Viscous relaxation as a prerequisite for tectonic resurfacing on Ganymede: Insights from numerical models of lithospheric extension

Viscous relaxation as a prerequisite for tectonic resurfacing on Ganymede: Insights from numerical models of lithospheric extension

Accepted Manuscript Viscous relaxation as a prerequisite for tectonic resurfacing on Ganymede: Insights from numerical models of lithospheric extensi...

32MB Sizes 0 Downloads 41 Views

Accepted Manuscript

Viscous relaxation as a prerequisite for tectonic resurfacing on Ganymede: Insights from numerical models of lithospheric extension Michael T. Bland, William B. McKinnon PII: DOI: Reference:

S0019-1035(17)30223-3 10.1016/j.icarus.2017.10.017 YICAR 12652

To appear in:

Icarus

Received date: Revised date: Accepted date:

17 March 2017 11 October 2017 12 October 2017

Please cite this article as: Michael T. Bland, William B. McKinnon, Viscous relaxation as a prerequisite for tectonic resurfacing on Ganymede: Insights from numerical models of lithospheric extension, Icarus (2017), doi: 10.1016/j.icarus.2017.10.017

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Highlights • We simulate grooved terrain formation with large-amplitude pre-existing topography • When initial topographic relief is ¿50 m, groove-like structures fail to form.

CR IP T

• Forming grooves directly from dark terrain by tectonic resurfacing is difficult. • Viscous relaxation of pre-existing topography enables tectonic resurfacing.

AC

CE

PT

ED

M

AN US

• Resurfacing occurred by viscous relaxation, tectonics, and cryovolcanism combined.

1

ACCEPTED MANUSCRIPT

CR IP T

Viscous relaxation as a prerequisite for tectonic resurfacing on Ganymede: Insights from numerical models of lithospheric extension

1 2

AN US

Michael T. Bland1 and William B. McKinnon2

U. S. Geological Survey, Astrogeology Science Center, Flagstaff AZ 86001

Department of Earth and Planetary Sciences and McDonnell Center for the Space

PT

ED

M

Sciences, Washington University in St. Louis, Saint Louis, MO 63130

AC

CE

65 Total pages, including 20 Figures Resubmitted to Icarus October 13, 2017

ACCEPTED MANUSCRIPT

Corresponding Author: Michael T. Bland U. S. Geological Survey, Astrogeology Science Center

AN US

2255 N. Gemini

CR IP T

Proposed Running Head: Viscous relaxation and tectonic resurfacing on Ganymede.

Flagstaff, AZ 86001 [email protected] phone: (928) 556-7080

AC

CE

PT

ED

M

fax: (928) 556-7014

2

ACCEPTED MANUSCRIPT

1

Abstract Ganymede’s bright terrain formed during a near-global resurfacing event (or events) that

3

produced both heavily tectonized and realatively smooth terrains. The mechanism(s) by

4

which resurfacing occurred on Ganymede (e.g., cryovolcanic or tectonic), and the relationship

5

between the older, dark and the younger, bright terrain are fundamental to understanding

6

the geological evolution of the satellite. Using a two-dimensional numerical model of litho-

7

spheric extension that has previously been used to successfully simulate surface deformation

8

consistent with grooved terrain morphologies, we investigate whether large-amplitude preex-

9

isting topography can be resurfaced (erased) by extension (i.e., tectonic resurfacing). Using

10

synthetically produced initial topography, we show that when the total relief of the initial

11

topography is larger than 25-50 m, periodic groove-like structures fail to form. Instead, ex-

12

tension is localized in a few individual, isolated troughs. These results pose a challenge to the

13

tectonic resurfacing hypothesis. We further investigate the effects of preexisting topography

14

by performing suites of simulations initialized with topography derived from digital terrain

15

models of Ganymede’s surface. These include dark terrain, fresh (relatively deep) impact

16

craters, smooth bright terrain, and a viscously relaxed impact crater. The simulations using

17

dark terrain and fresh impact craters are consistent with our simulations using synthetic

18

topography: periodic groove-like deformation fails to form. In contrast, when simulations

19

were initialized with bright smooth terrain topography, groove-like deformation results from

20

a wide variety of heat flow and surface temperature conditions. Similarly, when a viscously

21

relaxed impact crater was used, groove-like structures were able to form during extension.

22

These results suggest that tectonic resurfacing may require that the amplitude of the initial

24

AN US

M

ED

PT

CE

AC

23

CR IP T

2

topography be reduced before extension begins. We emphasize that viscous relaxation may be the key to enabling tectonic resurfacing, as the heat fluxes associated with groove terrain

25

formation are also capable of reducing crater topography through viscous relaxation. For

26

long-wavelength topography (large craters) viscous relaxation is unavoidable. We propose

27

that the resurfacing of Ganymede occurred through a combination of viscous relaxation,

28

tectonic resurfacing, cryovolcanism and, at least in a few cases, band formation. Variations 3

ACCEPTED MANUSCRIPT

29

in heat flow and strain magnitudes across Ganymede likely produced the complex variety of

30

terrain types currently observed.

CE

PT

ED

M

AN US

CR IP T

Keywords: Ganymede; Tectonics.

AC

31

4

ACCEPTED MANUSCRIPT

32

1

Ganymede’s terrains and evidence for resurfacing At large scales, Ganymede’s surface is composed of two types of terrain: relatively dark

34

terrain with higher crater densities, and relatively bright terrain with lower crater densities

35

(Fig. 1). The dark terrain likely represents an ancient, though not necessarily primordial

36

(e.g., crater densities are higher on neighboring Callisto (Zahnle et al., 2003; Schenk et al.,

37

2004)) crust. The bright terrain, in contrast, is a resurfaced unit formed during Ganymede’s

38

mid-life (∼2 Ga ago (Zahnle et al., 2003), although uncertainties are at least 1 Ga (Schenk

39

et al., 2004)). Understanding how the bright (and often “grooved”) terrain was emplaced and

40

the relationship between Ganymede’s dark and bright terrains are central to understanding

41

the overall geologic history of the satellite.

AN US

CR IP T

33

Ganymede’s dark terrain comprises roughly one-third of its surface (e.g., Patterson et al.,

43

2010) and is heavily cratered, with numerous knobs and massifs (Shoemaker et al., 1982;

44

Prockter et al., 1998) (Fig. 1A). The low albedo material is probably a relatively thin,

45

silicate-rich veneer that overlies an icier substrate (Prockter et al., 1998). This substrate is

46

apparently exposed in the rims of furrows (troughs) and craters, where the dark surface lag

47

has sloughed-off down-slope and accumulated in topographic lows. The icy substrate is also

48

exposed in crater palimpsests and the relatively bright ejecta associated with many dark

49

terrain craters (e.g., Osiris) (Pappalardo et al., 2004).

PT

ED

M

42

Within the dark terrain and far from the boundaries with the bright terrain (e.g., within

51

the the central portion of Galileo Regio), tectonic deformation is limited (Prockter et al.,

52

1998), consisting primarily of ancient, arcuate furrows up to ∼1 km deep, that may have

53

formed during ancient, basin-scale impacts (McKinnon and Melosh, 1980; Schenk and McK-

55

AC

54

CE

50

innon, 1987; Murchie et al., 1990; Prockter et al., 1998). Development and modification of the dark terrain’s furrows, massifs and knobs (possibly remnants of crater rims) likely

56

occurred during an older epoch, and in a different deformation style, than that of the bright

57

grooved terrain (Prockter et al., 1998, 2000). Nearer the dark-bright terrain boundary, more

58

substantial tectonic deformation has occurred. In some cases the dark terrain is heavily

59

fractured, and takes on a morphology similar to that of the bright terrain (Patterson et al., 5

ACCEPTED MANUSCRIPT

60

2010; and Fig. 1B). Because of their similar morphology and spatial association with bright

61

terrain, the tectonized dark terrain may be a precursor or transitional unit to the bright

62

terrain (Prockter et al., 2000; Patterson et al., 2010). The remaining two-thirds of Ganymede’s surface consists of a brighter patchwork of broad

64

polygons that are often (though not always) separated from the dark terrain by deep bound-

65

ing troughs (Pappalardo et al., 1998). Surface morphologies within each polygon are highly

66

variable, but range from highly tectonized “grooved terrain” (Fig. 1C) to relatively smooth

67

plains (Fig. 1D). The classic grooved terrain morphology consists of sets of periodically

68

spaced ridges and troughs with amplitudes of several hundred meters and ridge spacing of

69

3-17 km, varying from one set of grooves to the next (see Pappalardo et al., 1998, 2004,

70

for comprehensive reviews). High resolution Galileo data have also revealed a second set of

71

finer-scale lineations with a characteristic spacing of 1 km (Pappalardo et al., 1998). This

72

iconic morphology is found throughout the bright terrain; however, substantial variation ex-

73

ists, with isolated troughs and more subdued tectonic deformation found across the surface

74

(Patterson et al., 2010).

M

AN US

CR IP T

63

The bright terrain likely formed during extension of Ganymede’s ice lithosphere (Pap-

76

palardo et al., 1998), possibly during an epoch of global satellite expansion caused by differ-

77

entiation (e.g., Squyres, 1980; Mueller and McKinnon, 1988), or melting during resonance

78

passage (Showman et al., 1997; Bland et al., 2009). The iconic ridges and troughs of the

79

grooved terrain are thought to have formed via an extensional instability that deformed

80

the lithosphere into periodically spaced pinches and swells (Pappalardo et al., 1998; Collins

81

et al., 1998; Dombard and McKinnon, 2001; Bland and Showman, 2007), a process essentially

82

identical to boudinage formation but at lithospheric scale (Fletcher and Hallet, 1983). Re-

84

PT

CE

AC

83

ED

75

cent numerical models that include strain localization effects (e.g., via material weakening and/or non-associated plasticity) support this interpretation, permitting the formation of

85

large-amplitude, periodic deformation at relatively small strains (Bland et al., 2010; Bland

86

and McKinnon, 2015).

87

The lower crater density within the bright terrain suggests that these terrains formed

6

ACCEPTED MANUSCRIPT

at the expense of the darker, heavily cratered terrain during an epoch of near-planetary-

89

scale resurfacing; however, the process or processes by which such resurfacing has occurred

90

remain poorly understood. Resurfacing by cryovolcanism (i.e., the eruption of a fluid onto the

91

surface), tectonism (i.e., the disruption and erasure of preexisting terrains through tectonic

92

deformation alone), and Europa-like band formation (i.e., the emplacement of new material

93

during extension of the surface) have all been proposed.

CR IP T

88

Cryovolcanism has been invoked as a resurfacing mechanism on Ganymede since the

95

Voyager-era. In these models, an initial period of extension forms broad graben, which

96

are subsequently flooded with cryovolcanic material (water, warm ice, or slush) that buries

97

preexisting topography. The grooved terrain then forms during a continuing or later period

98

of extensional tectonism (Golombek and Allison, 1981; Allison and Clifford, 1987), possibly

99

with subsequent viscous relaxation or mass wasting of the resulting topography (Squyres,

100

1982; Parmentier et al., 1982). Evidence for cryovolcanic resurfacing comes from numerous,

101

unusually smooth regions on the surface (Fig. 2) as well as a number of unusual, caldera-like

102

features (Schenk et al., 2001). Although high-resolution imaging is limited, at least some of

103

the smooth regions appear to embay nearby topography (Fig. 2) (Allison and Clifford, 1987;

104

Schenk et al., 2001), and occur in topographic lows (Schenk et al., 2001).

ED

M

AN US

94

Cryovolcanic resurfacing of Ganymede faces several theoretical and observational chal-

106

lenges. Any cryovolcanic resurfacing mechanism proposed for icy satellites must address the

107

difficulty of erupting relatively higher density fluid (e.g., salty water) through the lower den-

108

sity ice shell. Such eruptions may be driven by gases entrained in the fluid (Crawford and

109

Stevenson, 1988) or by topographically induced pressure gradients that drive near-surface

110

fluid onto the surface (Showman et al., 2004), but the exact mechanism remains elusive.

112

CE

AC

111

PT

105

Additionally, there is a paucity of direct evidence for associated cryovolcanic source vents in higher-resolution Galileo images, although numerous enigmatic caldera-like features have

113

been identified in the Mummu Sulci region (previously identified as the Sippar Sulcus region)

114

and elsewhere (Schenk et al., 2001). Further, at least some regions that appeared smooth in

115

low-resolution Voyager and Galileo images (e.g., Harpagia Sulcus) appear lightly tectonized

7

ACCEPTED MANUSCRIPT

116

when seen in higher-resolution imaging (Pappalardo et al., 2004). Evidence for tectonic imbrication (Pappalardo et al., 1998; Collins et al., 1998), the iden-

118

tification of nascent rifts within the dark terrain (Prockter et al., 2000), and the challenges

119

faced by cryovolcanic resurfacing described above have led to the hypothesis that Ganymede

120

may have been resurfaced through tectonism alone (Head et al., 1997) (Fig. 3). In this view,

121

grooved terrain forms directly from dark terrain by fault imbrication and block rotation,

122

which completely erases dark terrain topography without large-scale cryovolcanic resurfac-

123

ing (Head et al., 1997; Pappalardo and Collins, 2005). Key to this process is the removal of

124

the dark surface lag by slumping of dark material into topographic lows (Patel et al., 1999).

125

The existence of geologic units that are morphologically and stratigraphically transitional

126

between “traditional” bright and dark terrain (e.g., the dark lineated units described above)

127

may indicate a progression from dark terrain to bright terrain. These transitional units

128

always occur adjacent to bright terrain and generally share its structural fabric (Patterson

129

et al., 2010). The tectonic resurfacing paradigm has also recently been reinforced by Cassini

130

observations of numerous fault cut craters on Enceladus (e.g., Crow-Willard and Pappalardo,

131

2015). Yet the tectonic resurfacing hypothesis is not without its challenges. Analog models

132

of Ganymede-like extension reproduce many of the distinctive characteristics of the grooved

133

terrain, including extensive faulting and block rotation (Sims et al., 2014). Models that in-

134

cluded preexisting craters, however, failed to result in resurfacing: in a blind test, 95% of the

135

initially imposed craters were still identifiable after 33% extension of the surface (Wyrick,

136

2012). Models with larger strains (e.g., Pappalardo and Collins (2005) observed Ganymede

137

craters deformed by >50% extension) have not yet been performed. Although supported

138

by mapping and structural analysis, the detailed mechanics of tectonic resurfacing requires

140

AN US

M

ED

PT

CE

AC

139

CR IP T

117

further investigation. Head et al. (2002) suggested Europa-like band formation as an alternative (or additional)

141

mechanism of resurfacing on Ganymede. Band formation is common on Europa (see Prockter

142

and Patterson, 2009, and references therein), and is generally thought to involve complete

143

separation of the lithosphere and the emplacement of new material as warmer ice wells

8

ACCEPTED MANUSCRIPT

upward from below (Prockter et al., 2002). The principal type locality for possible band

145

formation on Ganymede is Arbela Sulcus, a narrow swath of relatively smooth terrain that

146

can be “closed” to reconstruct the preexisting surface (Fig. 4). Such tectonic reconstructions

147

are rare on Ganymede (though see Pizzi et al., 2017, for potential additional examples), but

148

the importance of band formation as a resurfacing mechanism is difficult to assess given the

149

general lack of high-resolution imaging available for Ganymede. On Europa, band formation

150

may be associated with substantial amounts of surface strain (Bland and McKinnon, 2012).

151

In reality it is possible that all three mechanisms (cryovolcanism, tectonism, and band

152

formation) have operated on Ganymede at either some specific point in time, some particu-

153

lar locality, or in concert. In this paper we assess the plausibility and conditions necessary

154

for resurfacing Ganymede through tectonic deformation alone. That is, we seek to specif-

155

ically test the tectonic resurfacing hypothesis through numerical modeling. In doing so,

156

however, we also indirectly address the potential role of cryovolcanic resurfacing. We also

157

identify a process that appears critical to enabling tectonic resurfacing: viscous relaxation

158

of preexisting topography.

159

2

ED

M

AN US

CR IP T

144

Methods and previous modeling To numerically simulate lithospheric extension on Ganymede, we follow an identical ap-

161

proach to that of Bland and McKinnon (2015), with the exception of the initial topography

162

imposed. We use the viscoelastic-plastic finite element code Tekton (v2.3) (Melosh and Raef-

163

sky, 1980) in a two-dimensional Cartesian plane strain geometry. Many features of rifting

164

on terrestrial planets and icy worlds exhibit along-strike symmetry and are therefore con-

166

CE

AC

165

PT

160

ducive to a two-dimensional modeling approach. Indeed, the majority of investigations of rifting on Earth have utilized a two-dimensional geometry, which facilitates the investigation

167

of large parameter spaces at high resolution (e.g., Bassi, 1991; Govers and Wortel, 1995;

168

Beaumont et al., 1996; Buck et al., 1999; Gerbault et al., 1999; Lavier et al., 2000; Huis-

169

mans and Beaumont, 2002, 2003; Wijns et al., 2005; Nagel and Buck, 2007; Delescluse et al.,

9

ACCEPTED MANUSCRIPT

2008; Whitney et al., 2013; Gueydan and Pr´ecigout, 2014, and many others). Furthermore,

171

three-dimensional simulations of the rifting of terrestrial continental crust indicate that the

172

resulting deformation is fundamentally two-dimensional in nature (Sharples et al., 2015),

173

and most three-dimensional studies focus on along-strike rift propagation (e.g., Dunbar and

174

Sawyer, 1996; Van Wijk and Blackman, 2005; Allken et al., 2011, 2012). Despite this, the

175

interaction between extension, strain localization, and large-scale preexisting topography has

176

received little investigation, and strain localization during extension may be modified when

177

a fully three dimensional geometry is used. The hazards of modeling a three-dimensional

178

process with two-dimensional simulations are discussed more fully in section 4.6. The mod-

179

els presented here are therefore a necessary first-step to numerically investigating tectonic

180

resurfacing, but do not tell the complete story. Three-dimensional simulations of groove

181

formation are currently in progress (Bland and Wyrick, 2017).

AN US

CR IP T

170

Our two-dimensional simulation domain is initially 80 km long and 24 km deep, ensuring

183

edge and bottom boundary effects are negligible, and allowing numerous instability wave-

184

lengths to develop at the surface. The bottom boundary is fixed in the vertical and free slip

185

in the horizontal, and the sides are free slip in the vertical. Extension is imposed by keeping

186

the left edge of the domain fixed in the horizontal and imposing a constant velocity boundary

187

condition on the right edge of the domain. We impose up to 10% extension at a strain rate

188

of 10−13 s−1 (i.e., 1% extension every ∼3000 yrs). Decreasing the strain rate, which may

189

be more geologically plausible (Dombard and McKinnon, 2001; Stempel et al., 2005), has

190

only a small effect on our results, and generally leads to increased instability amplification

191

rates (Bland and McKinnon, 2015; Dombard and McKinnon, 2001). The top surface is free,

192

and we impose initial topography as described below. Each simulation is initialized with a

194

ED

PT

CE

AC

193

M

182

lithostatic stress state using a surface gravity of 1.4 m s−2 . The density of the ice is assumed to be 950 kg m−3 , consistent with cold, somewhat dirty ice.

195

The simulations include all relevant viscous flow mechanisms for ice I, including dislo-

196

cation creep (three regimes), grain boundary sliding (GBS), basal slip (BS), and diffusion

197

creep. Each flow mechanism can act simultaneously, although GBS and BS rate-limit each

10

ACCEPTED MANUSCRIPT

other such that the slower mechanism dominates (Goldsby and Kohlstedt, 2001). At the

199

stress, strain rate, and temperatures used in our simulations, deformation is predominantly

200

controlled by GBS flow and dislocation creep regime ‘B’ (see Durham and Stern, 2001). More

201

detail is provided in Bland and McKinnon (2015) and Bland and Showman (2007). We use

202

elastic parameters consistent with cold intact ice: a Young’s modulus (E) of 9.33 GPa and

203

a Poisson’s ratio of 0.325 (Gammon et al., 1983). A pervasively fractured lithosphere might

204

have a lower E (e.g. Pritchard and Stevenson, 2000; Klimczak et al., 2015), which would

205

generally suppresses amplification rates (Bland and Showman, 2007); however, said fractures

206

would also produce a compensating effect by lowering the conductivity and thus increasing

207

the thermal gradient and instability growth rate.

AN US

CR IP T

198

Brittle behavior is modeled using non-associative plasticity, which results in strong lo-

209

calization of brittle deformation (Rudnicki and Rice, 1975; Vermeer and de Borst, 1984;

210

Poliakov and Herrmann, 1994). Non-associated plasticity is appropriate for granular geo-

211

logic materials (including ice (Lade, 2002; Pritchard, 1988)) that undergo limited dilation

212

during shear failure (Vermeer and de Borst, 1984). Plastic failure occurs when the second in-

213

variant of the deviatoric stress reaches the defined yield criterion. We use a Drucker-Praeger

214

yield criterion (Owen and Hinton, 1980; Vermeer and de Borst, 1984; Iwashita and Oda,

215

1999) with a nominal cohesion of 1 MPa (Beeman et al., 1988) and an internal friction angle

216

of 30◦ . The Drucker-Praeger criterion is similar to a Mohr-Coulomb criterion but is more

217

numerically stable. We also investigate how using a lower cohesion (100 kPa) affects our

218

results. This lower cohesion is more consistent with the failure threshold inferred for cycloid

219

formation on Europa (Rhoden and Hurford, 2013). Our implementation of non-associative

220

plasticity is described in detail in Bland and McKinnon (2015).

222

ED

PT

CE

AC

221

M

208

Although Tekton does not include thermodynamics, we impose a thermal structure

through the temperature-dependent viscosity. For each simulation we choose a heat flux

223

(F ) and surface temperature (Ts ), and assume a temperature-dependent thermal conduc-

224

tivity appropriate for intact ice (k = 651 W m−1 /T ) (Petrenko and Whitworth, 1999) to

225

calculate the temperature of each element as a function of depth (z) assuming a conductive

11

ACCEPTED MANUSCRIPT

226

steady state:  T (z) = Ts exp F z/651 W m−1 .

(1)

We use the temperature of each element to calculate its rheologic parameters (i.e., the effec-

228

tive viscosity for each of the flow mechanism described above). For the simulations shown

229

here we use only constant (in time) thermal/viscosity structures (cf. Bland et al., 2017). The

230

temperature and viscosity field is not updated during the simulation itself, so isotherms are

231

advected with the material contours. Some consequences of this simplification are discussed

232

in section 4.3.1. We nominally use a surface temperature of 100 K, but investigate the effect

233

of surface temperatures between 70 K and 120 K, as appropriate for colder polar regions and

234

warmer equatorial dark terrain, respectively. We investigate heat fluxes ranging from 70 to

235

150 mW m−2 , consistent with previous inferences for grooved terrain formation (Dombard

236

and McKinnon, 2001; Bland and Showman, 2007). Note that in reality, the thermal con-

237

ductivity may be substantially lower than that of intact ice due to both micro and macro

238

porosity throughout Ganymede’s lithosphere; however, decreasing the conductivity has the

239

same effect as increasing the heat flux. Modifying either (or both) effectively changes the

240

thickness of the lithosphere, and the length scale over which the ice transitions from brittle

241

to ductile behavior (i.e., the rheologic contrast between the ice lithosphere and subjacent ice

242

asthenosphere).

PT

ED

M

AN US

CR IP T

227

The significant difference between the simulations described here and those described in

244

Bland and McKinnon (2015) is the initial topography imposed at the surface of the simulation

245

domain. In our previous successful simulations, we used semi-random topography created by

246

randomly phase shifting and co-adding sinusoidal perturbations with wavelengths from 1 to

248

AC

247

CE

243

80 km. This topography was generally normalized to have a maximum total relief (i.e., the difference between the highest and lowest point) of just 15 m. In these simulations, the goal

249

was to provide enough initial topography to allow an instability to initiate and naturally

250

establish a dominant wavelength, which is a function of the thickness of the deforming layer

251

(e.g., Fletcher and Hallet, 1983; Herrick and Stevenson, 1990; Dombard and McKinnon,

252

2001; Bland and Showman, 2007, and many others). An example of a successful simulation 12

ACCEPTED MANUSCRIPT

using low-amplitude topography is shown in Fig 5. Extension of the initial topography in

254

panel A results in the periodic groove-like deformation shown in panel B. The simulation

255

used Ts =70 K and F =100 mW m−2 . The resulting surface deformation is similar to the

256

topography of Ganymede’s archetypal grooved terrain as shown in panel D. Clearly, these

257

simulations can successfully reproduce the wavelength, amplitude, and general morphology of

258

Ganymede’s grooved terrain when low-amplitude initial topography is used. The simulations

259

are described more fully in Bland and McKinnon (2015)

CR IP T

253

As clearly illustrated in Fig. 5, however, the amplitude of the initial topography used in

261

our previous simulations (panel A) is inconsistent with the amplitude of dark terrain topog-

262

raphy that may have predated groove formation on Ganymede (panel C), which if resurfacing

263

of dark terrain occurred by tectonism, consisted of knobs, massifs, and hummocky terrain

264

with amplitudes of several hundred meters. Further, many groove lanes appear to disrupt

265

older groove sets, the ridges and troughs of which were also likely several hundred meters

266

in amplitude. In either case, groove formation had to modify large-amplitude preexisting

267

topography if tectonic resurfacing was the dominant resurfacing mechanism. In this work

268

we examine how such large amplitude deformation modifies the previous results of Bland

269

and McKinnon (2015). To do this we utilize two types (or cases) of initial topography. The

270

first case uses “synthetic” topography created using the methods described above, but at

271

a variety of initial amplitudes (Section 3). This permits us to systematically examine how

272

the amplitude of the initial topography affects groove formation. In the second case we use

273

topography extracted from digital terrain models (DTMs) of Ganymede’s surface (Section

274

4). This permits us to investigate how grooves may have formed in the presence of different

275

initial terrain types. In both cases, surface topography is imposed simply by varying the

277

M

ED

PT

CE

AC

276

AN US

260

position of surface and near-surface nodes, with the topographic perturbation decaying with depth such that “flat” material contours are reached at depth (i.e., vertical nodal position

278

is unperturbed). After describing the results of these two investigations, we discuss the

279

implications of our modeling for resurfacing on Ganymede (Section 5).

13

ACCEPTED MANUSCRIPT

280

281

3

Simulating tectonic resurfacing with synthetic initial topography In order to examine how initial topographic amplitude affects our simulations of groove

283

terrain formation, we performed a suite of simulations in which we kept the form of the

284

initial topography constant but varied its initial amplitude from 1 m to 200 m of total

285

relief (left column of Fig. 6). Each simulation was otherwise identical: a strain rate of

286

10−13 s−1 and 10% total extension, Ts =100 K, F =100 mW m−2 , and a cohesion of 1 MPa.

287

Figure 6 shows the effect of increasing initial topographic amplitudes. For reference, the

288

simulations described in Bland and McKinnon (2015) used an initial amplitude of 12 m.

289

At relatively small initial amplitudes (1 to 25 m) extension results in strongly periodic

290

deformation. Little, if any, signature of the pre-existing terrain remains. These results are

291

consistent with Bland and McKinnon (2015). As the initial amplitude becomes larger (50-

292

100 m), however, the deformation that results from extension becomes less periodic with

293

strain preferentially localizing in several individual troughs, which deepen at the expense of

294

more distributed deformation. At the largest initial amplitudes investigated (150-200 m),

295

extension results in the formation of several isolated troughs ∼500 m deep, with little or no

296

periodicity in the final topography.

ED

M

AN US

CR IP T

282

The decrease in the periodicity of the simulated final topography as the amplitude of the

298

initial topography is increased was quantified using Fourier analysis of the final simulated

299

deformation (Fig. 7). As expected, the initial topography has an extremely broad power

300

spectrum that includes contributions from many wavelengths (inset of Fig. 7). When small-

301

amplitude initial topography is used, the power spectrum of the final topography is strongly

303

CE

AC

302

PT

297

peaked with a wavelength of 6 km. The full width half maximum (FWHM) of the peak is 1.5 km. Note that because the resolution of the spectrum decreases with increasing wavelength,

304

the exact FWHM is poorly resolved. As the amplitude of the initial topography is increased,

305

the peak in the final topography power spectrum broadens. That is, a broader range of

306

wavelengths contribute to the topography, and the topography is less periodic. In the 200 m 14

ACCEPTED MANUSCRIPT

amplitude simulation, the resulting power spectrum has a peak with a FWHM of ∼10 km.

308

These values compare poorly to Fourier analysis of Ganymede’s actual grooved terrain,

309

which indicate that groove spacing is strongly periodic, with a power spectrum FWHM of

310

just 2-3 km (Grimm and Squyres, 1985; Patel et al., 1999). The observed power spectra are

311

therefore consistent with our simulations that use low-amplitude initial topography, but not

312

those that use larger initial amplitudes.

CR IP T

307

The influence of the initial topography is further illustrated in Fig. 8, which compares

314

the development of surface deformation for two simulations that used 25 m and 200 m

315

amplitude initial topography, respectively, but were otherwise identical. When initial topo-

316

graphic amplitudes are small, the dominant wavelength of the necking and/or localization

317

instability resulting from lithospheric extension controls the surface deformation. Horizon-

318

tal strain becomes partitioned periodically throughout the lithosphere and the surface is

319

consequently deformed into periodic structures with an amplitude of ∼150 m (a factor of

320

6 increase after 10% extension). In contrast, when the initial topography is larger in am-

321

plitude, strain preferentially becomes partitioned into preexisting structures as extension

322

occurs. In the simulation shown in Fig. 8B these structures are generally topographic lows.

323

As extension occurs, strain is accommodated nearly entirely within these structures (Fig.

324

8d), which deepen at the expense of surrounding regions. The final surface is composed

325

of isolated troughs a few kilometers wide and (in some cases) more than 500 m deep, each

326

separated by relatively undeformed regions (i.e., having similar small-scale topography as

327

initially present) 10-20 km wide. This morphology generally does not resemble Ganymede’s

328

archetypal grooved terrain.

330

331

M

ED

PT

CE

Figures 6, 7, and 8 indicate that simulations that utilize small-amplitude topography (up

AC

329

AN US

313

to 25-50 m) will result in periodic groove-like deformation, whereas simulations that utilize initial topography with a total relief >50 m may result in non-periodic deformation (deep

332

isolated troughs). To assess whether this result depends on our model assumptions, we per-

333

formed a suite of simulations that varied the heat flux (70-150 mW m−2 ), surface temperature

334

(70-120 K), and strain rate (10−13 s−1 - 10−14 s−1 ). Variations in the mechanical properties of 15

ACCEPTED MANUSCRIPT

the ice were also explored and are discussed further below. These simulations indicate that

336

the basic result described above is largely independent of the physical parameters assumed.

337

Figure 9 compares the results of pairs of identical simulations that used a variety of physical

338

conditions, but with initial topographic amplitudes (maximum relief) of 25 m and 150 m.

339

In general, cold surface temperatures result in larger deformation amplitudes compared to

340

warm surface temperatures, and high heat fluxes result in shorter wavelength (and lower

341

amplitude) deformation than lower heat fluxes. These results are consistent with those of

342

Bland and McKinnon (2015), where they are described in greater detail. More notably, in

343

every case the simulations initialized with low-amplitude topography resulted in periodic or

344

semi-periodic groove-like deformation whereas simulations initialized with large amplitude

345

topography resulted in a few (or in some cases single) deep, isolated troughs. We conclude

346

that the qualitative results described above are general, and do not depend on the physical

347

parameters assumed during groove formation.

AN US

CR IP T

335

The transition between relatively periodic deformation and the formation of isolated

349

troughs consistently occurs when initial topographic relief is 25-50 m in amplitude. That is, in

350

each case considered, simulations initialized with topography with relief greater than (and in

351

many cases equal to) 50 m result in isolated troughs rather than periodic structures. However,

352

some conditions lead to greater strain localization than others. Cold surface temperatures

353

(e.g., 70 K), and low heat fluxes (70 mW m−2 ) both result in strong localization such that

354

strain is often partitioned into a single isolated trough (e.g., Fig. 9D and H). As shown in

355

Fig. 9C, E, G, and I, simulations that produce large-amplitude periodic structures when

356

initialized with low-amplitude topography yield stronger localization when large-amplitude

357

initial topography is used. In contrast, warm surface temperatures (120 K, Fig. 9F) result

359

360

ED

PT

CE

AC

358

M

348

in relatively more periodic (although still somewhat localized) deformation even when large amplitude initial topography is used. This pattern reflects the tendency of the assumed parameters to result in large-amplitude deformation in general.

361

Fundamentally, localization of strain within troughs is the result of higher stresses, and

362

hence greater brittle failure, within the troughs. If one considers the lithosphere as a plate

16

ACCEPTED MANUSCRIPT

with variable thickness (due to topography) that is subjected to an axial force (Fa ) that

364

is constant with depth, the axial stress on a vertical surface through the plate is simply

365

Fa /h(x), where h(x) is the thickness at horizontal point x. Since the base of the lithosphere

366

is initially flat (or flatter), large-scale topography (like a trough) reduces h locally, and so the

367

stress is increased locally. The increase stress results in somewhat higher strain rates in these

368

regions, but more importantly increases the likelihood of brittle failure in those elements.

369

The higher strain rate associated with brittle failure, and the localization effects of non-

370

associated plasticity, results in additional local thinning, which further increases the local

371

stresses, such that a positive feedback results. This pattern has been verified by examining

372

the stress field early in the simulations. For low-amplitude topography this effect is small,

373

and other processes (e.g., periodic necking) dominate. When initial topography is large

374

however, such continuous necking (thinning in a single region) can occur. As described

375

above, conditions that favor brittle deformation (e.g., a cold lithosphere) more easily localize

376

strain in this manner.

AN US

CR IP T

363

Our simulations using synthetic topography challenge the tectonic model of resurfacing

378

Ganymede. They suggest that low-amplitude initial topography with .25–50 m total relief

379

can easily be “erased” and replaced by periodic ridges and troughs, whereas larger amplitude

380

topography is harder to resurface, resulting in localized isolated troughs. Even excluding

381

impact craters, typical dark terrain on Ganymede (from which the grooved terrain is thought

382

to form) includes topographic amplitudes of several hundred meters (see below). The sim-

383

ulations described above suggest that simply extending such terrain does not easily result

384

in groove-like deformation. Other resurfacing mechanisms, perhaps working in concert with

385

tectonic resurfacing, are apparently required. Below we use topography derived from a vari-

387

ED

PT

CE

AC

386

M

377

ety of actual terrains on Ganymede to further investigate the interaction between extension and initial topography.

17

ACCEPTED MANUSCRIPT

388

4

Simulating tectonic resurfacing with DTM-derived initial topography

389

To further evaluate how lithospheric extension has modified pre-existing structures on

391

Ganymede, we performed simulations of lithospheric extension that were initialized with to-

392

pography derived directly from Ganymede’s surface. The investigation focused on four types

393

of terrain or landforms: dark terrain, smooth terrain, fresh impact craters, and viscously re-

394

laxed impact craters. Each terrain type is discussed in detail below.

395

4.1

AN US

Forming grooves from dark terrain

CR IP T

390

As described above, the tectonic resurfacing hypothesis postulates that Ganymede’s

397

bright grooved terrain formed directly from dark terrain (Head et al., 1997; Prockter et al.,

398

2000; Pappalardo and Collins, 2005). Although generally uniform in appearance at low

399

spatial resolution, the dark terrain is quite rough when viewed at high resolution, and the

400

individual knobs, hummocks, and massifs constitute relatively large-amplitude topography.

401

As demonstrated in section 3, the formation of periodically spaced tectonic structures, in

402

which strain is broadly distributed, from such large-amplitude topography is challenging.

ED

M

396

To more directly assess how grooved terrain might have formed directly from dark terrain

404

we performed a suite of simulations that explicitly used dark terrain topography as their

405

initial condition. We extracted topographic profiles from a region of dark terrain in northern

406

Marius Regio. The region is bounded by Nippur Sulcus to the east and northeast, Byblus

407

Sulcus to the west and Philus Sulcus to the north. The DTM and topographic profiles derived

409

CE

AC

408

PT

403

from it are shown in Fig. 10. The DTM used was derived from Galileo-based photoclinometry (e.g., Schenk, 1989; Bierhaus and Schenk, 2010), and has a horizontal resolution of ∼1 km

410

and a vertical uncertainty of ∼5%. The profiles were intentionally selected to avoid large-

411

scale craters or crater forms. Modification of impact craters is discussed in detail in section

412

4.3. Based on these 14 profiles, topographic relief associated with the largest individual

413

topographic features (hummock or knobs) is typically 200 m, with one feature exceeding 18

ACCEPTED MANUSCRIPT

400 m. Many knobs are significantly smaller (∼100 m). Over longer wavelengths (tens to

415

hundreds of kilometers), total relief in this region of dark terrain can exceed 400 m (e.g.,

416

profiles 11, 13, and 14 in Fig. 10). Two topographic profiles were selected for use in our finite

417

element simulations (1 and 3 in Fig. 10). These profiles were selected because they appear

418

typical of the dark terrain (total relief in profile 1 and 3 is 325 m, and 300 m respectively)

419

and have limited long wavelength (∼50-100 km) slopes.

CR IP T

414

For both profiles, an 80 km-long portion was extracted, interpolated to the resolution

421

of the finite element model, and imposed on the surface of the model domain. Extension

422

was imposed as described above under a variety of conditions. Figure 11 shows the defor-

423

mation of the surface and lithosphere resulting from 10% extension with a Ts =100 K and

424

F =100 mW m−2 . The result is similar to that shown in Fig. 8B. For profile 1 (Fig. 11A),

425

extension has been primarily accommodated within two narrow (<5 km wide) troughs that

426

reach depths of 500 m and >1 km, respectively. These deep troughs originate in topographic

427

lows in the initial preexisting topography. Lower-amplitude (150-200 m) ridges and troughs

428

have formed between the deep troughs. Spectral analysis of the surface deformation suggests

429

limited periodicity: the power spectrum of the final topography is peaked at 12 km, but the

430

FWHM is greater than 10 km. Overall, the surface deformation, particularly the 1 km-deep

431

trough, is inconsistent with the morphology of Ganymede’s grooved terrain (although see

432

discussion in section 4.3).

PT

ED

M

AN US

420

The initial topography in profile 3 is more subdued and the resulting deformation is also

434

lower in in amplitude. Like the simulation using profile 1, the strain is primarily accommo-

435

dated in a several (3-6) troughs that have deepened at the expense of adjacent ridges and

436

troughs. The deepest trough in this case is 500-600 m deep. Unlike the deformation of profile

438

439

AC

437

CE

433

1, these troughs, while shallower, are separated by 10-15 km-wide regions of essentially flat terrain. The simulated surface deformation is unlike Ganymede’s archetypal grooved terrain (Fig. 5D).

440

Figure 12 illustrates (for profile 1) how the final deformation depends on the thermal

441

structure used in the simulation. Cold lithospheres (lower Ts and F ) generally result in 19

ACCEPTED MANUSCRIPT

stronger localization of deformation within a single trough, whereas warmer lithospheres

443

(higher Ts and F ) tend to result in more distributed deformation. As noted in section 3, this is

444

due to the predominance of brittle deformation, which readily leads to localization, in the cold

445

lithosphere. In the simulation with Ts =120 K and F =100 mW m−2 , the isolated troughs

446

have an amplitude of ∼600 m, which is still larger than, but considerably closer to, typical

447

groove amplitudes. The increased periodicity of the deformation for warmer lithospheres

448

results from the general decrease in the amplitude growth rate under warm conditions, which

449

has consistently been observed in numerical models of tectonic deformation on icy worlds

450

(Bland et al., 2010; Bland and McKinnon, 2012, 2013, 2015). In these simulations, the weaker

451

viscosity contrast between the relatively warm near-surface ice and the even warmer ice at

452

depth retards growth of the necking and localization instabilities, including the run-away

453

localization of strain within a few troughs. The result is lower amplitude but more periodic

454

deformation.

AN US

CR IP T

442

As with the simulations that used large-amplitude synthetic topography, the simulations

456

that were initialized with dark terrain topography generally fail to produce groove-like defor-

457

mation under thermal conditions typical of groove formation on Ganymede1 . Instead, strain

458

becomes localized within preexisting topographic lows, resulting in troughs ∼1 km deep that

459

are separated by lower amplitude ridges and troughs. These simulations suggest that forming

460

grooved terrain simply by “tectonically resurfacing” the dark terrain is unlikely. Further-

461

more, the simulations described above intentionally avoided large-scale crater forms, which,

462

as discussed below, provide an even greater challenge to the tectonic resurfacing mechanism.

463

We note, however, that our model has inherent limitations due to their two-dimensional

464

nature. Isolated knobs and depressions within the dark terrain act like ridges and troughs

466

ED

PT

CE

AC

465

M

455

in a two-dimensional profile, and localization may be less efficient when three-dimensional effects are accounted for (see discussion in section 4.6). 1

Dombard and McKinnon (2001) argue that strains up to 50% - 100% may be necessary to form “grooves”

in dark terrain

20

ACCEPTED MANUSCRIPT

467

4.2

Forming grooves from bright smooth terrain

The simulations described above suggest that forming grooved terrain by tectonically

469

resurfacing dark terrain terrain is challenging. An alternative end-member hypothesis is that

470

groove terrain formation occurred in several stages that included cryovolcanic resurfacing.

471

In this scenario, cryovolcanism locally provides a topographically smooth lane upon which

472

the ridges and troughs of the grooved terrain subsequently formed. To directly assess how

473

grooved terrain might have formed from such smooth terrain, we used the topography of

474

bright, relatively smooth sulci on Ganymede as the initial topography for our simulations of

475

lithospheric extension. Six topographic profiles were extracted from Nippur Sulcus and Elam

476

Sulci using the same DTM described in section 4.1 (Fig. 13), and interpolated for use in our

477

simulations. In these profiles, local relief was ∼50 m, and up to 250 m of relief is present at

478

long wavelengths (∼100 km). For our simulations we selected the first 80 km of profile 1, but

479

removed the regional slope using a second order polynomial. Our objective was to use the

480

smoothest surface available, which would be consistent with cryovolcanic resurfacing. This

481

also permits us to clearly delineate these simulations from those that utilized dark terrain

482

(described above). The initial topography used in our simulations is shown in Fig. 14A.

ED

M

AN US

CR IP T

468

Simulations of extension that utilize the “smooth” initial topography described above

484

result in strongly periodic groove-like structures. Figure 14 shows the final surface defor-

485

mation for three simulations that utilized a range of thermal structures (Ts =70-120 K, and

486

F =70-150 mW m−2 ). Each simulation resulted in periodic deformation. For Ts =100 K and

487

F =100 mW m−2 , the deformation has a wavelength of 6 km and an amplitude of 150-200 m

488

(Fig. 14B). Perhaps not surprisingly, the deformation is consistent with that resulting from

490

CE

AC

489

PT

483

low amplitude synthetic initial topography under identical surface temperature and heat flux (Fig. 6). Decreasing the surface temperature and heat flux increases the wavelength and

491

amplitude of the final deformation, but the surface remains periodic (λ ≈18 km for Ts =70 K,

492

and F =70 mW m−2 , Fig. 14C). Conversely, increasing surface temperature and heat flux

493

decreases the wavelength and amplitude of the deformation (λ ≈3 km for Ts =120 K, and

494

F =150 mW m−2 , Fig. 14D). The relationship between heat flux, surface temperature, and 21

ACCEPTED MANUSCRIPT

495

deformation morphology is consistent with previous modeling results (Bland et al., 2010;

496

Bland and McKinnon, 2015). The relative “ease” with which groove-like deformation can be produced from initially

498

smooth terrain compared with the difficulty of tectonically resurfacing dark terrain suggests

499

that cryovolcanic resurfacing may have played an important if not critical role in the forma-

500

tion of the grooved terrain. These simulations generally support the post-Voyager hypothesis

501

that bright grooved terrain formed through a multi-stage process including both volcanism

502

and tectonism a` la Golombek and Allison (1981) and Allison and Clifford (1987). The cry-

503

ovolcanic resurfacing hypothesis certainly has problems (as outlined above), but forming

504

broadly distributed extensional deformation from the relatively smooth canvas it provides

505

is more obviously feasible than directly modifying the large-scale topography of the dark

506

terrain.

507

4.3

AN US

CR IP T

497

M

Tectonic resurfacing of impact craters

In the dark-terrain simulations described above (section 4.1) we intentionally utilized

509

topographic profiles that did not include large impact craters. Impact craters provide one

510

of the greatest challenges to the tectonic resurfacing mechanism. Whereas elongated craters

511

have been documented within Ganymede’s dark terrain (Pappalardo and Collins, 2005),

512

few (if any) “ghost craters”, in which subtle evidence of crater rims remain but the crater

513

itself does not, have been observed within broad expanses of grooved terrain. This suggests

514

that the resurfacing that produced the bright terrain was extremely efficient. In contrast,

515

physical analog models of tectonic resurfacing fail to erase craters, even while producing

517

PT

CE

AC

516

ED

508

periodic graben-like structures (Wyrick, 2012). To investigate how the extending lithosphere interacts with impact craters, we performed

518

a suite of simulations that utilized impact crater topography derived from our DTM of Marius

519

Regio. The topographic profiles are shown in Fig. 15, along with those of a nearby viscously

520

relaxed crater (see below). The crater is 21 km in diameter, ∼900 m deep, and relatively

521

fresh. We use an 80 km portion of profile c1 (centered on the crater) and interpolate to the 22

ACCEPTED MANUSCRIPT

522

resolution of our finite element mesh. The disruption and erasure of craters in numerical simulations is challenging because their

524

substantial depth creates a locally thin region of the lithosphere that, as discussed in section

525

3, localizes horizontal stresses, brittle failure, and extensional stain. Figure 16 shows the

526

surface and lithospheric deformation for a simulation initialized with the crater topography

527

described above and a relatively cold and thick lithosphere (Ts =70 K, and F =70 mW m−2 ).

528

In this case, extension is more-or-less entirely accommodated within the crater itself, and the

529

crater actually deepens and narrows by a factor of two during extension, reaching a depth

530

of ∼2 km (with respect to the ground plane), and width of ≈10 km. The pervasive plastic

531

deformation within the crater may be sufficient to completely disrupt the crater-form through

532

imbricate faulting; however, the end result is a single deep trough rather than distributed,

533

periodic deformation. Here again, three-dimensional effects may be important (section 4.6).

534

In our two-dimensional simulations the crater effectively acts as a deep trough. In reality, a

535

crater is an isolated depression and strain localization may be modified. However, we note

536

that fractures and faults within Ganymede’s dark terrain (e.g., Nicholson Regio, Fig. 4), and

537

on Enceladus, often develop through the center of craters, suggesting that they do localize

538

brittle deformation in three-dimensions.

ED

M

AN US

CR IP T

523

When the lithosphere is warmer (Ts =100 K, and F =100 mW m−2 ) different behavior

540

results (Fig. 17). In this case the imposed impact crater undergoes rapid viscous relaxation

541

during the initial period of extension, such that by 1% extension most of the impact crater

542

topography is removed (Fig. 17A). We emphasize that flattening of the crater is not due to

543

tectonic resurfacing, but rather primarily to viscous relaxation (cf. Bland et al., 2017; Singer

544

et al., 2017). Despite the reduction in topography, the combination of viscous relaxation and

546

CE

AC

545

PT

539

extension leads to the formation of a weak zone at the location of the former crater (indicated by the colored portions of the lithosphere). As extension of the lithosphere continues, the

547

majority of the strain is accommodated within this zone (Fig. 17B and C). The result is

548

the formation of a relatively isolated, narrow, v-shaped trough nearly 800 m deep, similar to

549

the depth of the original crater (Fig. 17D). Critically, the regions adjacent to the crater are

23

ACCEPTED MANUSCRIPT

essentially undeformed: the initial, non-crater topography present at time zero still remains

551

after ∼5% extension. As in the colder simulation, extensive faulting is likely to disrupt the

552

crater (including its rim). This pervasive faulting is similar to tectonically disrupted, but still

553

visible craters in dark terrain (Pappalardo and Collins, 2005) (Fig. 3D). Whereas the crater

554

has effectively been removed (i.e., the rim and interior are likely disrupted), the resulting

555

topography is not consistent with grooved terrain morphologies.

CR IP T

550

Further increasing the lithospheric temperature yields similar behavior to the warm simu-

557

lation. Figure 18 illustrates this effect for simulations with F =150 mW m−2 and Ts ≥100 K.

558

For the two cases shown, the initially imposed crater is essentially flattened by relaxation

559

after just 1-2% extension. Again the majority of the strain is accommodated within the

560

weak zone associated with the initial crater, and extension results in an isolated, flat-floored

561

graben or v-shaped trough. The topography is generally inconsistent with grooved terrain

562

morphologies.

AN US

556

In the warm lithosphere case, the failure to form groove-like deformation is predominantly

564

the result of the weak zone that remains after viscous relaxation has occurred, rather than

565

the amplitude of the topography. Whereas viscous relaxation can result in plastic failure of

566

the near-surface (Dombard and McKinnon, 2006), the deep failure zone in our model results

567

from the concomitant extensional strain. However, it is plausible that rather than relaxation

568

and extension occurring simultaneously, viscous relaxation preceded extension. Such a two-

569

step scenario might prevent the development of a weak zone beneath the crater as long

570

as thermal steady state can be achieved. Further, micro- and macro-porosity that results

571

from brittle failure during viscous relaxation may thermally sinter or anneal (a process not

572

included here) under the warm conditions of the lithosphere, thus further healing any weak

574

ED

PT

CE

AC

573

M

563

zone that forms. Shoemaker et al. (1982) argued that at temperatures greater than ∼130 K, vapor transport and deposition through the regolith (or “damaged” zones) can re-cement

575

the ice lithosphere. Besserer et al. (2013) showed that viscous compaction is also capable of

576

removing porosity that might result from fracturing.

577

To further assess this scenario, we performed a suite of simulations using topography

24

ACCEPTED MANUSCRIPT

derived from a viscously relaxed crater within Marius Regio (Fig. 15). Like the fresh crater

579

described above, the relaxed crater is 21 km in diameter, with a central rebounded floor

580

that is at the level of the surrounding terrain. Surrounding the relaxed center is a circular

581

moat approximately 200 m deep. The crater retains obvious rims (short wavelength topog-

582

raphy is preserved during viscous relaxation) that stand 160-300 m above the surrounding

583

terrain. Figure 19 shows the post-extension deformation of the surface and lithosphere for

584

the relaxed topography. For intermediate and warm surface temperatures, the surface is

585

deformed into quasi-periodic ridges and troughs. This is especially true for the warmest

586

temperatures (Ts =120 K) in which plastic strain is distributed in numerous trough-like

587

features. The amplification of the surface deformation is quite limited: the total relief of the

588

final deformation is smaller after extension than before extension in the Ts =120 K simula-

589

tion. Further extension of the domain (the simulations shown here have been extended by

590

5%) results in similar surface deformation, although strain begins to localize more strongly

591

within individual troughs in the Ts =100 K simulation. These simulations illustrate that,

592

unsurprisingly, when topography is reduced by viscous relaxation before extension occurs,

593

subsequent groove terrain formation is more easily achieved. Tectonic resurfacing of impact

594

craters is likely enabled by concomitant viscous relaxation as the lithosphere begin to warm.

595

4.3.1

ED

M

AN US

CR IP T

578

PT

Thermal effects and deep crater topography

Because Tekton does not include thermodynamics, the simulations described in section

597

4.3 must be interpreted carefully. The simulations shown here use a laterally constant sur-

598

face temperature. Thus, beneath the crater the distance between the constant-temperature

599

surface and a “flat” isotherm at depth is reduced compared to adjacent regions. The litho-

601

AC

600

CE

596

spheric thermal gradient is therefore greater beneath the crater than in the surrounding region. This effect is negligible for small amplitude topography, but beneath a ∼1 km-deep

602

crater isotherms in our simulations are compressed by 20-50% (depending on the imposed

603

heat flux). Qualitatively, the increase is realistic (Seiferlin, 2009), but because we do not

604

actually solve the thermodynamics, it is only an approximation. 25

ACCEPTED MANUSCRIPT

Thermal effects are exacerbated as the simulation proceeds and extensional necking oc-

606

curs. Because isotherms (technically iso-viscous contours) are advected with the deforming

607

material, warm, low-viscosity material rises into the necked regions beneath the crater (e.g.,

608

Fig. 16). As long as the lithosphere remains several kilometers thick, the thermal diffusion

609

timescale is much longer than the simulation timescale and the advection of warm material

610

is realistic (i.e., the material would remain warm over the timescale simulated). For a litho-

611

spheric thickness of 1 km, however, the timescales for thermal diffusion and extension are

612

similar. In those cases (the simulation in Fig. 17 approaches this2 ) the material rising into

613

the necked region should cool, increase in viscosity, and act to stabilize the necked region.

614

Neglecting these thermal effects enhances the rate at which strain is localized during exten-

615

sion. This effect should be most significant when heat fluxes are large and the lithosphere

616

becomes very thin beneath the crater.

AN US

CR IP T

605

To further investigate these effects, we performed a suite of simulations that initialized

618

the temperature structure such that the heat flow was truly uniform (i.e., isotherms fol-

619

lowed the surface topography). In these simulations strain localization is, in fact, retarded,

620

and while the majority of the strain is accommodated within the crater, small amounts of

621

extensional strain is also partitioned quasi-periodically throughout the lithosphere. This pe-

622

riodic strain partitioning is insufficient to modify the surface, however, and the final surface

623

deformation, which is marked by a deep, isolated trough rather than periodic deformation,

624

is consistent with our nominal simulations. The primary difference between the results of

625

these simulations and our nominal simulations is the magnitude of the strain that is accom-

626

modated. For example, for conditions analogous to those shown in Fig. 17, a similar state

627

is reached in these simulations after 8-9% extension rather than 5% extension. Despite the

629

630

ED

PT

CE

AC

628

M

617

similarity in surface deformation, it is notable that the boundary between the ice lithosphere and asthenosphere is much flatter than in our nominal cases. These simulations indicate that our basic results are qualitatively insensitive to thermodynamic effects, at least for the 2

Viscous relaxation in Fig. 17 is faster than thermal adjustment to steady state: 1% extension occurs in

3000 yrs, whereas the thermal diffusion time for the 2-km thick lithosphere is about twice as long.

26

ACCEPTED MANUSCRIPT

631

lithospheric thicknesses we have modeled.

632

4.4

Effect of ice shell properties

The simulations discussed above have explored a range of physical parameters, focused

634

primarily on the heat flux and surface temperature. In addition, the properties of the ice

635

shell are likely critical to understanding the formation of Ganymede’s grooved terrain. The

636

simulations described here assume a cohesion of 1 MPa, consistent with laboratory data

637

(Beeman et al., 1988). However, it’s possible that at geologic scales the effective strength of

638

an ice lithosphere is much lower. The formation of cycloids on Europa, for example, seem

639

to require a relatively weak ice shell (Rhoden and Hurford, 2013), and models of mobile lid

640

convection in an ice shell generally must assume a low yield stress (Barr, 2008; Han et al.,

641

2012). Former numerical models of groove formation had to assume a high cohesion (high

642

strength contrast with deeper warm ice) in order to produce significant amplitude growth

643

(Bland et al., 2007, 2010); however the introduction of more realistic non-associated plasticity

644

in recent models has removed this requirement (Bland and McKinnon, 2015).

M

AN US

CR IP T

633

To assess how the assumed strength of the lithosphere affects groove formation in the

646

presence of large-amplitude initial topography, we repeated all of the simulations described

647

above using a cohesion of 100 kPa (one order of magnitude lower than our nominal model).

648

Figure 20 compares these low cohesion simulations to those that used our nominal value of

649

1 MPa for the case where dark terrain is used as the initial topography. In general, the

650

simulations that used a lower cohesion result in lower growth of the necking (or localizing)

651

instability. This also results in a reduction in the localization of strain within individual

653

PT

CE

AC

652

ED

645

troughs. Simulations with cold and intermediate lithospheric temperatures (Ts =70 K and 100 K) still result in deep troughs that dominate the deformation. In contrast, the simula-

654

tion with a warm lithosphere (Ts =120 K) results in semi-periodic structures of relatively

655

consistent amplitude. This deformation is consistent with Ganymede’s grooves.

656

Critically, the groove-like deformation in this simulation (Fig. 20C) does not result from

657

substantial changes in the amplitude of the initial topography. With a warm surface temper27

ACCEPTED MANUSCRIPT

658

ature and weak brittle ice shell, instability growth is limited (the strength contrast between

659

the lithosphere and deeper ice is reduced). Growth rates are large enough for a “dominant”

660

wavelength to develop, but insufficient to produce large-amplitude structures. These simula-

661

tions suggest that a weaker ice shell might assist tectonic resurfacing of Ganymede, although

662

the cohesion assumed is relatively unimportant for the cold polar regions. Reducing the cohesion also does not have an appreciable effect on the tectonic resurfacing

664

of relatively fresh, deep craters, and the qualitative results described in section 4.3 hold

665

whether a cohesion of 1 MPa or 100 kPa is used. In these cases, the depth of the trough

666

that results from extension is reduced when a lower cohesion is used, but the difference is

667

not substantial enough to allow direct resurfacing. Even when high heat fluxes are used,

668

reducing the cohesion of the lithosphere does not enable periodic structures to form from

669

preexisting cratered terrain.

AN US

CR IP T

663

In summary, assuming a low cohesion for Ganymede’s ice lithosphere does not further en-

671

able widespread tectonic resurfacing on Ganymede. However, a lower cohesion may prevent

672

strain from strongly localizing within a narrow trough. It therefore may play an impor-

673

tant role in resurfacing terrain that has already been “smoothed” by viscous relaxation or

674

cryovolcanism.

675

4.5

ED

M

670

PT

The role of small-scale faulting

The tectonic resurfacing mechanism depends critically on the development of numerous

677

imbricated faults and subsequent rotation of fault blocks (Head et al., 1997). The block

678

rotation exposes brighter subsurface ice, and contributes to the complete erasure of preex-

680

AC

679

CE

676

isting structures like crater rims. Our numerical simulations result in strong localization of plastic strain into narrow, linear, fault-like structures (see Bland and McKinnon, 2015,

681

for details). The width and spacing of these faults is limited, however, by the resolution

682

of our simulations, which in the examples shown here is 167 m. Any discrete, smaller-scale

683

faulting cannot be captured by our simulations. Despite this, fault-like slip along planes of

684

elements does occur, only limited by the resolution of the mesh (see discussion in Bland and 28

ACCEPTED MANUSCRIPT

685

McKinnon (2015). A full “test” of the tectonic resurfacing mechanism should include a more realistic im-

687

plementation of true fault-like behavior, perhaps by combining increased resolution with a

688

strain- or strain-rate-dependent reduction in brittle strength (e.g., Buck et al., 2005; Bland

689

et al., 2010). These concerns are somewhat mitigated, however, by physical analog mod-

690

els of the extension of ice lithospheres. As described above, these models provide much

691

insight into fault processes on Ganymede, and reproduce many of the fault geometries (in-

692

cluding rotation) observed (Sims et al., 2014). Despite the development of these geometries,

693

faulting and block rotation apparently does not erase preexisting impact crater topography

694

(Wyrick, 2012). These models are therefore in broad agreement with our numerical simula-

695

tions. The analog models, however, do not include viscous deformation, and it may be that

696

a combination of viscous flow that reduces long-wavelength topography (crater bowls), and

697

small-scale faulting and fracturing that removes short-wavelength topography (e.g., crater

698

rims), is necessary to completely resurface preexisting terrain. We therefore conclude that

699

our simulations may underestimate the degree to which short wavelength topography can

700

be disrupted by fracturing and faulting, but they provide a relatively realistic depiction of

701

lithospheric behavior at longer wavelengths.

702

4.6

ED

M

AN US

CR IP T

686

PT

Localization in three-dimensions

In our two-dimensional simulations, strain localizes relatively easily within topographic

704

lows or thin regions of the lithosphere. These are regions of high stress and thus increased

705

plastic (brittle) strain, which readily localizes via non-associated plasticity. As described

707

AC

706

CE

703

above, this localization often results in isolated troughs rather than widespread deformation, and thus inhibits tectonic resurfacing. Strain localization in three-dimensions is likely more

708

complex. For example, the two-dimensional cross section of an impact crater is essentially

709

a trough, which easily localizes strain. In reality, a crater is an isolated depression, and

710

although it may be a locally thin region of the lithosphere, the regions adjacent to the crater

711

are not. Thus, whereas the crater may be a locus of brittle failure (see, e.g., Fig. 4), fail29

ACCEPTED MANUSCRIPT

ure must propagate laterally to form a continuous trough. Normal faults generally form by

713

linkage of smaller-scale fractures and faults (e.g., Peacock and Sanderson, 1991; Anders and

714

Schlische, 1994; Dawers and Anders, 1995; Mansfield and Cartwright, 2001; Wyrick et al.,

715

2011; Cartwright et al., 2016). If fault linkage is inhibited, periodic necking (a viscous insta-

716

bility) may dominate and prevent strain from localizing within a single crater. It is therefore

717

plausible that such strain localization is less efficient in three dimensions, more readily per-

718

mitting the development of periodic structures even when the preexisting topography is large.

719

How strain localizes in three-dimensions to produce grooves that are continuous along strike

720

for hundred of kilometers has not yet been addressed, and a complete understanding of the

721

resurfacing of Ganymede will not be achieved until the three-dimensional nature of groove

722

terrain formation is understood. Fully three-dimensional simulations (Bland and Wyrick,

723

2017) and more robust analog modeling (Wyrick et al., 2017) are warranted.

724

5

AN US

CR IP T

712

M

A new scenario for resurfacing Ganymede Collectively, the simulations described in section 4 suggest several modifications may be

726

necessary to the prevailing picture of resurfacing on Ganymede. Our simulations with large-

727

amplitude initial topography, whether synthetically produced or derived from Ganymede’s

728

dark terrain, fail to produce distributed, periodic surface deformation. This is particularly

729

true when the preexisting surface includes impact craters hundreds of meters or kilometers

730

deep. Creating archetypical grooved terrain directly from Ganymede’s dark terrain therefore

731

appears challenging. Forming periodic ridges and troughs is much easier if the preexisting to-

732

pography is low-amplitude, either due to cryovolcanic resurfacing or viscous relaxation. This

734

PT

CE

AC

733

ED

725

does not imply, however, that tectonic imbrication, block rotation, and pervasive fracturing does not play a critical role in disrupting older bright terrains. There is good observational

735

evidence that such processes occur (Head et al., 1997; Prockter et al., 2000; Pappalardo and

736

Collins, 2005, and see Fig. 3). Rather, we suggest that tectonic resurfacing is a critical

737

resurfacing process on Ganymede, but is likely enabled by viscous relaxation either before or

30

ACCEPTED MANUSCRIPT

738

during extension. As noted above, the heat fluxes commonly associated with the formation

739

of grooved terrain will necessarily result in viscous relaxation of long-wavelength topography

740

(e.g., craters). Based on these results, we propose that Ganymede’s resurfacing has occurred in several

742

stages. Models of grooved terrain formation (Dombard and McKinnon, 2001; Bland and

743

Showman, 2007), viscous relaxation of impact craters (Bland et al., 2017), and Ganymede’s

744

interior evolution (Showman et al., 1997; Bland et al., 2009) all indicate that the global expan-

745

sion associated with resurfacing was accompanied by substantial warming of the lithosphere.

746

This warming may have been associated with incipient mobile lid convection (Hammond and

747

Barr, 2014), resulting in a non-uniform distribution of heat flux and surface strain. That is,

748

regions of high heat flux and high strain are correlated and localized (Hammond and Barr,

749

2014). Where heat fluxes were highest, long wavelength surface topography – principally

750

impact craters – likely underwent rapid viscous relaxation, reducing the amplitude of the

751

pre-existing topography. The subsequent onset of extensional strain within these regions

752

of high-heat flow resulted in the formation of the periodic, distributed deformation associ-

753

ated with the grooved terrain. Where heat flow was especially highly localized, complete

754

separation of the thin lithosphere might have occurred, resulting in band formation (e.g.,

755

Head et al., 2002). Differences in preexisting dark-terrain topography (e.g., predominant

756

topographic wavelengths – affecting viscous relaxation rates) may have also contributed to

757

differences in the style of dark-terrain breakup.

PT

ED

M

AN US

CR IP T

741

Away from the regions of localized extension and high heat fluxes, the dark terrain

759

would remain intact. These regions might correspond to convective down-wellings in the

760

mobile-lid model of Hammond and Barr (2014). Between these more quiescent regions and

762

AC

761

CE

758

the highly deforming regions (e.g., around the edges of a convective plume) transitional geologic units would be expected to develop with a variety of morphologies. Regions that

763

experienced relatively high strains, but comparatively low heat fluxes might develop heavily

764

tectonized but marginally intact impact craters, such as those shown in Fig. 3B and D,

765

and the simulation shown in Fig. 16. Unlike the now-gone craters that might have existed

31

ACCEPTED MANUSCRIPT

before bright terrain formation, these impact craters were preserved (though disrupted)

767

because heat fluxes were insufficient (or too short lived) to permit viscous relaxation before

768

or during extension. Extensional strain localized within these craters, rather than producing

769

the distributed strain patterns (and surface brighting) associated with the grooved terrain.

770

Where only moderate surface strain occurred, tectonism might have proceeded without fully

771

disrupting the dark surface lag deposits to produce surface brightening, resulting in the

772

formation of dark lineated terrains. In contrast, where surface strains were low but heat

773

fluxes were elevated, widespread viscous relaxation would be expected without tectonism.

774

Such widespread viscous relaxation is observed in regions such as northern Marius Regio

775

(Singer et al., 2017), which is situated adjacent to expansive regions of grooved and smooth

776

bright terrain.

AN US

CR IP T

766

The scenario described above depends only upon regional differences in heat flow and

778

surface strain. Mobile lid convection, as suggested by Hammond and Barr (2014), is one

779

mechanism to achieve such differences, but is not explicitly required. Mobile lid convection

780

implies large surface strains: extensional in the bright terrain and contractional in the dark

781

terrain. The former are not required by our simulations of groove formation, and little, if

782

any, evidence is observed for the latter. We do not dismiss the possibility of mobile lid

783

convection, but the mechanism, and its implications for Ganymede’s resurfacing, require

784

further investigation (e.g., with convection models that include plastic rheology).

PT

ED

M

777

The above discussion does not preclude a significant role for cryovolcanism during resur-

786

facing. The high heat flow and extension would have been conducive to the production of

787

melt, which could have been locally driven to the surface by exsolved gases (Crawford and

788

Stevenson, 1988) or topographic gradients (Showman et al., 2004). This is particularly true

790

AC

789

CE

785

if Ganymede’s ice shell was thin during the epoch of resurfacing (Bland et al., 2009). Cryovolcanism could have occurred at any point during the resurfacing process. If (or where)

791

it proceeded extension, it would have provided a bright, “clean canvas” for the formation of

792

periodic ridges and troughs (as demonstrated by the simulations shown in Fig. 14). This

793

might have been the case for regions such as Bubastis Sulcus, where the swaths of grooved

32

ACCEPTED MANUSCRIPT

terrain are broad, the ridges and troughs are both high-amplitude and strongly periodic,

795

evidence of embayment by smooth material is observed (Allison and Clifford, 1987), and the

796

cold temperatures make resurfacing by tectonism alone difficult even if viscous relaxation

797

occurred. In these regions, the tectonism that followed cryovolcanic resurfacing may have

798

erased most evidence of vents. In contrast, if (or where) cryovolcanism occurred either after

799

extension was complete, or where surface strain was limited, smooth bright terrain would

800

result. There, resurfacing would have occurred predominately by cryovolcanism alone, ei-

801

ther by covering the recently formed grooved swath in the former case, or low-lying dark

802

terrain in the latter. In these regions, we expect vent structures to be preserved, such as the

803

caldera-like features in the Mumma Sulci and elsewhere (Schenk et al., 2001).

AN US

CR IP T

794

This discussion also does not preclude a significant role for band formation as an agent

805

of resurfacing. Although not a focus of this paper, bands on Ganymede resemble pull-apart

806

bands on Europa (Prockter et al., 2002; Prockter and Patterson, 2009), and formation of

807

relatively smooth terrain during rifting and separation of preexisting dark and grooved ter-

808

rain blocks (e.g., Fig. 4) would clearly provide the subdued topography that would allow

809

full development of ridges and troughs (i.e., grooves) via extensional instability (Sec. 4.2).

810

Bands form with subdued tectonic or cryovolcanic fabrics along strike to begin with, but

811

high-amplitude grooves (Tiamat Sulcus, Fig. 1C, may be an example) as well as complex

812

grooved domain patterns (e.g., Uruk Sulcus, (Golombek and Allison, 1981)) require addi-

813

tional extension.

PT

ED

M

804

In summary, resurfacing on Ganymede was likely the result of a complex interplay be-

815

tween tectonism, viscous relaxation, band formation, and cryovolcanism. In broad terms,

816

each terrain type observed on Ganymede may simply reflect which process locally dominated

818

AC

817

CE

814

the evolution, with the dominant process determined by local heat flux and surface strain. Ganymede’s dark terrain clearly experienced the lowest heat fluxes and limited strain. Tec-

819

tonism and evidence for high heat fluxes there date to an early epoch of Ganymede’s history

820

(e.g., Prockter et al., 1998). On the other end of the spectrum, the bright grooved ter-

821

rain likely experienced heat fluxes high enough to viscously relax preexisting impact craters,

33

ACCEPTED MANUSCRIPT

enabling concomitant or subsequent extension to result in broadly distributed lithospheric

823

deformation and tectonic resurfacing. Between these extremes, the relative amount of strain

824

and heating resulted in dark lineated terrain, localized tectonic disruption of craters, and

825

viscous relaxation without tectonism. Throughout the process, cryovolcanism both enabled

826

groove terrain formation, and resurfaced both dark terrain and newly formed grooved swaths.

827

The end result is the complex patchwork of bright and dark, tectonized and smooth, old and

828

young terrains observed on Ganymede today. Despite the new insight provided by these sim-

829

ulations, the relative importance of tectonic resurfacing, cryovolcanism, and band formation

830

in resurfacing Ganymede will remain an open question until new data, including imaging,

831

altimetry, and radar sounding are acquired (e.g., by ESA’s JUICE mission).

832

Acknowledgments

AN US

CR IP T

822

This work was supported by NASA’s Outer Planets research program (NNH14AY69I to

834

MTB) and NASA’s Planetary Geology and Geophysics program (NNX11AP16G to WBM).

835

DTMs were generously provided by Paul Schenk. Comments from Chris Okubo, Francis

836

Nimmo, Geoff Collins, and an anonymous reviewer strengthened and clarified this work.

837

Any use of trade, firm, or product names is for descriptive purposes only and does not imply

838

endorsement by the U.S. Government.

AC

CE

PT

ED

M

833

34

ACCEPTED MANUSCRIPT

839

References

840

Allison, M. L., Clifford, S. M., 1987. Ice-covered water volcanism on Ganymede. J. Geophys.

841

Res. 92, 7865–7876. Allken, V., Huismans, R. S., Thieulot, C., 2011. Three-dimensional numerical modeling of up-

843

per crustal extensional systems. J. Geophys. Res. 116, B10409, doi:10.1029/2011JB008319.

844

Allken, V., Huismans, R. S., Thieulot, C., 2012. Factors controlling the mode of rift in-

845

teraction in brittle-ductile coupled systems: A 3D numerical study. Geochem. Geophys.

846

Geosys. 13, Q05010, doi:10.1029/2012GC004077.

850

851

852

AN US

849

of normal faults. J. Geo. 102, 165–179, doi:10.1086/629661.

Barr, A. C., 2008. Mobile lid convection beneath Enceladus’ south polar terrain. J. Geophys. Res. 113, 7009, doi:10.1029/2008JE003114.

M

848

Anders, M. H., Schlische, R. W., 1994. Overlapping faults, intrabasin highs, and the growth

Bassi, G., 1991. Factors controlling the style of continental rifting: insights from numerical modelling. Earth Planet. Sci. Lett. 105, 430–452, doi:10.1016/0012-821X(91)90183-I.

ED

847

CR IP T

842

Beaumont, C., Kamp, P. J. J., Hamilton, J., Fullsack, P., 1996. The continental collision

854

zone, South Island, New Zealand: Comparison of geodynamical models and observations.

855

J. Geophys. Res. 101, 3333–3360, doi:10.1029/95JB02401.

CE

Beeman, M., Durham, W. B., Kirby, S. H., 1988. Friction of ice. J. Geophys. Res. 93,

857

7625–7633.

858

AC

856

PT

853

859

860

Besserer, J., Nimmo, F., Roberts, J. H., Pappalardo, R. T., 2013. Convection-driven compaction as a possible origin of Enceladus’s long wavelength topography. J. Geophys. Res. 118, 908–915, doi:10.1002/jgre.20079.

35

ACCEPTED MANUSCRIPT

861

Bierhaus, E. B., Schenk, P. M., 2010. Constraints on Europa’s surface properties

862

from primary and secondary crater morphology. J. Geophys. Res. 115, E12004,

863

doi:10.1029/2009JE003451.

865

866

Bland, M. T., Beyer, R. A., Showman, A. P., 2007. Unstable extension of Enceladus’ lithosphere. Icarus 192, 92–105, doi:10.1016/j.icarus.2007.06.011.

CR IP T

864

Bland, M. T., McKinnon, W. B., 2012. Forming Europa’s folds:

Strain require-

867

ments for the production of large-amplitude deformation. Icarus 221, 694–709, doi:

868

10.1016/j.icarus.2012.08.029.

Bland, M. T., McKinnon, W. B., 2013. Does folding accommodate Europa’s contractional

870

strain? The effect of surface temperature on fold formation in ice lithospheres. Geophys.

871

Res. Lett. 40, 2534–2538, doi:10.1002/grl.50506.

AN US

869

Bland, M. T., McKinnon, W. B., 2015. Forming Ganymede’s grooves at smaller strain:

873

Toward a self-consistent local and global strain history for Ganymede. Icarus 245, 247–

874

262, doi:10.1016/j.icarus.2014.09.008.

ED

M

872

Bland, M. T., McKinnon, W. B., Showman, A. P., 2010. The effects of strain lo-

876

calization on the formation of Ganymede’s grooved terrain. Icarus 210, 396–410,

877

doi:10.1016/j.icarus.2010.06.008. Bland, M. T., Showman, A. P., 2007. The formation of Ganymede’s grooved ter-

CE

878

PT

875

rain:

880

doi:10.1016/j.icarus.2007.01.012.

881

882

Numerical modeling of extensional necking instabilities. Icarus 189, 439–456,

AC

879

Bland, M. T., Showman, A. P., Tobie, G., 2009. The orbital-thermal evolution and global expansion of Ganymede. Icarus 200, 207–221, doi:10.1016/j.icarus.2008.11.016.

883

Bland, M. T., Singer, K. N., McKinnon, W. B., Schenk, P. M., 2017. Viscous relax-

884

ation of Ganymede’s impact craters: Constraints on heat flux. Icarus 296, 275–288,

885

doi:10.1016/j.icarus.2017.06.012. 36

ACCEPTED MANUSCRIPT

886

Bland, M. T., Wyrick, D. Y., 2017. Simulating the formation of Ganymede’s grooved terrain

887

in three dimensions: Numerical approach and preliminary results. Lunar Planet. Sci. Conf.

888

48 , #2461.

890

891

892

Buck, W. R., Lavier, L. L., Poliakov, A. N. B., 1999. How to make a rift wide. Phil. Trans. R. Soc. London, Ser. A 357, 671–693.

CR IP T

889

Buck, W. R., Lavier, L. L., Poliakov, A. N. B., 2005. Modes of faulting at mid-ocean ridges. Nature 434, 719–723, doi:10.1038/nature03358.

Cartwright, J. A., Mansfield, C., Trudgill, B., 2016. The growth of normal faults by segment

894

linkage. In: Buchanan, P. G., Nieuwland, D. A. (Eds.), Modern developments in struc-

895

tural interpretation, validation, and modeling. GSA Special Publication No. 99. Geological

896

Society of America.

AN US

893

Collins, G. C., Head, J. W., Pappalardo, R. T., 1998. The role of extensional instability in

898

creating Ganymede grooved terrain: Insights from Galileo high-resolution stereo imaging.

899

Geophys. Res. Lett. 25, 233–236, doi:10.1029/97GL03772.

ED

901

Crawford, G. D., Stevenson, D. J., 1988. Gas-driven water volcanism in the resurfacing of Europa. Icarus 73, 66–79, doi:10.1016/0019-1035(88)90085-1.

PT

900

M

897

Crow-Willard, E. N., Pappalardo, R. T., 2015. Structural mapping of Enceladus

903

and implications for formation of tectonized regions. J. Geophys. Res. 120, 1–23,

904

doi:10.1002/2015JE004818.

906

Dawers, N. H., Anders, M. H., 1995. Displacement-length scaling and fault linkage. J. Struct.

AC

905

CE

902

Geo. 17, 607–614, doi:10.1016/0191-8141(94)00091-D.

907

Delescluse, M., Mont´esi, L. G. J., Chamot-Rooke, N., 2008. Fault reactivation

908

and selective abandonment in the oceanic lithosphere. Geophys. Res. Lett. 35,

909

doi:10.1029/2008GL035066.

37

ACCEPTED MANUSCRIPT

910

Dombard, A. J., McKinnon, W. B., 2001. Formation of grooved terrain on Ganymede:

911

Extensional instability mediated by cold, superplastic creep. Icarus 154, 321–336,

912

doi:10.1006/icar.2001.6728. Dombard, A. J., McKinnon, W. B., 2006. Elastoviscoplastic relaxation of impact crater

914

topography with application to Ganymede and Callisto. J. Geophys. Res. 111, E01001,

915

10.1029/2005JE002445.

CR IP T

913

Dunbar, J. A., Sawyer, D. S., Dec. 1996. Three-dimensional dynamical model of continen-

917

tal rift propagation and margin plateau formation. J. Geophys. Res. 101, 27845–27863,

918

doi:10.1029/96JB01231.

919

AN US

916

Durham, W. B., Stern, L. A., 2001. Rheological properties of water ice:

Applica-

920

tions to satellites of the outer planets. Ann. Rev. Earth Planet. Sci. 29, 295–330,

921

doi:10.1146/annurev.earth.29.1.295.

924

925

M

for basin-and-range structure. J. Geophys. Res. 88, 7457–7466.

ED

923

Fletcher, R. C., Hallet, B., 1983. Unstable extension of the lithosphere: A mechanical model

Gammon, P. H., Klefte, H., Clouter, M. J., 1983. Elastic constants of ice samples by brillouin spectroscopy. J. Phys. Chem. 87, 4025–4029.

PT

922

Gerbault, M., Burov, E. B., Poliakov, A. N. B., Daigni`eres, M., 1999. Do faults trigger

927

folding in the lithosphere? Geophys. Res. Lett.l 26, 271–274, doi:10.1029/1998GL900293.

928

Goldsby, D. L., Kohlstedt, D. L., 2001. Superplastic deformation of ice: Experimental ob-

930

931

932

933

servations. J. Geophys. Res. 106, 11017–11030, doi:10.1029/2000JB900336.

AC

929

CE

926

Golombek, M. P., Allison, M. L., 1981. Sequential development of grooved terrain and polygons on Ganymede. Geophys. Res. Lett. 8, 1139–1142. Govers, R., Wortel, M. J. R., Aug. 1995. Extension of stable continental lithosphere and the initiation of lithospheric scale faults. Tectonics , 1041–1055Doi:10.1029/95TC00500.

38

ACCEPTED MANUSCRIPT

934

935

Grimm, R. E., Squyres, S. W., 1985. Spectral analysis of groove spacing on Ganymede. J. Geophys. Res. 90, 2013–2021. Gueydan, F., Pr´ecigout, J., 2014. Modes of continental rifting as a function of

937

ductile strain localization in the lithospheric mantle. Tectonophys. 612, 18–25,

938

doi:10.1016/j.tecto.2013.11.029.

940

941

942

Hammond, N. P., Barr, A. C., 2014. Formation of Ganymede’s grooved terrain through convection-driven resurfacing. Icarus 227, 206–209, doi:10.1016/j.icarus.2013.08.024. Han, L., Tobie, G., Showman, A. P., 2012. The impact of a weak south pole on thermal convection in Enceladus’ ice shell. Icarus 218, 320–330, doi:10.1016/j.icarus.2011.12.006.

AN US

939

CR IP T

936

Head, J., Pappalardo, R., Collins, G., Belton, M. J. S., Giese, B., Wagner, R.,

944

Breneman, H., Spaun, N., Nixon, B., Neukum, G., Moore, J., 2002. Evidence for

945

Europa-like tectonic resurfacing styles on Ganymede. Geophys. Res. Lett. 29 (24), 4–1,

946

doi:10.1029/2002GL015961.

M

943

Head, J. W., Pappalardo, R. T., Collins, G., Greeley, R., 1997. Tectonic resurfacing on

948

Ganymede and its role in the formation of grooved terrain. Lunar Planet. Sci. XXVIII ,

949

535–536.

PT

951

Herrick, D. L., Stevenson, D. J., 1990. Extensional and compressional instabilities in icy satellite lithospheres. Icarus 85, 191–204, doi:10.1016/0019-1035(90)90110-U.

CE

950

ED

947

Huismans, R. S., Beaumont, C., 2002. Asymmetric lithospheric extension: The role of fric-

953

tional plastic strain softening inferred from numerical experiements. Geology 30, 211–214.

954

955

956

957

958

AC

952

Huismans, R. S., Beaumont, C., 2003. Symmetric and asymmetric lithospheric extension: Relative effects of frictional-plastic and viscous strain softening. J. Geophys. Res. 108, doi:10.1029/2002JB002026. Iwashita, K., Oda, M., 1999. The mechanics of granular material: An introduction. CRC Press, Rotterdam, Netherlands. 39

ACCEPTED MANUSCRIPT

959

Klimczak, C., Byrne, P. K., Solomon, S. C., 2015. A rock-mechanical assessment of Mercury’s

960

global tectonic fabric. Earth Planet. Sci. Lett. 416, 82–90, doi:10.1016/j.epsl.2015.02.003.

961

Lade, P. V., 2002. Characterization of the mechanical behavior of sea ice as a frictional

964

965

966

Lavier, L. L., Buck, W. R., Poliakov, A. N. B., 2000. Factors controlling normal fault offset

CR IP T

963

material. J. Geophys. Res. 107, 2377, doi:10.1029/2001JB000497.

in an ideal brittle layer. J. Geophys. Res. 105, 23431, doi:10.1029/2000JB900108.

Mansfield, C., Cartwright, J., 2001. Fault growth by linkage: observations and implications from analogue models. J. Struct. Geo. 23, 745–763, doi:10.1016/S0191-8141(00)00134-6.

AN US

962

967

McKinnon, W. B., Melosh, H. J., 1980. Evolution of planetary lithospheres: Evidence from

968

multiringed structures on Ganymede and Callisto. Icarus 44, 454–471, doi:10.1016/0019-

969

1035(80)90037-8.

972

973

M

971

Melosh, H. J., Raefsky, A., 1980. The dynamical origin of subduction zone topography. Geophys. J. R. Astr. Soc. 60, 333–354.

Mueller, S., McKinnon, W. B., 1988. Three-layered models of Ganymede and Callisto: Com-

ED

970

positions, structures, and aspects of evolution. Icarus 76, 437–464. Murchie, S. L., Head, J. W., Plescia, J. B., 1990. Tectonic and volcanic evolution of dark ter-

975

rain and its implications for the internal structure and evolution of Ganymede. J. Geophys.

976

Res. 95, 10743–10768, doi:10.1029/JB095iB07p10743.

978

979

980

981

CE

Nagel, T. J., Buck, W. R., 2007. Control of rheological stratification on rifting geometry:

AC

977

PT

974

a symmetric model resolving the upper plate paradox. Int. J. Earth Sci. 96, 1047–1057, doi:10.1007/s00531-007-0195-x.

Owen, D. R. J., Hinton, E., 1980. Finite elements in plasticity: Theory and practice. Pineridge Press Limited, Swansea, UK.

40

ACCEPTED MANUSCRIPT

982

983

Pappalardo, R. T., Collins, G. C., 2005. Strained craters on Ganymede. J. Struct. Geo. 27, 827–838. Pappalardo, R. T., Collins, G. C., Head, J. W., Helfenstein, P., McCord, T. B., Moore,

985

J. M., Prockter, L. M., Schenk, P. M., Spencer, J. R., 2004. Geology of Ganymede. In:

986

Bagenall, F., Dowling, T., McKinnon, W. B. (Eds.), Jupiter - The Planet, Satellites and

987

Magnetosphere. Cambridge University Press, Cambridge, UK, pp. 363–396.

CR IP T

984

Pappalardo, R. T., Head, J. W., Collins, G. C., Kirk, R. L., Neukum, G., Oberst, J.,

989

Giese, B., Greeley, R., Chapman, C. R., Helfenstein, P., Moore, J. M., McEwen, A.,

990

Tufts, B. R., Senske, D. A., Breneman, H. H., Klaasen, K., 1998. Grooved terrain

991

on Ganymede: First results from Galileo high-resolution imaging. Icarus 135, 276–302,

992

doi:10.1006/icar.1998.5966.

994

Parmentier, E. M., Squyres, S. W., Head, J. W., Allison, M. L., 1982. The tectonics of Ganymede. Nature 295, 290–293.

M

993

AN US

988

Patel, J. G., Pappalardo, R. T., Head, J. W., Collins, G. C., Hiesinger, H., Sun, J., 1999. To-

996

pographic wavelengths of Ganymede groove lanes from Fourier analysis of Galileo images.

997

J. Geophys. Res. 104, 24057–24074, doi:10.1029/1998JE001021.

PT

ED

995

Patterson, G. W., Collins, G. C., Head, J. W., Pappalardo, R. T., Prockter, L. M., Lucchitta,

999

B. K., Kay, J. P., 2010. Global geologic mapping of Ganymede. Icarus 207, 845–867,

1001

1002

1003

doi:10.1016/j.icarus.2009.11.035. Peacock, D. C. P., Sanderson, D. J., 1991. Displacements, segment linkage and relay ramps

AC

1000

CE

998

in normal fault zones. J. Struct. Geo. 13, 721–733, doi:10.1016/0191-8141(91)90033-F.

Petrenko, V. F., Whitworth, R. W., 1999. Physics of Ice. Oxford University Press, Oxford.

1004

Pizzi, A., Di Domenica, A., Kimatsu, G., Cofano, A., Mitri, G., Bruzzone, L., 2017. Spread-

1005

ing vs. rifting as modes of extensional tectonics on the globally expanded Ganymede.

1006

Icarus 288, 148–159, doi:10.1016/j.icarus.2017.01.034. 41

ACCEPTED MANUSCRIPT

1007

1008

Poliakov, A. N. B., Herrmann, H. J., 1994. Self-organized criticality of plastic shear bands in rocks. Geophys. Res. Lett. 21, 2143–2146, doi:10.1029/94GL02005. Pritchard, M. E., Stevenson, D. J., 2000. Thermal aspects of a lunar origin by giant impact.

1010

In: Canup, R. M., Righter, K. (Eds.), Origin of the Earth and Moon. University of Arizona

1011

Press, pp. 179–196.

1012

1013

CR IP T

1009

Pritchard, R. S., 1988. Mathematical characteristics of sea ice dynamic models. J. Geophys. Res. 93, 15609–15618.

Prockter, L. M., Figueredo, P. H., Pappalardo, R. T., Head, J. W., Collins, G. C., 2000.

1015

Geology and mapping of dark terrain on Ganymede and implications for grooved terrain

1016

formation. J. Geophys. Res. 105, 22519–22540, doi:10.1029/1999JE001179.

AN US

1014

Prockter, L. M., Head, J. W., Pappalardo, R. T., Senske, D. A., Neukum, G., Wagner,

1018

R., Wolf, U., Oberst, J. O., Giese, B., Moore, J. M., Chapman, C. R., Helfenstein, P.,

1019

Greeley, R., Breneman, H. H., Belton, M. J. S., 1998. Dark terrain on Ganymede: Geolog-

1020

ical mapping and interpretation of Galileo regio at high resolution. Icarus 135, 317–344,

1021

doi:10.1006/icar.1998.5981.

ED

M

1017

Prockter, L. M., Head, J. W., Pappalardo, R. T., Sullivan, R. J., Clifton, A. E., Giese, B.,

1023

Wagner, R., Neukum, G., 2002. Morphology of Europan bands at high resolution: A mid-

1024

ocean ridge-type rift mechanism. J. Geophys. Res. 107, 5028, doi:10.1029/2000JE001458.

1025

Prockter, L. M., Patterson, G. W., 2009. Morphology and evolution of Europa’s ridges

1026

and bands. In: Pappalardo, R. T., McKinnon, W. B., Khurana, K. K. (Eds.), Europa.

1028

CE

AC

1027

PT

1022

University of Arizona Press.

Rhoden, A. R., Hurford, T. A., 2013. Lineament azimuths on Europa: Implications for obliq-

1029

uity and non-synchronous rotation. Icarus 226, 841–859, doi:10.1016/j.icarus.2013.06.029.

1030

Rudnicki, J. W., Rice, J. R., 1975. Conditions for the localization of deformation in pressure-

1031

sensative dilatant materials. J. Mech. Phys. Solids 23, 371–394. 42

ACCEPTED MANUSCRIPT

1032

Schenk, P. M., 1989. Crater formation and modification on the icy satellites of Uranus and

1033

Saturn - Depth/diameter and central peak occurrence. J. Geophys. Res. 94, 3813–3832,

1034

doi:10.1029/JB094iB04p03813. Schenk, P. M., Chapman, C. R., Zahnle, K., Moore, J. M., 2004. Ages and interiors: The

1036

cratering record of the Galilean satellites. In: Bagnell, F., Dowling, T., McKinnon, W. B.

1037

(Eds.), Jupiter: The Planet, Satellites and Magnetosphere. Cambridge University Press,

1038

pp. 427–456.

1041

1042

1043

1044

72, 209–234, doi:10.1016/0019-1035(87)90126-6.

AN US

1040

Schenk, P. M., McKinnon, W. B., 1987. Ring geometry on Ganymede and Callisto. Icarus

Schenk, P. M., McKinnon, W. B., Gwynn, D., Moore, J. M., 2001. Flooding of Ganymede’s bright terrain by low-viscosity water-ice lavas. Nature 410, 57–60. Seiferlin, K., 2009. Topography effects on the surface heat flow and subsurface temperature. Euro. Planet. Sci. Cong. , #530.

M

1039

CR IP T

1035

Sharples, W., Moresi, L.-N., Jadamec, M. A., Revote, J., 2015. Styles of rifting and

1046

fault spacing in numerical models of crustal extension. J. Geophys. Res. 120, 4379–4404,

1047

doi:10.1002/2014JB011813.

PT

ED

1045

Shoemaker, E. M., Lucchitta, B. K., Wilhelms, D. E., Plescia, J. B., Squyres, S. W., 1982.

1049

The geology of Ganymede. In: Morrison, D. (Ed.), Satellites of Jupiter. University of

1050

Arizona Press, pp. 435–520.

1052

1053

1054

Showman, A. P., Mosqueira, I., Head, J. W., 2004. On the resurfacing of Ganymede by liquid

AC

1051

CE

1048

water volcanism. Icarus 172, 625–640, doi:10.1016/j.icarus.2004.07.011.

Showman, A. P., Stevenson, D. J., Malhotra, R., 1997. Coupled orbital and thermal evolution of Ganymede. Icarus 129, 367–383, doi:10.1006/icar.1997.5778.

43

ACCEPTED MANUSCRIPT

1055

Sims, D. W., Wyrick, D. Y., Ferrill, D. A., Morris, A. P., Collins, G. C., Pappalardo, R. T.,

1056

Colton, S. L., 2014. Physical models of grooved terrain tectonics on Ganymede. Geophys.

1057

Res. Lett. 41, 3774–3778, doi:10.1002/2014GL060359.

1060

1061

1062

1063

1064

1065

on Ganymede: Regional variations and high heat flow. Icarus , submitted.

CR IP T

1059

Singer, K. N., Bland, M. T., McKinnon, W. B., Schenk, P. M., 2017. Relaxed impact craters

Squyres, S. W., 1980. Volume changes in Ganymede and Callisto and the origin of grooved terrain. Geophys. Res. Lett. 7, 593–596.

Squyres, S. W., 1982. The evolution of tectonic features on Ganymede. Icarus 52, 545–559, doi:10.1016/0019-1035(82)90014-8.

AN US

1058

Stempel, M. M., Barr, A. C., Pappalardo, R. T., 2005. Model constraints on the opening rates of bands on Europa. Icarus 177, 297–304, doi:10.1016/j.icarus.2005.03.025. Van Wijk, J. W., Blackman, D. K., 2005. Dynamics of continental rift propagation: the

1067

end-member modes. Earth Planet. Sci. Lett. 229, 247–258, doi:10.1016/j.epsl.2004.10.039.

1068

Vermeer, P. A., de Borst, R., 1984. Non-associated plasticity for soils, concrete, and rocks.

1071

ED

Whitney, D. L., Teyssier, C., Rey, P., Buck, W. R., 2013. Continental and oceanic core

PT

1070

Heron 29, 1–64.

complexes. GSA Bull 125, 273–298, doi:10.1130/B30754.1.

CE

1069

M

1066

Wijns, C., Weinberg, R., Gessner, K., Moresi, L., 2005. Mode of crustal exten-

1073

sion determined by rheological layering. Earth Planet. Sci. Lett. 236, 120–134,

1074

1075

1076

1077

1078

AC

1072

doi:10.1016/j.epsl.2005.05.030.

Wyrick, D. Y., 2012. Analog models of tectonic resurfacing and impact structures on Ganymede. Geo. Soc. Am. Abstr. 44, 328. Wyrick, D. Y., Bland, M. T., Patterson, R., 2017. Physical analog models of Ganymede’s grooved terrain. Lunar Planet. Sci. Conf. 48, #2345. 44

ACCEPTED MANUSCRIPT

M

AN US

CR IP T

system. Icarus 163, 263–289, doi:10.1016/s0019-1035(03)00048-4.

ED

1082

Zahnle, K., Schenk, P., Levison, H. F., Dones, L., 2003. Cratering rates in the outer solar

PT

1081

on Mars. Icarus 212, 559–567, doi:10.1016/j.icarus.2011.01.011.

CE

1080

Wyrick, D. Y., Morris, A. P., Ferrill, D. A., 2011. Normal fault growth in analog models and

AC

1079

45

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

CE

Figure 1: Several of the basic terrain types identified on Ganymede. A. Dark terrain in Nicholson Regio. B. Dark lineated terrain near Harpagia Sulcus. C. A swath of bright

AC

grooved terrain (Tiamat Sulcus) disrupting the dark terrain of Marius Regio. D. Bright smooth terrains of Nippur Sulcus and Elam Sulci. A and B after Patterson et al. (2010),

which describes these and other terrain types in detail.

46

AC

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 2: Examples of possible cryovolcanic resurfacing on Ganymede. A. Voyager 2 mosaic

of a portion of Bubastis Sulci (near the south pole) showing smooth bright terrain interspersed with archetypal grooved terrain. Portions of the smooth material appears to embay adjacent ridges and troughs (white arrows). B. Portion of Sippar Sulcus showing a lane of

smooth material that embays adjacent reticulate terrain (white arrows). After Schenk et al. (2001), who also showed that the smooth material generally stands lower than adjacent terrain.

47

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

CE

Figure 3: Examples of possible tectonic resurfacing on Ganymede. A. The intersection of bright grooved terrain in Nippur Sulcus and Philus Sulcus, dark terrain in Marius Regio,

AC

and smooth terrain in Elam Sulci. The edges of both the dark and smooth terrain appear to be disrupted by imbricated faulting. An enlarged view of such imbrication is shown in B. C. A disrupted crater in Nicholson Regio (lower left). Whereas the crater is still obvious, its center (including its rim) are completely disrupted by a set of parallel, closely spaced fractures. If the bright scarps are interpreted as normal faults they imply the fracture set is down-dropped by several hundred meters relative to the crater floor. D. A pervasively faulted and fractured crater in Nicholson Regio (17 km in diameter, 9◦ N, 18.5◦ E). Only the bright crater rim is visible (white arrows). The image location is indicated in Fig. 4. C 48 and D are modified from Pappalardo and Collins (2005).

AC

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 4: Galileo image mosaic of Arbela Sulcus. Head et al. (2002) suggest Arbela (the north-south trending lane of smooth material) formed as a Europa-like band, involving complete separation of the lithosphere and emplacement of warm, clean ice from below. The white box indicates the location of Fig. 3D.

49

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 5: Comparison of the initial and final surface deformation from our “successful” model

ED

(A and B, respectively, with Ts =70 K, F =100 mW m−2 , ε˙ =10−13 s−1 , 10% extension), and typical dark terrain and grooved terrain topography (C and D, respectively). The

PT

final simulated topography (B) is similar to Ganymede’s grooved terrain (D), but the initial topography used (A, here the low-amplitude topography is barely visible. See Fig. 6 for an

AC

CE

enlarged view) is inconsistent with (not a good match to) the dark terrain (C).

50

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

AC

Figure 6: The surface deformation (right column) that results from extension of a model ice lithosphere initialized with topography of varying initial amplitude (left column). Each simulation used Ts =100 K, F =100 mW m−2 , a strain rate of 10−13 s−1 , and a cohesion of 1 MPa. The form of the initial topography used is identical in each case, only the total amplitude of the relief is changed. From top to bottom: 1 m, 10 m, 25 m, 50 m, 100 m, 150 m, 200 m. Note the difference in scale between left and right columns.

51

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 7: Power spectrum of the final deformation after 10% extension produced by simu-

CE

lations with initial topographic relief of 1 m (dark blue), 25 m (light blue), 100 m, (green), and 200 m (red) as labeled (simulations are those shown in Fig. 6). The black curve shows

AC

the spectrum of the initial topography, which is also shown in the inset for clarity. Note that spectral resolution decreases with increasing wavelength (power at long wavelengths becomes “smeared” out).

52

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

PT

Figure 8: Illustration of the differences in the development of surface deformation for two identical simulations (Ts =100 K, F =100 mW m−2 ) with different initial topographic relief

CE

(25 m in A and C; 200 m in B and D). A and B show the surface deformation as a function of strain. The top curve is the initial topography and each subsequent curve is

AC

shown at increments of 1% extension. The red curve indicates 10% extension. C and D show the distribution of strain (lower plot) and deformation of the surface (upper plot) at 10% extension (identical to red curves in A and B).

53

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

AC

Figure 9: Comparison of the surface deformation produced by simulations with initial topographic relief of 25 m (left column, initial topography shown in A) and 150 m (right

column, initial topography shown in B). The pairs of simulations used identical parameters (except where noted here, other parameters were identical to those used in Fig. 6): C and

D, Ts =70 K; E and F, Ts =120 K; G and H, Ts =70 K and F =150 mW m−2 ; I and J a strain rate of 10−14 s−1 . All simulations are shown after 10% extension except F, which is shown after 9%. 54

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 10: Digital terrain model (see text for details) of the northern portion of Marius

CE

Regio. The topographic profiles illustrate typical dark terrain, but intentionally avoid deep, fresh craters. Portions of profile 1 and profile 3 were used in the finite element simulations

AC

described here. In the DTM, blue is low and red is high. The total dynamic range is ∼1 km.

55

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 11: The initial topography (top), final topography (middle), and final lithospheric

PT

deformation (bottom) for simulations using Ts = 100 K, and F = 100 mW m−2 , and the

AC

CE

topography shown in profile 1 (A) and profile 3 (B) of Fig. 10.

56

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

CE

Figure 12: The effect of changing the surface temperature and heat flux on the resulting surface deformation for simulations initialized with dark terrain topography (top panel -

AC

profile 1 of Fig. 10). Each curve shows the deformation after 10% extension, and are offset for clarity. Red: Ts = 70 K, F = 100 mW m−2 ; gold: Ts = 100 K, F = 70 mW m−2 ;

teal: Ts = 100 K, F = 100 mW m−2 (identical to the simulation shown in Fig. 11A); blue: Ts = 100 K, F = 150 mW m−2 ; violet: Ts = 120 K, F = 100 mW m−2 .

57

AC

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 13: DTM and topographic profiles of Ganymede’s smooth terrain. The DTM is the same as that shown in Fig. 10, but profiles were extracted from Nippur Sulcus and Elam Sulci. Note the difference in vertical scale between the profiles shown here and those in Fig. 10.

58

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 14: Forming grooves from smooth initial topography. A. The initial topography (a portion of profile 1 in Fig. 13, but with the long wavelength slope removed), B. Ts =100 K,

AC

CE

PT

F =100 mW m−2 . C. Ts =70 K, F =70 mW m−2 . D. Ts =120 K, F =150 mW m−2 .

59

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

AC

Figure 15: DTM (as in Fig. 10) showing the locations where crater topography was extracted for both the fresh and viscously relaxed cases. The simulations shown in Figs. 16, 17, and 18 used a portion of profile c1. The simulations shown in Fig. 19 used a portion of profile r3.

60

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

CE

Figure 16: The initial topography (top), final topography (middle), and final lithospheric deformation (bottom) for a simulations with a relatively cold and thick lithosphere (Ts = 70 K,

AC

and F = 70 mW m−2 ) after 8% extension. Note the change in scale between the top and middle panels. The initial topography was derived from profile c1 in Fig. 15.

61

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 17: The topography (top) and lithospheric deformation (bottom) for a simulation

PT

with a relatively warm lithosphere (Ts = 100 K, and F = 100 mW m−2 ) after 0% (A), 1% (B), 3% (C), and 5% (D) extension. The initial topography was derived from profile c1 in

CE

Fig. 15, and is the same as that in Fig. 16. After 1% extension most of the initial crater topography is removed by viscous relaxation, but a residual zone of weakened lithosphere

AC

remains.

62

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 18: The initial topography (top), final topography (middle), and final lithospheric

CE

deformation (bottom) for simulations with a warm and thin lithosphere. A. Ts = 100 K, and F = 150 mW m−2 after 2% extension. B. As in A but after 8% extension. C. Ts = 120 K,

AC

and F = 150 mW m−2 after 1% extension. D. As in C but after 5% extension. In all cases, the initial topography was derived from profile c1 in Fig. 15 and is the same as that in Fig. 16 and Fig. 17A. Note the rapid removal of crater topography in A and C.

63

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

ED

Figure 19: The initial topography (top), final topography (middle), and final lithospheric deformation (bottom) for simulations initialized with topography derived from a viscously

PT

relaxed crater. A. Ts = 100 K, and F = 100 mW m−2 . B. Ts = 120 K, and F = 100 mW m−2 . Note the change in vertical scale in B. In both cases, the initial topography was derived from

AC

CE

profile r3 in Fig. 15. Simulations are shown after 5% extension.

64

AC

CE

PT

ED

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

Figure 20: Comparison of the final surface deformation after 10% extension for pairs of simulations of dark terrain extension that used a cohesion of 1 MPa (black curves) and 100 kPa (grey curves), but were otherwise identical. A. Ts = 70 K, and F = 150 mW m−2 ;

B. Ts = 100 K, and F = 100 mW m−2 ; C. Ts = 120 K, and F = 100 mW m−2 . Compare with Fig. 11.

65