [Lindal

Figure 1. Low-latitude branching networks in Cassini’s T13 and T43 swaths (249.5 E, 9.6 S).Whitearrows point to locations of visible valleys, and are oriented approximately parallel to flow direction.Black arrow points to perpendicular valley segments in a network that has been identified by Burr et al.[2009] as potentially reflecting tectonic influences.

Figure 2. (left) Branching networks that drain into a hydrocarbon lake near Titan’s North Pole (Swath T28;image centered at 108.5E, 75N). (right) The same region after mapping fluvial networks (white lines).

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

2 of 17

Page 3: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

et al., 1983; Fulchignoni et al., 2005; Niemann et al., 2005].This methane abundance creates the potential for methanesaturation, precipitation, and fluvial modification of thesurface [Perron et al., 2006; Turtle et al., 2011; Schneideret al., 2012]. However, photochemical breakdown andhydrogen loss to space means that the lifetime of methane inTitan’s atmosphere is in the range of 10–100 Myr [Niemannet al., 2005]. Therefore, in order to maintain the observedmethane pressures on Titan, there must be some source ofreplenishment. Over the course of geologic history, Yunget al. [1984] estimate that Titan may have lost a total of8 bars of methane. This assessment was confirmed byNiemann et al. [2005], who find that Titan’s primitiveatmosphere may have been several times as massive as atpresent. Mitri et al. [2007] determine that the observedatmospheric concentrations of methane could be supportedby evaporation from lakes covering 0.2–2% of Titan’s sur-face. Alternatively, the atmospheric concentrations ofmethane could provide evidence for transfer of subsurfacematerials to the surface and atmosphere of Titan [Lunine andLorenz, 2009]. Possible cryovolcanic features are not wide-spread on Titan’s exterior, however [Lopes et al., 2010;Moore and Pappalardo, 2011], suggesting that some of theother processes mentioned above, such as erosional modifi-cation, are important in the rejuvenation of the surface.[6] The history of Titan’s surface is thus critical to

understanding both its interior processes and hydrocarboncycle. While erosion of water-ice bedrock or organic sedi-mentary layers by methane-ethane rivers might seeminconsistent with the conditions of terrestrial erosion, theo-retical and experimental results imply that the rates anddominant mechanisms of fluvial erosion may be similar.Collins [2005] investigates scaling relationships betweenTitan and Earth. He argues that mechanical erosion pro-cesses can operate on Titan at rates that are not radicallydifferent from those for the corresponding processes onEarth. Burr et al. [2006] and Perron et al. [2006] concludethat it may be slightly easier to transport sediments on Titanthan on Earth or Mars, due to the greater buoyancy of sed-iment and the lower gravitational acceleration. Tomaskoet al. [2005] report that grains at the Huygens landing site,which are 2–20 cm in diameter [Keller et al., 2008], arerounded, suggesting that they have been transported over asufficient distance for rounding to occur. Solubility analysessuggest that chemical erosion of water-ice bedrock on Titanis unlikely [Lorenz and Lunine, 1996], and low surfacetemperatures and the lack of significant daily temperaturevariation would impede thermal erosion of water ice [Perronet al., 2006], leaving mechanical erosion as the most likelyincision and particle shaping mechanism for water ice bed-rock. As noted by Perron et al. [2006], some of the valleyheads imaged during the Huygens descent are located onisolated topographic highs, the networks have manytributaries, and valleys appear to widen downstream, all ofwhich are most consistent with mechanical erosion by run-off. Organic materials that have precipitated from Titan’satmosphere [Yung et al., 1984; McKay et al., 2001; Lavvaset al., 2008; Krasnopolsky, 2009] may be more solublethan water-ice bedrock in hydrocarbon liquids [Malaskaet al., 2011]. If significant thicknesses of organic falloutlayers do mantle areas of Titan, dissolution erosion may belocally important [Malaska et al., 2011]. However, we are

not aware of any extensive valley networks on Earth thathave formed mostly or entirely through dissolution.Mechanical erosion therefore appears to be the simplestexplanation for Titan’s valley networks.[7] In this study, we analyze the planform geometries of

Titan’s drainage networks in order to quantify erosionalexhumation, and thereby describe the evolution of the land-scape through time. Metrics of drainage network shape—such as drainage density, the width function, and channelsinuosity—are commonly applied on Earth to understand theway network structure relates to environmental variablessuch as precipitation or bedrock lithology [e.g., Rodriguez-Iturbe and Rinaldo, 1997]. Our measurements of purelyplanform geometry are unusual by terrestrial standards,however, because analyses of fluvial incision on Earth typ-ically also rely on elevation data to measure landscapecharacteristics such as slope, drainage area, and valley relief.In the case of Titan the limited available topographic datacurrently inhibit such measurements. Even on Earth, it ischallenging to assess the cumulative erosion that a landscapehas experienced without independent knowledge of the ini-tial conditions. The goal of this paper is to develop and applya method to assess the extent of erosional development of afluvial landscape, using only features that can be measuredin plan view.[8] We first present a geometrical statistic for quantifying

the shape of drainage networks, and demonstrate its useful-ness for evaluating the extent of erosional modification of asurface. We calibrate our statistic with a simple landscapeevolution model, and we validate it through comparisonswith terrestrial drainage networks with constrained exhu-mation histories. Finally, we map fluvial networks on Titanand apply our method to those networks. Our results suggestthat some of Titan’s most prominent drainage networks havecaused spatially averaged erosion of only a small fraction(<�10%) of the initial topographic relief in the regionswhere they have formed.

2. A Statistic Relating Drainage NetworkGeometries to Erosional Exhumation

2.1. Evolution of Drainage Networks

[9] With few impact craters and little ground-based orhistorical data available, Titan’s drainage networks provideone of the best sources of information about past erosion andresurfacing patterns. Over time, we expect the geometry ofthese networks to evolve. Topographic data that resolve thecross-sectional and longitudinal forms of fluvial valleys onTitan are currently limited. In the absence of widespreadtopographic data, as discussed above, we must focus ouranalysis on the planform geometry of drainage networks.Our conceptual model for drainage network development issimilar to that outlined by Horton [1945]. The basic stages inthis model are shown schematically in Figure 3, starting withinitial perturbations in a random surface (Figure 3a). Theseinitial perturbations develop into channels that compete forwater and therefore drainage area. Because the erosion ratein channels is proportional to flow discharge (see section 3),incipient channels that have smaller drainage areas andtherefore smaller discharge than their neighbors will growless rapidly than their faster-eroding neighbors. The morerapidly eroding networks will gradually integrate more

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

3 of 17

Page 4: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

drainage area and develop simple tributary channels(Figure 3b). At first, the channel networks will expandmainly through headward growth of the trunk channel, butonce they reach the drainage divide, tributaries will growlaterally and capture additional drainage area. As the sur-viving networks approach a steady form, they develop den-dritic drainage networks with evenly spaced trunk channelsand less-elongated drainage basins (Figure 3c).[10] By devising a method to quantify these trends, we can

use planform drainage network morphology to gauge theextent of erosional modification. Although this assessmentsounds like a simple task, it has not yet been completed forEarth—in part because elevation data and various lines of

geological evidence can often be used to constrain thedevelopment of terrestrial networks.[11] As the geometric sequence described above proceeds,

the drainage networks incise through the initial relief (theinitial range of elevations in a landscape over a given hori-zontal area), gradually overprinting the initial topography(Figures 3d–3f). We can define the relative “age” of a fluviallandscape as the extent to which it has modified an initialsurface. All else being equal, this sequence will proceedmore slowly at larger spatial scales with higher relief. Bynormalizing the total vertical erosion E (spatially averagedover the landscape) by the initial relief H0, we obtain adimensionless measure of the relative erosional “age” of thelandscape. Because landscapes can develop at different

Figure 3. (a–c) Schematic drawings illustrating the expected evolution of drainage networks in a landscapewith sinks at the top and bottom of the frame. Initially, headward advance dominates. Basins developroughly uniform width (Figure 3a). Once upslope area is exhausted, lateral tributary growth causes widening(Figure 3b). Interaction among basins concentrates area at intermediate distances from outlet (Figure 3c).(d–f) Three frames from a landscape evolutionmodel run illustrating the evolution predicted in Figures 3a–3c.Time increases by a multiple of five in each successive frame. Colors denote elevation, with blue to red show-ing lowest to highest elevations, respectively. Maximum initial elevations are �400 m, yielding averageslopes of �0.008. (g–i) Idealized illustration of the relationship between the cumulative width function(red line) and theDW statistic (shaded area) for the drainage networks shown in Figures 3a–3c. As shownin the figure, DW is proportional to the deviation of the cumulative width function from a uniform distri-bution of network links. Figures 3g–3i are schematic; actual width functions may not be symmetric abouttheir midpoints.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

4 of 17

Page 5: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

rates, the morphologic age does not correspond to an abso-lute calendar age. Henceforth, we will use the qualitativeterms “immature” and “mature” to refer to networks thatare in the early stages of fluvial dissection (small E/H0)and networks that have caused extensive erosion of aninitial surface (E/H0 approaching unity, or exceeding unityif uplift or base level lowering generate additional relief),respectively.

2.2. The Cumulative Width Function

[12] Given these general expectations for how a drainagenetwork should evolve over time, especially the develop-ment of dendritic tributaries and the trend toward moreequant and integrated basins, we can define a morphologicstatistic to quantify the relative maturity of a network.Rodriguez-Iturbe and Rinaldo [1997] define the widthfunction as the number of links in a drainage networkat a given distance from the drainage basin outlet. Thecumulative width function, as demonstrated in Figures 3g–3i,is the integral of the width function: the fraction of the net-work that lies within a given flow distance from the outlet.[13] We predict that the cumulative width functions of less

mature drainages (such as Figure 3a) will reflect a uniformdistribution of channel segments, because the drainage willnot have developed an equant shape with dendritic tributar-ies. Basins like those in Figure 3b–3c, which are furtheralong in their development, will have sigmoidal or concavecumulative width functions. In order to compare the cumu-lative width functions of various networks, we calculate theabsolute value of the area between each network’s cumula-tive width function and the straight line that represents auniform width function (Figures 3g–3i), and we definethis dimensionless quantity as DW. This statistic, whichmeasures the deviation of the measured width function froma uniform width function, is conceptually similar to theCramér-von Mises criterion, which measures the differencebetween a theoretical cumulative distribution function and ameasured distribution. TheDW statistic is distinct, however,from the Kolmogorov-Smirnov statistic, which only considersthe maximum difference between a measured cumulativedistribution and a theoretical one.[14] In summary, DW is a measure of how much the

distribution of segments within a fluvial network deviatesfrom a uniform distribution. In order to test our hypothesisthat DW should increase as drainage networks dissect asurface, and to quantify any such relationship between DWand cumulative erosion, we turn to a numerical model oflandscape evolution.

3. Landscape Evolution Model

[15] Numerical modeling provides a clear theoreticalframework for both testing and calibrating our predictions forhow the DW statistic should behave as drainage networksevolve. We identify several processes that are likely to bemost important in shaping fluvial landscapes on Titan andinclude them in our evolution model.[16] Drainage basins on Titan, like those on Earth, have

developed through a combination of fluvial incision, whichcreates networks of incised channels, and hillslope processesthat govern the response of the surrounding topography tochannel incision. The limited stereo topography currently

available for Titan confirms that many of the networksvisible in SAR are incised valleys with steep valley walls,and therefore bedrock erosion must have occurred duringtheir formation [Tomasko et al., 2005; Soderblom et al.,2007]. Images from the Huygens probe [Tomasko et al.,2005] show branched networks dissecting the landscapedown to a certain scale, but with channels that appear totruncate at a scale coarser than the image resolution. Thisobservation implies that at the finest scales, non-fluvial ero-sion mechanisms are more important than fluvial incision.However, because we are interested in the shapes of net-works which span 20 to 500 km, the relative contribution ofhillslope processes to network geometry is likely to be small.Cross sections through valleys near the Huygens landing siteshow nearly straight hillslope profiles with slopes of roughly30 degrees [Tomasko et al., 2005], suggesting that hillslopeerosion may be modeled with a critical angle, where anyhillslopes that steepen beyond this angle experience rapidmass wasting.[17] We therefore employed a simple model that includes

only bedrock river incision and mass wasting. The masswasting is driven by a slope threshold. When valley sideslopes exceed a critical gradient of 0.6 (31�) we assume thatfailure will occur, transporting material downslope to thechannel and then out of the system. We assume that theflow’s transport capacity exceeds the supply of landslide-derived sediment, such that the transport of sediment doesnot limit the rate of channel incision. Channel incision intobedrock is modeled with a stream power equation [Seidl andDietrich, 1992; Whipple and Tucker, 1999],

∂z∂tjchannel

¼ �KAm rzj j ð1Þ

which is derived from the assumption that the rate of inci-sion is linearly proportional to the rate of energy expenditureby the flow per unit area of the channel bed, also called the“unit stream power.” In equation (1), z is elevation, t is time,A is drainage area, m is a constant, and |rz| is the channelslope. We assume that there is no change in surface elevationdue to uplift or subsidence. K is the stream incision coeffi-cient, which represents the effects of bedrock erodibility,precipitation rates, drainage basin hydrology, and hydraulicgeometry on fluvial incision rates. We assume that K isspatially uniform, and therefore that bedrock erodibility,precipitation, and other components of K are also uniform.The coefficient m, which typically has a value of approxi-mately 0.5 [Howard and Kerby, 1983; Whipple and Tucker,1999], incorporates the relationships between erosion andstream power, channel geometry and discharge, and basingeometry and discharge [Howard, 1994; Whipple andTucker, 1999]. It influences the concavity of the channellongitudinal profile and the relative balance between drain-age area and channel slope in shaping the overall geometryof the network. Both theoretical arguments [Willgoose et al.,1991; Howard, 1994; Tucker and Bras, 1998] and empiricalobservations [e.g., Rosenbloom and Anderson, 1994]support the utility of equation (1) as a simplified generalmodel for bedrock river incision, and it has been used pre-viously in many models of landscape evolution [e.g.,Howard, 1994; Pelletier, 2004; Perron et al., 2008, 2009].

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

5 of 17

Page 6: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

[18] We modeled the evolution of the landscape bysolving equation (1) forward in time on a regular grid using afinite difference method. We began with random, self-affineinitial surfaces defined by the variance in elevation and theslope of the surface’s power spectrum. We assume there areno structures that would introduce strongly oriented initialtopography. The lowest 10% of elevations in the initial sur-face were set to a fixed elevation of zero to simulate a baselevel, as might be provided for high-latitude networks byTitan’s polar lakes. Initial relief for our model landscapes wasapproximately 400 m. The average relief of equatorialmountains reported by Radebaugh et al. [2007] was �400 mover horizontal profiles tens of kilometers long. The grid sizeis 400 by 400 points, withDx =Dy = 125 m.We selectedK =0.005 yr�1 with m = 0.5, and ran the model for 25 Myr.Because the magnitude of K is not known for Titan, thesemodel results are not tuned to be specifically applicable to therate of erosion on Titan. Rather, they form a test set for ourgeneral predictions of landscape evolution and the DW

statistic. There is a tradeoff between initial relief, K, hori-zontal scale, and the time interval over which the landscapeerodes. We are primarily interested in how drainage networkform changes as a function of how much of the initial reliefhas eroded. Because this evolution is independent of theabsolute rate of erosion, our results are valid without refer-ence to the value of K or elapsed time.[19] Our model runs employed a drainage area exponent

of m = 0.5 and a slope exponent of one, which implies thaterosion rate is linearly proportional to stream power[Whipple and Tucker, 1999]. Varying these exponents willshift the concavity and response time of the drainage profiles[Whipple and Tucker, 1999], but exploratory model runs notpresented here indicate that this has only a minor effect onDW. The 50 km by 50 km scale of the model runs is similarto that of most of the networks we analyzed on Titan, thoughsome of the networks are several times larger. In Figure 4,the imprint of initial topography is clearly visible even after25 Myr of landscape evolution. The self-affine random

Figure 4. Example 25 Myr model run, with one drainage network highlighted in red. We only analyzethe portion of the network that has eroded through >20% of the initial relief. TheDW increases over time,eventually approaching a steady value for a given drainage network, even though the topography does notreach a steady state.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

6 of 17

Page 7: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

surface that we use to initiate the models is qualitativelyconsistent with available data from Titan, which generallyshow gentle slopes over long baselines [Zebker et al., 2009],though higher resolution data from the Huygens landing sitecontain valley side slopes up to 30 degrees over short dis-tances [Tomasko et al., 2005].[20] Figures 3d–3f show three stages in a typical run of the

landscape evolution model. To investigate our hypothesizedrelationship between fluvial network morphology and theextent of erosion, we require a criterion for defining fluvialnetworks in our model topography that is comparable to thevisibility of fluvial networks in coarsely resolved CassiniSAR images. As we discuss in section 5, deeper valleysshould be much easier to detect in SAR imagery of Titan,and very shallow valleys may be invisible. Therefore, in ourmodel runs, we calculated theDW only for those portions ofthe network that have incised a depth greater than 20% of theinitial relief of the landscape. 10–25% thresholds all yieldsimilar trends, but lower thresholds correspond with lowerconfidence in our upper limits on erosion. We discuss thissensitivity in greater depth in section 6.2. In order to avoidfragmented networks, we included all portions of a networkdownstream from a segment that has eroded to this depth.This approach is comparable to our procedure for mappingdiscontinuous but clearly related fluvial features.[21] Four timesteps from a 25 Myr model run are shown in

Figures 4a–4d. In Figures 4e–4h, we isolate a single networkand show the portion that we analyzed after applying the“visibility” criterion described above. These frames showhow, for the most deeply eroded fluvial features, the evolu-tion of the network generally matches our predictions asshown in Figure 3. Because stream incision rate is propor-tional to drainage area, A, in equation (1), there is a positivefeedback between stream incision and drainage area. Themore drainage area a given network claims, the faster itincises, which allows it to lengthen or widen, thereby cap-turing more drainage area. Incision rate is also proportionalto land-surface slope, so networks grow more quickly in theupslope direction. As networks begin to form on an initiallyundissected surface, they will first grow headward, becauseheadward growth is unimpeded by competition with neigh-boring networks. This growth leads them to become elon-gated, as in Figures 4a, 4b, 4e, and 4f. Once networks beginto impinge on one another on opposite sides of a drainagedivide, however, they widen and become more equant, as inFigure 4c, 4d, 4g, and 4h.[22] These trends are reflected in DW, as shown in

Figures 4i–4l. On the x axis, instead of time, we express theprogressive development of the drainages in terms of E/H0,the normalized measure of average erosion defined in section2.1. Very early in the development of the networks, the DWis low. It rises quickly to a nearly steady value that variesfrom network to network. The precise final configuration ofeach drainage network depends on initial conditions [Perronand Fagherazzi, 2012]; the overall shape of the basin willcontrol the shape of the cumulative width function (of whichFigures 4i–4l show examples), and will therefore also influ-ence the ultimate value ofDW achieved by each network. Inorder to better estimate the distribution of possible outcomes,we employed a Monte Carlo procedure in which we per-formed twenty model runs with twenty different initial sur-faces that shared similar statistical characteristics. The largest

networks (drainage area >10% of the total grid) in these runswere used to construct a calibration data set for DW as afunction of E/H0. We binned these data according to E/H0.The resulting distributions ofDW for various values of E/H0

are plotted in Figure 5, with the mean trend with increasingE/H0 shown in Figure 6. Because the distributions for lowand high DW statistics are associated with different E/H0,we can use our results to calculate the statistical likelihoodof a given erosional exhumation for drainage networks onTitan or Earth.

4. Validation With Terrestrial DrainageNetworks

4.1. Glaciated And Unglaciated Landscapes

[23] To complement the theoretical predictions of land-scape evolution provided by numerical modeling, we soughtobservational validation of our hypotheses. We measured theDW for drainage networks in several landscapes in NorthAmerica. These terrestrial locations are summarized inTable 1. We selected four sites in North America with one oftwo histories: a long history of fluvial erosion, or a shorthistory of fluvial erosion following resurfacing by glacialprocesses. The first two sites, in the Allegheny Plateau ofPennsylvania and the Oregon Coast Range, have not beenglaciated in Pleistocene time. The other two sites, on thesouthern edge of Lake Superior inMichigan and near HudsonBay in northern Québec, were glaciated during the LastGlacial Maximum [Ehlers and Gibbard, 2004; Kobor andRoering, 2004]. Because continental ice sheets causedextensive erosion of the landscapes that we categorize asrecently glaciated, the onset of Northern Hemisphere degla-ciation around 19–20 ka [Clark et al., 2009] serves as theeffective surface age for the Michigan and Hudson Bay sites.[24] We also selected four sites that straddle the line

marking the maximum extent of the Wisconsonian glacia-tion. These sites—in Washington’s Puget Sound and north-western Pennsylvania—provide a range of surface ageswhile holding erodibility and climate generally constant.[25] As shown in Table 1, these eight sites all had rela-

tively similar mean DW statistics, compared to the broaderrange observed for individual terrestrial drainage networksand during model simulations. In part, this differencebetween the terrestrial and model data may be explained bythe absence of an incision threshold for the terrestrial data.Because initial relief is not known, we calculated the DW forall parts of a network, not only for those parts that had incisedmore than 20% of the initial relief. The eight terrestrial sites do,however, show a systematic pattern wherein recently glaciatedlandscapes are biased to lower DW statistics and older land-scapes are biased to higher DW statistics (Figure 7). Thispattern is consistent with the general prediction that drainagenetworks that have eroded a larger fraction of the initialtopographic relief will have a larger fraction of the networkconcentrated at intermediate distances from the basinoutlet (Figures 3 and 4).

4.2. Drainage Networks on Kaua’i

[26] In addition to this qualitative confirmation of thehypothesized relationship betweenDW and E/H0, we soughtto make a quantitative comparison of terrestrial fluvial net-works with the landscape evolution model predictions. The

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

7 of 17

Page 8: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

western coast of the Hawaiian island of Kaua’i, above theMana Plain, provides a unique opportunity to measure theDW statistic for a landscape with known amount of cumu-lative erosion. This portion of the island is relativelyhomogeneous, with a classic shield volcano structure thathas been partially eroded by drainage networks [Macdonaldet al., 1960]. We identified remnants of the original shieldsurface surrounding each basin and fit these remnants withspline functions in order to reconstruct the initial surface.We then estimated the volume of material that has beenremoved through erosion, allowing us to calculate E/H0 foreach basin. We calculated DW for 20 adjacent drainagenetworks (Haeleele, Hanakapiai, Hanakoa, Hikimoe, Hoea,Honopu, Hukipo, Kaawaloa, Kaaweiki, Kahelunui, Kahoa-loha, Kalalau, Kapilimao, Kauhao, Kaulaula, Kuapaa,Makaha, Milolii, Nahomalu, Niu, Ohaiula, Paua, Wailao,and Waipao), and obtained a relatively low DW = 0.061 �0.012.[27] In order to compare these measured combinations of

E/H0 and DW with data from our model, we performed a

Monte Carlo calculation (Figure 8). In each iteration, werandomly selected twenty networks from the numericalmodel results (the same as the number of measured Kaua’inetworks) for each of eight logarithmically spaced values ofE/H0, and calculated the meanDW for each group of twentymodel networks. This simulation was performed 100 times.If the mean of the simulated population was separated by atleast two standard errors from the mean of the Kaua’i popu-lation, then we counted this instance as a statistical differencein the mean. The number of times (out of 100 simulations)that the mean DW for the model was two standard errorshigher or lower than the mean DW for Kaua’i gives anestimate of the likelihood that the Kaua’i drainage basinshave eroded less or more, respectively, than each value ofE/H0.[28] Figure 8 plots these Monte Carlo percentiles for

Kaua’i, along with the E/H0 value we actually observe there.Based on an approximate initial relief of �1000 m in thisarea of Kaua’i, the Monte Carlo calculation suggests a veryhigh likelihood (95th percentile) that the landscape has

Figure 5. Six snapshots of the distribution ofDW from an ensemble of 20 model runs. These histogramsshow that mean DW increases with increasing erosion relative to initial relief, eventually approaching asteady value of DW.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

8 of 17

Page 9: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

experienced more than 4 m of spatially averaged erosion,and a moderate likelihood (50th percentile) of less than160 m. Our best estimate for the actual spatially averagederosion, based on current topography and the spline-fit initialsurface, is �80 m. The good agreement between theobserved erosion and the constraints based on DW suggeststhat theDW value of a drainage network is a robust indicatorof the approximate amount of cumulative fluvial erosion ina landscape.

5. Application to Fluvial Networks on Titan

[29] Analysis of terrestrial and model networks indicatesthe usefulness of DW as a proxy for the cumulative erosionexperienced by a landscape. In this section, we discuss theapplication of this statistic to Titan.

5.1. Mapping and Analysis

[30] We produced detailed maps of Titan’s fluvial networksusing SAR images from the Cassini spacecraft (see Figure 2for an example from Titan’s North Polar region). SAR ima-ges are radar backscatter images. Backscatter intensity isinfluenced by surface orientation relative to radar incidenceand azimuth angles of observation, roughness, dielectricconstant, and density of the targeted surface. The resolution ofSAR swaths is 350–720 m/pixel at the center of the swaths,and decreases by a factor of �3 at the two ends. Each SARswath ranges from 120 to 450 km wide, and images roughly1.1% of Titan’s surface [Elachi et al., 2004, 2005]. To date,approximately 40% of Titan’s surface has SAR coverage.[31] We mapped all fluvial networks visible in swaths Ta

through T71 [Drummond et al., 2011, 2012]. We identifiednetworks on the basis of five criteria: light-dark pairings,branching patterns, crosscutting of other radar features (suchas dunes or mountains), connectivity, and orientation withrespect to suspected topographic sinks (such as lakes). Light-dark pairings are created by differing illumination conditionson either side of a valley: the slope that is facing thespacecraft is illuminated by the radar and appears bright,whereas the other slope will appear dark. The networks wemapped ranged from first-order networks, which essentiallyconsist of only one visible link, up to fourth-order networkswith dozens of tributaries. These orders are apparent ordersthat describe the features that we have mapped, but becausemany fine scale features are probably invisible in the SARimagery, the actual orders of the drainage networks may bemuch higher. Drummond et al. [2011, 2012] provide a moredetailed description of the global mapping efforts.[32] For this paper, we focused on networks from swaths

T13, T43, T28, T29, and T39, which include some of themost extensive and most completely imaged networksfound on Titan so far. T13 and T43 are equatorial swathsnear Xanadu (Figure 1), T28 and T29 cover the lakesaround the north pole (Figure 2), and T39 is near the southpole. We also designated a probable sink for each largenetwork—a point through which we expect flow from therest of the network to drain. We identified these sinksbased on flow directions inferred from junction angles andapparent trends in valley width. Starting from the sink ineach network, we worked upstream to assign a flow direction

Table 1. DW Measurements From Titan and Eartha

Location Description (Swaths) CoordinatesNumber ofNetworks DW

CurrentRelief

Titan NP1 Titan Northern High-Latitude [T28/T29] 78.5E, 80.1N 19 0.07 � 0.016 >300 mTitan NP2 Titan Northern High-Latitude [T28/T29] 117.0E, 74.0N 21 0.05 � 0.012 >300 mTitan T13 Titan Low-Latitude [T13] 249.5E, 9.6S 4 0.07 � 0.035 �500 mTitan T39 Titan Southern High-Latitude [T39] 328.0E, 73.2S 8 0.06 � 0.015SW Pennsylvania Unglaciated 279.5E, 39.5N 4 0.107 � 0.041 �300 mNW Pennsylvania Glaciated 280.0E, 42.0N 6 0.066 � 0.016 �200 mNW Pennsylvania South of LGM 280.5E, 41.4N 4 0.084 � 0.030 �200 mWashington Glaciated 237.9E, 47.2N 7 0.068 � 0.012 �600 mWashington South of LGM 237.6E, 46.5N 4 0.092 � 0.023 �1000 mOregon Unglaciated 236.0E, 43.4N 8 0.091 � 0.026 �650 mMichigan Glaciated 271.4E, 46.5N 7 0.082 � 0.010 �350 mHudson Bay, Canada Glaciated 262.5E, 51.9N 5 0.076 � 0.019 �250 mWestern Kaua’i Mana/Na’Pali Coast 200.3E, 22.06N 20 0.061 � 0.012 �1000 m

aEstimates of current relief for Titan are based on SAR topography from Stiles et al. [2009]. “Glaciated” means that a given area was glaciated at the LastGlacial Maximum. Means for Titan are weighted by horizontal length of the networks. Uncertainties in DW are stated as 2 standard errors.

Figure 6. Compilation of binnedDW as a function of E/H0

for all twenty model runs. As erosion proceeds, drainagesbranch and become more equant, leading to higher valuesof DW. After erosion has reached approximately 10% ofthe initial relief, values of DW reach a steady state. Largeerror bars represent one standard deviation; small bold errorbars show one standard error.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

9 of 17

Page 10: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

to each pixel along the flow path, according to an eight-direction routing scheme.[33] We calculated DW for 52 fluvial networks. These

networks were located in four general regions, as shown inFigure 9 and in the auxiliary material, and we performed ouranalyses separately for each population.1 The mean DWstatistics for each region are weighted according to thelengths of the individual networks; larger networks aretherefore weighted more strongly. The results are summa-rized in Table 1. Below we compare these results with DWstatistics from numerically simulated landscapes in order toestimate the magnitude of fluvial erosion on Titan.

5.2. Estimating Erosion on Titan

[34] In order to place constraints on the amount of erosionthat has occurred on Titan, we performed a series of MonteCarlo calculations to compare the distributions of DWderived from the landscape evolution model (Figure 5)against the distributions of DW measured for Titan. TheseMonte Carlo calculations followed the same procedure weapplied to terrestrial networks on Kaua’i, as described insection 4.2, except that the number of randomly selectednetworks in each Monte Carlo iteration was chosen to matchthe number of mapped networks in each Titan region.Table 1 lists the number of networks we analyzed withineach study area on Titan.[35] Figure 9 shows percentiles for the upper and lower

bounds on E/H0 for each region, overlain on a map of Titan.Based on our Monte Carlo simulations, we estimate lowerlimits on erosion on Titan of 0.4% to 0.6% of initial relief inall our study regions. These numbers correspond to theregionally averaged erosion; erosion in fluvial valleys ismuch larger. As an analogy, for our landscape evolution

model runs, E/H0 = 0.4% corresponds to erosion within thevalleys of �25–30% of the initial relief, or �100–120 m.[36] For two of our study areas on Titan, the measured

DW was statistically indistinguishable from the uppervalues of DW obtained during the model runs (Figure 9),and therefore no upper limit on erosion could be determined.For the other two study areas, we were able to estimate upperlimits on erosion, though with less confidence than for ourestimates of the lower limits. Our constraints for an area ofTitan’s northern lakes region are most robust: we estimate an85% probability that fluvial erosion in this region is less than9% of the initial relief. Once again, this value is an averageover the entire landscape; within the valleys, the incisionmay be much deeper.

6. Discussion

6.1. Implications for Resurfacing and Long-TermErosion Rates on Titan

[37] In section 1, we reviewed several potential mechan-isms for resurfacing on Titan, including cryovolcanism,tectonic deformation, deposition of organic aerosols, possi-ble migration of observed dunes, and erosional exhumation.On the basis of the results presented in section 5.2, we canevaluate the potential effectiveness of fluvial erosion forresurfacing. The maximum relief in Titan’s landscapes atlow latitudes is approximately 500 to 2000 m along SAR-based topographic profiles 5 to 20 km long [Radebaughet al., 2007]; much greater relief would trigger flow at thebase of Titan’s dominantly water ice crust, resulting in oro-genic collapse [Perron and de Pater, 2004]. Elachi et al.[2006] also estimate maximum relief of �1000 m. Assum-ing that relief in landscapes on Titan does not generallyexceed �1000 m, we can calculate an order-of-magnitude

Figure 8. Monte Carlo simulations provide constraints onthe likely lower (dashed red line) and upper (solid blue line)bounds on erosion relative to initial relief. For Kaua’i, thepercentiles based on a measured DW of 0.061 � 0.012 fortwenty drainage networks are consistent with the actual E/H0, shown by the gray bar. The width of the gray bar repre-sents E/H0 values within two standard errors of the meanerosion among our 20 Kaua’i basins.

1Auxiliary materials are available with the HTML. doi:10.1029/2012JE004085.

Figure 7. DW for regions on Titan and Earth. DW forTitan is averaged over all four study regions, with a totalof 52 networks. Uncertainties represent two standard errorsof the mean.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

10 of 17

Page 11: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

estimate of erosion. For NP2 drainages near Ligeia Mare themeanDW is 0.05 � 0.012 (Table 1). This relatively uniformdistribution of network segments, in conjunction withnumerical model results (Figures 4 and 9) and an assumedH0 of �1000 m, implies an upper limit on areally averagederosion of �100 m. Fluvial erosion of <9% of the initialrelief (our estimate for the region of the NP2 drainages), mayhave produced a significant modification of the initial land-scape, in particular in valleys where incision may be muchdeeper than 9% of the initial topography.

[38] Impactors smaller than 1 km in diameter may failto penetrate Titan’s thick atmosphere, producing a dearthof microcraters [Artemieva and Lunine, 2003]. Erosionalexhumation of the magnitude that we predict might besufficient to erase some of the small craters that are created,especially those that have also experienced viscous relaxation[Parmentier and Head, 1981; Artemieva and Lunine, 2003].However, as is evident from the persistence of topographi-cally prominent features [e.g., Radebaugh et al., 2007], the

Figure 9. Available SAR swaths and probabilistic estimates of erosional exhumation for mapped regionsof Titan. North Pole Area 1 is outlined in light green, North Pole Area 2 is blue, the Equatorial region isred, and T39 in the southern high latitudes is purple. Ligeia Mare, one of Titan’s largest lakes, is visiblenext to the north polar networks. Four exterior panels show results of Monte Carlo calculations for eacharea. Dashed red curves show probability that E/H0 for a particular region on Titan is greater than a givenvalue on the x axis, and solid blue curves show the probability that E/H0 for Titan is less than a givenvalue.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

11 of 17

Page 12: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

magnitude of fluvial erosion has been insufficient to entirelylevel Titan’s surface.[39] If crater counting surface ages for Titan of 108–

109 years [Lorenz et al., 2007; Wood et al., 2010; Neish andLorenz, 2012] are accurate, we can use our constraints onmaximum exhumation to estimate a long-term erosion rate.A maximum of order 100 m of erosion in 108 years providesan upper limit on fluvial erosion rates of �10�6 m yr�1 forthe NP2 area of Titan’s northern high latitudes. The calcu-lated long-term erosion rates are even slower if we consideran initial relief in the landscape of less than order 103 meters,or surface ages of more than 108 years. In contrast, the typicalrange of fluvial erosion rates on Earth’s continents is �10�5

to 10�3 m yr�1 [Portenga and Bierman, 2011]. In otherwords, in order to be consistent with our results and with thesurface age predicted from crater counting, erosion rates forsome areas on Titan would have to be far slower than typicalterrestrial rates, despite the fact that rapid surface changesprovide signs of present-day methane precipitation at lowlatitudes on Titan [Turtle et al., 2011].[40] There are several possible explanations for the

apparently slow erosion rates on Titan. Collins [2005] haspreviously suggested that during bedrock incision on Titan,the lower kinetic energy of saltating grains is offset by thelower abrasion resistance of water ice at Titan surface tem-peratures. Assuming similar discharge rates, the erosionrates on Titan and Earth would then be comparable [Collins,2005]. However, recent experimental evidence shows thatice on Titan may in fact be much stronger than originallythought [Collins et al., 2012], with a tensile strength similarto welded tuff or quartzite, among the strongest of terrestrialcontinental materials [Sklar and Dietrich, 2001]. As a result,river incision on Titan could be slower than on Earth, givensimilar discharge rates.[41] On the other hand, typical discharge rates and the

frequency of high flow events on Titan may be very differentfrom those on Earth. Additional data on the intensity andduration of precipitation events, and also on the long-term

frequency of those events, would significantly improve ourknowledge of likely runoff discharge [Perron et al., 2006].If mechanical erosion also occurs through fallout deposits oforganic aerosols from the atmosphere, then additional dataon the strength of such materials would be important forunderstanding their erosional behavior.[42] A further possible explanation is that the surface age

for some regions is in fact much younger than crater countingseems to suggest. Surface ages on Titan may be heteroge-neous; in Figure 2 there are both mottled, slightly darkerareas and smooth, light-toned areas, each incised by drainagenetworks. These different surface appearances may corre-spond to different ages of resurfacing. If the northern polarregion is younger than 108 years, then our estimates for ero-sion rates on Titan would increase proportionally.[43] Finally, some fluvial networks may be relics, either

from an earlier period with higher atmospheric methaneconcentrations or from a previous topographic state that hasbeen altered tectonically or resurfaced. Methane photolysis[Yung et al., 1984] could engender gradual decreases inmethane concentration and precipitation over time, despitereports of present-day precipitation [Turtle et al., 2011].Examination of SAR imagery provides some localizedindications that changes in flow paths appear to have out-paced the rate at which fluvial erosion erases previousdrainage patterns. Figure 10 shows possible inconsistenciesin the flow paths of fluvial features near Ligeia Mare thatmay reflect relict earlier drainage patterns. Junction anglesare ambiguous, raising the possibility that the drainages inFigure 10 do not terminate into the lake, which we assume iscurrently a local topographic minimum. The diffuse darkband in Figure 10, which does connect to the lake, seems tocrosscut the narrower branching networks.[44] Depending on the dominant tectonic style on Titan,

our upper bounds on areally averaged erosion hint at morespeculative implications for Titan’s surface geology. In thecase where uplift of bedrock is driven purely by isostaticresponse to erosional exhumation, that uplift of bedrock

Figure 10. Apparent inconsistencies in drainage network flow paths located at 95E, 75N on the shores ofTitan’s Ligeia Mare (T28). Possible link between adjacent valleys via diffuse dark band could indicatestream capture (arrow A). Junction angles suggest that the large network entering the image from the lowerleft does not drain into Ligeia Mare (arrow B).

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

12 of 17

Page 13: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

cannot exceed the areally averaged erosion [England andMolnar, 1990]. Assuming <100 m of areally averaged ero-sion in the NP2 area, we conclude that fluvial erosion alonehas probably been insufficient to exhume material fromdeeper in Titan’s crust and to expose it at the surface, at leastduring the formation of the current landscape. Suchunroofing would require tectonic processes.[45] Given the apparent lack of waves on Titan’s lakes

[Stofan et al., 2007; Wye et al., 2009; Lorenz et al., 2010],the rarity of obvious deltas where fluvial systems drain intohydrocarbon lakes may provide additional evidence thaterosion and sediment transport are relatively slow comparedto changes in lake level. Over geologic timescales, sedimenttransport and deposition will gradually fill in topographicsinks. On the southwestern edge of Ontario Lacus, one fea-ture has been interpreted as a delta [Wall et al., 2010]. Whilecomparisons are problematic between terrestrial river widthsand the widths of valley features visible in SAR data[Burr et al., 2012], Wall et al. [2010] cite the small relativesize of the Ontario Lacus delta as evidence for low sedimentdischarge. The shorelines of many of the lakes near the northpole evoke drowned drainage networks [Stofan et al., 2007],

even when downstream of existing fluvial networks (e.g.,Figure 2). This observation suggests that sediment depositionduring current lake levels has been insufficient to allow sig-nificant delta progradation into the standing liquid that fillsthe drowned lower reaches of the networks. A digital eleva-tion model for the western edge of Kraken Mare, the largestof the northern polar lakes, does reveal a topographic benchthat may have been produced by depositional processes[Aharonson et al., 2012]. Aharonson et al. [2012] estimatethat the volume of material required to form this benchtranslates to �1 m of areally averaged erosion within MaydaInsula. This value is consistent with the areally averagederosion of between 0.4% and 9% of initial relief that we inferin this study for the areas near to Kraken Mare (Figure 11).[46] The few available topographic profiles based on stereo

Descent Imager/Spectral Radiometer imagery also confirmthat our erosion constraints are reasonable. In the highlandsnorth of the Huygens landing site, topographic profiles showvalleys�90 m deep incised into a landscape with�240 m ofrelief [Tomasko et al., 2005; Soderblom et al., 2007].Assuming that the ratio of valley depths to areally averagederosion is comparable to that in our model runs, this translates

Figure 11. Comparison of SAR images and topography for terrestrial drainage networks. (a) SAR imageand (b) shaded relief map of the region southwest of L’Anse Bay, Lake Superior, on Michigan’s upperpeninsula, centered at 46.57�N, 88.67�W. (c) SAR image and (d) shaded relief map of a section of theAllegheny Plateau east of the Ohio River in northern West Virginia and southwest Pennsylvania, centeredat 39.67�N, 80.70�W. SAR images are from the C-band antenna on board the Shuttle RADAR TopographyMission [Farr et al., 2007]. Topography is from the USGSNational Elevation Data set. Map projections aretransverse Mercator, with a pixel size of approximately 30 m.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

13 of 17

Page 14: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

to an average erosion of �1.5–2% of the initial relief, con-sistent with the range of erosion shown in Figure 9 for otherareas of Titan.

6.2. Effects of SAR Resolution

[47] Although SAR resolution constitutes one of the majorsources of uncertainty for mapping individual branchesof Titan’s drainage networks, small omissions do not signif-icantly affect the measurement because DW is a measure ofdrainage shape, not density. The Huygens Descent Imager/Spectral Radiometer revealed fluvial features much smallerthan the best resolution of the orbital SAR imagery [Tomaskoet al., 2005; Soderblom et al., 2007], indicating that the finestscale features on Titan’s surface are not resolved in SAR data.A comparison between terrestrial SAR and topographic data(Figure 11) also underscores the limitations of SAR-basedmapping. Any highly dissected areas on Titan that resemblethe fine-scale valley networks in Figures 11c and 11d may bedifficult to delineate. Fluvial features on Titan’s surface maybe far more extensive than can be detected on the basis of theSAR images [Burr et al., 2012]. Future missions to Titan maytherefore reveal landscapes that are even more diverse thanthose we can see currently.[48] It is possible that all regions of Titan imaged with

SAR have unresolved fine-scale fluvial dissection similar tothat at the Huygens probe landing site [Tomasko et al.,2005]. To evaluate the effects of mapping uncertaintyrelated to radar instrument resolution, we experimented on asmall data set of four model runs. We tested the sensitivity ofour results to image resolution by varying the drainage areathreshold that we use to define a stream link when calcu-lating DW (Figure 11a). Even for large variations in the finestructure of the network, the overall trend in DW remainsconsistent, as do DW values for low values of E/H0. How-ever, for higher values of E/H0, the inclusion of the lowest-order branches can shift the equilibrium value of DW. As a

result, the percentiles for the upper limits on E/H0 also shift(Figure 12b). We interpret this shift to mean that the truecertainty with which we know the upper limits may bemoderately lower than the stated percentile values. Thelower limits on erosion are robust and independent ofresolution.[49] As discussed in section 3, we apply a threshold to the

model topography such that only relatively deeply erodedvalleys are analyzed. This threshold is important becauseotherwise our analysis would include fluvial networks thatextend all the way to the drainage divides, whereas on Titanwe have only mapped the networks that are visible in SARimagery. Comparison of topography and SAR images(Figure 11) supports the use of a relief threshold for identi-fying the portion of the fluvial network to be analyzed. Asmentioned above, Figure 11 clearly shows that many ter-restrial fluvial features that are visible in the digital elevationmodels are invisible in the SAR images. We hypothesizethat one factor that contributes to this discrepancy is therelief across the valleys. In southwestern Pennsylvania, thedeepest and most easily visible valleys are �150–200 mdeeper than the surrounding ridges, or >50% of the totalrelief in the landscape (Table 1). In Michigan, many of thefeatures that disappear in the SAR imagery have relief of�20–50 m, or 5–15% of the relief in the landscape. Thiseffect is the basis for our relief filter for the model data; weanalyze only the portions of the networks that we thinkwould be visible in Titan’s SAR imagery.[50] The effects on network shape that result from

changing the relief filter are very similar to the effects ofchanging resolution (as described in the previous para-graph), and so it is unsurprising that these two experi-ments produce similar results on the modeled DW.Changing the sensitivity of this relief filter producesmodest shifts in the model calibration curve for DW (areduction of 25% in the relief threshold produces a shift of

Figure 12. Sensitivity of DW and erosion estimates to resolution of mapped fluvial networks. (a) As thefinest network links are excluded, the overall trend in measurements of DW is consistent. Curves areextracted from a set of four trial model runs. The effect on the steady value of DW at large cumulativeerosion values is small but significant. Curves are labeled with the minimum contributing area (as a fractionof the total area of the grid) used as a threshold for delineating the drainage network. This is illustrated in theinset with an example channel network. The green dashed curve denotes the value used elsewhere in thisstudy. (b) Monte Carlo calculations for Titan’s North Polar Area 2 using the synthetic data set in Figure 12a.Including or excluding the finest branches of the networks has a negligible effect on the predicted lowerlimits on exhumation. The magnitudes of the predicted upper limits on erosion are also relatively stable,but the percentiles shift, reflecting uncertainty in the specific percentiles for those estimates.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

14 of 17

Page 15: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

�3–4% in the mean DW). The sensitivity of our erosionestimates to such shifts mirrors the response to variable res-olution. Our estimates for the lower limits on Titan’s erosionremain robust, but the percentiles at which we ascertain theupper limits on erosion may decrease from the values shownin Figure 8.

6.3. Tectonics, Spatial Variability in Erosion and DW

[51] The similarity in our erosion estimates for equatorial,northern high-latitude, and southern high-latitude networkson Titan is interesting given the asymmetric distribution oflakes on Titan. Aharonson et al. [2009] suggest that thepreponderance of lakes in the northern hemisphere mayresult from nonlinear climate responses to the eccentricity ofSaturn’s orbit around the sun. Over timescales of �45 kyr,this asymmetry should reverse [Aharonson et al., 2009]. Thelack of asymmetry in our data may imply that the networkshave developed over times much longer than 45 kyr, whichis reasonable given the constraints on Titan’s surface age[Wood et al., 2010; Neish and Lorenz, 2012]. Althoughclouds have been observed most frequently at high latitudeson Titan [e.g., Rannou et al., 2006], Schaller et al. [2009]and Turtle et al. [2011] report storms in Titan’s tropics.The networks in swath T13 are hundreds of kilometersacross and are well developed (Figure 1), with high DW.The similar lower limits on E/H0 suggest that these largedrainages near Xanadu may have eroded as deeply as net-works close to the poles.[52] Within our study regions, the most distal branches of

some networks appear to originate in mottled light-and-dark“crenelated” terrains that evoke the appearance of highlydissected regions in SAR imagery for Earth (Figure 11). Thisobservation does not necessarily conflict with our interpre-tation that in some areas there has been relatively little are-ally averaged erosion. Our DW populations for Titan spanvast areas. If spatially non-uniform erosion has occurredwithin those areas, we may be averaging theDW statistics ofnetworks with disparate histories. It is also possible to gen-erate a spatially extensive imprint of fluvial erosion whileonly eroding through limited relief. For example, in ournumerical simulations, there is no bedrock uplift or loweringof sinks, implying that the maximum vertical erosion issimply the initial relief. Yet the model drainage networkspropagate throughout the landscape well before the relief iscompletely eroded. Finally, our results do not exclude thepossibility that some regions on Titan may have experiencedextensive erosion. Indeed, for two of our four study sites,including Xanadu, where there is abundant “crenelated”terrain, the DW statistic does not produce a reliable upperbound on erosion. Figure 6 demonstrates that the DW sta-tistic is only sensitive to the early stages of landscape evo-lution, when drainage networks are invading previouslyundissected terrain. After most of the drainage area has beencaptured, basin shapes no longer change appreciablyalthough erosion continues, and DW ceases to provide aunique upper limit on cumulative erosion. For two regionswhere we do provide upper bounds on erosion, our conclu-sion based on measurements of DW is that drainage net-works in these areas appear to have caused relatively littleexhumation.

[53] Some drainage networks on Titan, including drainagesfrom swaths T13 and T28 that are included in this study, havebeen classified as rectangular [Burr et al., 2009; Drummondet al., 2012]. This geometry may suggest that these regionshave been tectonically influenced [Burr et al., 2009;Radebaugh et al., 2011; Drummond et al., 2012]. However,unless this tectonic control is extensive enough to alter theshape of the entire drainage basin, it is unlikely to stronglyaffect measurements of DW or our estimates of cumulativeerosion. Sub-basins in which tributaries have preferredorientations will not necessarily impart an elongated shape tothe entire basin. Drummond et al. [2012] tabulate the geo-metric classification of all fluvial networks mapped throughCassini swath T71. Around half of the networks with morethan 10 links in swaths T28/T29 and T39 are rectangular; allof the networks of that size in swath T13 are rectangular[Drummond et al., 2012]. Given that the highest measuredDW is for the rectangular T13 networks (Table 1), it is dif-ficult to account for Titan’s overall lowDWvalues as a resultof tectonic influences. Limited erosion provides a simplerexplanation.[54] Several approaches could provide further insight into

Titan’s erosion rates, surface ages, and resurfacing history.Better constraints on precipitation rates during rain events,and the distribution and frequency of rain events, wouldpermit better estimates of long-term erosion rates. Studies ofnon-fluvial resurfacing mechanisms such as aerosol deposi-tion and tectonics would clarify the geologic controls onsurface topography. Improved crater counting estimateswould also contribute to our understanding of the chronol-ogy of landscape change on Titan.

7. Summary and Conclusions

[55] We develop a statistic that relates the planformgeometry of a fluvial network as it erodes into bedrock to theamount of cumulative erosion. We calibrate this statisticusing a simple landscape evolution model, and show that thewidth function of a fluvial network (the frequency distribu-tion of links in the network with respect to flow distancefrom the basin outlet) becomes systematically less uniformas fluvial erosion progresses. Comparisons of recently gla-ciated landscapes versus landscapes with a longer history offluvial erosion on Earth are qualitatively consistent with ourtheoretical prediction. Using fluvial networks on westernKaua’i, where an initial topographic surface with a knownage can be reconstructed, we demonstrate that the widthfunction can be used to probabilistically assess upper andlower bounds on the ratio of cumulative erosion to initialrelief.[56] We apply these methods to fluvial networks mapped

in four regions on Titan, and find that Titan’s north polarregion has undergone a minimum of 0.4–0.6% of spatiallyaveraged erosion of an initial surface, and for a subprovincewith better constraints we suggest that there has been at most9% erosion. Titan’s Xanadu region has experienced at least0.4% erosion of its initial relief, and the southern high lati-tudes show evidence for at least 0.5% erosion of the initialsurface. Our upper limit on cumulative erosion in Titan’snorth polar region implies that, if estimates of Titan’s surfaceage based on crater counting are correct, long-term erosion

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

15 of 17

Page 16: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

rates on Titan must be many times slower than typical con-tinental erosion rates on Earth. Alternatively, some land-scapes on Titan have been resurfaced in the geologicallyrecent past.

[57] Acknowledgments. The authors thank Trent Hare for providingGIS data and Tom Farr for processing the terrestrial SAR images inFigure 10. We thank Ken Ferrier, Scott Miller, and Paul Richardson forvaluable discussions. Comments from two anonymous reviewers improvedthe manuscript. Brenda Carbone and Kerin Willis provided administrativehelp. We acknowledge the Cassini Radar Team and all those who haveworked on the Cassini mission. This work was supported by a NASACassini Data Analysis Program award to DMB and JTP.

ReferencesAharonson, O., A. G. Hayes, J. I. Lunine, R. D. Lorenz, M. D. Allison, andC. Elachi (2009), An asymmetric distribution of lakes on Titan as a pos-sible consequence of orbital forcing, Nat. Geosci., 2, 851–854,doi:10.1038/ngeo698.

Aharonson, O., A. Hayes, R. Lopes, A. Lucas, P. Hayne, and J. T. Perron(2012), Titan’s surface geology, in Titan: Surface, Atmosphere and Mag-netosphere, edited by I. Mueller-Wodarg et al., Cambridge Univ. Press,Cambridge, U. K., in press.

Artemieva, N., and J. Lunine (2003), Cratering on Titan: Impact melt,ejecta, and the fate of surface organics, Icarus, 164, 471–480,doi:10.1016/S0019-1035(03)00148-9.

Brown, R. H., L. A. Soderblom, J. M. Soderblom, R. N. Clark, R. Jaumann,J. W. Barnes, C. Sotin, B. Buratti, K. H. Baines, and P. D. Nicholson(2008), The identification of liquid ethane in Titan’s Ontario Lacus,Nature, 454, 607–610, doi:10.1038/nature07100.

Burr, D. M., J. P. Emery, R. D. Lorenz, G. C. Collins, and P. A. Carling(2006), Sediment transport by liquid surficial flow: Application to Titan,Icarus, 181, 235–242, doi:10.1016/j.icarus.2005.11.012.

Burr, D. M., R. E. Jacobsen, D. L. Roth, C. B. Phillips, K. L. Mitchell, andD. Viola (2009), Fluvial network analysis on Titan: Evidence for subsur-face structures and west-to-east wind flow, southwestern Xanadu,Geophys. Res. Lett., 36, L22203, doi:10.1029/2009GL040909.

Burr, D. M., et al. (2012), Fluvial features on Titan, GSA Bull., in press.Clark, P. U., A. S. Dyke, J. D. Shakun, A. E. Carlson, J. Clark, B. Wohlfarth,J. X. Mitrovica, S. W. Hostetler, and A. M. McCabe (2009), The LastGlacial Maximum, Science, 325, 710–714, doi:10.1126/science.1172873.

Collins, G. C. (2005), Relative rates of fluvial bedrock incision on Titan andEarth, Geophys. Res. Lett., 32, L22202, doi:10.1029/2005GL024551

Collins, G. C., L. S. Sklar, K. L. Litwin, and P. J. Polito (2012), Do Titan’sriver channels carve into ice bedrock or loose regolith?, paper presentedat Titan through Time II Workshop, NASA Goddard Space Flight Cent.,Greenbelt, Md., 3–5 April.

Drummond, S. A., D. M. Burr, R. Cartwright, B. A. Black, and J. T. Perron(2011), Global mapping and morphologic classification of Titan fluvialfeatures, Lunar Planet. Sci., XLII, 1919.

Drummond, S. A., D. M. Burr, R. Cartwright, B. A. Black, and J. T. Perron(2012), Morphologic classification and geologic implications of Titanfluvial features, Lunar Planet. Sci., XLIII, 2868.

Ehlers, J., and P. L. Gibbard (2004), Quaternary Glaciations: Extent andChronology, Part II: North America, Elsevier, New York.

Elachi, C., et al. (2004), Radar: The Cassini Titan radar mapper, Space Sci.Rev., 115, 71–110, doi:10.1007/s11214-004-1438-9.

Elachi, C., et al. (2005), Cassini radar views the surface of Titan, Science,308, 970–974, doi:10.1126/science.1109919.

Elachi, C., Wall, S., Lopes, R., Soderblom, L., Lunine, J., Lorenz, R., Janssen,M., Kirk, R. (2006), Titan’s surface by radarlight, Bull. Am. Astron. Soc., 38,Abstract 4801.

England, P., and P. Molnar (1990), Surface uplift, uplift of rocks, and exhu-mation of rocks, Geology, 18, 1173–1177, doi:10.1130/0091-7613(1990)018<1173:SUUORA>2.3.CO;2.

Eshleman, V. R., G. F. Lindal, and G. L. Tyler (1983), Is Titan wet or dry,Science, 221, 53–55, doi:10.1126/science.221.4605.53.

Farr, T. G., et al. (2007), The shuttle radar topography mission, Rev.Geophys., 45, RG2004, doi:10.1029/2005RG000183.

Fulchignoni, M., et al. (2005), In situ measurements of the physical charac-teristics of Titan’s environment, Nature, 438, 785–791, doi:10.1038/nature04314.

Griffith, C. A., T. Owen, T. R. Geballe, J. Rayner, and P. Rannou (2003),Evidence for the exposure of water ice on Titan’s surface, Science, 300,628–630, doi:10.1126/science.1081897.

Hayes, A., et al. (2008), Hydrocarbon lakes on Titan: Distribution and inter-action with a porous regolith, Geophys. Res. Lett., 35, L09204,doi:10.1029/2008GL033409

Horton, R. E. (1945), Erosional development of streams and their drain-age basins: Hydrophysical approach to quantitative morphology, Geol.Soc. Am. Bull., 56, 275–370, doi:10.1130/0016-7606(1945)56[275:EDOSAT]2.0.CO;2.

Howard, A. D. (1994), A detachment-limited model of drainage-basin evo-lution, Water Resour. Res., 30, 2261–2285, doi:10.1029/94WR00757.

Howard, A. D., and G. Kerby (1983), Channel changes in Badlands, Geol.Soc. Am. Bull., 94, 739–752, doi:10.1130/0016-7606(1983)94<739:CCIB>2.0.CO;2.

Keller, H. U., B. Grieger, M. Kuppers, S. E. Schroder, Y. V. Skorov, andM. G. Tomasko (2008), The properties of Titan’s surface at the Huygenslanding site from Disr observations, Planet. Space Sci., 56, 728–752,doi:10.1016/j.pss.2007.11.020.

Khare, B. N., C. Sagan, E. L. Bandurski, and B. Nagy (1978), Ultraviolet-photoproduced organic solids synthesized under simulated Jovian condi-tions: Molecular analysis, Science, 199, 1199–1201, doi:10.1126/science.199.4334.1199.

Kobor, J. S., and J. J. Roering (2004), Systematic variation of bedrockchannel gradients in the Central Oregon Coast Range: Implications forrock uplift and shallow landsliding, Geomorphology, 62, 239–256,doi:10.1016/j.geomorph.2004.02.013.

Kouvaris, L. C., and F. M. Flasar (1991), Phase-equilibrium of methane andnitrogen at low temperatures: Application to Titan, Icarus, 91, 112–124,doi:10.1016/0019-1035(91)90131-C.

Krasnopolsky, V. A. (2009), A photochemical model of Titan’s atmosphereand ionosphere, Icarus, 201, 226–256, doi:10.1016/j.icarus.2008.12.038.

Kuiper, G. P. (1944), Titan: A satellite with an atmosphere, Astrophys. J.,100, 378–383, doi:10.1086/144679.

Langhans, M. H., et al. (2012), Titan’s fluvial valleys: Morphology, distri-bution, and spectral properties, Planet. Space Sci., 60, 34–51,doi:10.1016/j.pss.2011.01.020.

Lavvas, P. P., A. Coustenis, and I. M. Vardavas (2008), Coupling photo-chemistry with haze formation in Titan’s atmosphere, Part I: Modeldescription, Planet. Space Sci., 56, 27–66, doi:10.1016/j.pss.2007.05.026.

Lindal, G. F., G. E. Wood, H. B. Hotz, D. N. Sweetnam, V. R. Eshleman,and G. L. Tyler (1983), The atmosphere of Titan: An analysis of theVoyager-1 radio occultation measurements, Icarus, 53, 348–363,doi:10.1016/0019-1035(83)90155-0.

Lopes, R. M. C., et al. (2010), Distribution and interplay of geologic pro-cesses on Titan from Cassini radar data, Icarus, 205, 540–558,doi:10.1016/j.icarus.2009.08.010.

Lorenz, R. D., and J. I. Lunine (1996), Erosion on Titan: Past and present,Icarus, 122, 79–91, doi:10.1006/icar.1996.0110.

Lorenz, R. D., et al. (2006), The sand seas of Titan: Cassini radar observa-tions of longitudinal dunes, Science, 312, 724–727, doi:10.1126/science.1123257.

Lorenz, R. D., et al. (2007), Titan’s young surface: Initial impact crater sur-vey by Cassini radar and model comparison, Geophys. Res. Lett., 34,L07204, doi:10.1029/2006GL028971

Lorenz, R. D., C. Newman, and J. I. Lunine (2010), Threshold of wavegeneration on Titan’s lakes and seas: Effect of viscosity and implicationsfor Cassini observations, Icarus, 207, 932–937, doi:10.1016/j.icarus.2009.12.004.

Lunine, J. I., and R. D. Lorenz (2009), Rivers, lakes, dunes, and rain:Crustal processes in Titan’s methane cycle, Annu. Rev. Earth Planet.Sci., 37, 299–320, doi:10.1146/annurev.earth.031208.100142.

Lunine, J. I., D. J. Stevenson, and Y. L. Yung (1983), Ethane ocean onTitan, Science, 222, 1229–1230, doi:10.1126/science.222.4629.1229.

Macdonald, G. A., D. A. Davis, and D. C. Cox (1960), Geology and ground-water resources of the island of Kauai, Hawaii, Hawaii Div. Hydrogr. Bull.,23, 212 pp.

Malaska, M., J. Radebaugh, K. Mitchell, R. Lopes, S. Wall, and R. Lorenz(2011), Surface dissolution model for Titan Karst, in First InternationalPlanetary Caves Workshop: Implications for Astrobiology, Climate,Detection, and Exploration, LPI Contrib. 1640, pp. 15–16, Lunar andPlanet. Inst., Houston, Tex.

McCord, T. B., et al. (2006), Composition of Titan’s surface from CassiniVims, Planet. Space Sci., 54, 1524–1539, doi:10.1016/j.pss.2006.06.007.

McKay, C. P., A. Coustenis, R. E. Samuelson, M. T. Lemmon, R. D.Lorenz, M. Cabane, P. Rannou, and P. Drossart (2001), Physical propertiesof the organic aerosols and clouds on Titan, Planet. Space Sci., 49, 79–99,doi:10.1016/S0032-0633(00)00051-9.

Mitri, G., A. P. Showman, J. I. Lunine, and R. D. Lorenz (2007), Hydrocarbonlakes on Titan, Icarus, 186, 385–394, doi:10.1016/j.icarus.2006.09.004.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

16 of 17

Page 17: Estimating erosional exhumation on Titan from drainage network morphologyweb.mit.edu/perron/www/files/Black12.pdf · 2012-08-22 · Estimating erosional exhumation on Titan from drainage

Mitri, G., A. P. Showman, J. I. Lunine, and R. M. C. Lopes (2008), Resur-facing of Titan by ammonia-water cryomagma, Icarus, 196, 216–224,doi:10.1016/j.icarus.2008.02.024.

Moore, J. M., and A. D. Howard (2010), Are the basins of Titan’s HoteiRegio and Tui Regio sites of former low latitude seas?, Geophys. Res.Lett., 37, L22205, doi:10.1029/2010GL045234.

Moore, J. M., and R. T. Pappalardo (2011), Titan: An exogenic world?Icarus, 212, 790–806, doi:10.1016/j.icarus.2011.01.019.

Neish, C. D., and R. D. Lorenz (2012), Titan’s global crater population:A new assessment, Planet. Space Sci., 60, 26–33, doi:10.1016/j.pss.2011.02.016.

Niemann, H. B., et al. (2005), The abundances of constituents of Titan’satmosphere from the GCMS instrument on the Huygens probe, Nature,438, 779–784, doi:10.1038/nature04122.

Parmentier, E. M., and J. W. Head (1981), Viscous relaxation of impactcraters on icy planetary surfaces: Determination of viscosity variationwith depth, Icarus, 47, 100–111, doi:10.1016/0019-1035(81)90095-6.

Pelletier, J. D. (2004), Persistent drainage migration in a numerical land-scape evolution model, Geophys. Res. Lett., 31, L20501, doi:10.1029/2004GL020802.

Perron, J. T., and I. de Pater (2004), Dynamics of an ice continent on Titan,Geophys. Res. Lett., 31, L17S04, doi:10.1029/2004GL019802.

Perron, J. T., and S. Fagherazzi (2012), The legacy of initial conditions inlandscape evolution, Earth Surf. Proc. Landforms, 37, 52–63, doi:10.1002/esp.2205.

Perron, J. T., M. P. Lamb, C. D. Koven, I. Y. Fung, E. Yager, andM. Adamkovics (2006), Valley formation and methane precipitation rateson Titan, J. Geophys. Res., 111, E11001, doi:10.1029/2005JE002602.

Perron, J. T., W. E. Dietrich, and J. W. Kirchner (2008), Controls onthe spacing of first-order valleys, J. Geophys. Res., 113, F04016,doi:10.1029/2007JF000977.

Perron, J. T., J. W. Kirchner, and W. E. Dietrich (2009), Formation ofevenly spaced ridges and valleys, Nature, 460, 502–505, doi:10.1038/nature08174.

Portenga, E. W., and P. R. Bierman (2011), Understanding Earth’s erodingsurface with 10Be, GSA Today, 21, 4–10, doi:10.1130/G111A.1.

Radebaugh, J., R. D. Lorenz, R. L. Kirk, J. I. Lunine, E. R. Stofan, R. M. C.Lopes, S. D. Wall, and T. Cassini Radar (2007), Mountains on Titanobserved by Cassini radar, Icarus, 192, 77–91, doi:10.1016/j.icarus.2007.06.020.

Radebaugh, J., et al. (2011), Regional geomorphology and history ofTitan’s Xanadu province, Icarus, 211, 672–685, doi:10.1016/j.icarus.2010.07.022.

Rannou, P., F. Montmessin, F. Hourdin, and S. Lebonnois (2006), The lat-itudinal distribution of clouds on Titan, Science, 311, 201–205,doi:10.1126/science.1118424.

Rodriguez, S., S. Le Mouélic, C. Sotin, H. Clénet, R. N. Clark, B. Buratti,R. H. Brown, T. B. McCord, P. D. Nicholson, and K. H. Baines (2006),Cassini/Vims hyperspectral observations of the Huygens landing site onTitan, Planet. Space Sci., 54, 1510–1523, doi:10.1016/j.pss.2006.06.016.

Rodriguez-Iturbe, I., and A. Rinaldo (1997), Fractal River Basins: Chanceand Self-Organization, Cambridge Univ. Press, Cambridge, U. K.

Rosenbloom, N. A., and R. S. Anderson (1994), Hillslope and channel evo-lution in a marine terraced landscape, Santa-Cruz, California, J. Geophys.Res., 99, 14,013–14,029, doi:10.1029/94JB00048.

Sagan, C., and S. F. Dermott (1982), The tide in the seas of Titan, Nature,300, 731–733, doi:10.1038/300731a0.

Schaller, E. L., H. G. Roe, T. Schneider, and M. E. Brown (2009), Storms inthe tropics of Titan, Nature, 460, 873–875, doi:10.1038/nature08193.

Schneider, T., S. D. B. Graves, E. L. Schaller, and M. E. Brown (2012),Polar methane accumulation and rainstorms on Titan from simulationsof the methane cycle, Nature, 481, 58–61, doi:10.1038/nature10666.

Seidl, M. A., and W. E. Dietrich (1992), The problem of channel erosioninto bedrock, in Functional Geomorphology: Landform Analysis andModels, Catena Suppl. 23, edited by K. H. Schmidt and J. de Ploey,pp. 101–124, Catena, Cremlingen, Germany.

Sklar, L. S., and W. E. Dietrich (2001), Sediment and rock strength controlson river incision into bedrock, Geology, 29, 1087–1090, doi:10.1130/0091-7613(2001)029<1087:SARSCO>2.0.CO;2.

Soderblom, L. A., et al. (2007), Topography and geomorphology of theHuygens landing site on Titan, Planet. Space Sci., 55, 2015–2024,doi:10.1016/j.pss.2007.04.015.

Sotin, C., et al. (2005), Release of Volatiles from a Possible Cryovolcano fromnear-Infrared Imaging of Titan, Nature, 435, 786–789, doi:10.1038/nature03596.

Stiles, B. W., et al. (2009), Determining Titan surface topography fromCassini SAR data, Icarus, 202, 584–598, doi:10.1016/j.icarus.2009.03.032.

Stofan, E. R., et al. (2007), The lakes of Titan, Nature, 445, 61–64,doi:10.1038/nature05438.

Tomasko, M. G., et al. (2005), Rain, winds and haze during the Huygensprobe’s descent to Titan’s surface, Nature, 438, 765–778, doi:10.1038/nature04126.

Tucker, G. E., and R. L. Bras (1998), Hillslope processes, drainage density,and landscape morphology, Water Resour. Res., 34, 2751–2764,doi:10.1029/98WR01474.

Turtle, E. P., et al. (2011), Rapid and extensive surface changes near Titan’sequator: Evidence of April showers, Science, 331, 1414–1417,doi:10.1126/science.1201063.

Wall, S., et al. (2010), Active shoreline of Ontario Lacus, Titan: A morpho-logical study of the lake and its surroundings, Geophys. Res. Lett., 37,L05202, doi:10.1029/2009GL041821.

Whipple, K. X., and G. E. Tucker (1999), Dynamics of the stream-power river incision model: Implications for height limits of mountainranges, landscape response timescales, and research needs, J. Geophys.Res., 104, 17,661–17,674, doi:10.1029/1999JB900120.

Willgoose, G., R. L. Bras, and I. Rodrigueziturbe (1991), A coupledchannel network growth and hillslope evolution model: 1. Theory, WaterResour. Res., 27, 1671–1684, doi:10.1029/91WR00935.

Wood, C. A., R. Lorenz, R. Kirk, R. Lopes, K. Mitchell, E. Stofan, andR. T. Cassini (2010), Impact craters on Titan, Icarus, 206, 334–344,doi:10.1016/j.icarus.2009.08.021.

Wye, L. C., H. A. Zebker, and R. D. Lorenz (2009), Smoothness of Titan’sOntario Lacus: Constraints from Cassini radar specular reflection data,Geophys. Res. Lett., 36, L16201, doi:10.1029/2009GL039588

Yung, Y. L., M. Allen, and J. P. Pinto (1984), Photochemistry of theatmosphere of Titan: Comparison between model and observations,Astrophys. J. Suppl. Ser., 55, 465–506, doi:10.1086/190963.

Zebker, H. A., Y. Gim, P. Callahan, S. Hensley, and R. Lorenz (2009),Analysis and interpretation of Cassini Titan radar altimeter echoes,Icarus, 200, 240–255, doi:10.1016/j.icarus.2008.10.023.

BLACK ET AL.: EROSIONAL EXHUMATION ON TITAN E08006E08006

17 of 17