12.12.2012 Views

Reprint - Earth & Planetary Sciences - University of California, Santa ...

Reprint - Earth & Planetary Sciences - University of California, Santa ...

Reprint - Earth & Planetary Sciences - University of California, Santa ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

1 Saturn after Cassini/Huygens<br />

2<br />

3<br />

Revised Version April 10, 2009<br />

4<br />

Icy Satellites: Geological Evolution and Surface Processes<br />

5<br />

Ralf Jaumann 1,2,* , Roger N. Clark 3 , Francis Nimmo 4 , Amanda R. Hendrix 5 , Bonnie J. Buratti 5 6<br />

,<br />

Tilmann Denk 2 , Jeff M. Moore 6 , Paul M. Schenk 7 , Steve J. Ostro ✝5 and Ralf Srama 8<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

1<br />

DLR, Institute <strong>of</strong> <strong>Planetary</strong> Research. Rutherfordstrasse 2, 12489 Berlin, Germany;<br />

2<br />

Freie Universität Berlin, Institute <strong>of</strong> Geological <strong>Sciences</strong>, Malteserstr. 74-100, 12249 Berlin, Germany;<br />

3<br />

U.S. Geological Survey, Denver Federal Center, Denver CO 80225;<br />

4<br />

Dept. <strong>Earth</strong> & <strong>Planetary</strong> <strong>Sciences</strong>, <strong>University</strong> <strong>of</strong> <strong>California</strong> <strong>Santa</strong> Cruz, <strong>Santa</strong> Cruz, CA 95064, USA;<br />

5<br />

Jet Propulsion Laboratory, <strong>California</strong> Institute <strong>of</strong> Technology, Pasadena, CA 91109, USA;<br />

6<br />

NASA Ames Research Center, MS 245-3 M<strong>of</strong>fett Field, CA 94035-1000, USA;<br />

7<br />

Lunar and <strong>Planetary</strong> Institute Houston TX 77058, USA;<br />

8<br />

Max Planck Institut für Kernphysik, 69117, Heidelberg, Germany.<br />

18 *<br />

Corresponding author (Fax :+493067055 402 ; Email address: ralf.jaumann@dlr.de<br />

19<br />

20<br />

✝<br />

deceased<br />

21<br />

22<br />

Content<br />

Abstract<br />

23 21.1 Introduction<br />

24 21.2 Cassini’s Exploration <strong>of</strong> Saturn’s Icy Satellites<br />

25 21.3. Morphology, Geology and Topography<br />

26 21.4 Composition and Alteration <strong>of</strong> Surface Materials<br />

27 21.5 Constraints on top-meter Structure and Composition by Radar<br />

28 21.6 Geologic Evolution<br />

29<br />

30<br />

21.7 Conclusions<br />

31 Abstract<br />

32 The sizes <strong>of</strong> the Saturnian icy satellites range from ~1500 km in diameter (Rhea) to ~20 km<br />

33 (Calypso), and even smaller 'rocks' <strong>of</strong> only a kilometer in diameter are common in the system. All<br />

34 these bodies exhibit remarkable, unique features and unexpected diversity. In this chapter, we will<br />

35 mostly focus on the 'medium-sized icy objects' Mimas, Tethys, Dione, Rhea, Iapetus, Phoebe and<br />

36 Hyperion, and consider small objects only where appropriate, whereas Titan and Enceladus will be<br />

37 described in separate chapters. Mimas and Tethys show impact craters caused by bodies that were<br />

38 almost large enough to break them apart. Iapetus is unique in the Saturnian system because <strong>of</strong> its<br />

39 extreme global brightness dichotomy. Tectonic activity varies widely - from inactive Mimas<br />

40 through extensional terrains on Rhea and Dione to the current cryovolcanic eruptions on Enceladus<br />

41 – and is not necessarily correlated with predicted tidal stresses. Likely sources <strong>of</strong> stress include<br />

42 impacts, despinning, reorientation and volume changes. Accretion <strong>of</strong> dark material originating from<br />

43<br />

44<br />

outside the Saturnian system may explain the surface contamination that prevails in the whole<br />

satellite system, while coating by Saturn’s E-ring particles brightens the inner satellites.<br />

45 So far, among the surprising Cassini discoveries are the volcanic activity on Enceladus, the sponge-<br />

46 like appearance <strong>of</strong> Hyperion and the equatorial ridge on Iapetus - unique features in the solar<br />

47 system. The bright-ray system on Rhea was caused by a relatively recent medium impact which<br />

48 formed a ~40 km crater at 12°S latitude, 112°W longitude, while the wispy streaks on Dione and<br />

49 Rhea are <strong>of</strong> tectonic origin. Compositional mapping shows that the dark material on Iapetus is<br />

50<br />

51<br />

52<br />

composed <strong>of</strong> organics, CO2 mixed with H2O ice, and metallic iron, and also exhibits possible<br />

signatures <strong>of</strong> ammonia, bound water, H2 or OH-bearing minerals, and a number <strong>of</strong> as-yet<br />

unidentified substances. The spatial pattern, Rayleigh scattering effect, and spectral properties<br />

53 argue that the dark material on Iapetus is only a thin coating on its surface. Radar data indicate that<br />

54 the thickness <strong>of</strong> the dark layers can be no more than a few decimeters; this is also consistent with<br />

55 the discovery <strong>of</strong> small bright-ray and bright-floor craters within the dark terrain. Moreover, several<br />

56 spectral features <strong>of</strong> the dark material match those seen on Phoebe, Iapetus, Hyperion, Dione and<br />

57 Epimetheus as well as in the F-ring and the Cassini Division, implying that throughout the<br />

58 Saturnian system. All dark material appears to have a high content <strong>of</strong> metallic iron and a small<br />

59<br />

content <strong>of</strong> nano-phase hematite. However, the complete composition <strong>of</strong> the dark material is still<br />

1


60 unresolved, and additional laboratory work is required. As previously concluded for Phoebe, the dark<br />

61 material appears to have originated external to the Saturnian system.<br />

62 The icy satellites <strong>of</strong> Saturn <strong>of</strong>fer an unrivalled natural laboratory for understanding the geological<br />

63<br />

64<br />

diversity <strong>of</strong> different-sized icy satellites and their interactions within a complex planetary system.<br />

65 21.1 Introduction<br />

66 Covering a wide range <strong>of</strong> diameters, the satellites <strong>of</strong> Saturn can be subdivided into three major<br />

67 classes: ‘icy rocks,’ with radii < 10 km, 'small satellites' with radii


Siarnaq 17531 895.53 0.2960 46.002 20 2.3 0.0026 20.1R 0.06<br />

76<br />

77 Tab. 1 Physical and orbital characteristics <strong>of</strong> the small (


134 suggested endogenic activities. However, the nature <strong>of</strong> these activities was not well understood. For<br />

135 a more detailed discussion <strong>of</strong> our knowledge before Cassini, see Dougherty et al. (2009) and Orton<br />

136 et al. (2009). In terms <strong>of</strong> geology, the Cassini mission was designed to provide significantly better<br />

137 area coverage, spatial resolution, and spectral range, resulting in numerous new discoveries. A<br />

138 detailed description <strong>of</strong> the Cassini mission can be found in Dougherty et al. (2009) and Seal et al.<br />

139<br />

140<br />

(2009). Maps <strong>of</strong> the Saturnian satellites are included in Roatsch et al. (2009).<br />

141 21.2 Cassini’s Exploration <strong>of</strong> Saturn’s Icy Satellites<br />

142 The Cassini spacecraft is equipped with instruments tailored to investigate the surfaces,<br />

143 environments and interiors <strong>of</strong> icy satellites. The optical remote sensing (ORS) instrument suite<br />

144 includes cameras and spectrometers designed for high spatial and spectral resolution covering<br />

145 wavelengths between 0.06µm and 1000µm. With the imaging subsystem (ISS) (Porco et al., 2004),<br />

146 morphologic, stratigraphic and other geological surface properties can be observed at spatial<br />

147 resolutions down to a few meters (locally) or a few hundred meters (globally), depending on flyby<br />

148 distances. Moreover, as many as 33 color and polarization filter combinations permit mapping<br />

149 geologically diverse terrain spatially at wavelengths between ~0.3µm and ~1.1µm. The visible and<br />

150 infrared mapping spectrometer (VIMS; 0.4µm to 5.1µm) provides chemical and compositional<br />

151 spectral information with spatial resolutions from a few kilometers (globally) to better than one<br />

152 hundred meters (locally) (Brown et al., 2004; Jaumann et al., 2006). The composite infrared<br />

153 spectrometer (CIRS) determines global and regional surface temperatures and thermal properties on<br />

154<br />

155<br />

a kilometer scale (Flasar et al., 2004). The ultraviolet imaging spectrograph (UVIS) provides<br />

information about thin atmospheres and volcanic plume structures as well as about water ice and<br />

156 other minor constituents on the surface (Esposito et al., 2004), operating in the 60nm - 190nm<br />

157 wavelength range. A suite <strong>of</strong> magnetosphere and plasma science (MAPS) instruments characterizes<br />

158 the satellites’ environments by in-situ methods. Micron-sized dust grains and neutral molecules<br />

159 released from the surface by active or passive processes (volcanoes/sputtering) carry surface<br />

160 composition information to distances as far as hundreds <strong>of</strong> kilometers above the surface. Released<br />

161 material affects the plasma surrounding the satellite and is registered by variations in the magnetic<br />

162 field, ion density, and neutral gas and dust density as well as gas and dusty composition.<br />

163 The primary task <strong>of</strong> the radio science subsystem (RSS) is to determine the mass <strong>of</strong> the moons (Tab.<br />

164 1) by tracking deviations in Cassini's trajectory. RADAR data can provide unique information about<br />

165 the upper sub-surface (Elachi et al., 2004), though few close-up RADAR SAR observations were<br />

166 performed during satellite flybys. The lack <strong>of</strong> a scan platform for the remote sensing instruments<br />

167 prevents the radio science, RADAR and remote sensing systems from operating simultaneously<br />

168 during a flyby because the antenna and remote sensing instruments are oriented 90° apart on the<br />

169 spacecraft. Even joint surface scans by the ORS instruments require significant compromises<br />

170 between individual observations due to data rate and integration time constraints. While the ISS<br />

171 instrument needs short 'dwell' times for mosaicking, the VIMS depend on long 'dwells' for<br />

172 improved signal-to-noise ratios. On the other hand, the CIRS and UVIS instruments contain line-<br />

173 scanning devices, which depend on either slow or fast slews to scan the surfaces. Thus, reaching<br />

174 satisfactory compromises is a major challenge in the planning process.<br />

175 Since the Cassini spacecraft orbits the planet rather than individual satellites, the moons can only be<br />

176<br />

177<br />

observed at various distances and illumination conditions, ideally during very close-targeted flybys.<br />

Depending on the distance <strong>of</strong> a moon to Saturn, the flybys occur at very different velocities. In<br />

178 March 2008, for instance, Enceladus was passed at ~14 km/s, while the Iapetus flyby in September<br />

179 2007 took place at a leisurely 2.4 km/s. The flyby geometry can also vary significantly. The closest-<br />

180 ever approach was during the Enceladus flyby on October 9, 2008, when the spacecraft skimmed as<br />

181 low as 25 km over Enceladus’ surface. Other, more typical targeted flyby altitudes occur between<br />

182 100 km and 2000 km. A flyby can be polar, equatorial, or in between, and at the closest approach,<br />

183 the sub-spacecraft point over a moon can be located either over its illuminated or over its unlit side.<br />

184 The MAPS instruments measure densities or field gradients over time. Sputtering processes lead to<br />

185 high dust and neutral-gas densities above the surface, so that environmental in-situ instruments rely<br />

186 on measurements as close to the surface as possible. Further plasma density variations caused by<br />

187 satellites moving in the magnetosphere are empty flux tubes (wakes) and the drop <strong>of</strong> plasma along<br />

188 the related L-shell. Crossing those regions is <strong>of</strong> high value for in-situ plasma investigations that<br />

189 provide indirect information about the satellite surface and the density <strong>of</strong> its neutral and plasma<br />

190 environment.<br />

191 During the nominal mission, Cassini performed nine targeted as well as numerous close flybys <strong>of</strong><br />

192<br />

icy satellites (Tab. 2).<br />

4


193<br />

Target Orbit Flyby date Altitude<br />

[km]<br />

Solar Phase Angle 2 hrs before,<br />

at, and 2 hrs after C/A [deg]<br />

Phoebe 0 June 11, 2004 1997 85 / 25 / 90 T<br />

Dione B December 15, 2004 72100 108 / 85 / 62 nT<br />

Iapetus B/C December 31, 2004 123399 87 / 94 / 100 nT<br />

Enceladus 3 February 17, 2005 1176 25 / 114 / 158 nT<br />

Enceladus 4 March 9, 2005 500 47 / 42 / 130 T<br />

Hyperion 9 June 10, 2005 168000 25 / 35 / 47 nT<br />

Enceladus 11 July 14, 2005 170 47 / 63 / 132 T<br />

Mimas 12 August 2, 2005 62700 41 / 57 / 91 nT<br />

Tethys 15 September 24, 2005 1503 21 /68 / 155 T<br />

Hyperion 15 September 26, 2005 505 51 / 50 / 127 T<br />

Dione 16 October 11, 2005 500 23 / 66 / 156 T<br />

Iapetus 17 November 12, 2005 415400 89 / 91 / 93 nT<br />

Rhea 18 November 26, 2005 1264 19 / 87 / 161 T<br />

Rhea 22 March 21, 2006 82200 113 / 137 / 157 nT<br />

Iapetus 23 April 11, 2006 602400 123 / 124 / 125 nT<br />

Enceladus 27 September 9, 2006 39900 105 / 118 / 105 nT<br />

Dione 33 November 21, 2006 72300 125 / 144 / 119 nT<br />

Hyperion 39 February 16, 2007 175000 66 / 64 / 64 nT<br />

Tethys 47 June 27, 2007 18400 148 / 116 / 49 nT<br />

Tethys 49 August 29, 2007 55500 135 / 102 / 72 nT<br />

Rhea 49 August 30, 2007 5800 125 / 46 / 41 nT<br />

Iapetus 49 September 10, 2007 1620 139 / 59 / 27 T<br />

Dione 50 September 30, 2007 43400 89 / 49 / 19 nT<br />

Hyperion 51 October 21, 2007 122000 115 / 116 / 115 nT<br />

Mimas 53 December 3, 2007 84200 161 / 139 / 82 nT<br />

Enceladus 61 March 12, 2008 52 115 / 135 / 65 T<br />

Enceladus 80 August 11, 2008 50 108 / 110 / 73 T<br />

Enceladus 88 October 9, 2008 25 108 / 112 / 73 T<br />

Enceladus 91 October 31, 2008 200 108 / 113 / 73 T<br />

Tethys 93 November 16, 2008 57100 79 / 41 / 58 nT<br />

Tethys 94 November 24, 2008 24200 123 / 159 / 84 nT<br />

Tethys 115 July 26, 2009 68400 105 / 89 / 79 nT<br />

Rhea 119 October 13, 2009 40400 140 / 82 / 25 nT<br />

Mimas 119 October 14, 2009 44200 152 / 102 / 44 nT<br />

Hyperion 119 October 16, 2009 127000 155 / 140 / 126 nT<br />

Enceladus 120 November 2, 2009 100 177 / 90 / 5 T<br />

Enceladus 121 November 21, 2009 1600 146 / 87 / 36 T<br />

Rhea 121 November 21, 2009 24400 127 / 58 / 10 nT<br />

Tethys 123 December 26, 2009 52900 136 / 77 / 17 nT<br />

Dione 125 January 27, 2010 45100 158 / 106 / 53 nT<br />

Mimas 126 February 13, 2010 9500 173 / 99 / 21 nT<br />

Rhea 127 March 2, 2010 100 176 / 87 / 3 T<br />

Dione 129 April 7, 2010 500 167 / 79 / 11 T<br />

Enceladus 130 April 28, 2010 100 171 / 93 / 9 T<br />

Enceladus 131 May 18, 2010 200 153 / 108 / 29 T<br />

5<br />

Target [T]/Non<br />

Targeted [nT]


Tethys 132 June 3, 2010 52600 154 / 99 / 45 nT<br />

194<br />

195<br />

196<br />

Tab. 2 Mission characteristics <strong>of</strong> the Cassini icy satellites flybys; C/A ... closest approach<br />

197 Three <strong>of</strong> the targeted flybys were dedicated to Enceladus, two in 2005 and one in 2008. In addition,<br />

198 a very close but non-targeted Enceladus flyby happened in February 2005. As the data provided by<br />

199 the MAPS instruments in March 2005 had raised suspicions about potential internal activity (Spahn<br />

200 et al., 2006; Waite et al., 2006), the July 2005 flyby was lowered from 1000 km to 173 km to allow<br />

201 for increased sensitivity in MAPS measurements. This flyby supplied overwhelming evidence from<br />

202 multiple instruments <strong>of</strong> current activity on Enceladus, eventually confirmed by direct images <strong>of</strong><br />

203 plumes taken during a distant flyby in November 2005. The locations <strong>of</strong> the plumes were found to<br />

204 coincide with linear tectonic features near the south pole, dubbed 'tiger stripes'. The CIRS<br />

205 instrument also identified these 'tiger stripes' as unusually warm linear features (Spencer et al.,<br />

206 2006); the tiger strips contain relatively large ice particles identified by VIMS (Brown et al., 2006;<br />

207<br />

208<br />

Jaumann et al., 2008).<br />

All other moons had only one targeted flyby during the nominal mission, if any. The Tethys,<br />

209 Hyperion, and Dione flybys in the late summer/early fall <strong>of</strong> 2005 were performed within 3 weeks <strong>of</strong><br />

210 each other during the same orbit <strong>of</strong> Cassini around Saturn. Due to Hyperion's chaotic rotation, it<br />

211 was impossible to predict which part <strong>of</strong> its surface would be visible and how the surface might look.<br />

212 By coincidence, the viewing perspective during the approach was almost identical with the best<br />

213 Voyager 1 views. It showed a giant crater, dark material within numerous smaller craters and a very<br />

low mean density <strong>of</strong> only 0.54g/cm 3 214<br />

. The 'wispy streaks' <strong>of</strong> Dione were observed at very oblique<br />

215 viewing angles during the October 2005 flyby, and were confirmed to be tectonic in origin (Wagner<br />

216 et al., 2006). The images from this targeted Dione flyby are among the most spectacular pictures <strong>of</strong><br />

217 the mission. The Rhea flyby in November 2005 was dedicated to radio science, and the results from<br />

218 that flyby are reported in Matson et al. (2009). However, while the spacecraft itself was inertially<br />

219 pointed with its antenna towards <strong>Earth</strong>, the cameras slewed across the surface, obtaining the<br />

220 highest-resolved images <strong>of</strong> Rhea to date.<br />

221 In 2006, no targeted flybys took place, primarily because the spacecraft spent most <strong>of</strong> its time in<br />

222 highly inclined orbits. The very close but non-targeted Rhea flyby in August 2007 was dedicated to<br />

223 remote sensing and revealed small details <strong>of</strong> the prominent bright-ray crater and its ejecta<br />

224 environment. This might be the youngest large crater in the Saturnian system. The targeted Iapetus<br />

225 flyby in September 2007 over the anti-Saturnian and trailing hemispheres yielded very high-<br />

226 resolution observations <strong>of</strong> the highest parts <strong>of</strong> the equatorial ridge, the transition zone between the<br />

227 dark and the bright terrain, and the bright trailing side. Combined with data from a more distant,<br />

228 non-targeted flyby on December 31, 2004, and other distant observations, these data permitted<br />

229 developing a model <strong>of</strong> the formation <strong>of</strong> the global and unique brightness dichotomy <strong>of</strong> this unusual<br />

230 moon (Denk and Spencer, 2008). For more detailed descriptions <strong>of</strong> the Cassini mission, see<br />

231<br />

232<br />

Dougherty et al. (2009), Orton et al. (2009) and Seal et al. (2009).<br />

233 21.3. Morphology, Geology and Topography<br />

234 Voyager’s partial global mapping at 1 km resolution suggested that the morphology <strong>of</strong> the<br />

235 Saturnian satellites was dominated by impact craters (Smith et al., 1981). Indeed, Voyager’s best<br />

236<br />

237<br />

images, 500 m resolution views <strong>of</strong> Rhea, showed a landscape saturated with craters. Voyager also<br />

found evidence <strong>of</strong> limited volcanic and tectonic activity in the form <strong>of</strong> relatively smooth plains and<br />

238 linear features (Smith et al., 1981, 1982; Plescia and Boyce, 1982, 1983; Moore et al., 1985; Moore<br />

239 and Ahren, 1983), especially on Enceladus. The Enceladus discoveries aside, Cassini’s considerably<br />

240 improved resolution and global mapping coverage confirmed the abundance <strong>of</strong> impact craters on all<br />

241 satellites (Dones et al., 2009) and also revealed surprisingly varied degrees <strong>of</strong> past endogenic<br />

242 activity on Tethys, Rhea and especially Dione, indicating that these satellites were more dynamic<br />

243 than originally thought. Observed deformations range from linear grooves (Mimas) and nearly<br />

244 global-scale grabens (Tethys, Rhea) to repeated episodes <strong>of</strong> tectonics and resurfacing (Dione,<br />

245<br />

246<br />

Enceladus). To date, little or no direct evidence <strong>of</strong> past or present endogenic activity has been<br />

identified on Hyperion, Phoebe or Iapetus (other than its potentially endogenic equatorial ridge).<br />

247 This diversity is in itself a puzzle that has not yet been explained. In the following sections, we<br />

248<br />

249<br />

examine the features observed, their morphology, and implications for internal evolution.<br />

250<br />

Craters<br />

6


251 Comparison with craters on larger icy satellites (e.g. Ganymede) is illuminating because surface<br />

252 gravities differ by a factor <strong>of</strong> 5 to 10. Central-peak craters are common on Saturnian satellites.<br />

253 Central peaks in larger craters can be as high as 10 km, many <strong>of</strong> them rising above the local mean<br />

254 elevation. Cassini has revealed little evidence <strong>of</strong> terrace formation, and floor uplift dominates clearly<br />

255 over rim collapse in the formation <strong>of</strong> complex craters (Schenk and Moore, 2007). Pancake (formerly<br />

256 called pedestal) ejecta have been observed in smooth plains on Dione. This kind <strong>of</strong> ejecta deposit,<br />

257 once only known as a feature <strong>of</strong> the Galilean satellites, now appears to be common on icy satellites<br />

258 in general. The fact that it is not obvious on other satellites is most likely due to the difficulty <strong>of</strong><br />

259 identifying it on these rugged, heavily cratered surfaces.<br />

260 A transition to larger, more complex crater landforms (e.g. central pit craters, multi-ring basins) on<br />

261 these satellites is predicted to occur at much larger diameters (>150 km) than those seen on<br />

262 Ganymede and Callisto (Schenk, 1993), due to their weaker gravity. True central-pit craters like<br />

263 those on Ganymede or Callisto (Schenk, 1993) have not been observed, although the central<br />

264 complex <strong>of</strong> Odysseus features a central depression and closely resembles its counterparts on large<br />

265 icy moons. The dominant morphology at larger diameters (>150 km) is peak-ring or similar, with a<br />

266 prominent central peak surrounded by a ring or an elevated circular plateau. The number <strong>of</strong> such<br />

267 basins discovered by Cassini on Iapetus and Rhea was greater than expected. Indeed, at least 10<br />

268 basins larger than 300 km across have been seen on Iapetus (e.g. Giese et al., 2008).<br />

269 The most ancient large basins on Rhea and Tethys have been modified by relaxation and long-term<br />

270 impact bombardment (e.g. Schenk and Moore, 2007). The younger basins are generally deep and<br />

271<br />

272<br />

unrelaxed. However, the floor <strong>of</strong> Evander, is the largest basin identified on Dione, has been relaxed<br />

and uplifted to the level <strong>of</strong> the original surface, and although its rim and peak remain prominent<br />

273 (e.g. Schenk and Moore, 2007). Relaxation <strong>of</strong> this basin is consistent with elevated heat flows on<br />

274 Dione, as indicated by the relaxation <strong>of</strong> smaller (diameters <strong>of</strong> 25-50 km) craters within the extensive<br />

275 smooth plains to the north <strong>of</strong> Evander. Relaxation in this basin, which has a low superposed crater<br />

276 density, is consistent with elevated heat flows, as indicated by the smaller relaxed craters nearby and<br />

277 the extensive smooth plains to the north. The relaxation state <strong>of</strong> impact basins can be determined by<br />

278 comparing their depth to that <strong>of</strong> unmodified craters and examining their shape. The most ancient<br />

279 large basins on Rhea (e.g. Tirawa) and Tethys are roughly 10 to 20% shallower than predicted and<br />

280 feature prominent domical or conical central uplifts that approach or exceed the local elevation <strong>of</strong><br />

281 the satellite (Schenk and Moore, 2007), which suggests they have been modified to some degree by<br />

282 relaxation. The youngest basins appear to be deeper and possibly unrelaxed. The most prominent<br />

283 examples include Herschel on Mimas (diameter ~120 km, depth ~11 km) and Odysseus (diameter<br />

284 ~450 km, depth >8 km). In neither case do the central uplifts rise above the local topography. A<br />

285 prominent exception is the relatively young 400 km wide Evander basin on Dione, which may be<br />

286 either ~3Gyr (Wagner et al., 2006) or less than 2Gyr old (Kirch<strong>of</strong>f and Schenk, 2009). For a more<br />

287 detailed discussion <strong>of</strong> surface ages, see Dones et al. (2009).<br />

288 Iapetus’ leading side shows more relief than the other medium-sized Saturnian satellites (Thomas et<br />

289 al., 2007b; Giese et al., 2008). The lack <strong>of</strong> basin relaxation on Iapetus is consistent with the<br />

290 presence <strong>of</strong> a thick (50–100 km) lithosphere in Iapetus’ early history. Such a lithosphere might also<br />

291 support Iapetus’ prominent equatorial ridge. The ridge shows a complex morphology, probably the<br />

292 result <strong>of</strong> impact erosion in some places and <strong>of</strong> post-formation tectonic modifications in others.<br />

293<br />

294<br />

The abundance <strong>of</strong> large impact basins, even in Voyager’s limited view, suggested that such basins<br />

could have disruptive effects on these satellites (Smith et al., 1981). Several attempts to discover<br />

295 evidence <strong>of</strong> these effects (e.g. seismic shaking, global fracturing) have met with limited success (e.g.<br />

296<br />

297<br />

Moore et al., 2004), but Cassini-based studies are now in progress.<br />

298 Tectonics<br />

299 The Voyager 1 and 2 spacecraft revealed that the medium-sized icy Saturnian satellites have<br />

300 undergone varying degrees <strong>of</strong> tectonic deformation (Smith et al. 1981; 1982; Plescia and Boyce<br />

301 1982, 1983). Although such deformation can provide constraints on the interior structure and<br />

302 evolution <strong>of</strong> a satellite, the imaging resolution was <strong>of</strong>ten insufficient to permit sustainable<br />

303 conclusions. Post-Voyager advances include a vastly better understanding <strong>of</strong> tectonic deformation<br />

304 and thermal evolution in the Jovian system thanks to Galileo, and greatly improved experimental<br />

305 measurements <strong>of</strong> relevant ice properties (Schmitt et al., 1998). In this section, we combine this<br />

306 improved understanding with the results <strong>of</strong> the ongoing Cassini mission to discuss the current state<br />

307 <strong>of</strong> knowledge <strong>of</strong> icy-satellite tectonics in the Saturnian system. We also examine the extent to which<br />

308 cryovolcanic processes may be relevant. Most <strong>of</strong> the topics covered in this section are treated in<br />

309<br />

greater detail in Collins et al. (2009).<br />

7


310 Tectonics on icy satellites comprises three major aspects: (1) the various mechanisms by which ice<br />

311 deforms when subjected to tectonic stress; (2) the mechanisms by which tectonic stress may arise;<br />

312 and (3) the observational constraints on these deformation mechanisms.<br />

313 From a tectonics point <strong>of</strong> view, ice is rather similar to silicate materials: under surface conditions its<br />

314 behavior is rigid or brittle, while at greater depths it is ductile. The rheology <strong>of</strong> ice is complicated<br />

315 and cannot be discussed in depth here; useful discourses on its elastic, brittle and viscous properties<br />

316 may be found in Petrenko and Whitworth (1999), Beeman et al. (1988) and Durham and Stern<br />

317 (2001), respectively. Although a real ice shell will exhibit all three kinds <strong>of</strong> behavior, it can be<br />

318 modeled as a simple elastic layer (e.g. McNutt, 1984). This effective elastic thickness depends on<br />

319 the temperature gradient within the ice shell during deformation and can be inferred from<br />

320 topographical measurements (e.g. Giese et al., 2007). Thus, measuring the elastic thickness places a<br />

321 useful constraint on satellite thermal evolution (e.g. Nimmo et al., 2002). Pressures within Saturnian<br />

322<br />

323<br />

satellites are generally low (see Tab. 3).<br />

R<br />

(km)<br />

Ms<br />

(10 20 kg)<br />

a<br />

(10 3 km)<br />

e H<br />

(m)<br />

3eH<br />

(m)<br />

8<br />

Tidal<br />

(10 10 W)<br />

Rad.<br />

(10 10 W)<br />

Stress<br />

(kPa)<br />

Mimas 198 0.37 186 .0202 9180 556 78 0.013 8400 7.1 I<br />

Enceladus 252 1.08 238 .0045 3940 53 2 0.038 630 23 III<br />

Tethys 533 6.15 295 0. 7270 0 0 0.22 0 37 II<br />

Dione 562 11 377 .0022 2410 16 0.86 0.39 85 97 II<br />

Rhea 764 23 527 .001 1440 4.3 0.067 0.81 17 124 II<br />

Titan 2575 1346 1222 .0292 254 22 45 47 26 3280 -<br />

Iapetus 736 18 3561 .0283 5 0.4 2.6x10 -5 0.63 1.7 88 I/II<br />

324<br />

325<br />

326<br />

Tab. 3 Data from Yoder (1995) and Thomas et al. (2007). R, Ms, a and e are the radius, mass,<br />

semi-major axis, period and eccentricity, respectively. H is the permanent tidal bulge assuming<br />

327<br />

328<br />

329<br />

330<br />

h2=2.5. 3eH is the approximate magnitude <strong>of</strong> the diurnal tidal bulge – note that this is likely an<br />

overestimate for small satellites because their rigidity will reduce h2. “Tidal” is the tidal heat<br />

production assuming a homogeneous body with k2=1.5 (again, a likely overestimate for small<br />

satellites) and Q=100. “Rad” is the radiogenic heat production assuming a chondritic rate <strong>of</strong><br />

3.5x10 -12 331 W/kg. “Stress” is the approximate diurnal tidal stress given by EeH/R where E is<br />

332<br />

333<br />

Young’s modulus (assumed 9 GPa). Pcen is the central pressure, assuming uniform density. “MG” is<br />

the “metamorphic grade”, modified after Johnson (1998), which is a qualitative assessment <strong>of</strong> the<br />

334<br />

335<br />

degree <strong>of</strong> deformation: I=unmodified, II=intermediate, III=heavily modified<br />

336 Thus, although ice transmutes into high-pressure forms at pressures greater than about 0.2GPa (e.g.<br />

337 Sotin et al., 1998), this effect is relevant mainly on Titan.<br />

338 Tectonic features may be used to infer the direction and (sometimes) the magnitude <strong>of</strong> the driving<br />

339 stresses. Observations can thus place constraints on the mechanisms which gave rise to these<br />

340 tectonic features. We briefly discuss several mechanisms which are potential candidates for<br />

341 producing tectonic features.<br />

342 Satellites experience both rotational and tidal deformation. If the rotation rate or the magnitude <strong>of</strong><br />

343<br />

344<br />

the tide changes, the shape <strong>of</strong> the satellite will change as well, generating global-scale patterns <strong>of</strong><br />

stress. A recent compilation <strong>of</strong> the present-day shapes <strong>of</strong> the Saturnian satellites is given by Thomas<br />

345 et al. (2007b); a good general description <strong>of</strong> tidal and rotational satellite deformation is given in<br />

346 Murray and Dermott (1999).<br />

347 If a synchronously rotating satellite’s orbital eccentricity is zero, the tidal bulge will be at a fixed<br />

348 geographic point, and <strong>of</strong> constant amplitude. The maximum amplitude H <strong>of</strong> this static (or<br />

349 permanent) tidal bulge depends on the mass and radius <strong>of</strong> the satellite, the mass <strong>of</strong> the primary, the<br />

350<br />

351<br />

352<br />

distance between them, and the rigidity <strong>of</strong> the satellite. The latter is described by the h2 Love<br />

number, where h2 has a value <strong>of</strong> 5/2 for a fluid body <strong>of</strong> uniform density which, however, declines as<br />

the rigidity (or shear modulus) µ <strong>of</strong> the satellite increases (e.g. Murray and Dermott, 1999).<br />

353 In a synchronous satellite with non-zero orbital eccentricity, the amplitude and direction <strong>of</strong> the<br />

354 static tidal bulge will change slightly, resulting in a time-varying diurnal tide. Causing diurnal<br />

355 stresses, this tide has an amplitude expressed by 3eH, where e is the eccentricity and H is the<br />

356 equilibrium tidal bulge (e.g. Greenberg and Geissler, 2002). Obliquity librations are potentially an<br />

357 additional source <strong>of</strong> tidal stress (Bills and Ray, 2000); they can be large if their period is a small<br />

358<br />

multiple <strong>of</strong> the orbital period (Wisdom, 2004).<br />

Pcen<br />

(MPa)<br />

MG


359 Table 3 shows the diurnal tidal stresses <strong>of</strong> the icy Saturnian satellites at zero rigidity, demonstrating<br />

360 the general decrease in stress as the semi-major axis lengthens. For comparison, the diurnal stresses<br />

361 on Europa thought to be responsible for cycloidal features are a few tens <strong>of</strong> kPa (Hoppa et al.,<br />

362 1999a). Because the principal stresses rotate during each orbit, contraction, extension and mixed-<br />

363 mode horizontal principal stresses which promote strike-slip motions are all possible as a satellite<br />

364 orbits (Hoppa et al., 1999b). This motion can lead to both shear heating (Nimmo and Gaidos, 2002)<br />

365 and the monotonic accumulation <strong>of</strong> strike-slip <strong>of</strong>fsets (Hoppa et al., 1999b).<br />

366 A satellite surface may move longitudinally with respect to the static tidal bulge due to tidal<br />

367 (Greenberg and Weidenschilling, 1984) or atmospheric torques (Lorenz et al., 2008) or, in the case<br />

368 <strong>of</strong> floating ice shells, if the thermally equilibrated ice shell thickness variations are not compatible<br />

369 with rotational equilibrium (Ojakangas and Stevenson, 1989a). This non-synchronous rotation leads<br />

370 to surface stresses (Helfenstein and Parmentier, 1985). Stresses from non-synchronous rotation<br />

371 increase with the angular speed <strong>of</strong> rotation but will dissipate with time owing to viscous relaxation<br />

372 (Greenberg and Weidenschilling, 1984; Wahr et al., 2008). Maximum tensile stress depends on the<br />

373 amount <strong>of</strong> reorientation and the degree <strong>of</strong> flattening, f, which depends in turn on the satellite<br />

374 rotation rate (Leith and McKinnon, 1996). In non-rigid satellites, the stresses due to non-<br />

375 synchronous rotation will exceed the diurnal tidal stresses for a reorientation typically in excess <strong>of</strong><br />

376 roughly one degree. In certain circumstances, the surface <strong>of</strong> a satellite may reorient relative to its<br />

377 axis <strong>of</strong> rotation (Willemann, 1984; Matsuyama and Nimmo, 2008) and give rise to global fracture<br />

378 patterns (e.g. Schenk et al., 2008). This process <strong>of</strong> true polar wander is conceptually very similar to<br />

379<br />

380<br />

non-synchronous rotation (Melosh, 1980; Leith and McKinnon, 1996), and 'reorientation' will in the<br />

following be understood as implying either non-synchronous rotation or true polar wander. In<br />

381 general, rotation axis motion occurs roughly perpendicular to the tidal axis because this path is<br />

382 energetically favored (Matsuyama and Nimmo, 2007). True polar wander can occur because <strong>of</strong> ice<br />

383 shell thickness variations (Ojakangas and Stevenson, 1989b), long-wavelength density anomalies<br />

384 (Janes and Melosh, 1988; Nimmo and Pappalardo, 2006), volatile redistribution (Rubincam, 2003)<br />

385 or impacts, either directly (Chapman and McKinnon, 1986) or due to the creation <strong>of</strong> a new impact<br />

386 basin (Melosh, 1975; Murchie and Head, 1986; Nimmo and Matsuyama, 2007). Like non-<br />

387 synchronous rotation, the maximum stress developed by true polar wander depends on the amount<br />

388 <strong>of</strong> reorientation and the effective flattening f, with the value <strong>of</strong> f depending on whether any<br />

389 reorientation <strong>of</strong> the tidal axis has occurred (Matsuyama and Nimmo, 2008).<br />

390 Satellites that are initially rotating at a rate faster than synchronous will despin (i.e. reduce their<br />

391 rotation rate) to synchronous rotation in timescales that are generally very short compared to the age<br />

392 <strong>of</strong> the solar system (e.g. Murray and Dermott, 1999). Spin rate reductions result in shape changes<br />

393 and stresses – compressive at the equator as the equatorial bulge collapses, and tensile at the poles<br />

394 as the poles elongate. The maximum differential stress caused by despinning in a hydrostatic body<br />

395 depends mainly on the satellite radius and the difference between the initial and the final angular<br />

396 rotation velocities (Melosh, 1977). Spin-up may occur in some circumstances (e.g. if the satellite<br />

397 undergoes differentiation), and in this case the signs <strong>of</strong> all the stresses are reversed. Stresses along<br />

398 lines <strong>of</strong> longitude are always larger than the stresses along lines <strong>of</strong> latitude. Thus, irrespective <strong>of</strong><br />

399 whether a satellite is spinning down or spinning up, equatorial tectonic features are expected to be<br />

400 oriented in a north-south direction. This result is important when considering the origin <strong>of</strong> the<br />

401<br />

402<br />

global equatorial ridge on Iapetus.<br />

Synchronous satellites, which are evolving outwards due to tidal torques from the primary, will<br />

403 undergo both a reduction in their spin rate and a parallel reduction in the tidal bulge amplitude. This<br />

404 combination <strong>of</strong> despinning and tidal bulge reduction causes a stress pattern in which a region<br />

405 around the sub-planet point experiences compressive stress, middle latitudes experience horizontal<br />

406 shear stress, and the poles undergo tension (Melosh, 1980; Helfenstein and Parmentier, 1983). The<br />

407 maximum principal stress difference is similar to that produced by despinning alone (Melosh,<br />

408 1980).<br />

409 A large variety <strong>of</strong> mechanisms can lead to volume changes within a satellite, and thus to extensional<br />

410 or contractional features on the surface (Squyres and Cr<strong>of</strong>t, 1986; Kirk and Stevenson, 1987;<br />

411 Mueller and McKinnon, 1988). Volume changes generate isotropic stress fields on the surface. A<br />

412 potentially important source <strong>of</strong> expansion or contraction is the large density contrast between ice I<br />

413 and water. As water freezes to ice I, large surface extensional stresses may result (Cassen et al.,<br />

414 1979; Smith et al., 1981; Nimmo, 2004a). The reason for this effect is that, on a spherical body, the<br />

415 increased volume <strong>of</strong> ice relative to water drives the ice shell outwards; and this increase in radius<br />

416<br />

results in extension. A fractional change in radius <strong>of</strong> 0.1% gives rise to stresses <strong>of</strong> about 10MPa.<br />

9


417 This effect is much reduced if there are high-pressure ice phases freezing simultaneously (Squyres,<br />

418 1980; Showman et al., 1997).<br />

419 Similar but smaller effects are caused by thermal expansion or contraction (Ellsworth and Schubert,<br />

420 1983; Hillier and Squyres, 1991; Showman et al., 1997; Nimmo, 2004a). A global temperature<br />

change <strong>of</strong> 100K will cause stresses on the order <strong>of</strong> 100 MPa for a thermal expansivity <strong>of</strong> 10 -4 K -1 421 .<br />

422 Warming may also lead to silicate dehydration (e.g. Squyres and Cr<strong>of</strong>t, 1986), which in turn causes<br />

423 expansion.<br />

424 Satellites are quite likely to have non-axisymmetric structures, in which case some <strong>of</strong> the above-<br />

425 mentioned global mechanisms may lead to local deformation. For instance, local zones <strong>of</strong> weakness<br />

426 or pre-existing structures can result in enhanced tidal dissipation and deformation (e.g. Sotin et al.,<br />

427 2002; Nimmo, 2004b) and/or the alteration <strong>of</strong> local stress trajectories. Convection (thermal or<br />

428 compositional) and buoyancy forces due to lateral shell thickness variations can generate local<br />

429 deformation. They are discussed at greater length in Collins et al. (2009). We will not discuss these<br />

430 processes further here, as they do not tend to produce globally organized tectonic structures.<br />

431 Large impact events can also potentially create global fracture networks. This mechanism is implicit<br />

432 in the model presented by E. Shoemaker in Smith et al. (1981) for the catastrophic breakup and re-<br />

433 accretion <strong>of</strong> the medium-sized icy satellites <strong>of</strong> Saturn. Presumably, impacts not quite large enough<br />

434 to completely disrupt a satellite could do considerable damage in the form <strong>of</strong> global fracturing, and<br />

435 post-Voyager studies <strong>of</strong> these satellites focused in part on attempts to find evidence <strong>of</strong> an incipient<br />

436 break-up <strong>of</strong> this kind. For instance, the surface <strong>of</strong> Mimas is crossed by a number <strong>of</strong> linear-to-<br />

437<br />

438<br />

439<br />

arcuate, sub-parallel troughs (Fig. 1) that are thought to be the consequence <strong>of</strong> global-scale fractures<br />

during the Herschel impact event (McKinnon, 1985; Schenk, 1989a).<br />

440 Fig. 1 Linear troughs on Mimas, extending east to west. These troughs are interpreted as fractures,<br />

441 possibly related to the Herschel crater. Orthographic map projection at 400m/pixel is centered on<br />

442 the antipode to Herschel, the largest impact basin on Mimas.<br />

443<br />

444 Voyager and Cassini mapping indicates that these features, if related to Herschel, do not form a<br />

445 simple radial or concentric pattern. Nor has it been demonstrated that any <strong>of</strong> these features form on<br />

446 a plane intersecting Herschel, or whether they should do so. It thus remains possible that these are<br />

447 random fractures formed during freeze expansion <strong>of</strong> the Mimas interior. Detailed mapping remains<br />

448 to be done to test these hypotheses. Mimas does not otherwise exhibit endogenic landforms and<br />

449 will, therefore, not be discussed further.<br />

450 Enceladus shows the greatest tectonic deformation <strong>of</strong> any <strong>of</strong> the Saturnian satellites, and very<br />

451 significant spatial variations in surface age (Smith et al., 1982; Kargel and Pozio, 1996; Porco et al.,<br />

452 2006; Jaumann et al., 2008). Spencer et al. (2009) discusses the tectonic behavior <strong>of</strong> Enceladus in<br />

453 much greater detail; only a very brief summary is given here.<br />

454 Three different terrain types exist on Enceladus. Ancient cratered terrains cover a broad band from<br />

0 to 180 o longitude. Centred on 90 o and 270 o 455<br />

longitude there are younger, deformed terrains roughly<br />

90 o wide at the equator. The southern polar region south <strong>of</strong> 55 o 456<br />

S is heavily deformed and almost<br />

457 uncratered. The most prominent tectonic features <strong>of</strong> this region are the linear depressions called<br />

458 ‘tiger stripes’ (Porco et al., 2006). These tiger stripes are typically ~500m deep, ~2 km wide, up to<br />

459<br />

460<br />

~130 km long and spaced ~35 km apart. Crosscutting fractures and ridges with almost no<br />

superimposed impact craters characterize the area between the stripes. The tiger stripes are<br />

461 apparently the source <strong>of</strong> the geysers observed to emanate from the south polar region (Spitale and<br />

462 Porco, 2007). The energy source <strong>of</strong> these geysers is unknown, but it may be shear heating generated<br />

463 by strike-slip motion at the tiger stripes (Nimmo et al., 2007). The coherent (but latitudinally<br />

464 asymmetrical) global pattern <strong>of</strong> deformation on Enceladus is currently unexplained, but it is most<br />

465 probably due to shape changes, perhaps related to a hypothetical episode <strong>of</strong> true polar wander<br />

466 (Nimmo and Pappalardo, 2006).<br />

467<br />

468<br />

Nearly encircling the globe, Ithaca Chasma is the most prominent feature on Tethys (Fig. 2).<br />

469 Fig. 2 Tethys and Ithaca Chasma. Topography was derived from Cassini stereo images and has a<br />

470 horizontal resolution <strong>of</strong> 2-5 km and a vertical accuracy <strong>of</strong> 150-800m. Pr<strong>of</strong>iles have superimposed<br />

471<br />

472<br />

model flexural pr<strong>of</strong>iles (in red). Te is elastic thickness. (Reproduced from Giese et al. (2007))<br />

473 It extends approximately 270° around Tethys and is not a complete circle. It is narrowly confined to<br />

474 a zone which lies along a large circle whose pole is only ~20° from the center <strong>of</strong> Odysseus, the<br />

475<br />

relatively fresh largest basin on the satellite which is 450 km wide (Smith et al., 1982; Moore and<br />

10


476 Ahern, 1983). Smith et al. (1981) suggested that Ithaca Chasma was formed by freeze-expansion <strong>of</strong><br />

477 Tethys's interior. They noted that if Tethys was once a sphere <strong>of</strong> liquid water covered by a thin solid<br />

478 ice crust, freezing in the interior would have produced an expansion <strong>of</strong> the surface comparable to<br />

479 the area <strong>of</strong> the chasm (~10% <strong>of</strong> the total satellite surface area). This hypothesis fails to explain why<br />

480 the chasm occurs only within a narrow zone; besides, it is difficult to account for the geophysical<br />

481 energy needed for the presence <strong>of</strong> a molten interior. Expansion <strong>of</strong> the satellite's interior should have<br />

482 caused fracturing over the entire surface in order to effectively relieve stresses in a rigid crust<br />

483 (Moore and Ahern, 1983).<br />

484 The nearly concentric geometrical relationship between Odysseus and Ithaca Chasma prompted<br />

485 Moore and Ahern (1983) to suggest that the trough system was an immediate manifestation <strong>of</strong> the<br />

486 impact event, perhaps caused by a damped, whole-body oscillation <strong>of</strong> the satellite. Based on the<br />

487 deep topography <strong>of</strong> Odysseus, Schenk (1989b) also suggested that it had not undergone substantial<br />

488 relaxation, and that Ithaca Chasma was the equivalent <strong>of</strong> a ring graben formed during the impact<br />

489 event by a prompt collapse <strong>of</strong> the floor involving a large portion <strong>of</strong> the interior. In a Cassini-era<br />

490 study, Giese et al. (2007) doubted the Odysseus-related hypothesis (Moore and Ahern, 1983;<br />

491 Schenk, 1989b) on the basis <strong>of</strong> crater densities within Ithaca Chasma relative to Odysseus, from<br />

492 which they concluded that Ithaca Chasma is older than Odysseus and therefore could not have<br />

493 influenced its formation. They went on to measure photogrammetrically the raised flanks <strong>of</strong> Ithaca<br />

494 Chasma, which they attributed to flexural uplift. Assuming that these raised flanks were indeed due<br />

495 to flexural uplift, they derived a mechanical lithospheric thickness <strong>of</strong> ~20 km and surface heat<br />

fluxes <strong>of</strong> 18-30 mW/m 2 at assumed strain rates <strong>of</strong> 10 -17 to 10s -1 496<br />

497<br />

. They concluded that Ithaca Chasma<br />

is an endogenic tectonic feature but could not explain why it is confined to a narrow zone<br />

498 approximating a large circle.<br />

499 Several Voyager-era studies <strong>of</strong> Dione (Plescia, 1983; Moore, 1984) noted a near-global network <strong>of</strong><br />

500 tectonic troughs and the presence <strong>of</strong> a smooth plain crossed by both ridges and troughs located near<br />

501 the 90°W longitude and extending eastward. Cassini coverage revealed the westward extension <strong>of</strong><br />

502 this plain (Wagner et al., 2006) and demonstrated that the fractured and the cratered dark plains are<br />

503 spectrally distinct (Stephan et al., 2008). Like the eastern portion observed by Voyager, the western<br />

504 plain also exhibits both ridges and troughs (Fig. 3).<br />

505<br />

Fig. 3 (a) Image mosaic <strong>of</strong> Dione, centered on 180 o 506<br />

longitude. Red and white arrows point to Amata<br />

507 and Turnus impact structures, respectively. (b) Topography <strong>of</strong> Dione, from Schenk and Moore<br />

508 (2007). An ancient degraded impact basin centered just west <strong>of</strong> Amata can be identified by the two<br />

509 concentric rings in the DEM (white arrows). DEM has a dynamic range <strong>of</strong> 8 km (red is highest,<br />

510<br />

511<br />

magenta is lowest).<br />

512 However, the ridge system is much more pronounced. The western ridges are planimetrically<br />

513 narrower and morphologically better expressed (e.g. their relative lack <strong>of</strong> degradation). These<br />

514 western ridges also conform much more closely to the boundary between the plain and the older<br />

515 cratered rolling terrain to the west. The linear to arcuate troughs interpreted to have been created by<br />

516 tectonic extension form a complex network that divides Dione’s surface into large polygons (Smith<br />

517 et al., 1981, 1982; Plescia, 1983; Moore, 1984). The troughs <strong>of</strong>ten form parallel to sub-parallel sets,<br />

518<br />

519<br />

somewhat reminiscent <strong>of</strong> the grooved terrain on Ganymede. The global orientation <strong>of</strong> Dione’s<br />

tectonic fabric is non-random (Moore, 1984) and exhibits patterns consistent with a decline in<br />

520 oblateness due to either despinning or orbital recession (Melosh 1977, 1980).<br />

521 The topography derived from Cassini images <strong>of</strong> Dione shows one clear case <strong>of</strong> tectonic orientation<br />

522 regionally influenced by a preexisting but very degraded large impact basin centered on the smaller<br />

523 crater Amata (Schenk and Moore, 2007 (Figs. 3 and 4)).<br />

524 Fig. 4 Fracture zones on Dione. Global base map <strong>of</strong> Cassini imaging has a resolution <strong>of</strong> 400m. The<br />

525 circle represents the outer ridge corresponding to the rim <strong>of</strong> the degraded basin near Amata shown<br />

526<br />

527<br />

in Fig. 3.<br />

528 Moore et al (2004) identified another example <strong>of</strong> tectonic lineament orientation associated with an<br />

529 unambiguously identified impact feature, 90 km Turnus. Such crater-focused tectonics has been<br />

530 identified on Ganymede from Galileo data (e.g. Pappalardo and Collins, 2005). This crater-<br />

531 lineament relationship on Dione is the most pronounced among any <strong>of</strong> the middle-sized icy<br />

532 satellites except Enceladus (Spencer et al., 2009). Moore (1984) proposed a tectonic history <strong>of</strong><br />

533<br />

extension followed by (at least regional) compression, the latter being responsible for the ridges.<br />

11


534 Cassini-era images <strong>of</strong> the tectonic troughs (mostly grabens) show them to be morphologically fresh,<br />

535 implying that extension continued into geologically recent times (Wagner et al., 2006).<br />

536 After Voyager, Rhea was held to be the least endogenically evolved satellite larger than 1000 km in<br />

537 diameter. Moore et al. (1985) pointed out a few parallel lineaments and noted several large ridges or<br />

538 scarps (which they termed ‘megascarps') and speculated that these features were formed by a period<br />

539 <strong>of</strong> extension followed by compression. Cassini images, and digital terrain models derived from<br />

540 them, <strong>of</strong> the trailing hemisphere <strong>of</strong> Rhea seen from both middle northern and far southern latitudes<br />

541 clearly show that the wispy arcuate albedo markings seen in Voyager images are associated with a<br />

542 graben/extensional fault system dominantly trending north-south (Moore and Schenk, 2007;<br />

543<br />

544<br />

Wagner et al. 2007) (Fig. 5a).<br />

545 Fig. 5 (a) Cassini image <strong>of</strong> the northern trailing hemisphere <strong>of</strong> Rhea showing the bright scarps <strong>of</strong><br />

546 the northern extension <strong>of</strong> the graben system and their relationship to 'wispy' terrain. (b) Mosaic <strong>of</strong><br />

547 Cassini images showing curvilinear grabens (black arrows) and an orthogonal set <strong>of</strong> ridges and<br />

548<br />

549<br />

troughs (white arrow). Base mosaic resolution is 1 km/pixel.<br />

550 This is the first identification <strong>of</strong> unambiguous regional endogenic activity on Rhea. The digital<br />

551 terrain models <strong>of</strong> Rhea derived from Cassini images also show a set <strong>of</strong> roughly north-south trending<br />

552<br />

553<br />

ridges (Figs. 5b, and 6) that correspond to the 'megascarps' <strong>of</strong> Moore et al. (1985).<br />

554<br />

555<br />

556<br />

Fig. 6 Preliminary stereo topography map, from Schenk and Moore (2007). White arrows indicate<br />

graben systems. The DEM has a dynamic range <strong>of</strong> 10 km (red is highest, magenta is lowest).<br />

557 If these ridges are in fact due to compressional tectonics, their geometrical relation to the<br />

558 graben/scarp system may also indicate a genetic relationship. However, the poor morphological<br />

559 presentation <strong>of</strong> these ridges implies that they are old, while the 'wispy terrain' grabens are fresh and<br />

560 presumably much more recently formed.<br />

561 While the troughs and ridges suggest mainly extensional and (minor) compressional tectonics<br />

562 (Thomas, 1988), the en-echelon pattern <strong>of</strong> the scarps and troughs on the trailing hemisphere<br />

563 suggests shear stress (Wagner et al., 2007). This pattern may have been influenced by the possible<br />

564 presence <strong>of</strong> a large, degraded impact basin (basin C <strong>of</strong> Moore et al., 1985).<br />

565<br />

566<br />

The equatorial ridge <strong>of</strong> Iapetus (Fig. 7) is the most enigmatic feature on any <strong>of</strong> the medium-sized<br />

567 Fig. 7 Topography <strong>of</strong> Iapetus as derived from stereo Cassini imaging. Note the sharply defined<br />

568 equatorial ridge. The DEM at left is referenced to a sphere, the DEM at right to a biaxial ellipsoid.<br />

569<br />

570<br />

(Adapted from Giese et al., 2008)<br />

571 icy satellites (Porco et al., 2005a; Giese et al., 2008); it extends across more than half <strong>of</strong> Iapetus'<br />

572 circumference (Giese et al., 2008). Different parts show very different morphologies, from<br />

573 trapezoidal to triangular cross-sections with steep walls and a sharp rim to isolated mountains with<br />

574 moderate slopes and rounded tops (Giese et al., 2008). Although the ridge is variable in height, it<br />

575 rises up to 10 km above the local mean but also has a gently sloping flank. The ridge is abundantly<br />

576<br />

577<br />

cratered, indicating that it is comparable in age to other terrains on Iapetus (Schmedemann et al.,<br />

2008). Although the equatorial location <strong>of</strong> the ridge might be due to despinning (Castillo-Rogez et<br />

578 al., 2007), it represents such a large mass that, if it had formed in another location, it would almost<br />

579 certainly have reoriented to the equator anyway. However, as already mentioned above, irrespective<br />

580 <strong>of</strong> whether a satellite is spinning down or spinning up, equatorial tectonic features can be expected<br />

581 to be oriented north-south, making it difficult to explain the ridge with despinning. One possibility<br />

582 is that despinning stresses combined with lithospheric thickness variations due to convection may<br />

583 have resulted in equatorial extension and diking (Roberts and Nimmo, 2009; Melosh and Nimmo,<br />

584 2009). Other proposals suggest that the morphology <strong>of</strong> the ridge indicates that the surface was<br />

585 warped up by tectonic faulting (Giese et al., 2008), that it is a result <strong>of</strong> extensional forces acting<br />

586 above an ascending current <strong>of</strong> solid-state convection from a two-cell convection pattern<br />

587 (Czechowski and Leliwa-Kopystyński, 2008), or that it was formed by the impact <strong>of</strong> an ancient<br />

588 debris disk on Iapetus (Ip, 2006). The great apparent elastic thickness suggested by the survival <strong>of</strong><br />

589 the ridge may be used to place constraints on the thermal evolution <strong>of</strong> Iapetus (Castillo-Rogez et al.,<br />

590 2007).<br />

591 Although the global shape <strong>of</strong> Iapetus does suggest an earlier, more rapid spin rate (Castillo-Rogez et<br />

592<br />

al., 2007), despinning cannot explain the equatorial ridge (see above): stresses in the north-south<br />

12


593 direction are always smaller than those in the east-west direction, and will result in N-S oriented<br />

594 features (Melosh, 1977). Thus, the origin <strong>of</strong> the equatorial ridge is currently a mystery.<br />

595 The deformation <strong>of</strong> Saturnian satellites varies wildly, from currently active and heavily deformed<br />

596 Enceladus to nearby Mimas, which is tectonically uninteresting. Nor does the degree <strong>of</strong> deformation<br />

597 correlate in any straightforward way with predicted tidal stresses. However, there are at least two<br />

598 global yet somewhat contradictory aspects, which may be identified. On the one hand, coherent<br />

599 global tectonic patterns are evident on Dione, Tethys and Enceladus and to a lesser extent on Rhea<br />

600 and Iapetus. On the other, several satellites, especially Tethys and Enceladus, show very<br />

601 pronounced spatial variations in the extent <strong>of</strong> deformation. These two aspects suggest that global<br />

602 sources <strong>of</strong> stress are common, but also that lateral variations in the mechanical properties <strong>of</strong> the<br />

603 satellites’ ice shells play an important role.<br />

604 Regarding this second point, localization <strong>of</strong> deformation (e.g. at the south pole <strong>of</strong> Enceladus or at<br />

605 Ithaca Chasma) may arise naturally in ice shells (e.g. Sotin et al., 2002; Nimmo, 2004b) but is<br />

606 currently poorly understood and makes modeling tectonic deformation much harder. As far as<br />

607 global deformation patterns are concerned, several satellites (Enceladus, Iapetus, perhaps Dione)<br />

608 show patterns with simple symmetries about the rotational and tidal axes. This symmetry may be<br />

609 due either to feature reorientation or the operation <strong>of</strong> mechanisms (tides, despinning etc.), which<br />

610 have a natural symmetry. On Dione, Rhea and Iapetus at least, diurnal tidal stresses are rather small<br />

611 (Tab. 3), and most <strong>of</strong> the tectonic features are apparently ancient, which suggests that despinning,<br />

612 volume changes or reorientation are more likely to be responsible for the observed deformation than<br />

613<br />

614<br />

present-day tidal stresses.<br />

As in the Galilean satellites, extensional features are generally more common than compressional<br />

615 features on the other Saturnian satellites. This observation is readily explained if we assume that the<br />

616 satellites once contained subsurface oceans which progressively froze (e.g. Smith et al., 1981;<br />

617 Nimmo, 2004a); alternative explanations are that compressional features such as folds (Prockter and<br />

618 Pappalardo, 2000) are less easy to recognize than extensional features, or that larger stresses are<br />

619<br />

620<br />

required to form compressional features (see Pappalardo and Davis, 2007).<br />

621 Cryovolcanism<br />

622 Cryovolcanism can be defined as 'the eruption <strong>of</strong> liquid or vapor phases (with or without entrained<br />

623 solids) <strong>of</strong> water or other volatiles that would be frozen solid at the normal temperature <strong>of</strong> the icy<br />

624 satellite’s surface' (Geissler, 2000). Cryovolcanism therefore includes effusive eruptions <strong>of</strong> icy<br />

625 slurries and explosive eruptions consisting primarily <strong>of</strong> vapor (Fagents, 2003). Voyager-era data<br />

626 suggested that such activity might be common in the outer solar system, resulting in a flurry <strong>of</strong><br />

627 theoretical papers (e.g. Stevenson, 1982; Allison and Clifford, 1987; Crawford and Stevenson,<br />

628 1988; Kargel. 1991; Kargel et al., 1991). Higher-resolution images from Galileo generally revealed<br />

629 tectonic rather than cryovolcanic resurfacing processes, although some features may be due to<br />

630 cryovolcanic activity (e.g. Schenk et al., 2001; Fagents et al., 2000; Miyamoto et al., 2005). In the<br />

631 following, we discuss the current state <strong>of</strong> knowledge about cryovolcanism in the Saturnian<br />

632 satellites. Enceladus, which is definitely cryovolcanically active (Porco et al., 2006), and Titan,<br />

633 which may be (Lopes et al., 2007; Nelson et al., 2009a,b), are discussed by Spencer et al. (2009),<br />

634 Jaumann et al. (2009) and Sotin et al. (2009).<br />

635<br />

636<br />

The principal difference between cryovolcanism and silicate volcanism is that in the former, the<br />

melt is denser than the solid (by ~10%). This means that special circumstances must exist for an<br />

637 eruption <strong>of</strong> non-vapor cryovolcanic materials to occur (though <strong>of</strong> course on <strong>Earth</strong> it is not<br />

638 uncommon for denser basaltic lavas to erupt through lighter granitic material). Such circumstances<br />

639 include one or more <strong>of</strong> the following: (1) the melt includes dissolved volatiles, which exsolve and<br />

640 reduce the bulk melt density (Crawford and Stevenson, 1988); (2) the melt is driven upwards by<br />

641 pressure differences caused by tidal effects (Greenberg et al., 1998), surface topography (Showman<br />

642 et al., 2004) or freezing (Fagents, 2003; Manga and Wang, 2007); (3) compositional differences<br />

643 render the ice shell locally dense or the melt less dense (e.g. by the addition <strong>of</strong> ammonia) (Kargel<br />

644 1992, 1995). The latter mechanism can be nothing more than a local effect, since if the bulk ice<br />

645 shell is denser than the underlying material; it will sink if the ice shell is disrupted.<br />

646 Erupted material consisting primarily <strong>of</strong> water plus solid ice will simultaneously freeze and boil, the<br />

647 latter process ceasing once a sufficiently thick ice cover is formed (Allison and Clifford, 1987). The<br />

648 resulting flow velocity will depend on the viscosity <strong>of</strong> the ice-water slurry (Kargel et al., 1991) as<br />

649 well as on flow thickness and local slope (Wilson et al., 1997; Miyamoto et al., 2005). If the erupted<br />

650 material is dominated by the vapor phase, the vapor plume will expand outwards (Tian et al., 2007)<br />

651<br />

and escape entirely if the molecular thermal velocities exceed the escape velocity. Any associated<br />

13


652 solid material will follow ballistic trajectories, ultimately forming deposits centered on the eruption<br />

653 vents (Fagents et al., 2000).<br />

654 The plume properties measured on Enceladus by Cassini imply considerably smaller velocities for<br />

655 icy dust grains than for vapor. On hypothesis is that the dust grains condense and grow in vertical<br />

656 vents <strong>of</strong> variable width. Repeated wall collisions and re-acceleration by the gas in the vents lead to<br />

657 effective friction, which explains the observed plume properties (Schmidt et al., 2008). ). Most dust<br />

658 grains are growing by direct condensation <strong>of</strong> the ascending water vapor within the ice channels and<br />

659 contain almost pure water ice; they are free <strong>of</strong> atomic sodium. In 2008 the CDA instrument while<br />

660 analyzing the composition <strong>of</strong> E-ring particles discovered a second type <strong>of</strong> icy dust particles. These<br />

661 grains are rich in sodium salts (0.5 - 2% in mass) and there origin is explained by the existence <strong>of</strong><br />

662 an ocean <strong>of</strong> liquid water below the surface <strong>of</strong> Enceladus and which is in direct contact with its rock<br />

663 core (Postberg et al., 2009). This shows that cryovolcanism can provide valuable information about<br />

664 geological conditions <strong>of</strong> the interior.<br />

665 The plains <strong>of</strong> Dione are another candidate for cryovolcanic resurfacing (Fig. 8). Several Voyager-<br />

666<br />

667<br />

era studies concluded that these plains were<br />

668 Fig. 8 Cassini view <strong>of</strong> smooth plains on Dione. Older heavily cratered terrains lie to the west. Note<br />

669 the prominent north-south ridge and the subtle linear features to the east. Mosaic resolution is<br />

670<br />

671<br />

400m/pixel.<br />

672<br />

673<br />

formed in this way (e.g. Plescia and Boyce, 1982; Plescia, 1983; Moore, 1984). Moore (1984)<br />

suggested that several branching broad-rimmed troughs north <strong>of</strong> the equator at 60°W may have<br />

674 been the source vents. Surprisingly, the plains unit may be topographically higher than the cratered<br />

675 terrain (Moore and Schenk, 2007). This observation seems a little difficult to reconcile with the<br />

676 plains being emplaced as a low-viscosity cryovolcanic flow. One possibility is that the plains<br />

677 material was emplaced as a deformable plastic unit (like terrestrial glaciers), but this, or any other,<br />

678 explanation needs both further evidence and thought. The ridges discussed above bound (or at least<br />

679 occur near the edges) the plains unit in the west and southeast and may be related to their formation.<br />

680 A plains unit on Tethys is located antipodal to the very large Odysseus impact feature. Moore and<br />

681 Ahern (1983) proposed that the plains were formed by cryovolcanism. Alternatively, Moore et al.<br />

682 (2004) considered the possibility that it was extreme seismic shaking caused by energy focused<br />

683 there by the Odysseus impact that formed the antipodal plains. If seismic shaking was responsible,<br />

684 there would probably be a significant transition zone between that unit and the adjacent hilly and<br />

685 cratered terrain. Digital terrain models derived from Cassini images indicate that the transition<br />

686 between the cratered terrain and the plains is abrupt and the plains themselves are level, supporting<br />

687 the hypothesis that they are endogenic (Moore and Schenk, 2007).<br />

688 Both Tethys and Dione show evidence for weak plasma tori (Burch et al., 2007) somewhat<br />

689 analogous to the much more marked feature at Enceladus, which is caused by geysers (Porco et al.,<br />

690 2005). It is not yet clear whether surface activity or some other process is generating these tori.<br />

691 Similarly, Clark et al. (2008a) report a possible, but currently unconfirmed, detection <strong>of</strong> material<br />

692 around Dione, which may also indicate some surface activity. The Cassini magnetometer has also<br />

693 measured small amounts <strong>of</strong> mass loading near Dione (Burch et al., 2007).<br />

694<br />

695<br />

So far, none <strong>of</strong> the digital terrain models <strong>of</strong> Rhea developed in the Cassini era shows any regions<br />

that would qualify as plains, supporting the perception that Rhea shows no record <strong>of</strong> cryovolcanic<br />

696 resurfacing. However, unlike Dione, the diffuse nature <strong>of</strong> the wisps (the 'wispy terrain') on Rhea has<br />

697 yet to be explained as numerous smaller bright scarps associated with fracture and faulting, and may<br />

698 yet be shown to be an expression <strong>of</strong> cryo-pyroclastic mantling (Stevenson, 1982). Furthermore,<br />

699 Rhea, surprisingly, exhibits evidence for faint particle ring (Jones et al., 2008), the origin <strong>of</strong> which<br />

700 is unclear but may be related to endogenic or impact processes. On the other hand, Cassini VIMS<br />

701 observations found no evidence <strong>of</strong> a water vapor column around Rhea (Pitman et al., 2008). On<br />

702 Iapetus, no evidence <strong>of</strong> cryovolcanism has been detected thus far. If volcanism ever operated on<br />

703<br />

704<br />

these satellites, it occurred in ancient times and its evidence is now unseen.<br />

705 21.4 Composition and Alteration <strong>of</strong> Surface Materials<br />

706 The composition <strong>of</strong> the satellites <strong>of</strong> Saturn is determined by remote spectroscopy, by sampling<br />

707 materials that were sputtered from the satellite surfaces or, in the case <strong>of</strong> Enceladus, injected into space<br />

708 where they can be directly sampled by spacecraft instruments.<br />

709 It has long been known that the surfaces <strong>of</strong> Saturn’s major satellites, Mimas, Enceladus, Tethys,<br />

710<br />

Dione, Rhea, Hyperion, Iapetus and Phoebe, are predominantly icy objects (e.g. Fink and Larsen,<br />

14


711 1975; Clark et al., 1984, 1986; Roush et al., 1995; Cruikshank et al., 1998a; Owen et al., 2001;<br />

712 Cruikshank et al., 2005; Clark et al., 2005, 2008a; Filacchione et al., 2007, 2008). While the<br />

713 reflectance spectra <strong>of</strong> these objects in the visible range indicate that a coloring agent is present on<br />

714 all surfaces except Mimas, Enceladus and Tethys (Buratti 1984). Only Phoebe and the dark<br />

715 hemisphere <strong>of</strong> Iapetus display near-IR spectra markedly different from very pure water ice (e.g.<br />

716 Cruikshank et al., 2005; Clark et al., 2005).<br />

717 The major absorptions in icy satellite spectra shown in Figs 9 a-e are due to water ice and can be<br />

718<br />

719<br />

observed at 1.5µm., 2.0µm., 3.0µm., and 4.5µm.<br />

720 Fig. 9 VIMS spectra <strong>of</strong> the Saturnian icy satellites: (a) The major absorptions in icy satellite spectra<br />

721 are due to water ice and appear at 1.5, 2.0, 3.0, and 4.5µm; (b) Representative (A) bright and (B)<br />

722 dark region Dione spectra (the gaps in the bright region spectra (A) near 1µm are due to sensor<br />

723<br />

724<br />

725<br />

saturation); the inset shows an area <strong>of</strong> high concentration <strong>of</strong> CO2 (after Clark et al., 2008a); (c)<br />

Spectra <strong>of</strong> Hyperion showing an increase in blue reflectivity and absorptions <strong>of</strong> water ice and CO2<br />

(after Clark et al., 2008a); (d) Spectra <strong>of</strong> Iapetus showing a range <strong>of</strong> blue peak intensities,<br />

726<br />

727<br />

absorptions <strong>of</strong> different intensities by water ice and CO2 absorption bands (after Clark et al.,<br />

2008a); (e) Spectra <strong>of</strong> Iapetus showing a range <strong>of</strong> blue peak intensities and absorptions <strong>of</strong> different<br />

728<br />

729<br />

intensities by water ice and CO2 absorption bands (after Clark et al., 2008a). However, no CO2 was<br />

found on the small satellites Atlas, Pandora, Janus, Epimetheus, Telesto, and Calypso (Buratti et<br />

730<br />

731<br />

al., 2009a).<br />

732<br />

733 Weaker ice absorptions appear at 1.04µm and 1.25µm in the spectra <strong>of</strong> bright regions where the ice<br />

734<br />

735<br />

is purer. Trapped CO2 was first detected in Galileo near-infrared mapping spectrometer (NIMS)<br />

spectra <strong>of</strong> the Galilean satellites (Carlson et al., 1996) and in the Saturnian system on Iapetus<br />

736 (Buratti et al, 2005), Phoebe (Clark et al., 2005), Hyperion (Cruikshank et al., 2007), and Enceladus<br />

737<br />

738<br />

(Brown et al., 2006). Weak CO2 absorptions were also detected in VIMS data on Dione, Tethys,<br />

Mimas and Rhea (Clark et al., 2008a).<br />

739 Organic compounds were detected directly in the Enceladus plume by the INMS (Waite et al.,<br />

740 2006) and in trace amounts by the VIMS on Phoebe (Clark et al., 2005), Hyperion (Cruikshank et<br />

741 al., 2007) and Iapetus (Cruikshank et al., 2008), and possibly Enceladus (Brown et al., 2006). While<br />

742 the spectral signatures indicate aromatic hydrocarbon molecules, exact identifications have remained<br />

743 elusive, partly because the detections have low signal-to-noise ratios. Another spectral feature<br />

744 observed in VIMS data <strong>of</strong> dark material on Phoebe, Iapetus and Dione appears to be an absorption<br />

745 centered near 2.97µm (Fig. 9 d). The detection <strong>of</strong> this feature is complicated by the fact that VIMS<br />

746 has an order-sorting filter change at this wavelength. Clark et al. (2005) presented evidence that this<br />

747 feature is real and maps to geological patterns on Dione and Iapetus (Clark et al., 2008a, 2009).<br />

748 However, a better understanding <strong>of</strong> the instrument response is needed to increase confidence in the<br />

749 identification, including confirmation by a different instrument. If correct, the feature strength<br />

750 indicates approximately 1% ammonia in the dark material.<br />

751 The position shift <strong>of</strong> amorphous ice absorptions from their crystalline counterparts (Mastrapa et<br />

752 al., 2008 and references therein) can be used to determine whether amorphous or crystalline ice is<br />

753<br />

754<br />

present on a surface. Traditionally, amorphous ice displays almost no 1.65µm feature and the<br />

3.1µm Fresnel peak shifts, broadens and weakens. Newman et al. (2008) used variations in the<br />

755 3.1µm peak and the 1.65µm absorption to argue for amorphous ice around the 'tiger stripes' on<br />

756 Enceladus. However, Clark et al. (2009) showed that scattering and sub-micron grain sizes also<br />

757 influence the intensities <strong>of</strong> the 3.1-µm peak and the 1.65-µm absorption, and can be confused with<br />

758 temperature and crystallinity effects. Further examination <strong>of</strong> the ice feature positions in spectra <strong>of</strong><br />

759 Enceladus and other satellites in the Saturnian system showed that positions and strengths are<br />

760 consistent with those <strong>of</strong> fine-grained crystalline ice (Clark et al., 2009). Newman et al. (2009)<br />

761 showed that water ice on Dione exists exclusively in crystalline form.<br />

762 The spectra <strong>of</strong> the satellites in the visible wavelength range are smooth with no sharp spectral features,<br />

763 but all show an ultraviolet absorber affecting the blue end <strong>of</strong> the visible-wavelength spectral slope.<br />

764 Visible spectra can be classified by their spectral slope: from bluish Enceladus and Phoebe to redder<br />

765 Iapetus, Hyperion and Epimetheus. In the 1µm to1.3µm range, the spectra <strong>of</strong> Enceladus, Tethys,<br />

766 Mimas and Rhea are characterized by negative slope, consistent with a surface mainly dominated by<br />

767 water ice, while the spectra <strong>of</strong> Iapetus, Hyperion and Phoebe show considerable reddening, pointing<br />

768 out the relevant role played by the darkening materials present on the surface (Filacchione et al., 2007,<br />

769<br />

2008). In between these two classes are Dione and Epimetheus, which have a relatively flat global<br />

15


770 spectrum in this range. At wavelengths less than about 0.5µm, all satellite spectra display a UV<br />

771 absorber, which might be due to charge transfer in oxides such as nano-phase hematite (Clark et al.,<br />

772 2008a), with the apparent strength <strong>of</strong> the absorber generally indicating the amount <strong>of</strong> contaminant.<br />

773 Trends in the UV absorber are complicated by ice dust contributions to satellite surfaces transmitted<br />

774 from Enceladus via the E-ring. Satellites close to the rings, such as Janus, Atlas and Prometheus, show<br />

775 an intermediate red tint between that <strong>of</strong> the rings and Enceladus. Mimas and Tethys show a red tint<br />

776 between that <strong>of</strong> Janus, Atlas and Prometheus and that <strong>of</strong> Enceladus on the other (e.g. Cuzzi et al.,<br />

777 2009; Buratti et al., 2009a).<br />

778 Disk-integrated ultraviolet observations by the Cassini UVIS (Esposito et al., 2004; Hendrix and<br />

779<br />

780<br />

Burrati, 2009) (Fig. 10) show Tethys and Dione exhibit marked UV<br />

781 Fig. 10 Cassini UVIS disk-integrated spectra <strong>of</strong> the icy moons, all at phase angles between 10° and<br />

782<br />

783<br />

13.6°.<br />

784 leading-trailing differences whereas Mimas, Enceladus and Rhea do not, suggesting compositional<br />

785 variations among the satellites as well as spectrally active components in the ultraviolet. UVIS<br />

786 spectra <strong>of</strong> all the icy moons <strong>of</strong> Saturn are dominated by a pronounced water-ice absorption edge at<br />

787 ~0.165µm. Its strength varies with the abundance <strong>of</strong> water ice, and its wavelength depends on grain<br />

788 size. Nearly all UVIS spectra display a small absorption feature at ~0.185µm (Hendrix and Hansen,<br />

789 2008b) whose origin is not clear at present. It is possible that this feature is due to water ice as it is<br />

790<br />

791<br />

seen in published reflectance spectra <strong>of</strong> water ice (Pipes et al., 1974; Hapke et al., 1981). Published<br />

optical constants (Warren, 1982) do not exhibit the feature, but the constants have not been well<br />

792 measured in the ~0.17µm to 0.4µm range.<br />

793 Cassini spectral coverage <strong>of</strong> the icy moons is complete in the 0.4µm to 5.2µm and in the 0.1µm to<br />

794 0.19µm region, but there is a gap in the 0.19µm to 0.4µm region (except for the broad band spectral<br />

795 measurements <strong>of</strong> ISS). This is an important wavelength region for non-water ice absorptions, and the<br />

796 Cassini UVIS data in the far UV provide a critical tiepoint between the far UV and the visible band that<br />

797 is helpful in understanding the nature <strong>of</strong> the non-ice species present in the surfaces <strong>of</strong> the moons. The<br />

798 icy moons are dark in the far UV compared to their spectra in the Vis-NIR. For instance, the disk-<br />

799 integrated albedo <strong>of</strong> Enceladus at 0.19µm is ~0.4, while at 0.439µm (Verbiscer et al., 2005) it is ~1.2.<br />

800 Water ice is thought to have a relatively flat spectral signature in this wavelength range, which suggests<br />

801 that some other species darkens Enceladus in the FUV. Ammonia, which exhibits an abrupt absorption<br />

802 edge at 0.2µm (Pipes et al., 1974), is a candidate (and ammonia hydrate has been tentatively detected at<br />

803 Enceladus and Tethys in <strong>Earth</strong>-based spectra (Verbiscer et al., 2005; Verbiscer et al., 2008)), though at<br />

804<br />

805<br />

this time other species cannot be ruled out other materials.<br />

806 Dark Material<br />

807 Cassini (1672) was the first to infer the presence <strong>of</strong> dark material in the Saturnian system, which<br />

808 was verified by Murphy et al. (1972) and Zellner (1972). Dark material appears throughout the<br />

809 system, most predominantly on Dione, Rhea, Hyperion, Iapetus and Phoebe. The nature <strong>of</strong> the dark<br />

810 material especially on Iapetus has been studied by numerous authors, <strong>of</strong>ten with conflicting<br />

811 conclusions, including Cruikshank et al. (1983), Wilson and Sagan (1995, 1996) Vilas et al. (1996),<br />

812<br />

813<br />

Jarvis et al. (2000), Owen et al. (2001), Buratti et al. (2002) and Vilas et al. (2004).<br />

The dark material tends to darken and redden the satellites’ spectra, particularly in the visible. In<br />

814 dark-region spectra, additional absorptions are seen at 2.42µm and 4.26µm, with the latter due to<br />

815<br />

816<br />

CO2, as well as stronger water ice absorption near 3µm (Fig. 21.5.1d). The 2.42µm feature was first<br />

tentatively identified as a CN overtone band by Clark et al. (2005) in spectra <strong>of</strong> Phoebe, but<br />

817 evidence <strong>of</strong> CN fundamentals in the 4µm to 5µm region has not been confirmed. At present, the<br />

818 2.42µm feature is believed to be due to OH or hydrogen (Clark et al., 2008a, b, 2009), but<br />

819 complications with the VIMS calibration have limited any detailed definition <strong>of</strong> the absorption.<br />

820 Diverse spectral features <strong>of</strong> the dark material on Dione match those seen on Phoebe, Iapetus, Hyperion<br />

821 and Epimetheus as well as in the F-ring and the Cassini Division, implying that its composition is the<br />

822 same throughout the Saturnian system (Clark et al., 2008a, 2009). Models <strong>of</strong> the dark material on<br />

823 Iapetus have used tholins to explain the reddish visible spectral slope (e.g. Cruikshank et al., 2005).<br />

824 However, VIMS data, which better isolate the spectral signature <strong>of</strong> the dark material from that <strong>of</strong> icy<br />

825 regions thanks to the high spatial resolution afforded by close fly-bys, show that the dark material has<br />

826 a remarkably linear spectrum (Clark et al., 2008a, 2009) (Fig. 12) not explained by tholins which have<br />

827 curved spectral response. Tholins also have strong CH absorbers in the 3-4-μm region as well as other<br />

828<br />

marked absorptions at wavelengths longer than 2 μm that are not seen in, for example, spectra <strong>of</strong> the<br />

16


829 dark regions <strong>of</strong> Iapetus. The problem <strong>of</strong> matching one spectrum might be accomplished with a<br />

830 specific mixture <strong>of</strong> many compounds, but with spatially resolved spectra on the satellites we now<br />

831 see a large range <strong>of</strong> mixtures, so models must show consistency across that range <strong>of</strong> observed<br />

832 mixtures, and many models do not.<br />

833 The source <strong>of</strong> the dark material in the Saturnian system is unclear. Several researchers (e.g. Clark et<br />

834 al., 2005) suggest a source outside the system, perhaps the Kuiper belt or comets. Owen et al.<br />

835 (2001) suggested that an impact on Titan created a spray <strong>of</strong> dark, organic-rich material, part <strong>of</strong><br />

836 which was swept up by Iapetus’ leading hemisphere, another part being transported to the inner<br />

837 system to accrete onto the surfaces <strong>of</strong> the inner icy moons. The Dione VIMS spectra provide a key to<br />

838 solving some <strong>of</strong> the mysteries <strong>of</strong> the dark material. Clark et al. (2008a) showed that a pattern <strong>of</strong><br />

839 bombardment by fine particles below 0.5 μm impacted Dione from the trailing-side direction. Several<br />

840 lines <strong>of</strong> evidence point to an external origin <strong>of</strong> the dark material on Dione, including its global spatial<br />

841 pattern, local patterns including crater and cliff walls shielding implantations on slopes facing away<br />

842 from the trailing side, exposing clean ice, and slopes facing the trailing direction that show higher<br />

843 abundances <strong>of</strong> dark material.<br />

844 A blue scattering peaks is seen in the spectrum <strong>of</strong> the dark material and was initially observed in the<br />

845 spectra <strong>of</strong> Tethys, Dione and Rhea by Noland et al. (1974). Clark et al. (2008a) suggest that the<br />

846 blue peak is caused by particles less than 0.5μm in diameter embedded in the ice. The VIMS resolved<br />

847 areas on Phoebe, Iapetus and Hyperion also show blue peaks (Fig. 9 c, d, e), indicating that this is a<br />

848 common property <strong>of</strong> satellites in the Saturnian system (Clark et al., 2008a, 2009). The blue scattering<br />

849<br />

850<br />

851<br />

peak with a strong UV/visible absorption was modelled with ice plus 0.2 μm diameter carbon grains<br />

(to provide the Rayleigh scattering effect) plus nano-phase hematite (Fe2O3) to produce the UV<br />

absorber (Fig. 11). But carbon is inconsistent with the linear-red slope seen in the purest dark material<br />

852 spectra from 0.4µm to 2.5µm. Metallic iron is the only known compound to display this spectral<br />

853 property and nano-phase metallic iron plus nano-phase hemtatite mixed with ice show consistency<br />

854<br />

855<br />

with the mixture spectra observed on the icy satellites and in Saturn’s rings (Clark et al., 2009).<br />

856 Fig. 11 Spectrum <strong>of</strong> Dione compared to a spectrum <strong>of</strong> silicon carbide. (b) Spectrum <strong>of</strong> Dione<br />

857 compared to the spectrum <strong>of</strong> a laboratory mixture <strong>of</strong> 99.25wt% H2O ice, 0.5wt% carbon black, and<br />

858 0.25wt% nano-crystalline hematite (reagent by Sigma-Aldrich). The carbon black grains are 0.2µm<br />

859 in diameter. Both the carbon black and hematite grains create Rayleigh scattering in the sample, but<br />

860 at wavelengths shorter than about 0.5µm, absorption by hematite causes a downturn similar to that<br />

861<br />

862<br />

seen in the Dione spectra (after Clark et al., 2008a).<br />

863 Previous models <strong>of</strong> the dark material on Iapetus typically used tholins (e.g. Cruikshank et al., 2005).<br />

864 However, VIMS data, which better isolate the spectral signature <strong>of</strong> the dark material from that <strong>of</strong> icy<br />

865 regions thanks to the high spatial resolution afforded by close fly-bys, show that the dark material has<br />

866<br />

867<br />

a remarkably linear spectrum (Clark et al., 2008a, 2009) (Fig. 12).<br />

868 Fig. 12 Spectrum <strong>of</strong> the dark material <strong>of</strong> Iapetus compared to laboratory spectra <strong>of</strong> mixtures <strong>of</strong> iron<br />

869<br />

870<br />

powder, H2O, CO2 and NH3. The small Rayleigh scattering peak and UV absorber can be explained<br />

by traces <strong>of</strong> iron oxide nano-powder. The 3-micron absorption is best explained by a combination <strong>of</strong><br />

871<br />

872<br />

873<br />

bound water, ice and trace ammonia. Finer-grained iron would require higher abundances <strong>of</strong> water,<br />

ammonia and carbon dioxide to match the spectrum <strong>of</strong> Iapetus.<br />

874 Tholins have strong CH absorbers in the 3-4-μm region as well as other marked absorptions at<br />

875 wavelengths longer than 2μm that are not seen in, for example, spectra <strong>of</strong> the dark regions <strong>of</strong> Iapetus.<br />

876 Using VIMS data, Clark et al. (2008a) found the general composition <strong>of</strong> the dark material in the<br />

877<br />

878<br />

Saturnian system likely contains bound water, CO2 and, tentatively, ammonia. More recently, Clark et<br />

al. (2008b, 2009) argued that the signature <strong>of</strong> iron identifies the major coloring agent on the icy<br />

879 surfaces in the Saturnian system. Metallic iron does not exhibit sharp diagnostic spectral features,<br />

880 but it does have unique spectral properties in that its absorption increases linearly with decreasing<br />

881 wavelength (Clark et al., 2009).<br />

882 The theory <strong>of</strong> Clark et al. (2008b, 2009) envisages small-particle meteoroid dust contaminating the<br />

883 system as well as meteoroids with a high metallic-iron content. As one moves closer to the rings,<br />

884 the dark material diminishes (e.g. Mimas and Enceladus have little dark material). However, the<br />

885 competing process <strong>of</strong> accretion by Saturn’s E-ring, which originates on Enceladus, also contributes<br />

886 to the greater abundance <strong>of</strong> high-albedo material on Mimas and Enceladus (Buratti et al. 1990,<br />

887<br />

Vebiscer et al., 2007), The main rings do not show any strong signature <strong>of</strong> metallic iron particles,<br />

17


888 but the Cassini Division displays spectra that closely match those <strong>of</strong> Iapetus in Iapetus’ transition<br />

889 zone from dark to light material, implying mixtures <strong>of</strong> the same materials. The ring spectra are<br />

890 much redder in the 0.3µm to 0.6µm region than the spectra <strong>of</strong> dark material on Phoebe, Iapetus and<br />

891 Hyperion. Cosmic rays impacting the huge surface area <strong>of</strong> the rings break up water molecules, thus<br />

892 creating an atmosphere <strong>of</strong> hydrogen and oxygen. The hydrogen escapes, leaving an atmosphere <strong>of</strong><br />

893 oxygen around the rings (Johnson et al., 2006). Oxygen ions are very reactive, and when they<br />

894<br />

895<br />

encounter iron particles, they oxidize them and turn them into nano-phase hematite (Fe2O3). The<br />

spectra <strong>of</strong> nano-phase hematite mixed with ice and dark particles closely match the UV red slope<br />

896 seen in the spectra <strong>of</strong> Dione (Clark et al., 2008a) (Fig. 11). To explain the dark material in the<br />

897 Saturnian system, Clark et al. (2008b, 2009) suggest that oxidized sub-micron iron particles from<br />

898 meteorites, mixed with ice, might explain all the colors and spectral shapes observed in the system.<br />

899 A small amount <strong>of</strong> carbon might also be identified in the mixture, which would change the slope <strong>of</strong><br />

900 iron and further lower the albedo. Sub-micron particles also explain the origin <strong>of</strong> the observed<br />

901<br />

902<br />

Rayleigh scattering. Trace amounts <strong>of</strong> CO2, ammonia and organics would not significantly<br />

influence the visible and UV spectra down to 0.3µm. Finally, the inner satellites may be exposed to<br />

903 as much iron as the outer satellites, but E-ring ice particles mix with and cover up the iron particles.<br />

904 Species found by INMS, CAPS and MIMI in the magnetosphere are candidates for surface<br />

905 components as the magnetospheric material originates either directly from satellite surfaces or from<br />

906 ring particles. Compositional analyses <strong>of</strong> Enceladus’ plume are described in Spencer et al. (2009).<br />

907 The idea <strong>of</strong> iron in the Saturn system is supported by the Cassini CDA instrument, which has<br />

908<br />

909<br />

910<br />

discovered small (


947 late 2004 suggest external material emplacement (e.g. dark material on ram-facing crater walls at<br />

948 high latitudes) (Porco et al., 2005a), but this was later found to be due to photometric effects (Denk<br />

949 et al., 2008). The initial theory <strong>of</strong> an exogenically-created dark pattern (Cook and Franklin, 1970)<br />

950 suggested that pre-existing dark material was uncovered by meteoritic bombardment; this idea was<br />

951 extrapolated by Wilson and Sagan (1996). Researchers theorized (Soter, 1974) that the dark<br />

952 material is exogenically emplaced on Iapetus’ leading hemisphere as material is lost from the moon<br />

953 Phoebe (Burns, 1979). Retrograde Phoebe dust from a distance <strong>of</strong> 215 Saturn radii would travel<br />

954 inward from the effects <strong>of</strong> Poynting-Robertson drag and impact the leading hemisphere <strong>of</strong> Iapetus<br />

955 orbiting at 59 Saturn radii. However, Phoebe is spectrally neutral at visible wavelengths, while the<br />

956 Iapetus dark material is reddish (Cruikshank et al., 1983; Squyres et al., 1984). If the material does<br />

957 come from Phoebe, then some sort <strong>of</strong> chemistry or impact volatilization must occur to change the<br />

958 color and darken the material (Cruikshank et al., 1983; Buratti and Mosher, 1995). Another<br />

959 possibility is that the exogenic source <strong>of</strong> the dark material is Hyperion (Matthews, 1992; Marchi et<br />

960 al., 2002), Titan (Wilson and Sagan, 1995; Owen, 2001), Iapetus itself (Tabak and Young, 1989) or<br />

961 the debris remnants <strong>of</strong> a collision between a former outer moon and a planetocentric object (Denk<br />

962 and Neukum, 2000). Matthews (1992) suggested that the hypothetical impact that disrupted<br />

963 Hyperion created a debris cloud that subsequently hit Iapetus. Both Hyperion and Titan dark<br />

964 materials are spectrally reddish (Thomas and Veverka, 1985; McDonald et al., 1994; Owen et al.,<br />

965 2001), though not as dark as those <strong>of</strong> Iapetus. Buratti et al. (2002; 2005) suggested that both<br />

966 Hyperion's and Iapetus’ leading hemispheres are impacted by dark, reddish dust from recently<br />

967<br />

968<br />

discovered retrograde satellites exterior to Phoebe. Clark et al. (2008a) <strong>of</strong>fer a solution for the color<br />

discrepancy. Red, fine-grained dust (less than one-half micron in diameter) when mixed in small<br />

969 amounts with water ice produces a Rayleigh scattering effect, enhancing the blue response. This<br />

970 blue enhancement appears just enough on Phoebe to create an approximately neutral spectral<br />

971 response. An examination <strong>of</strong> the spectral signatures <strong>of</strong> Dione, Hyperion, Iapetus and Phoebe by<br />

972 Clark et al (2008a) showed a continuous mixing space on all these satellites, demonstrating variable<br />

973 mixtures <strong>of</strong> fine-grained dark material in the ice. Thus, the colors observed are simply variable<br />

974 amounts <strong>of</strong> dark material, with the redder slopes indicating greater abundance.<br />

975 Ground-based radar observations at 13cm (Black et al., 2004) and Cassini RADAR data at 2.2cm<br />

976 (Ostro et al., 2006) indicate that the dark terrain on Iapetus must be quite thin (one to several<br />

977 decimeters); an ammonia-water ice mixture may be present at a depth <strong>of</strong> several decimeters below<br />

978 the surface on both the leading and trailing hemispheres <strong>of</strong> Iapetus. The RADAR results appear to<br />

979 rule out any theories <strong>of</strong> a thick dark material layer (Matthews, 1992; Wilson and Sagan, 1996) and<br />

980<br />

981<br />

are consistent with the spectral coating indicated by VIMS data as described by Clark et al (2008a).<br />

3<br />

Cassini radio science results (Rappaport et al., 2005) indicate a bulk density for Iapetus <strong>of</strong> 1.1g/cm ,<br />

982 from which it can be inferred that the moon is composed primarily <strong>of</strong> water ice.<br />

983 Because the large craters within the dark terrain appear evenly colored by the dark material in<br />

984 Voyager data, no craters significantly break up the dark material to expose bright underlying terrain<br />

985 (Denk et al., 2000). This suggests that the emplacement <strong>of</strong> dark material by whatever mechanism is<br />

986 relatively new or ongoing.<br />

987 With the arrival <strong>of</strong> Cassini at Saturn and the first Iapetus flyby (Dec 2004), the idea <strong>of</strong> thermal<br />

988 segregation developed (Spencer et al., 2005; Hendrix and Hansen, 2008a). This idea was originally<br />

989<br />

990<br />

mentioned as one option by Mendis and Axford (1974). The September 2007 close flyby <strong>of</strong> Iapetus<br />

permitted testing <strong>of</strong> this hypothesis, resulting in the following theory. A small amount <strong>of</strong> exogenic<br />

991 dust from an unknown source created a 'global color dichotomy': the leading side became slightly<br />

992 redder and darker than the trailing with fuzzy boundaries close to the sub-Saturn and anti-Saturn<br />

993 meridians (Denk et al., 2008). A runaway thermal segregation process might have been initiated<br />

994 primarily at the low and middle latitudes <strong>of</strong> the leading side as well as in some local areas at low<br />

995 latitudes on the trailing side, especially on equatorward-facing slopes (Denk et al., 2008). Indeed,<br />

996 poleward-facing slopes even within the dark terrain (at middle latitudes) are bright. On these slopes<br />

997 the temperature never raises enough to allow the thermal process to act faster than the destruction<br />

998 process by micrometeorite gardening, which transports bright sub-surface material to the surface. In<br />

999 those locations that are now dark, the originally bright surface material becomes warm enough for<br />

1000 volatiles (mainly water ice) on the top part <strong>of</strong> the surface to grow unstable and migrate towards cold<br />

1001 traps, leaving behind residual dark material that was originally intimately mixed with the water ice<br />

1002 as a minor constituent. The slow 79.3day rotation is a crucial boundary condition for this process to<br />

1003 work in the Saturnian system. ISS images suggest that significant cold traps might be located at<br />

1004<br />

1005<br />

high latitudes on the trailing side, where lower-resolution images show bright white polar caps.<br />

19


1006 Compositional data from the UVIS and the VIMS (Hendrix and Hansen, 2009a; Hendrix and<br />

1007 Hansen, 2009b; Clark et al., 2009) indicate variable amounts <strong>of</strong> ice and volatiles in the dark<br />

1008 material. The water-ice absorption edge in the UV at 0.165µm was shown to increase with latitude,<br />

1009 consistent with decreasing temperatures and increasing numbers <strong>of</strong> visibly bright regions (Hendrix<br />

1010 and Hansen, 2009a). This is consistent with thermal segregation as a possible process explaining the<br />

1011 overall appearance <strong>of</strong> Iapetus and, more particularly, the observed correlations <strong>of</strong> the dark/bright<br />

1012 boundary with geographical location and topography (Fig. 15) and the small-scale dramatic<br />

1013<br />

1014<br />

brightness variations (Fig. 15).<br />

1015 Fig. 14 Cassini ISS images from the September 2007 flyby, demonstrating the large- and small-scale<br />

1016 albedo effects <strong>of</strong> thermal segregation. (Left) The central longitude <strong>of</strong> the trailing hemisphere is 24°<br />

1017 to the left <strong>of</strong> the mosaic's center. (Right) The mosaic consists <strong>of</strong> two image footprints across the<br />

1018 surface <strong>of</strong> Iapetus. The view is centered on terrain near 42° southern latitude and 209.3° western<br />

1019 longitude, on the anti-Saturn facing hemisphere. Image scale is approximately 32m/pixel; the crater<br />

1020<br />

1021<br />

in the center is about 8 km across.<br />

1022 The 3µm absorption observed in VIMS data (Clark et al., 2009) might be due to ice, ammonia or<br />

1023 bound water. As the 3µm absorption is not been well matched by those due to ice, bound water and<br />

1024 ammonia, a combination seems to be indicated, but a model showing a good match has not yet been<br />

1025 developed (Clark et al., 2009). The amount <strong>of</strong> ice contained in the dark materials from VIMS data<br />

1026<br />

1027<br />

1028<br />

could be on the order <strong>of</strong> 10%, but is estimated to be


1065 visible wavelength by the Hubble space telescope (Verbiscer et al., 2007). Their brightness changes<br />

1066 with their distance from the E-ring, which is consistent with the assumption that their albedo results<br />

1067 from E-ring coating. Hamilton and Burns (1994) predicted that the relative velocities between the<br />

1068 E-ring particles and the icy satellites might explain the large-scale longitudinal albedo patterns on<br />

1069 the icy satellites. Because <strong>of</strong> their higher relative velocity, the satellites exterior to the densest point<br />

1070 in the E-ring have brighter leading hemispheres, while those interior to it have brighter trailing<br />

1071 hemispheres because <strong>of</strong> the higher velocity <strong>of</strong> the particles. Indeed, Mimas and Enceladus are both<br />

1072 slightly darker on their leading than on their trailing hemispheres, while the leading hemispheres <strong>of</strong><br />

1073 Tethys, Dione and Rhea are brighter than their trailing hemispheres (Buratti et al., 1998). The<br />

1074 visible orbital phase curve <strong>of</strong> Enceladus shows that the leading hemisphere is ~1.2 times darker than<br />

1075 its trailing hemisphere. At visible wavelengths, the leading hemisphere <strong>of</strong> Tethys is ~1.1 times<br />

1076 brighter than its trailing hemisphere, Dione’s leading hemisphere is up to ~1.8 times brighter and<br />

1077 Rhea’s leading hemisphere is ~1.2 times brighter than the trailing hemisphere (Buratti and Veverka,<br />

1078 1984).<br />

1079 The bombardment <strong>of</strong> icy surfaces by charged particles causes both structural and chemical changes<br />

1080 in the ice (Johnson et al., 2004). Effects include the implantation <strong>of</strong> magnetospheric ions, sputtering<br />

1081 and grain size alteration. Low-energy electrons can cause chemical reactions on the surface, while<br />

1082 high-energy electrons and ions tend to cause radiation effects at depth. The latter can result in bulk<br />

1083 chemical changes in composition as well as structural damage. A primary effect <strong>of</strong> energetic<br />

1084 particle bombardment on ice is to cause defects in it (Johnson, 1997; Johnson and Quickenden,<br />

1085<br />

1086<br />

1997; Johnson et al., 1998). Extended defects, in the form <strong>of</strong> voids and bubbles, affect the lightscattering<br />

properties <strong>of</strong> the surface. They also act as reaction and trapping sites for gases, which can<br />

1087 produce spectral absorption features. Incident ions deposit energy over a small region around their<br />

1088 path through the ice, causing local damage. Depending on excitation density and cooling rate, this<br />

1089 can result in either local crystallization or amorphization (Strazzulla, 1998).<br />

1090 Energetic ions and electrons have been seen to brighten laboratory surfaces in the visible by<br />

1091 producing voids and bubbles (Sack et al., 1992). Depending on the size and density <strong>of</strong> these<br />

1092 features, a clear surface can become brighter due to enhanced scattering in the visible. The plasma-<br />

1093 induced brightening competes with annealing, which operates efficiently in regions where the ice<br />

1094 temperature exceeds ~100K and can make equatorial regions less light-scattering despite the plasma<br />

1095 bombardment.<br />

1096 A significant effect <strong>of</strong> the ion bombardment <strong>of</strong> ice is the creation <strong>of</strong> new species through radiolysis.<br />

1097<br />

1098<br />

Bombarding particles can decompose H2O ice, creating molecular hydrogen and oxygen. Miniatmospheres<br />

<strong>of</strong> trapped volatiles have been suggested to form in voids in the ice (Johnson and<br />

1099<br />

1100<br />

Jesser, 1997). As a result, species such as O3, H2O2, O2 and oxygen-rich trace species have been<br />

detected as gases trapped in the regoliths <strong>of</strong> the icy Galilean satellites (Nelson et al., 1987; Noll et<br />

1101 al., 1996; Hendrix et al., 1998, 1999; Spencer et al., 1995; Carlson et al., 1996, 1999; Hibbitts et al.,<br />

1102 2000, 2003). On these satellites, the hydrogen and oxygen produced by radiation decomposition<br />

1103 form tenuous atmospheres; in contrast, atmospheres have not been detected on satellites in the<br />

1104 Saturnian system (with the exception <strong>of</strong> the gaseous plume <strong>of</strong> Enceladus). There are several<br />

1105 possible reasons for this. Lower gravities might support the loss <strong>of</strong> volatiles; temperatures are<br />

1106<br />

1107<br />

1108<br />

lower, which might slow down the breakup <strong>of</strong> H2O and the loss <strong>of</strong> H; the particle environment is<br />

different (discussed below); the electron density and temperature is not effective in exciting and<br />

'lighting up' any gases around the moons; and sputtering is not as intense as in the Jupiter system<br />

1109 due to differences in the net charge and energy <strong>of</strong> the particle environment. On Saturn, the primary<br />

1110 surface alterations found thus far are due to the E-ring, dust (e.g. on Dione) and thermal processing<br />

1111 (especially apparent on Iapetus, as previously discussed).<br />

1112 Neutrals, rather than ions as on Jupiter, dominate the Saturnian system. The energetic heavy ions<br />

1113 (~10keV to MeV) that are critical at Europa and Ganymede are lost by charge exchange to the<br />

1114 neutral cloud <strong>of</strong> the Saturnian system as they diffuse inwards from beyond ~10Saturn radii<br />

1115 (Paranicas et al., 2008). However, although the energy fluxes due to trapped plasma and the solar<br />

1116 UV are relatively low on the icy Saturnian satellites (Paranicas et al., 2008), radiation processing<br />

1117 appears to be occurring. Radiation fluxes are known to cause decomposition in ice, possibly<br />

producing ozone in the icy surfaces <strong>of</strong> Dione and Rhea (Noll et al., 1997) and O2 + 1118<br />

1119<br />

in Saturn’s<br />

plasma (Martens et al., 2007). In addition, the role <strong>of</strong> low energy ions has been recently re-<br />

1120 examined using Cassini CAPS data (Sittler et al., 2008), indicating that sputtering by this<br />

1121<br />

1122<br />

component is not negligible (Johnson et al., 2008). Spencer (1998) searched for but did not find O2<br />

inclusions on Rhea and Dione. Such inclusions have been seen on Ganymede in the low latitudes <strong>of</strong><br />

1123<br />

the trailing hemisphere, and it was suggested that they might be related to charged-particle<br />

21


1124<br />

1125<br />

1126<br />

bombardment or UV photolysis. The lack <strong>of</strong> O2 on Rhea and Dione suggests that low temperatures<br />

preclude the formation <strong>of</strong> O2 bubbles in the ice. Hydrogen peroxide (H2O2) has been tentatively<br />

VIMS data (Newman et al., 2007) <strong>of</strong> the south pole tiger-stripes region <strong>of</strong> Enceladus, although it is<br />

1127 not clear whether exogenic processing is needed to create peroxide.<br />

1128 It is important to distinguish between hemispheric dichotomies in color, albedo, texture and<br />

1129 macroscopic structure because each exogenic alteration mechanism leaves a different signature on<br />

1130 the physical and optical properties <strong>of</strong> the surface. Photometric methods have been employed over<br />

1131 the past two decades to characterize the roughness, compaction state, particle size and phase<br />

1132 function, and single scattering albedo <strong>of</strong> planetary and satellite surfaces, including the Saturnian<br />

1133 satellites (Buratti, 1985; Verbiscer and Veverka, 1989, 1992, 1994; Domingue et al. 1995). An<br />

1134 indication <strong>of</strong> the scattering properties <strong>of</strong> the satellites’ surfaces can be obtained by extracting<br />

1135 photometric scans across their disks at small solar phase angles. Dark surfaces such as that <strong>of</strong> the<br />

1136 Moon, in which single scattering is dominant, are not significantly limb-darkened, whereas bright<br />

1137 surfaces in which multiple scattering is important could be expected to show limb darkening.<br />

1138 Physical properties such as roughness, compaction state and particle size give clues as to the past<br />

1139 evolution <strong>of</strong> the satellites’ surfaces and reveal the relative importance <strong>of</strong> exogenic alteration<br />

1140 processes. Accurate photometric modeling is also necessary to define the intrinsic changes in albedo<br />

1141 and color on a planetary surface; much, and in many cases most <strong>of</strong> the change in intensity is due to<br />

1142 variations in the radiance viewing geometry and is not intrinsic to the surface.<br />

1143 <strong>Planetary</strong> surface scattering models have been developed (e.g. Horak, 1950; Goguen, 1981; Lumme<br />

1144<br />

1145<br />

and Bowell, 1981a, 1981b; Hapke, 1981, 1984, 1986, 1990; Shkuratov et al., 1999, 2005) which<br />

express the radiation reflected from the surface in terms <strong>of</strong> the following physical parameters: single<br />

1146 scattering albedo, single particle phase function, the compaction properties <strong>of</strong> the optically active<br />

1147 portion <strong>of</strong> the regolith, and the scale and extent <strong>of</strong> macroscopically rough surface features.<br />

1148 Observations at small solar phase angles (0-10°) are particularly important for understanding the<br />

1149 compaction state <strong>of</strong> the upper regolith: this is the region <strong>of</strong> the well-known opposition effect, in<br />

1150 which the rapid disappearance <strong>of</strong> shadowing among surficial particles causes a nonlinear surge in<br />

1151 brightness as the object becomes fully illuminated. At the smallest phase angles (40°) are necessary for<br />

1153 investigating the macroscopically rough nature <strong>of</strong> surfaces, ranging in size from clumps <strong>of</strong> several<br />

1154 particles to mountains, craters and ridges. Such features cast shadows and alter the local incidence<br />

1155 and emission angles. Two formalisms have been developed to describe roughness for planetary<br />

1156 surfaces: a mean-slope model in which the surface is covered by a Gaussian distribution <strong>of</strong> slope<br />

1157 angles with a mean angle <strong>of</strong> θ (Hapke, 1984; Helfenstein and Shepard, 1998), and a crater model in<br />

1158 which the surface is covered by craters with a defined depth-to-radius ratio (Buratti and Veverka,<br />

1159 1985). The single particle phase function, which expresses the directional scattering properties <strong>of</strong> a<br />

1160 planetary surface, is an indicator <strong>of</strong> the physical character <strong>of</strong> individual particles in the upper<br />

1161 regolith, including their size, size distribution, shape, and optical constants. Small or transparent<br />

1162 particles tend to be more isotropically scattering because photons survive to be multiply scattered;<br />

1163 forward scattering can exist if photons exit in the direction away from the observer. The single<br />

1164 particle phase function is usually described by a 1 or 2-term Henyey-Greenstein phase function defined<br />

1165 by the asymmetry parameter g, where g = 1 is purely forward-scattering, g = -1 is purely<br />

1166<br />

1167<br />

backscattering, and g = 0 is isotropic. The opposition surge <strong>of</strong> airless surfaces has been described by<br />

parameters that define the compaction state <strong>of</strong> the surface and the degree <strong>of</strong> coherent backscatter <strong>of</strong><br />

1168 multiply scattered photons (Irvine, 1966; Hapke, 1986, 1990). One final physical photometric<br />

1169 parameter is the single scattering albedo (ώ), which is the probability that a photon will be scattered<br />

1170 into 4π steradians after one scattering.<br />

1171 Three other important physical parameters are the geometrical albedo (p), the phase integral (q) and the<br />

1172<br />

1173<br />

Bond albedo (AB). The geometrical albedo is the flux received from a planet at a solar phase angle <strong>of</strong><br />

0º compared with that <strong>of</strong> a perfectly diffusing disk <strong>of</strong> the same size, also at 0º. The phase integral (q) is<br />

1174 the flux, normalized to unity at a solar phase angle <strong>of</strong> 0º, integrated over all solar phase angles. The<br />

1175 Bond albedo is p·q, and when integrated over all wavelengths it yields the bolometric Bond albedo,<br />

1176 which is a measure <strong>of</strong> the energy balance on the planet or satellite. Except for the bolometric Bond<br />

1177 albedo, all these parameters are wavelength-dependent.<br />

1178 Buratti (1984) and Buratti and Veverka (1984) performed photometric analyses <strong>of</strong> visible-<br />

1179 wavelength Voyager data <strong>of</strong> the icy Saturnian moons based on Voyager disk-resolved data sets, and<br />

1180 disk-integrated Voyager observations at large phase angles to supplement existing ground-based<br />

1181 observations at smaller phase angles. These studies investigated limb darkening with implications<br />

1182<br />

for multiple scattering (significant limb darkening suggests that multiple scattering is important).<br />

22


1183 Physical photometric modeling <strong>of</strong> Saturnian moons based on Voyager and ground-based data was<br />

1184 done by Buratti (1985), Verbiscer and Veverka (1989, 1992, 1994), Domingue et al. (1995) and<br />

1185 Simonelli et al. (1999). In general, the surfaces <strong>of</strong> the Saturnian satellites are backscattering and<br />

1186 exhibit roughness comparable to that <strong>of</strong> the Moon, indicating that impact processes produce similar<br />

1187 morphological effects on rocky and icy bodies at low temperatures. The photometric and physical<br />

1188 properties <strong>of</strong> the satellites determined so far from both Voyager and Cassini data are summarized in<br />

1189<br />

1190<br />

Table 4.<br />

Parameter Mimas Enceladus Tethys Dione Rhea Phoebe<br />

Mean slope 30 (2)<br />

13±5 (4) 31±4 (6)<br />

angle θ (º)<br />

Asymmetry<br />

parameter g<br />

30±1 (5) 6±1 (3)<br />

33±3 (7)<br />

Single<br />

scattering<br />

albedo ώ<br />

Phase<br />

integral q<br />

Visual<br />

geometric<br />

albedo pv<br />

Bond<br />

albedo AB<br />

-0.3±0.05 (2)<br />

-0.213±0.005<br />

(5)<br />

0.93±0.03 (2)<br />

0.951±0.002<br />

(5)<br />

0.82±0.05 (1)<br />

0.78±0.05 (5)<br />

-0.35±0.03<br />

(2)<br />

-0.399±0.005<br />

(3)<br />

0.99±0.02 (2)<br />

0.998±0.001<br />

(3)<br />

0.86±0.1 (1)<br />

0.92±0.05 (3)<br />

0.77±0.15 (1) 1.04±0.15 (1)<br />

1.41±0.03 (8);<br />

0.6±0.1 (1)<br />

0.5±0.1 (5)<br />

0.9±0.1 (1)<br />

0.91±0.1 (3)<br />

0.7±0.1 (1)<br />

0.80±0.15<br />

(1)<br />

0.6±0.1 (1)<br />

23<br />

0.8±0.1 (1)<br />

0.55±0.15<br />

(1)<br />

0.45±0.1<br />

(1)<br />

-0.287±0.007 (4) -0.24 (6)<br />

-0.3±0.1 (plus<br />

a forward<br />

component)<br />

(7)<br />

0.861±0.008 (4) 0.068 (6)<br />

0.07±0.01 (7)<br />

0.70±0.06 (1) 0.24±0.05 (6)<br />

0.78±0.05 (4) 0.29±0.03 (7)<br />

0.65 (1) 0.081±0.002<br />

(6);<br />

0.082±0.002<br />

(7);<br />

0.45±0.1 (1) 0.020±0.004<br />

0.49±0.06 (4) (6)<br />

0.023±0.007<br />

(7)<br />

1191<br />

1192 Tab. 4 Photometric and physical parameters <strong>of</strong> the icy Saturnian satellites (references: (1) Buratti<br />

1193 and Veverka, 1984; (2) Buratti, 1985; (3) Verbiscer and Veverka, 1994; (4) Verbiscer and Veverka<br />

1194 1989; (5) Verbiscer and Veverka, 1992; (6) Simonelli et al., 1999; (7) Buratti et al., 2008; (8)<br />

1195 Verbiscer et al., 2005<br />

1196<br />

1197 Opposition surge data from 0.2 to 5.1µm were obtained for Enceladus, Tethys, Dione, Rhea and<br />

1198<br />

1199<br />

Iapetus. Fig. 16 shows the opposition curve <strong>of</strong> Iapetus in between 0.35µm and 3.6µm<br />

1200 Fig. 15 (Above) Cassini visual infrared mapping spectrometer (VIMS) disk-integrated observations<br />

1201 <strong>of</strong> the opposition surge on Iapetus between 0.35 and 3.6µm. (Lower left) The slope <strong>of</strong> the solar phase<br />

1202 curve (the 'phase coefficient') between 1º and 8º decreases as the albedo increases. This effect is<br />

1203 expected for shadow hiding, since bright surfaces have partly illuminated primary shadows. (Lower<br />

1204 right) On the other hand, the amplitude <strong>of</strong> the 'spike' under 1º (shown as a fractional increase above<br />

1205 the value <strong>of</strong> the phase curve at 1º) increases with the albedo, suggesting that it is a multiple scattering<br />

1206<br />

1207<br />

effect, such as coherent backscatter.<br />

1208 (Buratti et al., 2009b; Hendrix and Buratti, 2009). The curve is divided into two components: at<br />

1209 phase angles greater than 1º, brightness increases linearly with decreasing phase angle, while at<br />

1210 phase angles smaller than one degree the increase is exponential. Moreover, the albedo-dependence<br />

1211 <strong>of</strong> the two types <strong>of</strong> surge is different. For the first type, the magnitude <strong>of</strong> the surge decreases as the<br />

1212 albedo increases, while for the second type <strong>of</strong> surge the reverse is true. This result suggests that<br />

1213 different phenomena are responsible for the two types <strong>of</strong> surge. At angles greater than 1º, the surge<br />

1214<br />

1215<br />

is caused by mutual shadowing, in which shadows cast among particles <strong>of</strong> the regolith rapidly<br />

disappear as the face <strong>of</strong> the satellite becomes fully illuminated to the observer. Since multiply<br />

1216 scattered photons, which partly illuminate primary shadows, become more numerous as the albedo<br />

1217 increases, this effect becomes less pronounced for bright surfaces. On the other hand, the huge<br />

1218 increase at very small solar phase angles can be explained by coherent backscatter, a phenomenon<br />

1219 in which photons following identical but reversed paths in a surface interfere constructively in exactly<br />

1220<br />

the backscattering direction, causing brightness to increase by up to a factor <strong>of</strong> two (Hapke, 1990).


1221 The Cassini flyby in June 2004 prior to Saturn orbit insertion afforded views at large phase angles –<br />

1222 important for modeling roughness - and over a full excursion in geographical longitudes. Phoebe<br />

1223 exhibits a substantial forward-scattering component to its single particle phase function. This result<br />

1224 is consistent with a substantial amount <strong>of</strong> fine-grained dust on the surface <strong>of</strong> the satellite generated<br />

1225 by particle infall, as suggested by Clark et al. (2005, 2008a). As observed in Cassini images (Porco<br />

1226 et al., 2005a), Phoebe also shows extensive macroscopic roughness, hinting at the violent collisional<br />

1227 history predicted by Nesvorny et al. (2003).<br />

1228 ISS images <strong>of</strong> the low-albedo hemisphere <strong>of</strong> Iapetus were fitted to the crater roughness model<br />

1229 (Buratti and Veverka, 1985), yielding a markedly smooth value for the degree <strong>of</strong> macroscopic<br />

1230 roughness on this side with a mean slope angle <strong>of</strong> only a few degrees (Lee et al., 2009). This result<br />

1231 suggests that small-scale rough features on the dark side have been filled in (features which<br />

1232 probably dominate the photometric effects; see Helfenstein and Shepard, 1998), which is consistent<br />

1233 with the idea <strong>of</strong> Spencer and Denk (2009) that thermal migration <strong>of</strong> water ice leaves behind a fluffy<br />

1234 residue <strong>of</strong> dark material. The low-albedo side <strong>of</strong> Iapetus shows a roughness similar to Enceladus,<br />

1235 which is coated with micron-sized particles from its plume and the E-ring (Verbiscer and Veverka,<br />

1236<br />

1237<br />

1994; see Table 21.4.1). Model-fits for the low albedo hemisphere are shown in Fig. 14.<br />

1238 Fig. 16 A macroscopic roughness model for the low-albedo hemisphere <strong>of</strong> Iapetus, based on the<br />

1239 crater roughness model by Buratti and Veverka (1985). The best-fit roughness function corresponds<br />

1240 to a depth-to-radius <strong>of</strong> 0.14, or a mean slope angle <strong>of</strong> 11º, which is much lower than the ~30º typical<br />

1241<br />

1242<br />

1243<br />

<strong>of</strong> other icy satellites (see Tab. 4). For these fits, a surface phase function f(α) <strong>of</strong> 0.023 was derived.<br />

From Lee et al., 2009<br />

1244 Finally, observations at very large solar phase angles (155º-165º) have been used to search for<br />

1245 cryovolcanic activity on satellites. By comparing the brightness <strong>of</strong> Rhea to Enceladus in this phase<br />

1246 angle range at 2.02µm, Pitman et al. (2008) were able to place an upper limit on water vapor<br />

column density based on a possible plume <strong>of</strong> 1.52 x 10 14 to 1.91 x 10 15 cm -2 1247 , two orders <strong>of</strong><br />

1248<br />

1249<br />

magnitude below the observed plume density <strong>of</strong> Enceladus (Fig. 17).<br />

1250 Fig. 17 Cassini VIMS disk-integrated brightness as a function <strong>of</strong> solar phase angle at 2.23µ.<br />

1251 Symbols: Normalized Rhea and Enceladus VIMS brightness (every fifth data point plotted). Solid<br />

1252 line: Third-order polynomial fit to Rhea data. Dashed lines: (polynomial fit to Rhea data).<br />

1253 Enceladus's plume can be seen as a peak at a solar phase angle <strong>of</strong> 159°. Adapted from Pitman et al.<br />

1254<br />

1255<br />

(2008)<br />

1256 The microphysical structure <strong>of</strong> the E-ring grain coatings <strong>of</strong> the inner icy moons embedded in the E-<br />

1257 ring can be probed by investigating the opposition surge. Verbiscer et al. (2007) found that the<br />

1258 amplitude and angular width <strong>of</strong> the opposition effect on the inner icy moons are correlated with the<br />

1259<br />

1260<br />

moons’ position relative to the E-ring.<br />

1261 21.5 Constraints on the Top-Meter Structure and Composition by Radar<br />

1262 The wavelength <strong>of</strong> Cassini's 13.8GHz RADAR instrument, 2.2cm, is about six times shorter than<br />

1263<br />

1264<br />

the only ground-based radar wavelength available to study the satellites (13cm, at Arecibo) and 22<br />

times longer than the millimeter wavelengths at the limit <strong>of</strong> the CIRS.<br />

1265 The echoes result from volume scattering, and their strength is sensitive to ice purity. Therefore,<br />

1266 they provide unique information about near-surface structural complexity as well as about<br />

1267 contamination with non-ice material. This section summarizes the most basic aspects <strong>of</strong> Cassini-<br />

1268 and ground-based radar observations <strong>of</strong> the satellites, the theoretical context for interpreting the<br />

1269 echoes, and the inferences drawn about subsurface structure and composition.<br />

1270 Cassini measures echoes in the same linear polarization as transmitted, whereas most ground-based<br />

1271 radar astronomy consists <strong>of</strong> Arecibo 13cm or 70cm observations and Goldstone 3.5cm observations<br />

1272 that almost always use transmission <strong>of</strong> a circularly polarized signal and simultaneous reception <strong>of</strong><br />

1273 echoes in same-circular (SC) and opposite-circular (OC) polarizations. A radar target's radar albedo<br />

1274 is defined as its radar cross section divided by its projected area. The radar cross section is the<br />

1275 projected area <strong>of</strong> a perfectly reflective isotropic scatterer which, if observed at the target's distance<br />

1276 from the radar and using the same transmitted and received polarizations, would return the observed<br />

1277 echo power. The total-power radar albedo is the sum <strong>of</strong> the albedos in two orthogonal polarizations,<br />

1278 TP = SL + OL = SC + OC. Here, we will use 'SL-2' to denote the 2cm radar albedo in the SL<br />

1279<br />

polarization, 'TP-13' to denote the 13cm total-power albedo, etc.<br />

24


1280 For a smooth, homogeneous target, single back reflections would be entirely OC when the<br />

1281 transmission is circularly polarized, or entirely SL when the transmission is linearly polarized, so<br />

1282 the polarization ratios SC/OC and OL/SL would equal zero. Saturn's icy satellites and the icy<br />

1283 Galilean satellites not only have radar albedos an order <strong>of</strong> magnitude greater than those <strong>of</strong> the Moon<br />

1284 and most other solar system targets but also unusually large polarization ratios.<br />

1285 These strange radar signatures were discovered in observations <strong>of</strong> Europa, Ganymede and Callisto<br />

1286 in the mid-1970s. It took a decade and a half to realize that the signatures are the outcome <strong>of</strong><br />

1287 'coherent backscattering,' which is phase-coherent, multiple scattering within a dielectric medium<br />

1288 that is both heterogeneous and non-absorbing (e.g. MacKintosh and John, 1989; Hapke, 1990).<br />

1289 During the next decade, Peters (1992) produced a vector formulation <strong>of</strong> coherent backscattering that<br />

1290 led to predictions <strong>of</strong> radar albedos and polarization ratios, and Black (1997) and Black et al. (2001b)<br />

1291 built on Peters' work with models <strong>of</strong> scatterer properties.<br />

1292 Extremely clean ice is insufficient for anomalously large radar albedos and circular polarization<br />

1293 ratios, which arise from multiple scattering within a random, disordered medium whose intrinsic<br />

1294 microwave absorption is very low. Thus, perfectly clean ice that is structurally homogeneous<br />

1295 (constant density at millimeter scales) would not give anomalous echoes. The regoliths on solar<br />

1296 system objects are structurally very heterogeneous, with complex density variations from the<br />

1297 distribution <strong>of</strong> particle sizes and shapes produced by impacts <strong>of</strong> projectiles <strong>of</strong> different sizes and<br />

1298 impact energies.<br />

1299 It is to be expected that the uppermost few meters <strong>of</strong> Saturn's icy satellite regolith are basically<br />

1300<br />

1301<br />

similar to those <strong>of</strong> the icy Galilean satellites or the Moon's in terms <strong>of</strong> particle size distribution,<br />

porosity and density as functions <strong>of</strong> depth, and lateral/vertical heterogeneity. It is the combination<br />

1302 <strong>of</strong> naturally complex density variations and the extreme transparency <strong>of</strong> pure water ice that gives<br />

1303 icy regolith its exotic radar properties.<br />

1304 The icy Galilean satellites were observed in dual-circular-polarization experiments from the ground<br />

1305 at 3.5, 13, and 70cm; experiments measuring SL-13 and OL-13 were done only for Europa,<br />

1306 Ganymede and Callisto. Arecibo obtained estimates <strong>of</strong> SC-13 and OC-13 for Enceladus, Tethys,<br />

1307 Dione, Rhea and Iapetus. The Cassini radar measured SL-2 for Mimas, Hyperion and Phoebe. The<br />

1308 satellites' albedos show different styles and are location- and wavelength-dependent, as discussed<br />

1309 below. The results described in this chapter make use <strong>of</strong> 73 Cassini SL-2 measurements, more than<br />

1310 twice the number reported by Ostro et al. (2006).<br />

1311 If a radar target is illuminated by a uniform antenna beam (that is, if the antenna gain is constant<br />

1312 across the disk), then it is a simple matter to use the appropriate 'radar equation' to convert the<br />

1313 measured echo power spectrum (power as a function <strong>of</strong> Doppler frequency) into a radar cross<br />

1314 section (e.g. Ostro, 1993). For Cassini measurements, except for the SAR imaging <strong>of</strong> Iapetus<br />

1315 discussed below, the antenna beam width is within a factor <strong>of</strong> several times the target's angular<br />

1316 width, so the gain varies across the target's disk, usually significantly. Data reduction yields an<br />

1317 estimate <strong>of</strong> echo power, which is the sum <strong>of</strong> contributions from different parts <strong>of</strong> the surface, with<br />

1318 each contribution weighted by the two-way gain. However, the distribution <strong>of</strong> echo power as a<br />

1319 function <strong>of</strong> location on the target disk is not known a priori. Thus, we do not know how much the<br />

1320 echo power from any given surface element has been 'amplified' by the antenna gain. To estimate a<br />

1321 radar albedo from our data we must make an assumption about the homogeneity and angular<br />

1322<br />

1323<br />

dependence <strong>of</strong> the surface's radar scattering. For the purpose <strong>of</strong> obtaining a radar albedo estimate<br />

from an echo power spectrum, one assumes uniform, azimuthally isotropic, cosine scattering laws,<br />

1324 and uses least squares to estimate the value <strong>of</strong> a single parameter that defines the echo's incidence-<br />

1325<br />

1326<br />

angle dependence. In general, only modest departures from Lambert limb darkening are found.<br />

1327 Radar-Optical Correlations<br />

1328<br />

1329<br />

Fig. 18 shows all measurements <strong>of</strong> the satellites' individual SL-2 albedos along with their<br />

1330 Fig. 18 Individual Cassini measurements <strong>of</strong> SL-2 (small symbols) and average optical geometrical<br />

1331 albedos, plotted for the satellites vs. their distances from Saturn. Optical albedos <strong>of</strong> Mimas,<br />

1332 Enceladus, Tethys, Dione and Rhea are from Verbiscer et al. (2007); those <strong>of</strong> Iapetus and Phoebe<br />

1333 are from Morrison et al. (1986) and involve a V filter; Hyperion’s is from Cruikshank and Brown<br />

1334 (1982). Note the similar distance dependences. Hyperion's radar albedo is large and appears to<br />

1335 break the pattern, perhaps (Ostro et al. 2006) because Hyperion's optical coloring may be due to a<br />

1336 thin layer <strong>of</strong> exogenously derived material that, in contrast to the other satellites, has never been<br />

1337<br />

1338<br />

worked into the icy regolith, which remains very clean and radar-bright.<br />

25


1339 average optical geometrical albedos, plotted vs. distance from Saturn. A very pronounced radar-<br />

1340 optical correlation means that variations in optical and radar albedos involve the same or related<br />

1341 contaminant(s) and/or surface processes. Absolutely pure water ice at satellite surface temperatures<br />

1342 is extremely transparent to radar waves. Contamination <strong>of</strong> water ice with no more than trace<br />

1343 concentrations <strong>of</strong> virtually any other substance (including earth rocks, chondritic meteorites, lunar<br />

1344 soil, ammonia, metallic iron, iron oxides and tholins) can decrease its radar transparency (and hence<br />

1345 the regolith's radar albedo) by one to two orders <strong>of</strong> magnitude. Phenomena that depend on the<br />

1346 distance from Saturn include primordial composition, meteoroid flux and the influx <strong>of</strong> dark material<br />

1347 (perhaps from several sources), the balance between the influx <strong>of</strong> ammonia-containing particles and<br />

1348 the removal <strong>of</strong> ammonia by a combination <strong>of</strong> ion erosion and micrometeoroid gardening, variation<br />

1349 in E-ring fluxes, and the systematics <strong>of</strong> thermal volatiles redistribution.<br />

1350 Modeling <strong>of</strong> icy Galilean satellite echoes (Black et al., 2001b) suggests that penetration to about<br />

1351 half a meter should be adequate to produce the range <strong>of</strong> measured values <strong>of</strong> SL-2.<br />

1352 The explanation by Clark et al. (2008b, 2009) for dark Saturnian system material, which suggests<br />

1353 that it consists <strong>of</strong> sub-micron iron particles from meteorites which oxidized and became mixed with<br />

1354 ice, would also suffice to explain, at least qualitatively, the overall radar-optical albedo correlation<br />

1355 in Fig. 18 if the contamination <strong>of</strong> the ice extends some one to several decimeters below the optically<br />

1356 visible surface, as expected by Clark et al. (2008b, 2009). The tentative detection <strong>of</strong> ammonia in<br />

1357 VIMS spectra by Clark et al. (2008a) opens up the possibility that contamination by this material<br />

1358 may play an important role in inter- and intra-satellite radar albedo variations. If so, this<br />

1359<br />

1360<br />

contamination would affect the radar but not the optical albedo.<br />

Almost all Cassini radar data on the icy satellites consist <strong>of</strong> real-aperture scatterometry, which<br />

1361 means that the beam size determines the resolution. Cassini has no real-aperture resolution <strong>of</strong><br />

1362 Mimas, Enceladus, Hyperion or Phoebe, but does have abundant measurements for Tethys, Dione,<br />

1363 Rhea and Iapetus with beams smaller than, and in many cases less than half, the size <strong>of</strong> the disk<br />

1364 (Ostro et al., 2006).<br />

1365 The SL-2 distributions <strong>of</strong> the individual satellites can be summarized briefly as follows. As regards<br />

1366 Mimas, Herschel may or may not contribute to the highest value <strong>of</strong> the disc-integrated albedo.<br />

1367 Enceladus shows no dramatic leading/trailing asymmetry and no noteworthy geographical<br />

1368 variations except for a low albedo in an observation <strong>of</strong> the Saturn-facing side. Tethys' leading side<br />

1369 is brighter both optically and in SL-2. Dione's highest-absolute-latitude views give the highest radar<br />

1370 albedos. Rhea's leading side has a lower SL-2 in the north, where there is much cratering, and<br />

1371 higher in the south, where there are optically bright rays. Its trailing and Saturn-facing sides, which<br />

1372 look younger, show a relatively high SL-2. Iapetus displays an extremely marked correlation<br />

1373 between optical brightness and SL-2.<br />

1374 In general, a first-order understanding <strong>of</strong> the SL-2 distribution is that higher SL-2 means cleaner<br />

1375 near-surface ice, but the nature <strong>of</strong> the contaminant(s) and the depth <strong>of</strong> the contamination are open<br />

1376 questions. The latter can be constrained by measuring radar albedos at more than one wavelength.<br />

1377 Fortunately, there are Arecibo 13-cm observations <strong>of</strong> Enceladus, Tethys, Dione, Rhea (Black et al.,<br />

1378 2004) and Iapetus (Black et al., 2007).<br />

1379 Also fortunately, the wavelength dependence observed for Europa, Ganymede and Callisto (EGC)<br />

1380 gives us a context for discussing the wavelength dependence observed for Saturn's satellites.<br />

1381<br />

1382 Wavelength Dependence<br />

1383 EGC total-power radar albedos and circular polarization ratios at 3.5cm and 13cm are<br />

1384 indistinguishable. This means that their regolith structure (distribution <strong>of</strong> density variations) and ice<br />

1385 cleanliness look identical at the two wavelengths, for each object. That is, near-surface structures<br />

1386 are 'self-similar' at 3.5cm and 13cm scales; there are no depth-dependent phenomena, like<br />

1387 contamination levels or structural character, perceptible in EGC's 3.5cm and 13cm radar signatures.<br />

1388 At 70cm, moreover, EGC SC/OC ratios show little wavelength dependence, so subsurface multiple<br />

1389 scattering remains the primary source <strong>of</strong> the echo even on this longer scale. However, the radar<br />

1390 albedos (TP-70) plummet for all three objects. The implication is that at 70cm, scattering<br />

1391 heterogeneities are relatively sparse, and the self-similarity <strong>of</strong> the regolith structure disappears on<br />

1392 the larger scale. On Europa at least, with the most striking albedo drop from TP-13 to TP-70, the<br />

1393 regolith may simply be too young for it to be heterogeneous at 70cm.<br />

1394 Within the framework <strong>of</strong> the Black et al. (2001b) model, this implies that at 70cm 'the scattering<br />

1395 layer is disappearing in the sense that there are too few scatterer at this scale and they are confined<br />

1396 to a layer too shallow to permit numerous multiple scatterings. At even longer wavelengths, <strong>of</strong> the<br />

1397<br />

order <strong>of</strong> several meters, the reflectivity from the surfaces <strong>of</strong> these moons, and Europa in particular,<br />

26


1398 may drop precipitously or even vanish as the layers responsible for the centimeter-wavelength radar<br />

1399 properties become essentially transparent.'<br />

1400 Now, given that the 3.5cm and 13cm polarization properties <strong>of</strong> EGC show no significant<br />

1401 wavelength dependence, and that the 13cm OL/SL ratios <strong>of</strong> those objects are very similar, let us<br />

1402 assume that the linear polarization properties <strong>of</strong> the Saturnian satellites are wavelength-invariant.<br />

1403 Let us then use EGC's mean OL/SL value to estimate total-power 2cm albedos: TP-2 = (1.52 +/-<br />

1404 0.13) SL-2 and compare them to the Arecibo 13cm total-power albedos (TP-13).<br />

1405<br />

1406<br />

Fig. 19 shows relevant Cassini and Arecibo information.<br />

1407 Fig. 19 Wavelength dependence <strong>of</strong> total-power radar albedos. Individual symbols are estimates <strong>of</strong><br />

1408 TP-2 obtained from individual measurements <strong>of</strong> SL-2 using TP-2 = 1.52 SL-2 +/- 0.13 SL. Arecibo<br />

1409<br />

1410<br />

13-cm albedos are horizontal lines. Dashed lines represent EGC.<br />

1411 Moving outward from Saturn, the magnitude <strong>of</strong> 2cm to-13cm wavelength dependence is<br />

1412 unremarkable for Enceladus' leading side, extreme for Enceladus' trailing side, very large for<br />

1413 Tethys, large for Dione, unremarkable for Rhea, large for Iapetus' leading side, and extreme for<br />

1414 Iapetus' trailing side.<br />

1415 Whereas compositional variation (ice cleanliness) almost certainly does not contribute to the<br />

1416 wavelength dependence seen in EGC, it may well play a key role in the wavelength dependence<br />

1417 seen in Saturn's satellites. In this complicated system, the tapestry <strong>of</strong> satellite radar properties may<br />

1418<br />

1419<br />

involve many factors, including inter- and intra-satellite variations in E-ring flux, variations in the<br />

meteoroid or ionizing flux, variations in the concentration <strong>of</strong> radar absorbing materials (including<br />

1420 optically dark material and possibly ammonia) as a function <strong>of</strong> depth, the systematics <strong>of</strong> thermal<br />

1421 volatile redistribution (especially for Iapetus), and/or constituents with wavelength- or temperature-<br />

1422 dependent electrical properties. Here, we will present what we consider reasonable hypotheses<br />

1423 about what the most important factors are.<br />

1424 The unremarkable 2- to-13-cm wavelength dependence seen on Enceladus' leading side and also on<br />

1425 Rhea is reminiscent <strong>of</strong> Europa's 3.5- to-13-cm wavelength invariance. The simplest interpretation is<br />

1426 that at least the top few decimeters are uniformly clean and old enough for meteoroid bombardment<br />

1427 to have created density heterogeneities that are able to produce coherent backscattering and extreme<br />

1428 radar brightness.<br />

1429 Enceladus' leading-side 2cm and 13cmcm TP albedos resemble Europa's 3.5cm and 13cm TP<br />

1430 albedos, whereas Rhea's 2cm and 13cm TP albedos resemble Ganymede's 3.5cm and 13cmcm TP<br />

1431 albedos, reflecting much greater contamination <strong>of</strong> Rhea's upper regolith.<br />

1432 Dione's and Rhea's TP-2 albedos are similar, but Dione is dimmer at 13cm. Dione VIMS data show<br />

1433 evidence <strong>of</strong> a bombardment <strong>of</strong> apparently dark particles on the trailing side (Clark et al., 2008a). If<br />

1434 the dark material contains ammonia, and the ejecta systematics <strong>of</strong> meteoroid bombardment<br />

1435 efficiently distribute the dark material over the surface <strong>of</strong> the object, multiple scattering will be cut<br />

1436 <strong>of</strong>f at a depth that may be too shallow for TP-13 to be enhanced.<br />

1437 The striking wavelength dependence <strong>of</strong> Enceladus' trailing side (TP-2 >>TP-13) mirrors Europa's<br />

1438 (TP-13 >>TP-70). It is easily understood if we assume that the top meter or so <strong>of</strong> ice was deposited<br />

1439 so recently that there has not been time for the regolith to be matured by meteoroid bombardment.<br />

1440<br />

1441<br />

That is, there are small-scale heterogeneities that produce coherent backscattering at 2cm, but none<br />

<strong>of</strong> the larger-scale heterogeneities needed to produce coherent backscattering at 13cm.<br />

1442 This idea is consistent with the fact that Enceladus' trailing side is optically brighter than its leading<br />

1443 side, apparently because <strong>of</strong> the greater E-ring flux it receives (Buratti, 1988; Showalter et al.,<br />

1444 1991b; Buratti et al., 1998). So all <strong>of</strong> Enceladus is very clean, but the uppermost layer <strong>of</strong> one or<br />

1445 several meters on the trailing side is also extremely young, resulting in a striking hemispheric<br />

1446 dichotomy at 13cm that is the most intense seen at that wavelength in the icy-satellite systems <strong>of</strong><br />

1447 either Jupiter or Saturn.<br />

1448 At least some <strong>of</strong> the decrease in average SL-2 (Fig. 18) and the decrease in average wavelength<br />

1449 dependence outward from Enceladus (Fig. 19) is probably due to the outward decrease in E-ring<br />

1450 flux (Verbiscer et al., 2007).<br />

1451 Finally, we get to Iapetus, where TP-2 is strongly correlated with the optical albedo: low for the<br />

1452 optically dark leading-side material and high for the optically bright trailing-side material.<br />

1453 However, Iapetus' TP-13 values show much less <strong>of</strong> a dichotomy and are several times lower than<br />

1454 TP-2, being indistinguishable from the weighted mean <strong>of</strong> TP-13 for main-belt asteroids, 0.15 ± 0.10<br />

1455<br />

1456<br />

(Tab. 5).<br />

27


1457 Dark/Leading Bright/Trailing<br />

1458 2.2-cm 0.32 + 0.09 0.58 + 0.18<br />

1459 13-cm 0.13 + 0.04 0.17 + 0.04<br />

1460<br />

1461<br />

Tab. 5 Total-power radar albedos<br />

1462 Therefore, whereas Iapetus is anomalously bright at 2cm (similar to Callisto) and mimics the optical<br />

1463 hemispheric dichotomy, at 13cm it looks like a typical main-belt asteroid, with minimal evidence <strong>of</strong><br />

1464 mimicking the optical hemispheric dichotomy.<br />

1465 The Iapetus results are understandable if (1) the leading side's optically dark contaminant is present<br />

1466 to depths <strong>of</strong> at least one to several decimeters, so that SL-2 can see the optical dichotomy, and (2)<br />

1467 ammonia and/or some other radar-absorbing contaminant is globally much less abundant within the<br />

1468 upper one to several decimeters than at greater depths, which would explain the wavelength<br />

1469 dependence, that is, the virtual disappearance <strong>of</strong> the object's 2-cm signature at 13cm. The first<br />

1470 conclusion is consistent with the dark material being a shallow phenomenon, as predicted from the<br />

1471 thermal migration hypothesis discussed elsewhere in this chapter.<br />

1472 Moreover, ammonia contamination may be uniform within the top few meters, so that the<br />

1473 wavelength dependence arises because ammonia's electrical loss is much greater at 13cm<br />

1474 (2.38GHz) than at 2cm (13.8GHz). Unfortunately, the likelihood <strong>of</strong> this possibility cannot yet be<br />

1475 judged because measuring the complex dielectric constant <strong>of</strong> ammonia-contaminated water ice as a<br />

1476 function <strong>of</strong> concentration, frequency and temperature is so difficult that, although needed, it has not<br />

1477<br />

1478<br />

been done yet. This is a pity, because the lack <strong>of</strong> this information undermines the interpretation <strong>of</strong><br />

the 2cm to13cm wavelength dependence seen in other Saturnian satellites as well as <strong>of</strong> the possible<br />

1479 influence <strong>of</strong> ammonia and its 'modulation' by magnetospheric bombardment (which varies with<br />

1480 distance from Saturn) on the dependence <strong>of</strong> SL-2 on the distance from Saturn.<br />

1481 Verbiscer et al. (2006) suggested that, while their spectral data do not contain an unambiguous<br />

1482 detection <strong>of</strong> ammonia hydrate on Enceladus, their spectral models do not rule out the presence <strong>of</strong> a<br />

1483 modest amount <strong>of</strong> it on both hemispheres. However, they note that the trailing hemisphere is also<br />

1484 exposed to increased magnetospheric particle bombardment, which would preferentially destroy the<br />

1485<br />

1486<br />

ammonia. In any case, ammonia does not appear to be a factor in Enceladus’ radar albedos.<br />

1487 Iapetus Radar Image<br />

1488 During the Iapetus 49 flyby, RADAR obtained an SAR image at 2km to12km surface resolution<br />

1489 that covers much <strong>of</strong> the object's leading (optically darker) side. From a comparison with optical<br />

1490 imaging (Fig. 20), we see that, as expected, the SAR reliably reveals the surface features seen in<br />

1491<br />

1492<br />

ISS images.<br />

1493 Fig. 20 RADAR SAR image <strong>of</strong> the leading side <strong>of</strong> Iapetus (Sept. 2007, left) and the ISS mosaic<br />

1494 PIA 08406 (Jan. 2008), using similar projections and similar scales (the large crater (on the right in<br />

1495<br />

1496<br />

the ISS image is about 550 km in diameter).<br />

1497 The pattern <strong>of</strong> radar brightness contrast is basically similar to that seen optically, except for the<br />

1498 large basin at 30°N, 75°W, which has more SL-2-bright than optically bright highlights. This<br />

1499<br />

1500<br />

suggests that the subsurface <strong>of</strong> these highlights at decimeter scales may be cleaner than in other<br />

parts <strong>of</strong> the optically dark terrain, which is consistent with the idea that the contamination by dark<br />

1501 material might be shallow, at least in this basin. Conversely, many optically bright crater rims are<br />

1502<br />

1503<br />

SL-2-dark, so the exposure <strong>of</strong> clean ice cannot be very deep.<br />

1504 21.6 Geological Evolution<br />

1505 The densities <strong>of</strong> the Saturnian satellites constrain their composition to be about 2/3 ice and 1/3 rock.<br />

1506 Accretional energy, tidal despinning and radioactive decay might have heated and melted the<br />

1507 moons. Due to the very low melting temperatures <strong>of</strong> various water clathrates, it is likely that the<br />

1508 larger bodies differentiated early in their evolution while smaller satellites may still be composed<br />

1509 and structured in their primitive state. For a more detailed discussion <strong>of</strong> the aspects <strong>of</strong> origin and<br />

1510 thermal evolution, see Johnson et al. (2009) and Matson et al. (2009).<br />

1511 The heavily cratered solid surfaces in the Saturnian system indicate the role <strong>of</strong> accretional and post-<br />

1512 accretional bombardement in the geological evolution <strong>of</strong> the satellites. Observed crater densities are<br />

1513 summarized in Dones et al. (2009). However, the source as well as the flux <strong>of</strong> bombarding bodies is<br />

1514 still under discussion, and Dones et al. (2009) came to the conclusion that, since no radiometric<br />

1515<br />

sample data exist from surfaces in the outer solar system and the size-frequency distribution <strong>of</strong><br />

28


1516 comets is less well understood than that <strong>of</strong> asteroids, the chronology <strong>of</strong> the moons <strong>of</strong> the giant<br />

1517 planets remains in a primitive state. Therefore, we refer the reader to Dones et al. (2009) for the<br />

1518 crater chronology discussion. In interpreting the geological evolution <strong>of</strong> the Saturnian satellites, we<br />

1519 will rely on relative stratigraphic relations. Nevertheless, absolute ages are given in Dones et al.<br />

1520 (2009).<br />

1521 In a stratigraphic sense, three major surface unit categories can be distinguished on all icy satellites:<br />

1522 (1) Cratered plains, which are characterized by dense crater populations that are superimposed by<br />

1523 all other geological features and, thus, are the relatively oldest units.<br />

1524 (2) Tectonized zones <strong>of</strong> narrow bands and mostly sub-parallel fractures, faults, troughs and ridges<br />

1525 <strong>of</strong> global extension indicating impact-induced or endogenic crustal stress. These units dissect<br />

1526 cratered plains and partly dissect each other, indicating different relative ages.<br />

1527 (3) Resurfaced units consisting <strong>of</strong> terrain that has undergone secondary exogenic and endogenic<br />

1528 processes such as mass wasting, surface alterations (due to micrometeorite gardening, sputtering,<br />

1529 thermal segregation and crater ray formation by recent impacts), contamination by dark material<br />

1530 and cryovolcanic deposition mostly resulting in smoothed terrain. Superimposed on cratered plains<br />

1531 and tectonized regions, these are the youngest surface units.<br />

1532<br />

1533<br />

Heavily cratered surfaces are found on all the icy rocks and small satellites (Fig. 21) as well as on<br />

1534 Fig. 21 Phoebe, 213 km in average diameter, was the first one <strong>of</strong> the nine classic moons <strong>of</strong> Saturn to<br />

1535 be captured by the ISS cameras aboard Cassini in a close flyby about three weeks prior to Cassini’s<br />

1536<br />

1537<br />

Saturn Orbit Insertion (SOI) in July 2004. The mosaic shows a densely cratered surface. Higherresolution<br />

data revealed that the high frequency <strong>of</strong> craters extends with a steep slope down to craters<br />

1538 only tens <strong>of</strong> meters in diameter (Porco et al., 2005). Image source: http://photojournal.jpl.nasa.gov,<br />

1539<br />

1540<br />

image identification PIA06064.<br />

1541 the medium-sized moons Mimas, Enceladus, Tethys, Dione, Rhea, Titan, Hyperion, Iapetus and<br />

1542 Phoebe, indicating an intense post-accretional collision process relatively early in the geological<br />

1543 history <strong>of</strong> the satellites. Cassini observations added to the number <strong>of</strong> large basins and ring structures<br />

1544 mainly on Rhea, Dione and Iapetus. Differences in the topographical expression <strong>of</strong> these impact<br />

1545 structures suggest post-impact processes such as relaxation, which may reflect differences in the<br />

1546 thermal history and/or crustal thickness <strong>of</strong> individual satellites (e.g. Schenk and Moore, 2007; Giese<br />

1547 et al., 2008). Although age models are based on different impactor populations, the cratered<br />

1548 surfaces on Iapetus seem to be the oldest in the Saturnian system, whereas cratered plains on the<br />

1549 other satellites seem to be younger (see Dones et al., 2009). One large (48 km in diameter)<br />

1550<br />

1551<br />

stratigraphically young ray crater was found on Rhea (Wagner et al., 2007) (Fig.22).<br />

1552 Fig. 22 A prominent bright ray crater with a diameter <strong>of</strong> 48 km is located approximately at 12.5° S,<br />

1553 112° W on Saturn’s second-largest satellite Rhea. The crater is stratigraphically young, indicated by<br />

1554 the low superimposed crater frequency on its floor and continuous ejecta (Wagner et al., 2008). The<br />

1555 high-resolution mosaic (right) <strong>of</strong> NAC images, shown in context <strong>of</strong> a WAC frame, was obtained by<br />

1556 Cassini ISS during orbit 049 on Aug. 30, 2007. The global view to the left showing the location <strong>of</strong><br />

1557 the detailed view is a single NAC image from orbit 056.<br />

1558<br />

1559 While the small satellites lack tectonic features, the medium-sized bodies experienced rotation-,<br />

1560 tide- and impact-induced deformation and volume changes, which created extensional or<br />

1561 contractional tectonic landforms. Parallel lineaments on Mimas seem to be related to the Herschel<br />

1562 impact event (McKinnon, 1985; Schenk, 1989a), although freezing expansion cannot be ruled out as<br />

1563<br />

1564<br />

the cause (Fig. 23).<br />

1565 Fig. 23 The anti-Saturnian hemisphere <strong>of</strong> Mimas, centered approximately at lat. 22° S, long. 185°<br />

1566 W, is shown in a single image (NAC frame N1501638509) at a resolution <strong>of</strong> 620 m/pixel, taken<br />

1567 during a non-targeted encounter in August 2005. The densely cratered landscape implies a high<br />

1568<br />

1569<br />

surface age. Grooves and lineaments infer weak tectonic on this geologically unevolved body.<br />

1570 Enceladus exhibits the greatest geological diversity in a complex stress and strain history that is<br />

1571 predominantly tectonic and cryovolcanic in origin, with both processes stratigraphically correlated<br />

1572 in space and time, although the energy source is so far not completely understood (e.g. Smith et al.,<br />

1573 1982; Kargel and Pozio, 1996; Porco et al., 2006; Nimmo and Pappalardo, 2006; Nimmo et al.,<br />

1574<br />

2007; Jaumann et al., 2008). Enceladus is discussed in detail in Spencer et al. (2009).<br />

29


1575 Located in densely cratered terrain, the nearly globe-encircling Ithaca Chasma rift system on Tethys<br />

1576 (Fig. 2) is 100 km wide, consists <strong>of</strong> at least two narrower branches towards the south and was<br />

1577 caused either by expansion when the liquid-water interior froze (Smith et al., 1981) or by<br />

1578 deformation by the large impact which formed Odysseus (Moore and Ahern, 1983). However,<br />

1579 crater densities derived from Cassini images suggest Ithaca Chasma to be older than Odysseus<br />

1580 (Giese et al., 2007), so that the rift system may be an endogenic feature.<br />

1581 Dione exhibits a nearly global network <strong>of</strong> tectonic features such as troughs, scarps, ridges and<br />

1582<br />

1583<br />

lineaments partly superimposed on each other (Fig. 24), which correlate with smooth plains<br />

1584 Fig. 24 This image <strong>of</strong> Tethys, taken by the Cassini ISS NA camera in May 2008 in visible light,<br />

1585 shows the old, densely cratered surface <strong>of</strong> Tethys and a part <strong>of</strong> the extensive graben system <strong>of</strong><br />

1586 Ithaca Chasma. The graben is up to 100 km wide and is divided into two branches in the region<br />

1587 shown in this image. Using digital elevation models infers flexural uplift <strong>of</strong> the graben flanks and a<br />

1588 total topography ranging several kilometers from the flanks to the graben floor (Giese et al., 2007).<br />

1589 Image details: ISS frame number N1589080288, scale 1 km/pixel, centered at lat. 42° S, long. 50°<br />

1590<br />

1591<br />

W in orthographic projection.<br />

1592 (e.g. Plescia, 1983; Moore, 1984; Stephan et al., 2008). The global orientation <strong>of</strong> Dione’s tectonic<br />

1593 pattern is non-random, suggesting a complex endogenic history <strong>of</strong> extension followed by<br />

1594 compression. Cassini images indicate that some <strong>of</strong> the grabens are fresh, which implies that tectonic<br />

1595<br />

1596<br />

activity may have persisted into geologically recent times (Wagner et al., 2006). Cassini also<br />

revealed a crater-lineament relationship on Dione (Moore et al., 2004; Schenk and Moore, 2007),<br />

1597 indicating exogenic-induced tectonic activity in the early geological history <strong>of</strong> the moon. Thus,<br />

1598 Dione’s crust not only experienced stress caused by different exogenic and endogenic processes,<br />

1599 such as ancient impacts, despinning and volume change over a long period <strong>of</strong> time, but also more<br />

1600 recent tidal stress. As on Dione, wispy accurate albedo markings on Rhea are associated with an<br />

1601 extensional fault system (Moore and Schenk et al., 2007; Wagner et al., 2007), although Rhea<br />

1602<br />

1603<br />

seems less endogenically evolved than the other moons (Fig. 25).<br />

1604 Fig. 25 Map <strong>of</strong> Dione’s trailing hemisphere showing an age sequence <strong>of</strong> troughs, graben and<br />

1605 lineaments, inferred from various crosscutting relationships (Wagner et al., 2009). Three age groups<br />

1606 <strong>of</strong> troughs and graben can be distinguished: Carthage and Clusium Fossae are oldest, cut by the<br />

1607 younger systems <strong>of</strong> Eurotas and Palatine Chasmata, which in turn are truncated by Padua Chasmata<br />

1608 which are the youngest. Several systems <strong>of</strong> lineaments with various trends, either parallel or radial<br />

1609 (the latter previously assumed to be a bright ray crater named Cassandra) can be mapped which<br />

1610<br />

1611<br />

appear to be younger than the troughs and graben.<br />

1612 Parallel north-south trending lineaments are thought to be the result <strong>of</strong> extensional stress followed<br />

1613 by compression (Moore et al., 1985) and appear in Cassini observations as poorly developed ridges.<br />

1614 This suggests that they are older than the grabens identified in the wispy terrain that appear to have<br />

1615 formed more recently (Wagner et al., 2006).<br />

1616 Iapetus’ equatorial ridge (Porco et al., 2005a; Giese et al., 2008) (Fig. 7) is densely cratered, making<br />

1617<br />

1618<br />

it comparable in age to any other terrains on this moon (Schmedemann et al., 2008). Models explain<br />

the origin <strong>of</strong> the ridge either by endogenic shape and/or spin state changes (e.g. Porco et al., 2005a;<br />

1619 Giese et al., 2008) or exogenic causes, with the ridge accumulating from ring material left over<br />

1620 from the formation <strong>of</strong> proto-Iapetus (Ip, 2006). On Iapetus, diurnal tidal stress is rather low,<br />

1621 indicating that the ridge is old, and its preservation suggests the formation <strong>of</strong> a thick crust early in<br />

1622 the moon's history (Castillo-Rogez et al., 2007).<br />

1623 Enceladus, where cryovolcanism is definitely active (Porco et al., 2006), is discussed in Spencer et<br />

1624 al. (2009). Titan, where some evidence <strong>of</strong> cryovolcanism exists (Lopes et al., 2007a; Nelson et al.,<br />

1625 2009a,b) is discussed in Jaumann et al. (2009) and Sotin et al. (2009).<br />

1626 There is no direct evidence <strong>of</strong> cryovolcanic processes (Geissler, 2000) on the other Saturnian<br />

1627 satellites. Although it has been suggested in pre-Cassini discussions that the plains on Dione<br />

1628 (Plescia, and Boyce, 1982; Plescia, 1983; Moore, 1984) and Tethys (Moore and Ahern, 1984) are<br />

1629 endogenic in origin, Cassini has not directly observed any cryovolcanic activity on these satellites.<br />

1630 The plains correlate with tectonic features and exhibit distinct boundaries with more densely<br />

1631 cratered units (Moore and Schenk, 2007), which indicate that the plains are a stratigraphically<br />

1632 younger formation. In addition, there is evidence <strong>of</strong> material around Dione, Tethys and even Rhea<br />

1633<br />

(Burch et al., 2007; Jones et al., 2008; Clark et al., 2008a), but its origin is unclear. There is no<br />

30


1634 evidence <strong>of</strong> cryovolcanism on Iapetus. Although the plains on Dione and Tethys are younger than<br />

1635 the surrounding terrain, cryovolcanic activity must have occurred in ancient times, assuming that<br />

1636 such processes caused these surface units.<br />

1637 All medium-sized icy satellites appear to have a debris layer, as may be inferred from (1) the non-<br />

1638 pristine morphology <strong>of</strong> most craters; (2) the presence <strong>of</strong> groups <strong>of</strong> large boulders and debris lobes<br />

1639 seen in the highest-resolution ISS images; and (3) derived thermal-physical properties consistent<br />

1640 with loose particulate materials. With the exception <strong>of</strong> Enceladus, which is addressed in Spencer et<br />

1641 al. (2009), these debris layers were generated by impacts and their ejecta on the other satellites. The<br />

1642 upper limiting case for a given satellite is crustal disruption from extreme global impact-induced<br />

1643 seismic events. A way to evaluate the depth <strong>of</strong> fracture in such events is to assume that these events<br />

1644 were energetic enough to open fractures in the brittle ice (brittle on the timescale <strong>of</strong> these events)<br />

1645 composing the interiors <strong>of</strong> the target satellite, and that, during the impact event, the lithostatic<br />

1646 pressure everywhere was less than the brittle tensile strength <strong>of</strong> cold ice. Brittle fracture caused by<br />

1647 global seismicity, all else being equal, extends deeper into smaller satellites, given that larger<br />

1648 satellites are sheltered from dynamic rupture by their interior pressure. These being transient events,<br />

1649 a simple estimate <strong>of</strong> the realm where brittle fracture may be expected to be induced, and visibly<br />

1650 sustained, can be obtained from global seismicity by applying an expression (e.g. Turcott and<br />

1651<br />

1652<br />

Schubert, 2002) <strong>of</strong> the lithostatic pressure at radius Rp, and comparing this to the brittle tensile<br />

strength <strong>of</strong> water ice at timescales in which extreme global impact-induced seismic events produce<br />

1653 energy sufficient to break ice mechanically. The determination <strong>of</strong> the fracture depth is based on<br />

1654<br />

1655<br />

laboratory-derived dynamic values <strong>of</strong> ~1MPa (Lange and Ahrens, 1987; Hawkes and Mellor, 1972).<br />

-3<br />

With ρ the (bulk) density <strong>of</strong> the body (approximated as 1000kg/m ), G the gravitational constant,<br />

1656 and R the radius <strong>of</strong> the satellite, the depth <strong>of</strong> the fracture can be constrained. Based on this, the<br />

1657<br />

1658<br />

value <strong>of</strong> R - Rp for Mimas is ~20 km (~10% <strong>of</strong> its radius), ~10 km (~2% <strong>of</strong> radius) for Tethys, and<br />

~5 km (~0.5% <strong>of</strong> radius) for Rhea. This is a measure <strong>of</strong> the depth to which it is reasonable to expect<br />

1659 the surface layer (i.e., the megaregolith) to be intensely fractured. Of course, a body may be<br />

1660 fractured to greater depths by large impacts, and the resulting pore space may not easily be<br />

1661 eliminated. Deep porosity is most easily maintained on smaller and presumably colder bodies such<br />

1662 as Mimas (Eluszkiewicz, 1990). Thus, for Mimas we expect that relatively recent major impact<br />

1663 events, such as Herschel, might exploit pre-existing fracture patterns in the 'megaregolith' which<br />

1664 might penetrate deep into the body rather than focus energy, as has been suggested for the<br />

1665 formation <strong>of</strong> the plains <strong>of</strong> Tethys (e.g. Bruesch and Asphaug, 2004; Moore et al., 2004).<br />

1666 Consequently, the lack <strong>of</strong> antipodal focusing or axial symmetry about a radius through Herschel in<br />

1667 the lineament pattern on Mimas may be unsurprising, if that pattern is indeed an expression <strong>of</strong><br />

1668 tectonic movements initiated by the Herschel impact.<br />

1669 The impact-gardened regolith <strong>of</strong> most middle-sized icy satellites can be assumed to be many meters<br />

1670 thick if the regolith in the lunar highlands is any guide. Calculations <strong>of</strong> impact-generated regolith,<br />

1671 based on crater densities observed in Voyager images, yielded mean depths <strong>of</strong> 500m for Mimas,<br />

1672 740m for Dione, 1600m for Tethys and 1900m for Rhea (Veverka et al., 1986).<br />

1673 This regolith will be further distributed by gravity-driven mass wasting (non-linear creep, i.e.<br />

1674 disturbance-driven sediment transport that increases nonlinearly with the slope due to granular<br />

1675 creep (e.g. Roering et al., 2001; Forsberg-Taylor et al., 2004)) which reduces high-standing relief on<br />

1676<br />

1677<br />

craters (e.g. lowering <strong>of</strong> crater rims and central peaks) and tectonic features (e.g. the brinks <strong>of</strong><br />

scarps) as well as filling depressions (e.g. Howard, 2007). The process <strong>of</strong> non-linear creep, when<br />

1678 not competing with other processes, produces a landscape, which has a repetitive consistent<br />

1679 appearance. Impact-induced shaking and perhaps diurnal and seasonal thermal expansion facilitate<br />

1680 such mass wasting. Note that this common expression <strong>of</strong> mass wasting on regolith-covered slopes<br />

1681 assumes no topography-maintaining precipitation <strong>of</strong> frosts on local topographical highs, as<br />

1682 discussed below.<br />

1683 Bedrock erosion might be caused by the sublimation <strong>of</strong> mechanically cohesive interstitial volatiles.<br />

1684 The icy bedrock would crumble, move down slope through nonlinear creep and form mantling<br />

1685 slopes, perhaps leaving only topographical highs (crests) composed <strong>of</strong> bedrock exposed. This<br />

1686 situation would occur if the initial bedrock landscape disaggregates to form debris at a finite rate, so<br />

1687 that exposure <strong>of</strong> bedrock gradually diminishes, lasting longest at local crests. Such a situation might<br />

1688 explain the landscape <strong>of</strong> Hyperion. If there is re-precipitation <strong>of</strong> volatiles in the form <strong>of</strong> frosts on<br />

1689 local topographical highs, these frosts would armor the topographical highs against further<br />

1690 desegregation. It has been proposed that landscapes evolved in this fashion on Callisto (Moore et al,<br />

1691 1999; Howard and Moore, 2008). Thermally driven global and local frost migration and<br />

1692<br />

redistribution may occur on Iapetus (Spencer et al., 2005; Giese et al., 2008), as previously<br />

31


1693 discussed. This surficial frost redistribution may occur regardless <strong>of</strong> whether volatiles in the<br />

1694<br />

1695<br />

bedrock <strong>of</strong> Iapetus sublimate and contribute to bedrock erosion.<br />

1696 21.7 Conclusions<br />

1697 The icy satellites <strong>of</strong> Saturn show great and unexpected diversity. In terms <strong>of</strong> size, they cover a range<br />

1698 from ~1500 km (Rhea) to ~270 km (Hyperion) and even smaller 'icy rocks' <strong>of</strong> less than a kilometer<br />

1699 in diameter. The icy satellites <strong>of</strong> Saturn <strong>of</strong>fer an unrivalled natural laboratory for understanding the<br />

1700 geology <strong>of</strong> diverse satellites and their interaction with a complex planetary system.<br />

1701 Cassini has executed several targeted flybys over icy satellites, and more are planned for the future.<br />

1702 In addition, there are numerous other opportunities for observation during non-targeted encounters,<br />

1703 usually at greater distances. Cassini used these opportunities to perform multi-instrument<br />

1704 observations <strong>of</strong> the satellites so as to obtain the physical, chemical and structural information<br />

1705 needed to understand the geological processes that formed these bodies and governed their<br />

1706 evolution.<br />

1707 All icy satellites exhibit densely cratered surfaces. Although there have been some attempts to<br />

1708 correlate craters and potential impactor populations, no consistent interpretation <strong>of</strong> the crater<br />

1709 chronology in the Saturnian system has been developed so far (Dones et al., 2009). However, most<br />

1710 <strong>of</strong> the surfaces are old in a stratigraphical sense. The number <strong>of</strong> large basins discovered by Cassini<br />

1711 was greater than expected (e.g. Giese et al., 2008). Differences in their relaxation state provide<br />

1712 some information about crustal thickness and internal heat flows (Schenk and Moore, 2007). The<br />

1713<br />

1714<br />

absence <strong>of</strong> basin relaxation on Iapetus is consistent with a thick lithosphere (Giese et al., 2008),<br />

whereas the relaxed Evander basin on Dione exhibits floor uplift to the level <strong>of</strong> the original surface<br />

1715 (Schenk and Moore, 2007), suggesting higher heat flow.<br />

1716 The wide variety in the extent and timing <strong>of</strong> tectonic activity on the icy satellites defies any simple<br />

1717 explanation, and our understanding <strong>of</strong> Saturnian satellite tectonics and cryovolcanism still is at an<br />

1718 early stage. The total extent <strong>of</strong> deformation varies wildly, from Iapetus and Mimas (barely<br />

1719 deformed) to Enceladus (heavily deformed and currently active), and there is no straightforward<br />

1720 relationship to predicted tidal stresses (e.g. Nimmo and Pappalardo, 2006; Matsuyama and Nimmo,<br />

1721 2007, 2008; Schenk et al., 2008; Spencer et al., 2009). Extensional deformation is common and<br />

1722 <strong>of</strong>ten forms globally coherent patterns, while compressional deformation appears rare (e.g. Nimmo<br />

1723 and Pappalardo, 2006; Matsuyama and Nimmo, 2007, 2008; Schenk et al., 2008; Spencer et al.,<br />

1724 2009). These global patterns are suggestive <strong>of</strong> global mechanisms, such as despinning or<br />

1725 reorientation, which probably occurred early in the satellites’ histories (e.g. Thomas et al., 2007);<br />

1726 present-day diurnal tidal stresses are unlikely to be important, except on Enceladus (e.g. Spencer et<br />

1727 al., 2009) and (perhaps) Mimas and Dione (e.g. Moore and Schenk, 2007). Ocean freezing gives<br />

1728 rise to isotropic extensional stresses; modulation <strong>of</strong> these stresses by pre-existing weaknesses (e.g.<br />

1729 impact basins) may explain some <strong>of</strong> the observed long-wavelength tectonic patterns (e.g. Moore<br />

1730 and Schenk, 2007). Except for Enceladus, there is as yet no irrefutable evidence <strong>of</strong> cryovolcanic<br />

1731 activity in the Saturnian system, either from surface images and topography or from remote sensing<br />

1732 <strong>of</strong> putative satellite atmospheres. The next few years will hopefully see a continuation <strong>of</strong> the flood<br />

1733 <strong>of</strong> Cassini data, and will certainly see a concerted effort to characterize and map in detail the<br />

1734 tectonic structures discussed here. Understanding the origins <strong>of</strong> these structures and the histories <strong>of</strong><br />

1735<br />

1736<br />

the satellites will require both geological and geophysical investigations and probably provide a<br />

challenge for many years to come.<br />

1737 Although the surfaces <strong>of</strong> the Saturnian satellites are predominately composed <strong>of</strong> water ice,<br />

1738 reflectance spectra indicate coloring agents (dark material) on all surfaces (e.g. Fink and Larsen<br />

1739 1975; Clark et al., 1984, 1986; Roush et al., 1995; Cruikshank et al., 1998a; Owen et al., 2001;<br />

1740 Cruikshank et al., 2005; Clark et al., 2005, 2008a, 2009; Filacchione et al. 2007, 2008). Recent<br />

1741 Cassini data provide a greater range in reflected solar radiation at greater precision, show new<br />

1742 absorption features not previously seen in these bodies, and allow new insights into the nature <strong>of</strong> the<br />

1743 icy-satellite surfaces (Buratti et al., 2005b; Clark et al., 2005; Brown et al., 2006; Jaumann et al.,<br />

1744<br />

1745<br />

2006; Cruikshank et al., 2005, 2007, 2008; Clark et al., 2008a, b, 2009). Besides H2O ice, CO2<br />

absorptions are found on all medium-sized satellites (Buratti et al., 2005b; Clark et al., 2005;<br />

1746<br />

1747<br />

Cruikshank et al., 2007; Brown et al., 2006; Clark et al., 2008a), which make CO2 a common<br />

molecule on icy surfaces as well. Various spectral features <strong>of</strong> the dark material match those seen on<br />

1748 Phoebe, Iapetus, Hyperion, Dione and Epimetheus as well as in the F-ring and the Cassini Division,<br />

1749 implying that the material has a common composition throughout the Saturnian system (Clark et al.,<br />

1750 2008a, 2009). Dark material diminishes closer to Saturn, which might be due to surface<br />

1751<br />

contamination <strong>of</strong> the inner moons by E-ring particles and some chemical alteration processes (Clark<br />

32


1752 et al., 2008a). Although the general composition <strong>of</strong> the dark material in the Saturnian system has not<br />

1753<br />

1754<br />

yet been determined, it is known that it at least contains CO2, bound water, organics and, tentatively,<br />

ammonia (Clark et al., 2008a). In addition, a blue scattering peak that dominates the dark spectra<br />

1755 (Clark et al., 2008a, 2009) can be modelled by nano-phase hematite mixed with ice and dark<br />

1756 particles. This led Clark et al. (2008b, 2009) to infer that sub-micron iron particles from meteorites,<br />

1757 which oxidize (explaining spectral reddening) and become mixed with ice could best explain all the<br />

1758 colors and spectral shapes observed in the system. Iron has also been detected in E-ring particles<br />

1759 and in particles escaping the Saturnian system (Kempf et al., 2005). Previous explanations for the<br />

1760 dark material were based on tholins (e.g. Cruikshank et al., 2005), but this is inconsistent with the<br />

1761 lack <strong>of</strong> marked absorptions due to CH in the 3μm to 4μm region as well as other major absorptions at<br />

1762 wavelengths longer than 2μm. Iron particles are consistent with previous Cassini studies that<br />

1763 suggested an external origin for the dark material on Phoebe, Iapetus and Dione (Clark et al., 2005,<br />

1764 2008a,b, 2009). Another remarkable issue relating to dark material is the albedo dichotomy on Iapetus,<br />

1765 whose leading hemisphere is the darkest surface in the Saturnian system. In contrast to the outermost<br />

1766 satellite, Phoebe, also covered by dark material, Iapetus' dark hemisphere is spectrally reddish, making<br />

1767 it doubtful whether Phoebe is the source <strong>of</strong> the dark material. Clark et al. (2008a, 2009) explained the<br />

1768 color discrepancy by red fine-grained dust (


1810<br />

1811<br />

1812<br />

1813<br />

1814<br />

1815<br />

1816<br />

1817<br />

1818<br />

material has been observed throughout the system, and there is some evidence that this<br />

contamination might also be a relatively recent process.<br />

Acknowledgements: We gratefully acknowledge the long years <strong>of</strong> work by the entire Cassini team that allowed these<br />

data <strong>of</strong> the Saturnian satellites to be obtained. We should also like to thank NASA, ESA, ASI, DLR, CNES and JPL for<br />

providing support for the international Cassini team. We thank K. Stephan, R. Wagner and M. Langhans for data<br />

processing. The discussions <strong>of</strong> reviewers C. Chapman and D. Domingue are highly appreciated. Part <strong>of</strong> this work was<br />

performed at the DLR Institute <strong>of</strong> <strong>Planetary</strong> Research, with support provided by the Helmholz Alliance ’<strong>Planetary</strong><br />

Evolution and Life’ and the Jet Propulsion Laboratory under contract to the National Aeronautics and Space<br />

1819 Administration.<br />

We dedicate this work to Steve Ostro, who provided outstanding scientific contributions to the exploration <strong>of</strong> the<br />

1820<br />

1821<br />

1822<br />

Saturnian System.<br />

1823 References<br />

1824 Allison, M.L. and Clifford, S.M., 1987, Ice-covered water volcanism on Ganymede, J. Geophys.<br />

1825<br />

1826<br />

Res. 92, 7865-7876.<br />

1827<br />

1828<br />

Beeman, M., Durham, W.B., Kirby, S.B., 1988, Friction <strong>of</strong> ice, J. Geophys. Res. 93, 7625-7633.<br />

1829 Bell, J.F., Clark, R.N., and McCord, T.B.,1981, Reflection Spectra <strong>of</strong> Pluto and Three Distant<br />

1830<br />

1831<br />

Satellites, Bull <strong>of</strong> the Am. Astro. Soc., 11, 570 (abstract).<br />

1832 Bell, J.F., D.P. Cruikshank, and M.J. Gaffey, 1985, The Composition and Origin <strong>of</strong> the Iapetus<br />

1833 Dark Material, Icarus 61, 192-207.<br />

1834<br />

1835 Bills, B.G. and Ray, R.D., 2000, Energy dissipation by tides and librations in synchronous satellites,<br />

31 st 1836<br />

1837<br />

Lunar and <strong>Planetary</strong> Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #1709.<br />

1838 Black, G.J., 1997, Chaotic rotation <strong>of</strong> Hyperion and modeling the radar properties <strong>of</strong> the icy<br />

1839 Galilean satellites as a coherent backscatter effect. Ph.D. thesis, Cornell <strong>University</strong>.<br />

1840<br />

1841 Black, G. J., Campbell, D. B., Ostro, S. J., 2001a, Icy Galilean satellites: 70 cm radar results from<br />

1842<br />

1843<br />

Arecibo, Icarus 151, 160-166.<br />

1844 Black, G. J., Campbell, D. B., Nicholson, P. D., 2001b, Icy Galilean satellites: Modeling radar<br />

1845<br />

1846<br />

reflectivities as a coherent backscatter effect, Icarus 151, 167-180.<br />

1847 Black, G. J., Campbell, D. B., Carter, L. M., Ostro, S. J., 2004, Radar detection <strong>of</strong> Iapetus, Science<br />

1848<br />

1849<br />

304, 553.<br />

1850 Black, G. J., Campbell, D. B., Carter, 2007, Arecibo radar observations <strong>of</strong> Rhea, Dione, Tethys, and<br />

1851 Enceladus, Icarus 191, 702–711.<br />

1852<br />

1853 Brown, R.H., Baines, K.H., Bellucci, G., Bibring, J.-P., Buratti, B.J., Bussoletti, E., Capaccioni, F.,<br />

1854 Cerroni, P., Clark, R.N., Coradini, A., Cruikshank, D.P., Drossart, P., Formisano, V., Jaumann, R.,<br />

1855 Langevin, Y., Matson, D.L., McCord, T.B., Miller, E., Nelson, R.M., Nicholson, P.D., Sicardy, B.,<br />

1856 and Sotin, C., 2004, The Cassini Visual and Infrared Mapping Spectrometer Investigation, Space<br />

1857 Science Reviews 115, 111-168.<br />

1858<br />

1859 Brown, R. H., Clark, R.N., Buratti, B.J., Cruikshank, D.P., Barnes, J.W., Mastrapa, R.M.E., Bauer,<br />

1860 J., Newman, S., Momary, T., Baines, K.H., Bellucci, G., Capaccioni, F., Cerroni, P., Combes, M.,<br />

1861 Coradini, A., Drossart, P., Formisano, V., Jaumann, R., Langevin, Y., Matson, D.L., McCord, T.B.,<br />

1862 Nelson, R.M., Nicholson, P.D., Sicardy, B., Sotin, C., 2006, Composition and Physical Properties<br />

1863 <strong>of</strong> Enceladus Surface, Science, 311, 1425-1428.<br />

1864<br />

1865<br />

1866<br />

Brown, R.H., Lebreton, J.-P., Waite, H., (eds.), 2009, Titan from Cassini-Huygens, Springer 2009.<br />

1867<br />

1868<br />

Buratti, B. J., 1984, Voyager disk resolved photometry <strong>of</strong> the Saturnian satellites, Icarus 59, 426-435.<br />

34


1869 Buratti, B. J., 1985, Application <strong>of</strong> a radiative transfer model to bright icy satellites, Icarus 61, 208-<br />

1870 217.<br />

1871<br />

1872 Buratti, B. J., and Veverka, J., 1984, Voyager photometry <strong>of</strong> Rhea, Dione, Tethys, Enceladus, and<br />

1873 Mimas, Icarus 58, 254-264.<br />

1874<br />

1875 Buratti, B. J., and Veverka, J. 1985, Photometry <strong>of</strong> rough planetary surfaces: The role <strong>of</strong> multiple<br />

1876 scattering, Icarus 64, 320-328.<br />

1877<br />

1878 Buratti, B.J. 1988, Enceladus: Implications <strong>of</strong> its unusual photometric properties, Icarus 75, 113-<br />

1879 126.<br />

1880<br />

1881 Buratti, B.J., Mosher, J., Johnson, T.V. 1990, Albedo and color maps <strong>of</strong> the Saturnian satellites,<br />

1882<br />

1883<br />

Icarus, 87, 339-357.<br />

1884 Buratti, B.J., and Mosher, J., 1995, The dark side <strong>of</strong> Iapetus: New evidence for an exogenous origin,<br />

1885<br />

1886<br />

Icarus 115, 219-227.<br />

1887 Buratti, B.J., Mosher, J.A., Nicholson, P.D., McGhee, C., French, R., 1998, Photometry <strong>of</strong> the<br />

1888 Saturnian satellites during Ring Plane Crossing, Icarus 136, 223-231.<br />

1889<br />

1890 Buratti, B. J., Hicks, M. D., Tryka, K. A., Sittig, M. S. & Newburn, R. L., 2002, High-resolution<br />

1891 0.33-0.92 µm spectra <strong>of</strong> Iapetus, Hyperion, Phoebe, Rhea, Dione, and D-type asteroids: How are<br />

1892 they related? Icarus 155, 375-381.<br />

1893<br />

1894 Buratti, B., Hicks, M., Davies, A., 2005a, Spectrophotometry <strong>of</strong> the small satellites <strong>of</strong> Saturn and<br />

1895<br />

1896<br />

their relationship to Iapetus, Phoebe, and Hyperion, Icarus 175, 490-495.<br />

1897 Buratti, B. J., Cruikshank, D.P., Brown, R.H., Clark, R.N., Bauer, J.M., Jaumann, R., McCord,<br />

1898 T.B., Simonelli, D.P., Hibbitts, C.A., Hansen, G.A., Owen, T.C., Baines, K.H., Bellucci, G.,<br />

1899 Bibring, J.-P., Capaccioni, F., Cerroni, P., Coradini, A., Drossart, P., Formisano, V., Langevin, Y.,<br />

1900 Matson, D.L., Mennella, V., Nelson, R.M., Nicholson, P.D., Sicardy, B., Sotin, C., Roush, T.L.,<br />

1901 Soderlund, K., and Muradyan, A., 2005b, Cassini Visual and Infrared Mapping Spectrometer<br />

1902<br />

1903<br />

Observations <strong>of</strong> Iapetus: Detection <strong>of</strong> CO2, Astrophysical Journal, 622, L149-L152.<br />

1904<br />

1905<br />

Buratti, B.J., Soderlund, K., Bauer, J., Mosher, J.A., Hicks, M.D., Simonelli, D.P., Jaumann, R.,<br />

Clark, R. N., Brown, R. H., Cruikshank, D. P., Momary, T., 2008, Infrared (0.83-5.1µm)<br />

1906 Photometry <strong>of</strong> Phoebe from the Cassini Visual Infrared Mapping Spectrometer, Icarus 193, 309-<br />

1907<br />

1908<br />

322.<br />

1909 Buratti, B.J., Bauer, J., Hicks, M.D., Mosher, J. A., Filacchione, G., Momary, T. R., Baines, K.<br />

1910 Brown, R. H ., Clark, R. N., Nicholson, P. D. 2009a, Cassini spectra and photometry 0.25-5.1 µm <strong>of</strong><br />

1911<br />

1912<br />

the small inner satellites <strong>of</strong> Saturn, Icarus, submitted.<br />

1913<br />

1914 Buratti, B. J., Hillier, J., Mosher, J., DeWet, S., Abramson, L., Akhter, N., Clark, R. N., Brown,<br />

1915 R.H., Baines, K., Nicholson, P., 2009b, Opposition surges on the icy satellites <strong>of</strong> Saturn as seen by<br />

1916 Cassini, Icarus, submitted.<br />

1917<br />

1918 Bruesch, L.S., Asphaug, E., 2004, Modeling global impact effects on middle-sized icy bodies:<br />

1919<br />

1920<br />

Applications to Saturn’s moons, Icarus 168, 457-466<br />

1921 Burns, J. A., Lamy, P.L., Soter, S., 1979, Radiation forces on small particles in the solar system,<br />

1922<br />

1923<br />

Icarus 40, 1-48.<br />

1924 Burch, J.L, Goldstein, J., Lewis, W. S., Young, D.T., Coates, A.J., Dougherty, M.K., André, N.,<br />

1925 2007, Tethys and Dione as sources <strong>of</strong> outward-flowing plasma in Saturn’s magnetosphere, Nature<br />

1926<br />

1927<br />

447, 833-835. doi:10.1038/nature05906<br />

35


1928 Cassen, P., Reynolds, R.T., Peale, S.J., 1979, Is there liquid water on Europa? Geophys. Res. Lett.<br />

1929 6, 731-734.<br />

1930<br />

1931 Cassini, G. D., 1672, Phil. Trans. 12, 831. Quoted in Alexander, A. F. O’D., 1962, The Planet<br />

1932 Saturn, McMillan, New York, 474 p.<br />

1933<br />

1934 Carlson, R., Smythe, W., Baines, K., Barbinas, E., Burns, R., Calcutt, S., Calvin, W., Clark, R.,<br />

1935 Danielson, G., Davies, A., Drossart, P., Encrenaz, T., Fanale, F., Granahan, J., Hansen, G., Hererra,<br />

1936 P., Hibbitts, C., Hui, J., Irwin, P., Johnson, T., Kamp, L., Kieffer, H., Leader, F., Lopes-Gautier, R.,<br />

1937 Matson, D., McCord, T., Mehlman, D., Ocampo, A., Orton, G., Roos-Serote, M., Segura, M.,<br />

1938 Shirley, J., Soderblom, L., Stevenson, A., Taylor, F., Weir, A., Weissman, P., 1996, Near-Infrared<br />

1939 Spectroscopy and Spectral Mapping <strong>of</strong> Jupiter and the Galilean Satellites: First Results from<br />

1940<br />

1941<br />

Galileo's Initial Orbit, Science, 274, 385-388.<br />

1942 Carlson, R.W., Anderson, M.S., Johnson, R.E., Smythe, W.D., Hendrix, A.R., Barth, C.A.,<br />

1943 Soderblom, L.A., Hansen, G.B., McCord, T.B., Dalton, J.B., Clark, R.N., Shirley, H.J., Ocampo,<br />

1944<br />

1945<br />

A.C., Matson, D.L., 1999, Hydrogen peroxide on the surface <strong>of</strong> Europa, Science, 283, 2062-2064.<br />

1946 Castillo-Rogez, J.C., Matson, D.L., Sotin, C., Johnson, T.V., Lunine, J.I., Thomas, P.C., 2007,<br />

1947 Iapetus’ geophysics: Rotation rate, shape and equatorial ridge, Icarus 190, 179-202.<br />

1948<br />

1949 Chapman C.R., and McKinnon, W.B., 1986, Cratering <strong>of</strong> planetary satellites. In: Satellites (Burns,<br />

1950 J.A., Matthews, M.S. (eds.)), Univ. <strong>of</strong> Arizona Press, Tucson, 492-580.<br />

1951<br />

1952 Chapman, C.R., 2004, Space weathering <strong>of</strong> asteroid surfaces. Annual. Rev. <strong>Earth</strong> Planet Science 32,<br />

1953 539-567.<br />

1954<br />

1955 Chen, E.M.A., and Nimmo, F., 2008, Implications from Ithaca Chasma on the thermal and orbital<br />

1956<br />

1957<br />

history <strong>of</strong> Tethys, Geophys. Res. Lett. 35, L19203.<br />

1958 Clark, R.N., Brown, R.H., Owensby, P.D., Steele, A., 1984, Saturn’s satellites: near-infrared<br />

1959 spectrophotometry (0.65–2.5µm) <strong>of</strong> the leading and trailing sides and compositional implications,<br />

1960<br />

1961<br />

Icarus 58, 265–281.<br />

1962 Clark, R.N., Fanale, F.P., Gaffey, M.J., 1986, Surface composition <strong>of</strong> natural satellites. In: Satellites<br />

1963<br />

1964<br />

(Burns, J.A., Matthews, M.S. (eds.)), Univ. <strong>of</strong> Arizona Press, Tucson, 437–491.<br />

1965 Clark, R.N., Brown, R.H., Jaumann, R., Cruikshank, D.P., Nelson, R.M., Buratti, B.J., McCord,<br />

1966 T.B., Lunine, J., Baines, K.H., Bellucci, G., Bibring, J.-P., Capaccioni, F., Cerroni, P., Coradini, A.,<br />

1967 Formisano, F., Langevin, Y., Matson, D.L., Mennella, V., Nicholson, P.D., Sicardy, B., Sotin, C.,<br />

1968 Hoefen, T. M., Curchin, J.M., Hansen, G., Hibbits, K., Matz, K.-D., 2005, Compositional maps <strong>of</strong><br />

1969 Saturn's moon Phoebe from imaging spectroscopy, Nature 435, 66-69, doi:10.1038/nature03558.<br />

1970<br />

1971 Clark, R.N., Curchin, J.M., Jaumann, R., Cruikshank, D.P., Brown, R.H., Hoefen, T.M., Stephan,<br />

1972 K., Moore, J.M., Buratti, B.J., Baines, K.H., Nicholson, P.D., Nelson, R.M., 2008a, Compositional<br />

1973 Mapping <strong>of</strong> Saturn's Satellite Dione with Cassini VIMS and Implications <strong>of</strong> Dark Material in the<br />

1974 Saturn System, Icarus, 193, 372-386.<br />

1975<br />

1976 Clark, R. N., Cruikshank, D.P., Jaumann, R., Filacchione, G., Nicholson, P.D., Brown, R.H.,<br />

1977 Stephan, K., Hedman, M., Buratti, B.J., Curchin, J.M., T.M., Hoefen, Baines, K.H., and Nelson, R.,<br />

1978 2008b, Compositional Mapping <strong>of</strong> Saturn's Rings and Icy Satellites with Cassini VIMS, Saturn<br />

1979<br />

1980<br />

After Cassini-Huygens, London, July, 2008.<br />

1981 Clark R.N., Cruikshank, D.P., Jaumann, R., Brown, R.H., Cruchin, J.M., Hoefen, T.M., Stephan,<br />

1982 K., Dalle Ore, C., Buratti, J.B., Filacchione, G., Baines, K.H., Nicholson, P.D., 2009, The<br />

1983<br />

1984<br />

Composition <strong>of</strong> Iapetus: Mapping Results from Cassini VIMS, Icarus, submitted.<br />

36


1985 Collins, G.C., McKinnon, W.B., Moore, J.M., Nimmo, F., Pappalardo, R.T., Prockter, L.M.,<br />

1986 Schenk, P.M., 2009, Tectonics <strong>of</strong> the outer planet satellites, in <strong>Planetary</strong> Tectonics, (R.A. Schultz<br />

1987 and T.R. Watters (eds.)), Cambridge Univ. Press, submitted.<br />

1988<br />

1989 Cook, A. F. and Franklin, F., 1970, An explanation for the light curve <strong>of</strong> Iapetus, Icarus 13, 282-<br />

1990 291.<br />

1991<br />

1992 Crawford, G.D. and Stevenson, D.J., 1988, Gas-driven water volcanism and the resurfacing <strong>of</strong><br />

1993 Europa, Icarus 73, 66-79.<br />

1994<br />

1995 Cruikshank, D. P., and Brown, R. H., 1982, Surface composition and radius <strong>of</strong> Hyperion, Icarus 50,<br />

1996<br />

1997<br />

82-87.<br />

1998 Cruikshank, D. P., Bell, J. F., Gaffey, M. J., Brown, R. H., Howell, R., Beerman, C., and Rognstad,<br />

1999<br />

2000<br />

M. 1983, The Dark Side <strong>of</strong> Iapetus, Icarus 53, 90-104.<br />

2001 Cruikshank, D.P., Veverka, J., Leb<strong>of</strong>sky. L.A., 1984, Satellites <strong>of</strong> Saturn: The optical properties, In<br />

2002<br />

2003<br />

Saturn, T. Gehrels and M.S. Matthews (eds.), <strong>University</strong> <strong>of</strong> Arizona Press, 640-667.<br />

2004 Cruikshank, D.P., Brown, R.H., Calvin, W., Roush, T.L., Bartholomew, M.J., 1998a, Ices on the<br />

2005<br />

2006<br />

satellites <strong>of</strong> Jupiter, Saturn, and Uranus. In: Solar System Ices, Schmitt, B., de Bergh, C., Festou,<br />

M. (eds.), Kluwer Academic, Dordrecht, 579–606.<br />

2007<br />

2008 Cruikshank, D.P., Owen, T.C., Ore, C.D., Geballe, T.R., Roush, T.L., de Bergh, C., Sandford, S.A.,<br />

2009 Poulet, F., Benedix, G.K., and Emery, J.P., 2005, A spectroscopic study <strong>of</strong> the surfaces <strong>of</strong> Saturn’s<br />

2010<br />

2011<br />

large satellites: H2O ice, tholins, and minor constituents, Icarus 175, 268-283.<br />

2012 Cruikshank, D.P., Dalton, J.B., Dalle Ore, C.M., Bauer, J., Stephan, K., Filacchione, G., Hendrix,<br />

2013 C.J., Hansen, C.J., Coradini, A., Cerroni, P., Tosi, F., Capaccioni, F., Jaumann, R., Buratti, B.J.,<br />

2014 Clark, R.N., Brown, R.H., Nelson, R.M., McCord, T.B., Baines, K.H., Nicholson, P.D., Sotin, C.,<br />

2015 Meyer, A.W., Bellucci, G., Combes, M., Bibring, J.-P., Langevin, Y., Sicardy, B., Matson, D.L.,<br />

2016 Formisano, V., Drossart, P., and Mennella, V., 2007, Surface composition <strong>of</strong> Hyperion, Nature 448,<br />

2017<br />

2018<br />

54-57, doi:101038/nature05948.<br />

2019 Cruikshank, D.P., Wegryn, E., Dalle Ore, C.M., Brown, R.H., Baines, K.H., Bibring, J.-P., Buratti,<br />

2020 B.J., Clark, R.N., McCord, T.B., Nicholson, P.D., Pendleton, Y.J., Owen, T.C.,<br />

2021 Filacchione, G., and the VIMS Team, 2008, Hydrocarbons on Saturn’s Satellites Iapetus and<br />

2022<br />

2023<br />

Phoebe, Icarus 193, 334-343, doi:10.1016/j.icarus.2007.04.036.<br />

2024 Cuzzi, J.N., Dones, L., Clark, R.N., French, R.G., Johnson, T.V., Marouf, E.A., Spilker, L.J.,<br />

2025 2009, Ring composition, in Saturn after Cassini-Huygens, Saturn from Cassini-Huygens, (M.<br />

2026 Dougherty, L. Esposito and S. Krimigis, (eds.)), Springer, 2009.<br />

2027<br />

2028 Czechowski, L., and Leliwa-Kopystyński, J., 2008, The Iapetus's ridge: Possible explanations <strong>of</strong> its<br />

2029 origin, Adv. Space Res. 42, 61-69.<br />

2030<br />

2031 Dougherty, M.K., Khurana, K.K., Neubauer, F.M., Russell, C.T., Saur, J., Leisner, J.S., and<br />

2032 Burton, M.E., 2006, Identification <strong>of</strong> a Dynamic Atmosphere at Enceladus with the Cassini<br />

2033<br />

2034<br />

Magnetometer, Science 311, 1406-1409.<br />

2035 Dougherty, M.K., Esposito, L.W., Krimigis, S.M., 2009, Overview, Saturn from Cassini-Huygens,<br />

2036<br />

2037<br />

(M. Dougherty, L. Esposito and S. Krimigis, (eds.)), Springer, 2009<br />

2038 Denk, T., Matz, K.-D., Roatsch, T. Wolf, U., Wagner, R.J., Neukum, G., and Jaumann, R., 2000,<br />

2039 Iapetus (1): Size, Topography, Surface Structures, Craters. 31th Lunar and <strong>Planetary</strong> Science<br />

2040<br />

2041<br />

Conference, Houston, abstract #1596.<br />

2042 Denk, T., and Neukum, G., 2000, Iapetus (2): Dark-side Origin. 31th Lunar and <strong>Planetary</strong> Science<br />

2043<br />

Conference, Houston, abstract #1660.<br />

37


2044<br />

2045 Denk, T., Neukum, G., Roatsch, T., McEwen, A.S., Turtle, E.P., Thomas, P.C., Helfenstein, P.,<br />

2046 Wagner, R.J., Porco, C.C., Perry, J.E., Giese, B., Johnson, T.V., Veverka, J., and the Cassini ISS,<br />

2047 2005, Team: First Imaging Results From the Iapetus B/C Flyby <strong>of</strong> the Cassini Spacecraft. 36th<br />

2048 Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #2268.<br />

2049<br />

2050 Denk, T., G. Neukum, T. Roatsch, B. Giese, R.J. Wagner, P. Helfenstein, J.A. Burns, E.P. Turtle,<br />

2051 T.V. Johnson, C.C. Porco, 2006, Iapetus: Leading/ trailing side color dichotomy, and the origin <strong>of</strong><br />

2052 the brightness dichotomy. EGU06-A-08352, Vienna (Austria).<br />

2053<br />

http://www.cosis.net/abstracts/EGU06/08352/EGU06-J-08352.pdf<br />

2054<br />

2055 Denk, T., Neukum, G., Schmedemann, N., Roatsch, T., Wagner, R.J., Giese, B., Perry, J.E.,<br />

2056 Helfenstein, P., Turtle, E.P., Porco, C.C., 2008, Iapetus imaging during the targeted flyby <strong>of</strong> the<br />

2057<br />

2058<br />

Cassini spacecraft, 39th Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #2533.<br />

2059 Denk, T., and Spencer, J.R., 2008, Iapetus: A Two-step Explanation for its Unique Global<br />

2060 Appearance. BAAS 40 (3), 61.04.<br />

2061 http://www.abstractsonline.com/viewer/viewAbstract.asp?CKey={105D4425-F206-476D-BDF1-<br />

2062 D155267B27BD}&MKey={35A8F7D5-A145-4C52-8514-<br />

2063 0B0340308E94}&AKey={AAF9AABA-B0FF-4235-8AEC-<br />

2064<br />

2065<br />

74F22FC76386}&SKey={2B16A2FD-1D5E-4F2A-B234-3EDCA2413097}<br />

2066 Domingue, D. L., Lockwood, G. W., Thompson, D. T., 1995, Surface textural properties <strong>of</strong> icy<br />

2067 satellites: A comparison between Europa and Rhea, Icarus 115, 228-249.<br />

2068<br />

2069 Dones, L., Chapman, C., Melosh, H.J., McKinnon, W.B., Neukum, G., Zahnle, K., 2009, Impact<br />

2070 cratering and age determination, In Saturn after Cassini Huygens, Saturn from Cassini-Huygens,<br />

2071<br />

2072<br />

(M. Dougherty, L. Esposito and S. Krimigis, (eds.)), Springer, 2009.<br />

2073 Durham, W.B. and Stern, L.A., 2001, Rheological properties <strong>of</strong> water ice – applications to satellites<br />

2074<br />

2075<br />

<strong>of</strong> the outer planets, Ann. Rev. <strong>Earth</strong> Planet. Sci. 29, 295-330.<br />

2076 Elachi, C., Allison, M., Borgarelli, L., Encrenaz, P., Im, E., Janssen, M. A., Johnson, W. T. K.,<br />

2077 Kirk, R. L., Lorenz, R. D., Lunine, J. I., Muhleman, D. O., Ostro, S. J., Picardi, G., Posa, F., Roth,<br />

2078 L. E., Seu, R., Soderblom, L. A., Vetrella, S., Wall, S. D., Wood, C. A., and Zebker, H. A., 2004,<br />

2079<br />

2080<br />

RADAR: The Cassini Titan Radar Mapper, Space Science Reviews, 115, 71–110.<br />

2081 Ellsworth, K. and Schubert, G., 1983, Saturn’s icy satellites – thermal and structural models, Icarus<br />

2082<br />

2083<br />

54, 490-510, 1983.<br />

2084<br />

2085<br />

Eluszkiewicz, J.,1990, Compaction and internal structure <strong>of</strong> Mimas, Icarus 84, 215-225.<br />

2086<br />

2087<br />

Esposito, L. W., Barth, C. A.; Colwell, J. E.; Lawrence, G. M.; McClintock, W. E.; Stewart, A. I.<br />

F.; Keller, H. U.; Korth, A.; Lauche, H.; Festou, M. C.; Lane, A. L.; Hansen, C. J.; Maki, J. N.;<br />

2088 West, R. A.; Jahn, H.; Reulke, R.; Warlich, K.; Shemansky, D. E.; Yung, Y. L., 2004, The Cassini<br />

2089 Ultraviolet Imaging Spectrograph Investigation, Space Science Reviews, 115, 299-361,<br />

2090<br />

2091<br />

doi:10.1007/s11214-004-1455-8.<br />

2092 Fagents, S.A., Greeley, R.; Sullivan, R. J.; Pappalardo, R. T.; Prockter, L. M.; The Galileo SSI<br />

2093 Team, 2000, Cryomagmatic mechanisms for the formation <strong>of</strong> Rhadamanthys Linea, triple band<br />

2094<br />

2095<br />

margins, and other low-albedo features on Europa, Icarus 144, 54-88, doi:10.1006/icar.1999.6254.<br />

2096 Fagents, S.A., 2003, Considerations for effusive cryovolcanism on Europa: the post-Galileo<br />

2097<br />

2098<br />

perspective, J. Geophys. Res. 108, 5139.<br />

2099 Filacchione, G., Capaccioni, F., McCord, T.B., Coradini, A., Cerroni, P., Bellucci, G., Tosi, F.,<br />

2100 D’Aversa, E., Formisano, V., Brown, R.H., Baines, K.H., Bibring, J.-P., Buratti, B.J., Clark, R.N.,<br />

2101 Combes, M., Cruikshank, D.P., Drossart, P., Jaumann, R., Langevin, Y., Matson, D.L., Mennella,<br />

2102<br />

V., Nelson, R.M., Nicholson, P.D., Sicardy, B., Sotin, C., Hansen, G., Hibbitts, K., Showalter, M.,<br />

38


2103 and Newmann, S., 2007, Saturn’s icy satellites investigated by Cassini VIMS. I. Full-disk<br />

2104<br />

2105<br />

properties: 0.35 – 5.1 µm reflectance spectra and phase curves, Icarus 186, 259-290.<br />

2106 Filacchione, G., Capaccioni, F., Tosi, F., Cerroni, P., McCord, T.B., Baines, K.H., Bellucci, G.,<br />

2107 Brown, R.H., Buratti, B.J., Clark, R.N., Cruikshank, D.P., Cuzzi, J.N., Jaumann, R., Stephan,K.,<br />

2108 Matson, D.L., Nelson, R.M., Nicholson, P.D., 2008, Analysis <strong>of</strong> the Saturnian icy satellites full-disk<br />

2109<br />

2110<br />

spectra by Cassini-VIMS, Saturn After Cassini-Huygens, London, July, 2008.<br />

2111 Fink, U., and Larson, H.P., 1975, Temperature dependence <strong>of</strong> the water-ice spectrum between 1 and<br />

2112<br />

2113<br />

4 microns: application to Europa, Ganymede, and Saturn’s rings, Icarus 24, 411–420.<br />

2114 Flasar, F.M., V.G. Kunde, M.M. Abbas, R.K. Achterberg, P. Ade, A. Barucci, B.Bézard, G.L.<br />

2115 Bjoraker, J.C. Brasunas, S. Calcutt, R. Carlson, C.J. Césarsky, B.J. Conrath, A. Coradini, R.<br />

2116 Courtin, A. Coustenis, S. Edberg, S. Edgington, C. Ferrari, T. Fouchet, D. Gautier, P.J. Gierasch, K.<br />

2117 Grossman, P. Irwin, D.E. Jennings, E. Lellouch, A.A. Mamoutkine, A. Marten, J.P. Meyer, C.A.<br />

2118 Nixon, G.S. Orton, T.C. Owen, J.C. Pearl, R. Prangé, F. Raulin, P.L. Read, P.N. Romani, R.E.<br />

2119 Samuelson, M.E. Segura, M.R. Showalter, A.A. Simon-Miller, M.D. Smith, J.R. Spencer, L.J.<br />

Spilker , 2120 F.W.Taylor, Exploring The Saturn System In The Thermal Infrared: The Composite<br />

2121 Infrared Spectrometer, 2004, Space Science Reviews 115, 169-297, DOI 10.1007/s11214-004-<br />

2122 1454-9.<br />

2123<br />

2124 Forsberg-Taylor, N.K., A.D. Howard, and R.A. Craddock (2004), Crater degradation in the Martian<br />

2125 highlands: Morphometric analysis in the Sinus Sabaeus region and simulation modeling suggest<br />

2126 fluvial processes, J. Geophys. Res 109, doi:10.1029/2004JE002242.<br />

2127<br />

2128 Gehrels, T., and Matthews, M.S., 1984, Saturn, <strong>University</strong> <strong>of</strong> Arizona Press.<br />

2129<br />

2130 Geissler, P.E., 2000, Cryovolcanism in the outer solar system, in Encyclopedia <strong>of</strong> Volcanoes, ed. H.<br />

2131 Sigurdsson, 785-800, Academic, San Diego.<br />

2132<br />

2133 Giese, B., Wagner, R., Neukum, G., Helfenstein, P., Thomas, P.C., 2007, Tethys: Lithospheric<br />

2134 thickness and heat flux from flexurally supported topography at Ithaca Chasma, Geophys. Res. Lett.<br />

2135<br />

2136<br />

34, L21203.<br />

2137 Giese, B., Denk, T., Neukum, G., Roatsch, T., Helfenstein , P., Thomas, P.C., Turtle, E.P.,<br />

2138<br />

2139<br />

McEwen, A., Porco, C.C., 2008, The topography <strong>of</strong> Iapetus’ leading side, Icarus 193, 359–371.<br />

2140 Goguen, J. 1981, A theoretical and experimental investigation <strong>of</strong> the photometric functions <strong>of</strong><br />

2141<br />

2142<br />

particulate surfaces, Ph.D. Thesis Cornell Univ., Ithaca, NY.<br />

2143 Greenberg, R. and Weidenschilling, S.J., 1984, How fast do Galilean satellites spin? Icarus 58, 186-<br />

2144 196.<br />

2145<br />

2146 Greenberg, R. Geissler, P.; Hoppa, G.; Tufts, B. R.; Durda, D. D.; Pappalardo, R.; Head, J. W.;<br />

2147 Greeley, R.; Sullivan, Robert; Carr, M. H., 1998, Tectonic processes on Europa: Tidal stresses,<br />

2148 mechanical response, and visible features, Icarus 135, 64-78, doi:10.1006/icar.1998.5986<br />

2149<br />

2150 Greenberg, R. and Geissler, P., 2002, Europa’s dynamic icy crust, Meteorit. Planet. Sci. 37, 1685-<br />

2151 1710.<br />

2152<br />

2153 Hamilton, D. P., and Burns, J. A., 1994, Origin <strong>of</strong> Saturn’s E-ring: Self sustained, naturally, Science<br />

2154<br />

2155<br />

264, 550-553.<br />

2156 Hansen, C.J., Esposito, L., Stewart, A.I.F., Colwell, J., Hendrix, A., Pryor, W., Shemansky, D., and<br />

2157<br />

2158<br />

West, R., 2006, Enceladus' Water Vapor Plume, Science 311, 1422-1425<br />

2159<br />

2160<br />

Hapke, B. 1981, Bidirectional reflectance spectroscopy, 1. Theory, J. Geophys. Res. 86, 4571-4586.<br />

39


2161 Hapke, B. 1984, Bidirectional reflectance spectroscopy, 3. Correction for macroscopic roughness,<br />

2162 Icarus 59, 41-59.<br />

2163<br />

2164 Hapke, B. 1986, Bidirectional reflectance spectroscopy, 4. The extinction coefficient and the<br />

2165 opposition effect, Icarus 67, 264-280.<br />

2166<br />

2167 Hapke, B. 1990, Coherent backscatter and the radar characteristics <strong>of</strong> outer planet satellites, Icarus 88,<br />

2168 407-417.<br />

2169<br />

2170 Hartmann, W. K., 1980, Surface evolution <strong>of</strong> two component stone/ice bodies in the Jupiter region,<br />

2171 Icarus 44, 441-453.<br />

2172<br />

2173 Hawkes, I. and Mellor, M.,1972, Deformation and fracture <strong>of</strong> ice under uniaxial stress, Journal <strong>of</strong><br />

2174<br />

2175<br />

Glaciology 11, 61, 103-131.<br />

2176 Helfenstein P. and Parmentier, E.M., 1983, Patterns <strong>of</strong> fracture and tidal stresses on Europa, Icarus<br />

2177<br />

2178<br />

53, 415-430.<br />

2179 Helfenstein P. and Parmentier, E.M., 1985, Patterns <strong>of</strong> fracture and tidal stresses due to<br />

2180 nonsynchronous rotation – implications for fracturing on Europa, Icarus 61, 175-184.<br />

2181<br />

2182 Helfenstein, P., and Shepard, M.K.,1998, Submillimeter-scale topography <strong>of</strong> the lunar Regolith,<br />

2183 Icarus 124, 262-267.<br />

2184<br />

2185 Hendrix, A. R. Barth, C. A. and Hord, C. W. and Lane, A. L., 1998, Europa: Disk-Resolved<br />

2186 Ultraviolet Measurements Using the Galileo Ultraviolet Spectrometer, Icarus 135, 79-94.<br />

2187<br />

2188 Hendrix, A. R., Barth, C. A., Hord, C. W., 1999, Ganymede's ozone-like absorber: Observations by<br />

2189<br />

2190<br />

the Galileo Ultraviolet Spectrometer, J. Geophys. Res. 104, 14169-14178.<br />

2191 Hendrix, A. R. and C. J. Hansen, 2008a, The albedo dichotomy <strong>of</strong> Iapetus measured at UV<br />

2192<br />

2193<br />

wavelengths, Icarus 193, 344-351.<br />

2194 Hendrix, A. R. and C. J. Hansen, 2008b, Ultraviolet observations <strong>of</strong> Phoebe from the Cassini UVIS,<br />

2195<br />

2196<br />

Icarus 193, 323-333.<br />

2197 Hendrix, A. R. and C. J. Hansen., 2009a. The UV albedo <strong>of</strong> Enceladus: Implications for surface<br />

2198 composition, Icarus submitted<br />

2199<br />

2200<br />

th<br />

Hendrix, A. R. and Hansen, C.J., 2009b. Iapetus: New results from Cassini UVIS, 40 Lunar and<br />

2201<br />

2202<br />

<strong>Planetary</strong> Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #2200.<br />

2203 Hendrix, A. R. and Buratti, B.J., 2009, Multi-wavelength photometry <strong>of</strong> icy Saturnian satellites: A<br />

fist look, 40 th 2204<br />

Lunar and <strong>Planetary</strong> Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract<br />

2205 2438#.<br />

2206<br />

2207<br />

2208<br />

Hibbitts, C.A., McCord, T.B., Hansen, G.B., 2000, Distributions <strong>of</strong> CO2 and SO2 on the surface <strong>of</strong><br />

Callisto, J. Geophys. Res. 105, 22541-22557.<br />

2209<br />

2210 Hibbitts, C.A., Pappalardo, R.T., Hansen, G.B., McCord, T.B., 2003, Carbondioxide on Ganymede,<br />

2211<br />

2212<br />

J. Geophys. Res. 108, doi:10.1029/2002JE001956.5036.<br />

2213 Hillier, J. and Squyres, S.W., 1991, Thermal stress tectonics on the satellites <strong>of</strong> Saturn and Uranus,<br />

2214<br />

2215<br />

J. Geophys. Res. 96, 15665-15674.<br />

2216 Hoppa, G.V., Tufts, B.R., Greenberg, R., Geissler, P.E., 1999a, Formation <strong>of</strong> cycloidal features on<br />

2217<br />

2218<br />

Europa, Science 285, 1899-1902.<br />

40


2219 Hoppa, G., Tufts, B.R., Greenberg, R., Geissler, P., 1999b, Strike-slip faults on Europa: Global<br />

2220 shear patterns driven by tidal stress, Icarus 141, 287-298.<br />

2221<br />

2222 Horak, H. G. 1950, Diffuse reflection by planetary atmospheres, Ap. J. 112, 445-463.<br />

2223<br />

2224 Horányi, M., Juhász, A., Morfill, E. G., 2008, Large-scale structure <strong>of</strong> Saturn's E-ring, Geophys.<br />

2225 Res. Lett., 35, L04203, doi:10.1029/2007GL032726.<br />

2226<br />

2227 Howard, A.D., 2007, Simulating the development <strong>of</strong> Martian highland landscapes through the<br />

2228 interaction <strong>of</strong> impact cratering, fluvial erosion, and variable hydrologic forcing, Geomorphology, 9,<br />

2229 332–363.<br />

2230<br />

2231 Howard, A.D. and Moore, J.M., 2008, Sublimation-driven erosion on Callisto: A landform<br />

2232<br />

2233<br />

simulation model test, Geophys. Res. Let. 35, L03203, doi:101029/2007GL032618.<br />

2234 Hurford, T.A., Helfenstein, P., Hoppa, G.V., Greenberg, R., Bills, B.G., 2007, Eruptions arising<br />

2235<br />

2236<br />

from tidally controlled periodic openings <strong>of</strong> rifts on Enceladus, Nature 447, 292-294.<br />

2237 Ip, W.-H., 2006. On a ring origin <strong>of</strong> the equatorial ridge <strong>of</strong> Iapetus. Geophys. Res. Lett. 33,<br />

2238 doi:10.1029/2005GL025386. L16203.<br />

2239<br />

2240<br />

2241<br />

Irvine, W. 1966, The shadowing effect in diffuse radiation, J. Geophys. Res. 71, 2931-2937.<br />

2242 Jacobson, R. A. and French, R. G., 2004, Orbits and Masses <strong>of</strong> Saturn's Coorbital and F-ring<br />

2243<br />

2244<br />

Shepherding Satellites, Icarus 172, 382-387.<br />

2245 Jacobson, R.A. Antreasian, P. G., Bordi, J. J., Criddle, K. E., Ionasescu, R., Jones, J. B., Meek, M.<br />

2246 C., Owen, W. M., Jr., Roth, D. C., Roundhill, I. M., Stauch, J. R., Cassini Navigation, 2004, The<br />

2247 Orbits <strong>of</strong> the Major Saturnian Satellites and the Gravity Field <strong>of</strong> the Saturnian System, Bulletin <strong>of</strong><br />

2248<br />

2249<br />

the American Astronomical Society 36(3), 1097.<br />

2250 Jacobson, R. A., Antreasian, P. G., Bordi, J. J., Criddle, K. E., Ionasescu, R., Jones, J. B.,<br />

2251 Mackenzie, R. A., Pelletier, F. J., Owen Jr., W. M., Roth, D. C., and Stauch, J. R., 2005, The<br />

2252<br />

2253<br />

gravity field <strong>of</strong> the Saturnian system and the orbits <strong>of</strong> the major Saturnian satellites, BAAS 37, 729.<br />

2254<br />

2255<br />

Jacobson, R.A., 2006, SAT252 - JPL satellite ephemeris.<br />

2256 Jacobson, R.A., 2007, SAT270, SAT271 - JPL satellite ephemerides.<br />

2257<br />

2258 Janes, D.M. and Melosh, H.J., 1988, Sinker tectonics – an approach to the surface <strong>of</strong> Miranda, J.<br />

2259<br />

2260<br />

Geophys. Res. 93, 3217-3143.<br />

2261 Jarvis, K.S., Vilas, F., Larson, S.M., and Gaffey, M.J., 2000, Are Hyperion and Phoebe Linked to<br />

2262<br />

2263<br />

Iapetus? Icarus, 146, 125-132.<br />

2264 Jaumann, R., Stephan, K., Brown, R.H., Buratti, B.J., Clark, R.N., McCord, T.B., Coradini, A.,<br />

2265 Capaccioni, P., Filacchione, G., Cerroni, P., Baines, K.H., Bellucci, G., Bibring, J.P., Combes, M.,<br />

2266 Cruikshank, D.P., Drossart, P., Formisano, V., Langevin, Y., Matson, D.L., Nelson, R.M.,<br />

2267 Nicholson, P.D., Sicardy, B., Sotin, C., Soderblom, L.A., Griffith, C., Matz, K.D., Roatsch T.,<br />

2268 Scholten, F., and Porco, C.C., 2006, High-resolution Cassini-VIMS mosaics <strong>of</strong> Titan and the icy<br />

2269<br />

2270<br />

Saturnian satellites, <strong>Planetary</strong> and Space Science 54, 1146-1155.<br />

2271 Jaumann, R., Stephan, K., Hansen, G.B., Clark, R.N., Buratti, B.J., Brown, R.H., Baines, K.H.,<br />

2272 Bellucci, G., Coradini, A., Cruikshank, D.P., Griffith, C.A., Hibbits, C.A., McCord, T.B., Nelson,<br />

2273<br />

R.M., Nicholson, P.D., Sotin, C., and Wagner, R., 2008, Distribution <strong>of</strong> Icy Particles across<br />

41


2274 Enceladus’ Surface as derived from Cassini-VIMS Measurements, Icarus, 193, 407-419,<br />

2275<br />

2276<br />

doi:10.1016/j.icarus.2007.09.013.<br />

2277 Jaumann, R., Kirk, R.L., Lorenz, R.D., Lopez, R.M.G., St<strong>of</strong>an, E.R., Turtle, E.P., Keller, H.U.,<br />

2278 Wood, C.A., Sotin, C., Soderblom, L.A., Tomasko, M., 2009, Geology and surface Processes on<br />

2279 Titan, Titan from Cassini-Huygens, (R. Brown, J.-P. Lebreton and H. Waite,<br />

2280<br />

2281<br />

(eds.)), Springer, 2009.<br />

2282<br />

2283<br />

Johnson, R. E., 1997, Polar “caps” on Ganymede and Io revisited, Icarus 128, 469-471.<br />

2284<br />

2285<br />

2286<br />

Johnson, R. E., and Jesser, W.A., 1997, O2/O3 microatmospheres in the surface <strong>of</strong> Ganymede,<br />

Astrophys. J. 480, L79-L82.<br />

2287 Johnson, R. E., and Quickenden, T.I., 1997, Photolysis and radiolysis <strong>of</strong> water ice on outer solar<br />

2288<br />

2289<br />

system bodies, J. Geophys. Res. 102, 10985-10996.<br />

2290 Johnson, R. E., Killen, R.M., Waite, J.H., Lewis, W.S., 1998, Europa’s surface composition and<br />

2291<br />

2292<br />

sputter-produced ionosphere, Geophys. Res. Lett. 25, 3257-3260.<br />

2293 Johnson, R. E., Carlson, R.W., Cooper, J.F., Paranicas, C., Moore, M.H., Wong, M.C., 2004,<br />

2294<br />

2295<br />

Radiation effects on the surfaces <strong>of</strong> the Galilean satellites, In Jupiter- the Planet, Satellites and<br />

Magnetosphere, F. Bagenal, T. Dowling and W. B. McKinnon, (eds.) Cambridge <strong>University</strong>,<br />

2296<br />

2297<br />

Cambridge.<br />

2298 Johnson, R. E., Fama, M., Liu, M., Baragiola, R. A., Sittler, E. C., Smith, H. T., 2008, Sputtering <strong>of</strong><br />

2299 ice grains and icy satellites in Saturn’s inner magnetosphere, <strong>Planetary</strong> and Space Science 56, 1238-<br />

2300<br />

2301<br />

1243.<br />

2302 Johnson, T.V., 1998, Introduction to icy satellite geology, in Schmitt, B., C. De Bergh and M.<br />

2303<br />

2304<br />

Festou, eds., Solar System Ices, Kluwer, Dordrecht, 511-524.<br />

2305 Johnson, T.V., Estrada, E, Lissauer, J., 2009, Origin <strong>of</strong> the Saturnian System, Saturn from Cassini-<br />

2306<br />

2307<br />

Huygens, (M. Dougherty, L. Esposito and S. Krimigis, (eds.)), Springer, 2009.<br />

2308 Jones, G.H., Roussos, E., Krupp, N., Paranicas, C., Woch, J., Lagg, A., Mitchell, D.G., Krimigis,<br />

2309 S.M., and Dougherty, M.K., 2006, Enceladus' Varying Imprint on the Magnetosphere <strong>of</strong> Saturn,<br />

2310<br />

2311<br />

Science 311, 1412-1415.<br />

2312 Jones, G.H., E. Roussos, N. Krupp, U. Beckmann, A. J. Coates, F. Crary, I. Dandouras, V. Dikarev,<br />

2313 M. K. Dougherty, P. Garnier, C. J. Hansen, A. R. Hendrix, G. B. Hospodarsky, R. E. Johnson, S.<br />

2314 Kempf, K. K. Khurana, S. M. Krimigis, H. Krüger, W. S. Kurth, A. Lagg, H. J. McAndrews, D. G.<br />

2315 Mitchell, C. Paranicas, F. Postberg, C. T. Russell, J. Saur, M. Seiß, F. Spahn, R. Srama, D. F.<br />

2316<br />

2317<br />

Strobel, R. Tokar, J.-E. Wahlund, R. J. Wilson, J. Woch, and D. Young, 2008, The dust halo <strong>of</strong><br />

Saturn’s largest icy moon, Rhea, Science 319, 1380-1384. doi:10.1126/science.1151524.<br />

2318<br />

2319 Kargel, J.S., 1991, Brine volcanism and the interior structures <strong>of</strong> asteroids and icy satellites, Icarus<br />

2320 95, 368-390.<br />

2321<br />

2322 Kargel, J.S., Cr<strong>of</strong>t, S.K., Lunine, J.I., and Lewis, J.S., 1991, Rheological properties <strong>of</strong> ammonia-<br />

2323<br />

2324<br />

water liquids and crystal-liquid slurries: Planetological implications, Icarus 89, 93-112.<br />

2325 Kargel, J.S., 1992, Ammonia-water volcanism in icy satellites: Phase relations at 1 atmosphere,<br />

2326 Icarus 100, 556-574.<br />

2327<br />

2328<br />

2329<br />

Kargel, J.S., 1995, Cryovolcanism on the icy satellites, <strong>Earth</strong> Moon Planets 67, 101-113.<br />

2330 Kargel, J.S., and Pozio, S., 1996, The volcanic and tectonic history <strong>of</strong> Enceladus, Icarus 119, 385-<br />

2331<br />

2332<br />

404.<br />

42


2333 Kempf, S., Srama, R., Postberg, F., Burton, M., Green, S. F., Helfert, S. Hillier, J. K., McBride, N.,<br />

2334 McDonnell, J. A. M., Moragas-Klostermeyer, G., Roy, M., and Gru¨n, E., 2005, Composition <strong>of</strong><br />

2335 Saturnian Stream Particles, Science, 307, 1274-1276.<br />

2336<br />

2337 Kempf, S., R. Srama, R, U. Beckmann, and J. Schmidt, J., Saturn's E ring as seen by the Cassini<br />

2338 dust detector, 2008, AGU Fall Meeting, http://www.agu.org/meetings/fm08/fm08-<br />

2339 sessions/fm08_P32A.html.<br />

2340<br />

2341 Kirch<strong>of</strong>f, M, and Schenk, P., 2008, Cratering Records <strong>of</strong> Saturnian Satellites, 39th Lunar and<br />

2342<br />

<strong>Planetary</strong> Science Conference, Houston, abstract # 2234.<br />

2343<br />

2344 Kirch<strong>of</strong>f, M, and Schenk, P., 2009, Cratering Records <strong>of</strong> Saturnian Satellites, Icarus submitted.<br />

2345<br />

2346 Kirk, R.L. and Stevenson, D.J., 1987, Thermal evolution <strong>of</strong> a differentiated Ganymede and<br />

2347<br />

implications for surface features, Icarus 69, 91-134.<br />

2348<br />

2349<br />

2350<br />

Klavetter, J. J., 1989 Rotation <strong>of</strong> Hyperion, I: Observations Astron. J. 97, 570-579.<br />

2351 Lange, M.A and Ahrens, T.J., 1987, The dynamic tensile strength <strong>of</strong> ice and ice-silicate mixtures, J.<br />

2352 Geophys. Res. 88, 1197-1208.<br />

2353<br />

2354 Lanzerotti, L. J., Brown, W. L., Marcantonio, K. J., Johnson, R. E., 1984, Production <strong>of</strong> ammonia-<br />

2355<br />

2356<br />

depleted surface layers on the saturnian satellites by ion sputtering, Nature 312, 139-140.<br />

2357 Lee, J., Buratti, B. J., Mosher, J. A., Hicks, M. D, 2009, The Relative Surface Roughness <strong>of</strong> the<br />

2358<br />

2359<br />

dark sides <strong>of</strong> Iapetus from the 2005 flyby, Icarus submitted.<br />

2360 Leith, A. C., and McKinnon, W.B., 1996, Is there evidence for polar wander on Europa? Icarus 120,<br />

2361<br />

2362<br />

387– 398.<br />

2363 Lopes, R.M.C., Mitchell, K.L., St<strong>of</strong>an, E.R., Lunine, J.I., Lorenz, R., Paganelli, F., Kirk, R.L.,<br />

2364 Wood, C.A., Wall, S.D., Robshaw, L.E., Fortes, A.D., Neish, C.D., Radebaugh, J., Reffet, E., Ostro,<br />

2365 S.J., Elachi, C., Allison, M.D., Anderson, Y., Boehmer, R., Boubin, G., Callahan, P., Encrenaz, P.,<br />

2366 Flamini, E., Francescetti, G., Gim, Y., Hamilton, G., Hensley, S., Janssen, M.A., Johnson, W.T.K.,<br />

2367 Kelleher, K., Muhleman, D.O., Ori, G., Orosei, R., Picardi, G., Posa, F., Roth, L.E., Seu, R.,<br />

2368 Shaffer, S., Soderblom, L.A., Stiles, B., Vetrella, S., West, R.D., Wye, L., and Zebker, H.A., 2007,<br />

2369 Cryovolcanic features on Titan's surface as revealed by the Cassini Titan Radar mapper, Icarus 187,<br />

2370<br />

2371<br />

395–412, doi:10.1016/j.icarus.2006.09.006.<br />

2372 Lorenz, R.D., B. Stiles, R.L. Kirk, M. Allison, P. Persi del Marmo, L. Iess, J. I. Lunine, S. J. Ostro<br />

2373 and S. Hensley, 2008, Titan's Rotation Reveals an Internal Ocean and Changing Zonal Winds,<br />

2374 Science 319, 1649-1651.<br />

2375<br />

2376 Lumme, K., and Bowell, E., 1981a, Radiative Transfer in the Surfaces <strong>of</strong> Atmosphereless Bodies, I.<br />

2377<br />

2378<br />

Theory. Astron. J. 86, 1694-1704.<br />

2379 Lumme, K., and Bowell, E., 1981b, Radiative Transfer in the Surfaces <strong>of</strong> Atmosphereless Bodies,<br />

2380 II. Interpretation <strong>of</strong> Phase Curves. Astron. J. 86, 1705-1721.<br />

2381<br />

2382 MacKintosh, F. C., and John, S., 1989, Diffusing-wave spectroscopy and multiple scattering <strong>of</strong> light<br />

2383<br />

2384<br />

in correlated random media, Phys. Rev. B 40, 2383-2406.<br />

2385 Manga, M. and Wang, C.Y., 2007, Pressurized oceans and the eruption <strong>of</strong> liquid water on Europa<br />

2386<br />

2387<br />

and Enceladus, Geophys. Res. Lett. 34, L07202.<br />

2388 Marchi, S., Barbieri, C., Dell'Oro, A., and Paolicci, P., 2002, Hyperion-Iapetus: Collisional<br />

2389<br />

2390<br />

Relationships, Astronomy & Astrophysics 381, 1059-1065.<br />

43


2391 Martens, H. R., Reisenfeld, D. B., Williams, J. D., Johnson, R. E., Smith, H. T., Baragiola, R. A.,<br />

2392 Thomsen, M. F., Young, D. T., Sittler, E. C. 2007, Molecular oxygen ions in Saturn’s inner<br />

2393 magnetosphere for the first 24 Cassini orbits, AGU abstract #P43A-1032.<br />

2394<br />

2395 Mastrapa, R.M., Bernstein, M.P., Sandford, S.A., Roush, T.L., Cruikshank, D.P., Dalle Ore, C.M.,<br />

2396 2008, Optical constants <strong>of</strong> amorphous and crystalline H2O-ice in the near infrared from 1.1 to 2.6<br />

2397 µm, Icarus 197, 307-320, DOI:10.1016/j.icarus.2008.04.008, 307-320.<br />

2398<br />

2399 Matson, D.L., Castillo-Roget, J.C, McKinnon, Schubert, G., Sotin, C., 2009, Thermal evolution and<br />

2400 internal structure, Saturn from Cassini-Huygens, (M. Dougherty, L. Esposito and S. Krimigis,<br />

2401 (eds.)), Springer, 2009<br />

2402<br />

2403 Matsuyama, I. and Nimmo, F., 2007, Rotational stability <strong>of</strong> tidally deformed planetary bodies, J.<br />

2404<br />

2405<br />

Geophys. Res. 112, E11003.<br />

2406 Matsuyama, I. and Nimmo, F., 2008,Tectonic patterns on reoriented and despun planetary bodies,<br />

2407<br />

2408<br />

Icarus 195, 459-473.<br />

2409<br />

2410<br />

Matthews, R. A. J., 1992, The darkening <strong>of</strong> Iapetus and the origin <strong>of</strong> Hyperion, Q. J. R. Astron.<br />

Soc. 33, 253-258.<br />

2411<br />

2412<br />

2413<br />

Matthews, R.A.J., 1992, The Darkening <strong>of</strong> Iapetus and the Origin <strong>of</strong> Hyperion, Quarterly Journal <strong>of</strong><br />

the Royal Astronomical Society 33, 253-258.<br />

2414<br />

2415 McDonald, G. D. Thompson, W. R., Heinrich, M., Khare, B. N., Sagan, C., 1994, Chemical<br />

2416 investigation <strong>of</strong> Titan and Triton tholins, Icarus 108, 137-145.<br />

2417<br />

2418 McKinnon, W.B. and Melosh, H.J., 1980, Evolution <strong>of</strong> planetary lithospheres – evidence from<br />

2419 multi-ringed structures on Ganymede and Callisto, Icarus 44, 454-471.<br />

2420<br />

2421 McKinnon,W.B., 1985, Geology <strong>of</strong> icy satellites, in: Klinger, J., Benest, D.,Dollfus, A.,<br />

2422<br />

2423<br />

Smoluchowski, R. (eds.), Ices in the Solar System, Reidel, Dordrecht, 820–856.<br />

2424 McNutt, M.K., 1984, Lithospheric flexure and thermal anomalies, J. Geophys. Res. 89, 1180-1194.<br />

2425<br />

2426 Melosh H.J., 1975, Large impact basins and the Moon’s orientation, <strong>Earth</strong> Planet Sci. Lett. 26, 353-<br />

2427<br />

2428<br />

360.<br />

2429<br />

2430<br />

Melosh, H.J.,1977, Global tectonics <strong>of</strong> a despun planet, Icarus 31, 221-243.<br />

2431<br />

2432<br />

Melosh, H.J., 1980, Tectonic patterns on a tidally distorted planet, Icarus 43, 334-337.<br />

2433<br />

2434<br />

th<br />

Melosh, H.J., and F. Nimmo, 2009, An intrusive dike origin for Iapetus’ enigmatic ridge? 40<br />

Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #2478.<br />

2435<br />

2436 Mendis, D. A., and Axford, W. I., 1974, Satellites and magnetospheres <strong>of</strong> the outer planets, Annual<br />

2437 Reviews <strong>of</strong> <strong>Earth</strong> and Plantary <strong>Sciences</strong> 2, 419-475.<br />

2438<br />

2439 Miyamoto, H., Mitri, G., Showman, A.P., Dohm, J.M., 2005, Putative ice flows on Europa:<br />

2440 Geometric patterns and relation to topography collectively constrain material properties and<br />

2441<br />

2442<br />

effusion rates, Icarus, 177, 413-424.<br />

2443 Moore, J.M. and Ahern, J.L., 1983, The Geology <strong>of</strong> Tethys, JGR, 83, A577-A584.<br />

2444<br />

2445<br />

2446<br />

Moore, J.M., 1984, The tectonic and volcanic history <strong>of</strong> Dione, Icarus 59, 205-220.<br />

2447 Moore, J.M., Horner, V.M., and Greeley, R., 1985, The geomorphology <strong>of</strong> Rhea: Implications for<br />

2448<br />

2449<br />

geologic history and surface processes, J. Geophys. Res., 90, C785-C795.<br />

44


2450 Moore, J.M., Asphaug, E., Morrison, D., Klemaszewski, J.E., Sullivan, R.J., Chuang, F., Greeley,<br />

2451 R., Bender, K.C., Geissler, P.E., Chapman, C.R., Helfenstein, P., Pilcher, C.B., Kirk, R.L., Giese,<br />

2452 B., Spencer, J.R., 1999, Mass movement and landform degradation on the icy Galilean satellites:<br />

2453 Results from the Galileo nominal mission, Icarus 140, 294-312.<br />

2454<br />

2455 Moore, J.M., Schenk, P.M., Bruesch, L.S., Asphaug, E., McKinnon, W.B., 2004, Large Impact<br />

2456 Features on Middle-Sized Icy Satellites, Icarus 171, 421-443.<br />

2457<br />

2458 Morrison, D., Johnson, T. V., Shoemaker, E. M., Soderbloom, L. A., Thomas, P., Veverka, J., and<br />

2459 Smith, B. A. 1984, Satellites <strong>of</strong> Saturn: Geological Perspective, In Saturn (T. Gehrels and M. S.<br />

2460 Matthews Eds.), 609-63, Univ. <strong>of</strong> Arizona Press, Tucson.<br />

2461<br />

2462 Morrison, D., Owen, T., Soderblom, L.A., 1986, The Satellites <strong>of</strong> Saturn. In Satellites (Eds. J.A.<br />

2463<br />

2464<br />

Burns & M.S. Matthews, Univ. Arizona Press, Tucson), 764-801.<br />

2465 Mueller, S. and McKinnon, W.B., 1988, 3-layered models <strong>of</strong> Ganymede and Callisto –<br />

2466<br />

2467<br />

compositions, structures and aspects <strong>of</strong> evolution, Icarus 76, 437-464.<br />

2468 Murchie, S.L. and Head, J.W., 1986, Global reorientation and its effect on tectonic patterns on<br />

2469 Ganymede, Geophys. Res. Lett. 13, 345-348.<br />

2470<br />

2471 Murphy, R. E., Cruikshank, D. P., Morrison, D., 1972, Radii, albedos, and 20-micron brightness<br />

2472 temperatures <strong>of</strong> Iapetus and Rhea, Astrophys. J. Lett 177, L93-L96.<br />

2473<br />

2474 Murray, C.D., and Dermott, S.F., 1999, Solar system dynamics, Cambridge. Univ. Press.<br />

2475<br />

2476 Nelson, R.M., Lane, A. L., Matson, D.L., Veeder, G.J., Buratti, B.J., and Tedesco, E.F., 1987,<br />

2477 Spectral geometric albedos <strong>of</strong> the Galilean satellites from 0.24 to 0.34 micrometers: Observations<br />

2478 with the International Ultrviolett Explorer, Icarus 72, 358 – 380.<br />

2479<br />

2480 Nelson, R.M., Kamp, L.W., Lopes, R.M.C., Matson, D.L., Kirk, R.L., Hapke, B.W., Wall, S.D.,<br />

2481 Boryta, M.D., Leader, F.E., Smythe, W.D.,. Mitchell, K.L., Baines, K.H., Jaumann, R., Sotin, C.,<br />

2482 Clark, C.N., Cruikshank, D.P., Drossart, P., Lunine, J.I., Combes, M., Bellucci, G., Bibring, J.-P.,<br />

2483 Capaccioni, F., Cerroni, P., Coradini, A., Formisano, V., Filacchione, G., Langevin, Y., McCord,<br />

2484 T.B., Mennella, V., Nicholson, P.D., Sicardy, B., Irwin, P.G.J., Pearl, J.C., 2009a, Photometric<br />

2485 changes on Saturn’s Titan: Evidence for active Cryovolcanism, Geophys. Res. Lett. 36, L04202,<br />

2486<br />

2487<br />

doi:10.1029/2008GL036206.<br />

2488 Nelson, R.M., Kamp, L.W., Matson, D.L., Irwin, P.G.J., Baines, K.H., Boryta, M.D., Leader, F.E.,<br />

2489 Jaumann, R., Smythe, W.D., Sotin, C., Clark, C.N., Cruikshank, D.P., Drossart, P., Pearl, J.C.,<br />

2490 Habke, B.W., Lunine, J.I., Combes, M., Bellucci, G., Bibring, J.-P., Capaccioni, F., Cerroni, P.,<br />

2491 Coradini, A., Formisano, V., Filacchione, G., Langevin, R.Y., McCord, T.B., Mennella, V.,<br />

2492<br />

2493<br />

Nicholson, P.D., Sicardy, B., 2009b, Saturn’s Titan: Surface change, ammonia, and implications for<br />

atmospheric and tectonic activity, Icarus 199, 429-441.<br />

2494<br />

2495 Nesvorny, D., Alverellos, J.L., Dones, L., Levison, H.F., 2003, Orbital and Collisional Evolution <strong>of</strong><br />

2496 the Irregular Satellites, Astron. J. 126, 398-429.<br />

2497<br />

2498 Neukum, G., 1985, Cratering records <strong>of</strong> the satellites <strong>of</strong> Jupiter and Saturn, Adv. Space Res. 5,<br />

2499<br />

2500<br />

107-116.<br />

2501 Neukum, G., Wagner, R.J., Denk, T., Porco, C.C., and the Cassini ISS Team, 2005, The cratering<br />

2502 record <strong>of</strong> the Saturnian satellites Phoebe, Tethys, Dione and Iapetus in comparison: first results<br />

from analysis <strong>of</strong> the Cassini ISS imaging data, 36 th 2503 Lunar and <strong>Planetary</strong> Science Conference,<br />

2504<br />

2505<br />

Houston, abstract #2034.<br />

2506 Neukum, G., R. Wagner, T. Denk, and C. C. Porco (2006), Cratering chronologies and ages <strong>of</strong> the<br />

2507<br />

2508<br />

major Saturnian satellites, Geophys. Res. Abstracts 8, abstr. No. EGU06-A-09252.<br />

45


2509 Newman, S. F., Buratti, B.J., Jaumann, R., Bauer, J.M., Momary, T.M., 2007, Hydrogen peroxide<br />

2510 on Enceladus, Ap. J. 670, L143-L146.<br />

2511<br />

2512 Newman, S. F., Buratti, B. J., Brown, R. H., Jaumann, R., Bauer, J., Momary, T. 2008, Photometric<br />

2513 and spectral analysis <strong>of</strong> the distribution <strong>of</strong> crystalline and amorphous ices on Enceladus as seen by<br />

2514 Cassini, Icarus 193, 397-406.<br />

2515<br />

2516 Newman, S. F., Buratti, B. J., Brown, R. H., Jaumann, R., Bauer, J., Momary, T. 2009, Water ice<br />

2517 crystallinity and grain sizes on Dione Icarus, submitted.<br />

2518<br />

2519 Nimmo, F., and Gaidos, E., 2002, Thermal consequences <strong>of</strong> strike-slip motion on Europa. J.<br />

2520<br />

2521<br />

Geophys. Res. 107, 5021, 10.1029/2000JE001476.<br />

2522 Nimmo, F., Pappalardo, R.T., Giese, B., 2002, Elastic thickness and heat flux estimates on<br />

2523<br />

2524<br />

Ganymede, Geophys. Res. Lett. 29, 1158.<br />

2525 Nimmo, F., 2004a, Stresses generated in cooling viscoelastic ice shells: Application to Europa, J.<br />

2526<br />

2527<br />

Geophys. Res. 109, E12001, doi:10.1029/2004JE002347.<br />

2528 Nimmo, F., 2004b, Dynamics <strong>of</strong> rifting and modes <strong>of</strong> extension on icy satellites, J. Geophys.<br />

2529<br />

2530<br />

Res.109, E01003, doi:10.1029/2004JE002347.<br />

2531 Nimmo, F. and Pappalardo, R.T., 2006, Diapir-induced reorientation <strong>of</strong> Saturn’s moon Enceladus,<br />

2532 Nature 441, 614-616.<br />

2533<br />

2534 Nimmo, F., Spencer, R.J., Pappalardo, R.T., Mullen, M.E., 2007, Shear heating as the origin <strong>of</strong> the<br />

2535<br />

2536<br />

plumes and heat flux on Enceladus, Nature 447, 289-291.<br />

2537 Nimmo, F. and Matsuyama, I., 2007, Reorientation <strong>of</strong> icy satellites by impact basins, Geophys. Res.<br />

2538 Lett. 34, L19203.<br />

2539<br />

2540 Noland, M., Veverka, J., Morrison, D., Cruikshank, D.P., Lazarewicz, A.R., Morrison, N.D., Elliot,<br />

2541 J.L., Goguen, J., and Burns J.A., 1974, Six-Color Photometry <strong>of</strong> Iapetus, Titan, Rhea, Dione, and<br />

2542<br />

2543<br />

Tethys, Icarus 23, 334-354.<br />

2544 Noll, K.S., Johnson, R.E., Lane, A.L., Domingue, D.L., Weaver, H.A., 1996, Detection <strong>of</strong> ozone on<br />

2545<br />

2546<br />

Ganymede, Science 273, 341-343.<br />

2547 Noll, K. S., Roush, T.L., Cruikshank, D.P., Johnson, R.E., Pendleton, Y.J., 1997, Detection <strong>of</strong><br />

2548<br />

2549<br />

ozone on Saturn’s satellites Rhea and Dione, Nature 388, 45-47.<br />

2550 Dougherty, M.K., Esposito, L.W., Krimigis, S.M., 2009, Overview, Saturn from Cassini-Huygens,<br />

2551<br />

2552<br />

(M. Dougherty, L. Esposito and S. Krimigis, (eds.)), Springer, 2009<br />

2553 Orton, G.S., Beines, K.H., Cruikshank, D.P., Krimigis, S.M., Miller, E., 2009, Kowledge <strong>of</strong> the<br />

2554 Saturn system prior to Cassini, Saturn from Cassini-Huygens, (M. Dougherty, L. Esposito and S.<br />

2555 Krimigis, (eds.)), Springer, 2009<br />

2556<br />

2557 Ojakangas G. W., and Stevenson D. J., 1989a, Thermal state <strong>of</strong> an ice shell on Europa, Icarus 81,<br />

2558<br />

2559<br />

220-241.<br />

2560 Ojakangas G.W., and Stevenson D. J., 1989b, Polar wander <strong>of</strong> an ice shell on Europa, Icarus 81,<br />

2561 242-270.<br />

2562<br />

2563 Ostro, S.J., Campbell, D.B., Simpson, R.A., Hudson, R.S., Chandler, J.F., Rosema, K.D., Shapiro,<br />

2564 I.I., Standish, E.M., Winkler, R., Yeomans, D.K., 1992, Europa, Ganymede, and Callisto—New<br />

2565<br />

2566<br />

radar results from Arecibo and Goldstone, J. Geophys. Res. 97, 18227–18244.<br />

2567<br />

Ostro, S.J., 1993, <strong>Planetary</strong> radar astronomy, Rev. Mod. Phys. 65, 1235–1279.<br />

46


2568<br />

2569 Ostro, S.J., West, R. D.; Janssen, M. A.; Lorenz, R. D.; Zebker, H. A.; Black, G. J.; Lunine, J. I.;<br />

2570 Wye, L. C.; Lopes, R. M.; Wall, S. D.; Elachi, C.; Roth, L.; Hensley, S.T., Kelleher, K.; Hamilton,<br />

2571 G. A.; Gim, Y.; Anderson, Y. Z.; Boehmer, R. A.; Johnson, W. T. K.; the Cassini RADAR Team,<br />

2572 2006. Cassini Radar observations <strong>of</strong> Enceladus, Tethys, Dione, Rhea, Iapetus, Hyperion, and<br />

2573<br />

Phoebe, Icarus 183, 479–490. doi:10.1016/j.icarus.2006.02.019.<br />

2574<br />

2575 Ostro, S. J., R. D. West, L. C. Wye, H. A. Zebker, M. A. Janssen, B. Stiles, K. Kelleher, Y. Z.<br />

2576 Anderson, R. A. Boehmer, P. Callahan, Y. Gim, G. A. Hamilton, W. T. K. Johnson, C.<br />

2577 Veeramachaneni, R. D. Lorenz, and the Cassini RADAR Team, 2009. New Cassini RADAR<br />

2578 Results for Saturn's Icy Satellites, Icarus, submitted.<br />

2579<br />

2580 Owen, T. C., Cruikshank, D. P., Dalle Ore, C. M., Geballe, T. R., Roush, T. L., de Bergh, C.,<br />

2581 Pendleton, Y. J., and Khare, B. N., 2001, Decoding the domino: The dark side <strong>of</strong> Iapetus, Icarus<br />

2582<br />

2583<br />

149, 160-172.<br />

2584 Pang, K.D., Voge, C.C., Rhoads J.W., and Ajello, J.M., 1984, The E-ring <strong>of</strong> Saturn and satellite<br />

2585<br />

2586<br />

Enceladus. J. Geophys. Res. 89, 9459–9470.<br />

2587 Pappalardo, R.T., and Collins, G.C., 2005, Strained craters on Ganymede, J. Struct. Geol. 27, 827-<br />

2588<br />

2589<br />

838.<br />

2590 Pappalardo, R.T. and Davis, D.M., 2007, Where’s the compression? Explaining the lack <strong>of</strong><br />

2591<br />

2592<br />

contractional structures on icy satellites, Ice, Oceans and Fire, 6080.<br />

2593 Paranicas, C., Mitchell, D.G., Krimigis, S.M., Hamilton, D.C., Roussos, E., Krupp, N., Jones, G.H.,<br />

2594 Johnson, R.E., Cooper, J.F., and Armstrong T.P., 2008, Sources and losses <strong>of</strong> energetic protons in<br />

2595<br />

2596<br />

Saturn’s magnetosphere, Icarus 197, 519-525.<br />

2597 Peters, K. J., 1992, The coherent backscatter effect: A vector formulation acounting for polarization<br />

2598<br />

2599<br />

and absorption effects and small or large scatterers, Phys. Rev. B 46, 801–812.<br />

2600<br />

2601<br />

Petrenko V.F., and Whitworth, R.W., 1999, Physics <strong>of</strong> Ice. Oxford Univ. Press.<br />

2602<br />

2603<br />

2604<br />

Pipes, J.G., Browell, E.V., Anderson, R.C., 1974, Reflectance <strong>of</strong> amorphous-cubic NH3 frosts and<br />

amorphous-hexagonal H2O frosts at 77K from 1400 to 3000A. Icarus 21, 283-291.<br />

2605 Pitman, K.M., Buratti, B. J., Mosher, J. A., Bauer, J. M., Momary, T. W., Brown, R. H., Nicholson,<br />

2606 P. D., Hedman, M. M., 2008, First high solar phase angle observations <strong>of</strong> Rhea using Cassini<br />

2607 VIMS: upper limits on water vapour and geologic activity, Astrophys. J. 160, L65,<br />

2608<br />

2609<br />

doi:10.1086/589745.<br />

2610<br />

2611<br />

2612<br />

Plescia, J.B. and Boyce, J.M., 1982, Crater densities and geological histories <strong>of</strong> Rhea, Dione,<br />

Mimas and Tethys, Nature 295, 285-290.<br />

2613 Plescia, J.B. and Boyce, J.M., 1983, Crater numbers and geological histories <strong>of</strong> Iapetus, Enceladus,<br />

2614<br />

2615<br />

Tethys and Hyperion, Nature 301, 666-670.<br />

2616<br />

2617<br />

Plescia, J.B., 1983, The geology <strong>of</strong> Dione, Icarus 56, 255–277.<br />

2618 Porco, C.C., West, R.A., Squyres, S., McEwen, A., Thomas, P., Murray, C.D., Del Genio, A.,<br />

2619 Ingersoll, A.P., Johnson, T.V., Neukum, G., Veverka, J., Dones, L., Brahic, A., Burns, J.A.,<br />

2620 Haemmerle, V., Knowles, B., Dawson, D., Roatsch, T., Beurle, K., and Owen, W., 2004, Cassini<br />

2621 imaging science: Instrument characteristics and capabilities and anticipated scientific investigations<br />

2622<br />

2623<br />

at Saturn, Space Sci. Rev., 115, 363-497, doi:10.1007/s11214-004-1456-7, 2004.<br />

2624 Porco, C.C., E. Baker, J. Barbara, K. Beurle, A. Brahic, J. A. Burns, S. Charnoz, N. Cooper, D. D.<br />

2625 Dawson, A. D. Del Genio, T. Denk, L. Dones, U. Dyudina, M. W. Evans, B. Giese, K. Grazier, P.<br />

2626<br />

Helfenstein, A. P. Ingersoll, R. A. Jacobson, T. V. Johnson, A. McEwen, C. D. Murray, G.<br />

47


2627 Neukum, W. M. Owen, J. Perry, T. Roatsch, J. Spitale, S. Squyres, P. C. Thomas, M. Tiscareno, E.<br />

2628 Turtle, A. R. Vasavada, J. Veverka, R. Wagner, and R. West, 2005a, Cassini imaging science:<br />

2629<br />

2630<br />

initial results on Phoebe and Iapetus, Science 307, 1237-1242, doi:10.1126/science.1107981.<br />

2631 Porco, C. C., Baker, E., Barbara, J., Beurle, K., Brahic,A., Burns, J.A., Charnoz, S., Cooper, N.,<br />

2632 Dawson, D.D., Del Genio, A.D., Denk, T., Dones, L., Dyudina, U., Evans, M.V., Giese, B.,<br />

2633 Grazier, K., Helfenstein, P., Ingersoll, A.P., Jacobson, R.A., Johnson, T.V., McEwen, A., Murray,<br />

2634 C.D., Neukum, G., Owen, W.M., Perry, J., Roatsch, T., Spitale, J., Squyres, S., Thomas, P.,<br />

2635 Tiscareno, M., Turtle, E., Vasavada, A.R., Veverka, J., Wagner, R., West, R. et al., 2005b, Cassini<br />

2636<br />

2637<br />

Imaging Science: Initial Results on Saturn's Rings and Small Satellites, Science 307, 1226-1236.<br />

2638 Porco, C.C. Helfenstein, P., Thomas, P.C., Ingersoll, A.P., Wisdom, J., West, R., Neukum, G.,<br />

2639 Denk, T., Wagner, R., Roatsch, T., Kieffer, S., Turtle, E., McEwen, A., Johnson, T.V., Rathbun, J.,<br />

2640 Veverka, J., Wilson, D., Perry, J., Spitale, J., Brahic, A., Burns, J.A., DelGenio, A.D., Dones, L.,<br />

2641 Murray, C.D., and Squyres, S., 2006, Cassini observes the active south pole <strong>of</strong> Enceladus, Science<br />

2642<br />

2643<br />

311, 1393-1401.<br />

2644 Postberg, F., Kempf, S., Hillier, J. K., Srama, R., Green, S. F., McBride, N., Grün, E., 2008, The E-ring<br />

2645 in the vicinity <strong>of</strong> Enceladus II, Probing the moon's interior - The composition <strong>of</strong> E-ring particles,<br />

2646 Icarus,1 93, 438-454.<br />

2647<br />

2648 Prockter, L.M., and Pappalardo, R.T., 2000, Folds on Europa: implications for crustal cycling and<br />

2649 accommodation <strong>of</strong> extension, Science 289, 941-943.<br />

2650<br />

2651 Rappaport, N. J., Iess, L., Tortora, P., Asmar, S. W., Somenzi, L., Anabtawi, A., Barbinis, E.,<br />

2652 Fleischman, D. U., Goltz, G. L., 2005, Gravity Science in the Saturnian system: The masses <strong>of</strong><br />

2653 Phoebe, Iapetus, Dione and Enceladus, DPS abstract #39.02.<br />

2654<br />

2655 Roatsch, T., Jaumann, R., Stephan, K., Thomas, P.C., 2009, Cartographic Mapping <strong>of</strong> the Icy<br />

2656 Satellites using ISS and VIMS data, Saturn from Cassini-Huygens, (M. Dougherty, L. Esposito and<br />

2657 S. Krimigis, (eds.)), Springer, 2009<br />

2658<br />

2659 Roberts, J.H., and F. Nimmo, 2009, Tidal dissipation due to despinning and the equatorial ridge on<br />

Iapetus, 40 th 2660<br />

2661<br />

Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #1927.<br />

2662 Roering, J.J., Kirchner, J.W., Sklar, L.S., Dietrich, W.E., 2001, Hillslope evolution by nonlinear<br />

2663<br />

2664<br />

creep and landsliding: An experimental study, Geology 29, 143-146<br />

2665 Roush, T.L., Cruikshank, D.P., Owen, T.C., 1995, Surface ices in the outer Solar System. In:<br />

2666 Farley, K.A. (ed.), Volatiles in the <strong>Earth</strong> and Solar System, Am. Inst. Phys. Conf. Proc. 341, 143–<br />

2667<br />

2668<br />

153.<br />

2669<br />

2670<br />

2671<br />

Rubincam, D. P., 2003, Polar wander on Triton and Pluto due to volatile migration, Icarus, 163,<br />

469–478.<br />

2672 Sack, N. J., Johnson, R.E., Boring, J.W., Baragiola, R.A., 1992, The effect <strong>of</strong> magnetospheric ion<br />

2673<br />

2674<br />

bombardment on the reflectance <strong>of</strong> Europa’s surface, Icarus 100, 534-540.<br />

Schenk, P.M., 1989a, Mimas grooves, the Herschel impact, and tidal stresses, 20 th 2675<br />

Lunar and<br />

2676<br />

2677<br />

<strong>Planetary</strong> Science Conference, Houston, abstract #960.<br />

2678 Schenk, P.M., 1989b, Crater morphology and modification on the icy satellites <strong>of</strong> Uranus and<br />

2679<br />

2680<br />

Saturn: depth/diameter and central peak occurrence, J. Geophys. Res. 94, 3815–3832.<br />

2681 Schenk, P.M., 1993, Central pit and dome craters - Exposing the interiors <strong>of</strong> Ganymede and<br />

2682<br />

2683<br />

Callisto, J. Geophys. Res. 98, 7475-7498.<br />

2684 Schenk, P.M., McKinnon, W.B., Gwynn, D., Moore, J.M., 2001, Flooding <strong>of</strong> Ganymede’s bright<br />

2685<br />

terrains by low-viscosity water-ice lavas, Nature 410, 57-60.<br />

48


2686<br />

2687 Schenk P.M., Moore, J.M., 2007, Impact crater topography and morphology on Saturnian mid-sized<br />

satellites, 38 th 2688<br />

2689<br />

Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #2305.<br />

2690 Schenk, P.M., Matsuyama, I., Nimmo, F., 2008, True polar wander on Europa from global-scale<br />

2691<br />

2692<br />

small-circle depressions, Nature 453, 368-371.<br />

2693 Schmedemannn, N., Neukum, G., Denk, T., Wagner, R., 2008, Stratigraphy and surface ages on<br />

Iapetus and other Saturnian satellites, 39 th 2694<br />

Lunar and <strong>Planetary</strong> Science Conference, Houston,<br />

2695<br />

2696<br />

abstract #2070.<br />

2697 Schmidt, J., Brilliantov, N., Spahn, F., Kempf, S., 2008, Slow dust in Enceladus' plume from<br />

2698 condensation and wall collisions in tiger stripe fractures, Nature 451, 685-688,<br />

2699<br />

2700<br />

doi:10.1038/nature06491.<br />

2701<br />

2702<br />

Schmitt, B., C. De Bergh M. Festou, editors, 1998, Solar System Ices, Kluwer, Dordrecht.<br />

2703 Seal, D., Brown, R.H., Buffington, B., Dougherty, M.K., Esposito, L.W., 2009, Cassini Huygens<br />

2704 Extended Mission, Saturn from Cassini-Huygens, (M. Dougherty, L. Esposito and S. Krimigis,<br />

2705 (eds.)), Springer, 2009<br />

2706<br />

2707 Shkuratov, Yu., Starukhina, L., H<strong>of</strong>fmann, H., Arnold, G., 1999, A model <strong>of</strong> spectral albedo <strong>of</strong><br />

2708<br />

2709<br />

particulate surfaces: implication to optical properties <strong>of</strong> the Moon, Icarus 137, 235-246.<br />

2710 Shkuratov, Y. G., Stankevich, D.G., Petrov, D.V., Pinet, P.C, Cord, A.M., Daydou, Y.H., and<br />

2711 Chevrel, S.D., 2005, Interpreting photometry <strong>of</strong> regolith-like surfaces with different topographies:<br />

2712<br />

2713<br />

Shadowing and multiple scattering, Icarus 173, 3-15.<br />

2714 Showman A.P., Stevenson D.J., and Malhotra, R., 1997, Coupled orbital and thermal evolution <strong>of</strong><br />

2715<br />

2716<br />

Ganymede, Icarus 129, 367-383.<br />

2717 Showman A. P., Mosqueira I., Head J. W., 2004, On the resurfacing <strong>of</strong> Ganymede by liquid-water<br />

2718<br />

2719<br />

volcanism, Icarus 172, 625-640.<br />

2720 Showalter, M.R. 1991. Visual Detection <strong>of</strong> 1981 S13, Saturn's Eighteenth Satellite, and its Role in<br />

2721<br />

2722<br />

the Encke Gap, Nature 351, 709-713.<br />

2723 Showalter, M. R., Cuzzi, J. N., Larson, S. M., 1991. Structure and particle properties <strong>of</strong> Saturn's E<br />

2724<br />

2725<br />

Ring, Icarus 94, 451-473.<br />

2726 Simonelli, D. P., Kay, J., Adinolfi D., Veverka, J., Thomas, P.C., Helfenstein, P., 1999, Phoebe:<br />

2727 Albedo map and photometric properties, Icarus 138, 249-258.<br />

2728<br />

2729 Sittler, E. C., Andre, N., Blanc, M.; Burger, M., Johnson, R. E., Coates, A., Rymer, A., Reisenfeld,<br />

2730 D., Thomsen, M. F.; Persoon, A., Dougherty, M., Smith, H. T., Baragiola, R. A., Hartle, R. E.,<br />

2731 Chornay, D., Shappirio, M. D., Simpson, D., McComas, D. J., Young, D. T., 2008, Ion and neutal<br />

2732 sources and sinks within Saturn’s inner magnetosphere: Cassini results, <strong>Planetary</strong> Space Sci. 56, 3-<br />

2733<br />

2734<br />

18, doi:10.1016/j.pss.2007.06.006.<br />

2735 Smith, B.A., Soderblom, L., Beebe, R., Boyce, J., Briggs, G., Bunker, A., Collins, S.A., Hansen,<br />

2736 C.J., Johnson, T.V., Mitchell, J.L., Terrile, R.J., Carr, M., Cook, II, A.F., Cuzzi, J., Pollack, J.B.,<br />

2737 Danielson, G.E., Ingersoll, A., Davies, M.E., Hunt, G.E., Masursky, H., Shoemaker, E., Morrison,<br />

2738 D., Owen, T., Sagan, C., Veverka, J., Strom, R., and Suomi, V.E., 1981, Encounter with Saturn:<br />

2739<br />

2740<br />

Voyager 1 imaging science results, Science 212, 163-191, 1981, doi:10.1126/science.212.4491.163.<br />

2741 Smith, B.A. Soderblom, L., Batson, R., Bridges, P., Inge, J., Masursky, H., Shoemaker, E., Beebe,<br />

2742 R., Boyce, J., Briggs, G., Bunker, A., Collins, S.A., Hansen, C.J., Johnson, T.V., Mitchell, J.L.,<br />

2743 Terrile, R.J., Cook, II, A.F., Cuzzi, J., Pollack, J.B., Danielson, G., Ingersoll, A.P., Davies, M.E.,<br />

2744<br />

Hunt, G.E., Morrison, D., Owen, T., Sagan, C., Veverka, J., Strom, R., and Suomi, V.E., 1982, A<br />

49


2745 new look at the Saturn system: The Voyager 2 images, Science 215, 504-536,<br />

2746<br />

2747<br />

doi:10.1126/science.215.4532.504.<br />

2748<br />

2749<br />

Soter, S., 1974, Brightness <strong>of</strong> Iapetus, Presented at IAU Colloq. 28, Cornell Univ..<br />

2750 Spahn, F., Schmidt, J., Albers, N., Hörning, M., Makuch, M., Seiß, M., Kempf, S., Srama, R.,<br />

2751 Dikarev, V., Helfert, S., Moragas-Klostermeyer, G., Krivov, A.V., Sremcevic, M., Tuzzolino, A.J.,<br />

2752 Economou, T., and Grün, E., 2006, Cassini Dust Measurements at Enceladus and Implications for<br />

2753<br />

2754<br />

the Origin <strong>of</strong> the E Ring, Science 311, 1416-1418.<br />

2755 Spencer, J.R., Calvin, W.M., Person, M.J., 1995, CCD spectra <strong>of</strong> the Galilean satellites: Molecular<br />

2756<br />

2757<br />

Oxygen on Ganymede. J. Geophys. Res. 100, 19049-19056.<br />

2758<br />

2759<br />

2760<br />

Spencer, J. R., 1998, Upper limits for condensed O2 on Saturn’s icy satellites and rings, Icarus 136,<br />

349-352.<br />

2761 Spencer, J. R., Pearl, J. C., Segura, M., and The Cassini CIRS Team, 2005, Cassini CIRS<br />

2762 Observations <strong>of</strong> Iapetus' Thermal Emission, 36th Lunar and <strong>Planetary</strong> Science Conference,<br />

2763<br />

2764<br />

Houston, abstract #2305.<br />

2765<br />

2766<br />

Spencer, J.R., Pearl, J.C., Segura, M., Flasar, F.M., Mamoutkine, A., Romani, P., Buratti, B.J.,<br />

Hendrix, A.R., Spilker, L.J., and Lopes, R.M.C., 2006, Cassini Encounters Enceladus: Background<br />

2767<br />

2768<br />

and the Discovery <strong>of</strong> a South Polar Hot Spot, Science 311, 1401-1405.<br />

2769 Spencer, J.R., Pearl, J., Segura, M., Cassini CIRS Team (2007): New Cassini CIRS Observations <strong>of</strong><br />

2770 Temperatures on Iapetus and other Saturnian Satellites. BAAS 39 (3), 06.08.<br />

2771 http://www.abstractsonline.com/viewer/viewAbstract.asp?CKey={5073DF52-17A2-477D-A759-<br />

2772 A053DFADCDC6}&MKey={ADDC1E2E-9B5A-4D9F-B9E3-<br />

2773 2D6700638A29}&AKey={AAF9AABA-B0FF-4235-8AEC-74F22FC76386}&SKey={9C26F12B-<br />

2774<br />

2775<br />

AE46-4E71-B581-D3FEAD6294AD}<br />

2776 Spencer, J., Barr, A., Esposito, L., Helfenstein, P., Ingersoll, A., Jaumann, R., Kieffer, S., McKay,<br />

2777 C., Nimmo, F., Porco, C.C., Waite, W., 2009, Enceladus: a cryovolcanic active satellite, Saturn<br />

2778<br />

2779<br />

from Cassini-Huygens, (M. Dougherty, L. Esposito and S. Krimigis, (eds.)), Springer, 2009.<br />

2780 Spencer, J.R., and Denk, T., 2009, Formation <strong>of</strong> the Extreme Albedo Dichotomy <strong>of</strong> Iapetus by<br />

2781<br />

2782<br />

Exogenically-Triggered Thermal Migration <strong>of</strong> Water Ice, Science, submitted.<br />

2783 Spitale, J. N., Jacobson, R. A., Porco, C. C., and Owen, Jr., W. M. , 2006, The Orbits <strong>of</strong> Saturn's<br />

2784 Small Satellites Derived from Combined Historic and Cassini Imaging Observations, Astronomical<br />

2785<br />

2786<br />

Journal 132, 692.<br />

2787<br />

2788<br />

2789<br />

Spitale, J.N., and Porco, C.C., 2007, Association <strong>of</strong> the jets <strong>of</strong> Enceladus with the warmest regions<br />

on its south-polar fractures, Nature 449, 695-697.<br />

2790 Sotin C., Grasset O., and Beauchesne, S., 1998, Thermodynamical properties <strong>of</strong> high pressure ices.<br />

2791 Implications for the dynamics and internal structure <strong>of</strong> large icy satellites, In Solar System Ices (B.<br />

2792<br />

2793<br />

Schmitt, C. De Bergh, and M. Festou, eds.),79-96. Kluwer, Dordrecht, The Netherlands.<br />

2794 Sotin, C., Head, J.W., and Tobie, G., 2002, Europa: Tidal heating <strong>of</strong> upwelling thermal plumes and<br />

2795<br />

2796<br />

the origin <strong>of</strong> lenticulae and chaos melting, Geophys. Res. Lett., 29, doi:10.1029/2001GL013844.<br />

2797 Sotin, C., Stevenson, D.J., Rappaport, N.J., Mitri, G., Schubert, G., 2009, Interior structure, Titan<br />

2798<br />

2799<br />

from Cassini-Huygens, (R. Brown, J.-P. Lebreton and H. Waite, (eds.)), Springer, 2009<br />

2800 Spahn, F., Schmidt, J., Albers, N., Hörning, M., Makuch, M., Seiß, M., Kempf, S., Srama, R.,<br />

2801 Dikarev, V., Helfert, S., Moragas-Klostermeyer, G., Krivov, A. V., Sremevi, M., Tuzzolino, A. J.,<br />

2802 Economou, T., Grün, E., 2006, Cassini Dust Measurements at Enceladus and Implications for the<br />

2803<br />

Origin <strong>of</strong> the E Ring, Science, 311, 1416-1418, DOI:10.1126/science.1121375.<br />

50


2804<br />

2805 Squyres, S.W., 1980, Volume changes in Ganymede and Callisto and the origin <strong>of</strong> grooved terrain,<br />

2806<br />

2807<br />

Geophys. Res. Lett. 7, 593-596.<br />

2808<br />

2809<br />

Squyres, S.W. and Sagan, C., 1983, Albedo Asymmetry <strong>of</strong> Iapetus, Nature 303, 782-785.<br />

2810 Squyres, S. W. Buratti, B., Veverka, J., Sagan, C., 1984, Voyager photometry <strong>of</strong> Iapetus, Icarus 59,<br />

2811<br />

2812<br />

426-435.<br />

2813 Squyres, S.W. and Cr<strong>of</strong>t, S.K., 1986, The tectonics <strong>of</strong> icy satellites, in Satellites, ed. M.S.<br />

2814<br />

2815<br />

Matthews, 293-341, Univ. Ariz. Press, Tucson.<br />

2816 Stephan, K., Jaumann, R., Wagner, R., Roatsch, Th., Clark., R.N., Cruikshank, D.P., Hibbitts, C.A.,<br />

2817<br />

2818<br />

Hansen, G.B., Buratti, B.J., Filiacchione, G., Baines, K.H., Nicholson, P.D., McCord, T.B., and<br />

th<br />

Brown, R.H, 2008, Relationship <strong>of</strong> Dione’s spectral properties to geological surface units, 39<br />

2819<br />

2820<br />

Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #1717.<br />

2821<br />

2822<br />

Stevenson, D., 1982,Volcanism and igneous processes in small icy satellites, Nature 298, 142-144.<br />

2823 Strazzulla, G., 1998, Chemistry <strong>of</strong> ice induced by bombardment with energetic charged particles, in<br />

2824<br />

2825<br />

Solar System Ices (ed. B. Schmitt et al.), 281-302.<br />

2826 Tabak, R.G. and Young, W.M., 1989, Cometary Collisions and the Dark Material on Iapetus. <strong>Earth</strong>,<br />

2827<br />

2828<br />

Moon, and Planets 44, 251-264.<br />

2829 Tian, F., Stewart, A.I.F., Toon, O.B., Larsen, K.W., Esposito, L.W., 2007, Monte Carlo simulations<br />

2830<br />

2831<br />

<strong>of</strong> the water vapor plumes on Enceladus, Icarus 188, 154-161.<br />

2832 Thomas, P. C., Veverka, J., Morrison, D., Davies, M., and Johnson, T.V., 1983, Saturn's Small<br />

2833<br />

2834<br />

Satellites Voyager Imaging Results, J. <strong>of</strong> Geophys. Res. 88, 8743-8754.<br />

2835<br />

2836<br />

Thomas, P. and Veverka, J. 1985. Hyperion: Analysis <strong>of</strong> Voyager observations, Icarus 64: 414-424.<br />

2837 Thomas P.G., 1988, The tectonic and volcanic history <strong>of</strong> Rhea as inferred from studies <strong>of</strong> scarps,<br />

2838<br />

2839<br />

ridges, troughs and other lineaments, Icarus 74, 554-567.<br />

2840<br />

2841<br />

Thomas, P.C. 1989, Shapes <strong>of</strong> Small Satellites, Icarus 77, 248-274.<br />

2842 Thomas, P.C., Armstrong, J., Asmar, S. W., Burns, J. A., Denk, T., Giese, B., Helfenstein, P., Iess,<br />

2843 L., Johnson, T. V., McEwen, A.S., Nicolaisen, L., Porco, C.C., Rappaport, N., Richardson, J.,<br />

2844 Somenzi, L., Tortora, P., Turtle. E. P., Veverka, J., 2007a, Hyperion's sponge-like appearance,<br />

2845 Nature 448, 50-53, doi: 10.1038/nature05779.<br />

2846<br />

2847 Thomas, P.C., Burns, J. A., Helfenstein, P., Squyres, S., Veverka, J., Porco, C., Turtle, E. P.;<br />

2848 McEwen, A., Denk, T., Giese, B., Roatsch, T., Johnson, T. V., Jacobson, R. A., 2007b, Shapes <strong>of</strong><br />

2849 the Saturnian icy satellites and their significance, Icarus 190, 573-584.<br />

2850<br />

2851<br />

doi:10.1016/j.icarus.2007.03.012.<br />

2852 Tokar, R.L., Johnson, R.E., Hill, T.W., Pontius, D.H., Kurth, W.S., Crary, F.J., Young, D.T.,<br />

2853 Thomsen, M.F., Reisenfeld, D.B., Coates, A.J., Lewis, G.R., Sittler, E.C., and Gurnett, D.A., 2006,<br />

2854<br />

2855<br />

The Interaction <strong>of</strong> the Atmosphere <strong>of</strong> Enceladus with Saturn's Plasma, Science 311, 1409-1412.<br />

2856 Turcott, D.L. and Schubert, G., 2002, Geodynamics, 2nd Ed. Cambridge Univ. Press, Cambridge.<br />

2857<br />

2858 Verbiscer, A. J., and Veverka, J. 1989, Albedo dichotomy <strong>of</strong> Rhea: Hapke analysis <strong>of</strong> Voyager<br />

2859<br />

2860<br />

photometry, Icarus 82, 336-353.<br />

2861 Verbiscer, A. J., and Veverka, J., 1992, Mimas: Photometric roughness and albedo map, Icarus 99,<br />

2862<br />

63-69.<br />

51


2863<br />

2864 Verbiscer, A. J., and Veverka, J., 1994, A photometric study <strong>of</strong> Enceladus, Icarus 110, 155-164.<br />

2865<br />

2866 Verbiscer, A. J., French, R. G., McGhee, C. A., 2005, The opposition surge <strong>of</strong> Enceladus: HST<br />

2867 observations 0.338-1.022 µm, Icarus 173, 66-83.<br />

2868<br />

2869 Verbiscer, A. J., Peterson, D. E., Skrutskie, M. F., Cushing, M., Helfenstein, P., Nelson, M.<br />

2870 J.,Smith, J. D.,Wilson, J. C., 2006, Near-infrared spectra <strong>of</strong> the leading and trailing hemispheres <strong>of</strong><br />

2871 Enceladus, Icarus, 182, 211-223.<br />

2872<br />

2873 Verbiscer, A., French, R., Showalter, M., Helfenstein, P., 2007, Enceladus: Cosmic graffiti artist<br />

2874<br />

2875<br />

caught in the act, Science 315, 815.<br />

2876 Verbiscer, A.J., Peterson, D.E., Skrutskie, M.F., Cushing, M., Helfenstein, P., Nelson, M.J., Smith,<br />

2877 J.D., Wilson, J.C., 2008, Ammonia hydrate on Tethys’ Trailing Hemisphere, Science <strong>of</strong> Solar<br />

2878<br />

2879<br />

System Ices 2008, # 9064.<br />

2880 Veverka, J., Thomas, P.C., Johnson, T.V., Matson, D., and Housen, K.R., 1986, The physical<br />

2881 characteristics <strong>of</strong> satellites surfaces. In: Burns, J.A., Matthews, M.S. (eds.), In Satellites, Univ. <strong>of</strong><br />

2882 Arizona Press, Tucson, 242-402.<br />

2883<br />

2884 Vilas, F., Larson, S. M., Stockstill, K. R., and Gaffey, M. J., 1996, Unraveling the zebra: Clues to<br />

2885 the Iapetus dark material composition, Icarus 124, 262-267.<br />

2886<br />

2887 Vilas, F., Jarvis, K.S., Barker, E.S., Lederer, S.M., Kelly, M.S., and Owen, T., 2004, Iapetus dark<br />

2888 and bright material: giving compositional interpretation some latitude, Icarus, 170, 125-130.<br />

2889<br />

2890 Wahr, J., Selvans, Z.A., Mullen, M.E., Barr, A.C., Collins, G.C., Selvans, M.M., Pappalardo, R.T.,<br />

2891 2009, Modeling stresses on satellites due to non-synchronous rotation and orbital eccentricity using<br />

2892 gravitational potential theory, Icarus, 200, 188-206.<br />

2893<br />

2894 Wagner R. et al., Neukum, G., Giese, B., Roatsch, T., Wolf, U., Denk, T., and the Cassini ISS<br />

2895 Team, 2006, Geology, ages and topography <strong>of</strong> Saturn’s satellite Dione observed by the Cassini ISS<br />

camera, 37 th 2896<br />

Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #1805<br />

2897<br />

2898 Wagner, R.J., Neukum, G., Giese, B., Roatsch, T., Wolf, U., 2007, The global geology <strong>of</strong> Rhea:<br />

Preliminary implications from the Cassini ISS data, 38 th 2899 Lunar and <strong>Planetary</strong> Science Conference,<br />

2900<br />

2901<br />

Houston, abstract #1958.<br />

2902 Wagner, R. J., Neukum, G., Giese, B., Roatsch, T., Denk, T., Wolf, U., Porco, C.C., 2008, Geology<br />

2903 <strong>of</strong> Saturn’s satellite Rhea on the basis <strong>of</strong> the high-resolution images from the targeted flyby 049 on<br />

2904<br />

th<br />

Aug. 30, 2007, 39 Lunar and <strong>Planetary</strong> Science Conference, Houston, abstract #1930.<br />

2905<br />

2906 Wagner, R. J., Neukum, G., Stephan, K., Roatsch, T., Wolf, U., Porco, C.C., 2009, Stratigraphy <strong>of</strong><br />

tectonic features on Saturn’s satellite Dione derived from Cassini ISS camera data, 40 th 2907<br />

Lunar and<br />

2908 <strong>Planetary</strong> Science Conference, Houston, abstract #2142.<br />

2909<br />

2910 Waite, J.H., Jr., Combi, M.R., Ip, W.H., Cravens, T.E., McNutt, R.L., Jr., Kasprzak, W., Yelle, R.,<br />

2911 Luhmann, J., Niemann, H., Gell, D., Magee, B., Fletcher, G., Lunine, J., and Tseng, W.L., 2006,<br />

2912 Cassini Ion and Neutral Mass Spectrometer: Enceladus Plume Composition and Structure, Science<br />

2913<br />

311, 1419-1422.<br />

2914<br />

2915 Warren, S. G., 1982. Optical properties <strong>of</strong> snow, Rev. Geophys. Space Phys. 20, 67-89.<br />

2916<br />

2917<br />

2918<br />

Willemann, R.J., 1984, Reorientation <strong>of</strong> planets with elastic lithospheres, Icarus 60, 701-709.<br />

2919 Wilson, P. D., and Sagan, C., 1995, Spectrophotometry and organic matter on Iapetus, 1.<br />

2920<br />

2921<br />

Composiiton models, J. Geophys. Res. 100, 7531-7537.<br />

52


2922 Wilson, P. D., and Sagan, C., 1996, Spectrophotometry and organic matter on Iapetus, 2: Models <strong>of</strong><br />

2923 interhemispheric asymmetry, Icarus 122, 92-106.<br />

2924<br />

2925 Wilson L., Head J. W., Pappalardo R. T., 1997, Eruption <strong>of</strong> lava flows on Europa: Theory and<br />

2926 application to Thrace Macula, J. Geophys. Res. 102, 9263-9272.<br />

2927<br />

2928 Wisdom, J., Peale, S.J., Mignard, F., 1984, The chaotic rotation <strong>of</strong> Hyperion, Icarus 58, 137-152.<br />

2929<br />

2930 Wisdom, J., 2004, Spin-orbit secondary resonance dynamics <strong>of</strong> Enceladus, Astron. J. 128, 484-491.<br />

2931<br />

2932 Yoder, C.F., 1995, Astrometric and geodetic properties <strong>of</strong> <strong>Earth</strong> and the Solar System, in Global<br />

2933<br />

2934<br />

<strong>Earth</strong> Physics, T.J. Ahrens, ed., 1-31, Amer. Geophys. Union.<br />

2935 Zahnle, K., Schenk, P., Levison, H., and Dones, L., 2003, Cratering rates in the outer solar system,<br />

2936<br />

2937<br />

Icarus 163, 263-289.<br />

2938<br />

2939<br />

Zellner, B., 1972. On the nature <strong>of</strong> Iapetus, Astrophys. J. Lett. 174, L107-L109.<br />

2940<br />

2941<br />

Figures<br />

2942 Chap.21_Icy Sat_Revised Version March 31, 2009<br />

2943<br />

2944<br />

2945<br />

2946<br />

2947<br />

Fig. 1<br />

53


2948<br />

2949<br />

2950<br />

2951<br />

2952<br />

Fig. 2<br />

54


2953<br />

2954<br />

2955<br />

2956<br />

2957<br />

2958<br />

2959<br />

2960<br />

2961<br />

2962<br />

2963<br />

Fig. 3<br />

Fig. 4<br />

Fig. 5<br />

55


2964<br />

2965<br />

2966<br />

2967<br />

2968<br />

2969<br />

2970<br />

2971<br />

Fig. 6<br />

Fig. 7<br />

56


2972<br />

2973<br />

2974<br />

2975<br />

2976<br />

2977<br />

2978<br />

2979<br />

Fig. 8<br />

Fig. 9a<br />

57


2980<br />

2981<br />

2982<br />

2983<br />

2984<br />

2985<br />

2986<br />

2987<br />

Fig. 9b<br />

Fig. 9c<br />

58


2988<br />

2989<br />

2990<br />

2991<br />

2992<br />

2993<br />

2994<br />

2995<br />

Fig. 9d<br />

Fig. 9e<br />

59


2996<br />

2997<br />

2998<br />

2999<br />

3000<br />

3001<br />

3002<br />

3003<br />

Fig. 10<br />

Fig. 11<br />

60


3004<br />

3005<br />

3006<br />

3007<br />

3008<br />

3009<br />

3010<br />

3011<br />

3012<br />

Fig. 12<br />

Fig. 13<br />

61


3013<br />

3014<br />

3015<br />

3016<br />

3017<br />

3018<br />

3019<br />

3020<br />

Fig. 14 (left)<br />

Fig. 14 (right)<br />

62


3021<br />

3022<br />

3023<br />

3024<br />

3025<br />

3026<br />

3027<br />

3028<br />

3029<br />

3030<br />

Fig. 15 (above)<br />

Fig. 15 (lower left)<br />

63


3031<br />

3032<br />

3033<br />

3034<br />

3035<br />

3036<br />

3037<br />

3038<br />

Fig. 15 (lower right)<br />

Fig. 16 (left)<br />

64


3039<br />

3040<br />

3041<br />

3042<br />

3043<br />

3044<br />

3045<br />

3046<br />

3047<br />

3048<br />

3049<br />

Fig. 16 (right)<br />

Fig. 17<br />

65


3050<br />

3051<br />

3052<br />

3053<br />

3054<br />

3055<br />

3056<br />

Fig. 18<br />

Fig. 19<br />

66


3057<br />

3058<br />

3059<br />

3060<br />

3061<br />

3062<br />

3063<br />

3064<br />

Fig. 20<br />

Fig. 21<br />

67


3065<br />

3066<br />

3067<br />

3068<br />

3069<br />

3070<br />

3071<br />

3072<br />

Fig. 22<br />

Fig. 23<br />

68


3073<br />

3074<br />

3075<br />

3076<br />

3077<br />

3078<br />

3079<br />

3080<br />

3081<br />

Fig. 24<br />

Fig. 25<br />

69

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!