1

1 The and Proterozoic History of Peninsular : Tectonic 2 Framework for Precambrian Sedimentary Basins in India 3

4 Joseph G. Meert1 and Manoj K. Pandit2

5 1Department of Geological Sciences, 241 Williamson Hall, University of Florida, Gainesville, FL 32611 USA 6 2Department of Geology, University of Rajasthan, Jaipur 302004, Rajasthan, India 7 2

8 Abstract

9 The Precambrian geologic history of peninsular India covers nearly 3.0 billion years of time. The Peninsula 10 is an assembly of five different cratonic nuclei known as the Aravall-Bundelkhand, Eastern Dharwar, Western 11 Dharwar, Bastar and Singhbhum along with the Southern Granulite Province. Final amalgamation of these 12 elements occurred either by the end of the Archaean (2.5 Ga) or the end of the Paleoproterozoic (~1.6 Ga). Each of 13 these nuclei contains one or more sedimentary basins (or metasedimentary basins) of Proterozoic age. This paper 14 provides an overview of each of the cratons and a brief description of the Precambrian sedimentary basins in India 15 that form the focus of the remainder of this book. In our view, it appears that basin formation and subsequent clo- 16 sure can be grossly constrained to three separate intervals that are related to the assembly and disaggregation of the 17 supercontinents Columbia, Rodinia and Gondwana. The oldest Purana-I basins developed during the 2.5-1.6 Ga in- 18 terval, Purana-II basins formed during the 1.6-1.0 Ga interval and the Purana-III basins formed during the 19 Neoproterozoic-Cambrian interval. 20 21 Introduction

22 Peninsular India represents an amalgam of ancient cratonic nuclei that formed and

23 stabilized during Archaean to Paleoproterozoic (3.8-1.6 Ga). Modern-day Peninsular India

24 is comprised of five Archaean nuclei known as the Banded Complex (Aravalli re-

25 gion), Bundelkhand, Singhbhum, Bastar and Dharwar cratons (figure 1). The southern

26 granulite province is a polycyclic region that also contains Archaean-age crustal elements

27 that are dissected by younger orogenic belts. This review paper will focus on the Archean-

28 Paleoproterozoic history of each of the major cratonic elements of Peninsular India on

29 which large sedimentary basins were formed.

30 We make every effort to provide updated geochronologic data for each of the re-

31 gions, but note that there is still a paucity of reliable U-Pb, Pb-Pb, Ar-Ar and Sm-Nd ages.

32 Older K-Ar and Rb-Sr ages are used only when there are no other data available and only to

33 provide some constraints on the development of the region in question. The paper begins

34 with an overview of the Aravalli-Bundelkhand sectors and then moves eastward to the

35 Singhbhum ; the central Indian Bastar craton and concludes with a look at the east-

36 ern and western Dharwar and regions.

37 Each of the aforementioned cratonic elements contains Precambrian- Early Paleozo-

38 ic sedimentary (or metasedimentary) sequences that form the basinal infrastructure of

39 Peninsular India. Among the best-preserved sedimentary sequences are the so-called

40 “Purana” or ancient basin. These include the large Cuddapah, Chhattisgarh and Vindhyan

41 basins along with several smaller regional basins known as the Indravati, Khariar,

42 Prahnita-Godavari, Kaladgi, Bhima, , Kurnool and Marwar. There appear to be sev- 3

43 eral key intervals of basinal development (and closure) within Peninsular India. Purana-I

44 basins began development in the Paleoproterozoic (2.5-1.6 Ga); Purana-II basins formed

45 during the Mesoproterozoic (1.6-1.0 Ga); and the development of Purana-III basins is con-

46 fined to the Ediacaran-Cambrian interval. Basinal development and closure may be tempo-

47 rally related to the formation/breakup of the supercontinents Columbia (fig 2a; Purana-I),

48 Rodinia (fig. 2b; Purana-II) and Gondwana (fig. 2c; Purana-III). As a general reminder, we

49 note that the present-day basinal outcrops/areal extent of the sedimentary sequences and

50 may not accurately reflect the extent of those sequences at the time of deposition. We also

51 wish to note that we use the term ‘closure’ to indicate a period of time when the basin

52 stopped receiving sediments. Defining the cause of basinal closure in Proterozoic basins

53 can be difficult and may be related to the creation of tectonic barriers to sedimentation,

54 burial by younger (and now eroded) sedimentary sequences, sea-level change and many

55 other factors. If discussed in our overview, we provide only a best estimate for when clo-

56 sure occurred and offer some speculation for why sedimentation ceased.

57 Aravalli Banded Gneiss Complex and Bundelkhand Cratons

58 The Aravalli Banded Gneiss Complex (BGC)-Bundelkhand protocontinent occupies

59 the north-central region of the Indian sub-continent (Fig. 1). The Great Boundary Fault

60 (GBF) marks a present-day physiographic divide between the two blocks with the BGC

61 cratonic block to the west of the GBF and the Bundelkhand-Gwalior block to the east of the

62 GBF. These cratons are bounded to the Northeast by the Mesoproterozoic-aged Vindhyan

63 basin and the Indo-Gangetic alluvium and to the south by the northern edge of the Deccan

64 Traps volcanic rocks. The Western Margin Fault forms the wester boundary of the BGC.

65 The Bundelkhand and BGC regions are also separated from the Bastar and Singhbhum

66 cratons by the Central Indian Tectonic Zone (Fig. 1; Goodwin, 1991; Naqvi and Rogers,

67 1987; Meert et al., 2010).

68 There is considerable debate regarding the nature and age of the basement rocks in

69 the BGC. In part this was due to a lack of high-quality geochronological data for the differ-

70 ent metamorphic complexes and in part due to the obscured nature of the contacts be- 4

71 tween the separate regions (Ramakrishan and Vaidyanadhan, 2008; Buick et al., 2006; Roy

72 et al., 2012). The current status of this debate is described below.

73 The Banded Gneiss Complex was a catch-all term for the gneissic rocks that were lo-

74 cated to the west of the Great Boundary Fault (Heron, 1953) although Gupta (1934) had

75 recognized some differences between the eastern and western side of the Banas lineament

76 (Fig. 3). Gupta subdivided these two regions into the BGC-1 and BGC-2 domains (see also

77 Roy and Kroner, 1996; Wiedenbeck and Goswami, 1994; Gopalan et al., 1990). BGC-1 met-

78 amorphic basement included the Mewar whereas the BGC-2 region is composed of

79 rocks known as the Sandmata and Mangalwar Complexes.

80 The gneissic rocks of the BGC are dated to between 3300-2900 Ma (Roy and Kroner,

81 1996; Wiedenback and Goswami, 1994; Gopalan et al., 1990) and are dominated by

82 tonalite-trondjhemite gneisses (TTG’s) that are intruded by late Archaean granitoids dated

83 to between 2600-2500 Ma (i.e. the Untala, Gingla and Berach ; Wiedenbeck et al.,

84 1996a,b; Roy and Kroner, 1996; Meert et al., 2010).

85 Deb (1999) reported a 2075-2150 Ma Ga Pb-Pb age for galena, presumably

86 syngenetic with the basal Aravalli volcanics. In the absence of any direct geochronologic ev-

87 idence this age is taken to represent the initiation of Aravalli sedimentation. Further sup-

88 port is provided by Pandit et al. (2008) and de Wall et al. (2012) who describe the paleosol

89 below the Aravalli Supergroup to have developed during the Great Oxidation Event (GOE).

90 Intrusion of 1850 Ma Darwal has generally been accepted as the closing age for

91 deposition of the Aravalli Supergroup.

92 The Mangalwar Complex (BGC-2) is located between the Banas & Delwara linea-

93 ments (Fig. 3). It has been subdivided into a number of formations (Ramakrishan and

94 Vaidyanadhan, 2008), but the relationships between the various formations that make up

95 the Mangalwar are very poorly known. The rocks of the Mangalwar Complex include

96 TTG’s, gneisses, , , and quartzites. Commonly cited ages for

97 the Mangalwar Complex ranged between 2900-2600 Ma (Ramakrishan and Vaidyanadhan,

98 2008).

99 The Sandmata Complex (BGC-2) is sandwiched between the Delwara and Kaliguman

100 lineaments (Fig. 3) and is composed of TTG’s, metapelites, metapsammites, metagabbros,

101 charnockites and granulites (Buick et al., 2006). Ages from the Sandmata Complex also 5

102 cluster between 1.7-1.8 Ga (Buick et al., 2006). Roy et al. (2012) reported older ages from

103 the Sandmata Complex between 2.9-1.9 Ga. Those authors reaffirmed the problematic

104 metamorphic history recorded in both the Sandmata and Mangalwar Complexes, but ar-

105 gued that an Archaean-age component was present in the BGC-2 rocks albeit considerably

106 younger than the oldest ages recorded in the Mewar gneisses to the southeast.

107 Bhowmik et al. (2012) attempted to synthesize the available geochronological data

108 into a plate-tectonic framework for the development of the BGC region that we have modi-

109 fied slightly. During the interval from 3.3-2.5 Ga, the basement rocks of the craton were

110 formed that included both the BGC-1 rocks (TTG gneisses) and the BGC-2 rocks

111 (Mangalwar and Sandmata Complexes). The basement gneisses were then intruded by a

112 series of granitoid rocks that completed the stabilization of the craton (Sinha-Roy et al.,

113 1998; Roy and Jakhar, 2002). Between 2.4-2.1 Ga an ocean basin formed and the Aravalli

114 Supergroup sediments were deposited (Deb and Thorpe, 2004, Deb, 1999). Between 1.85-

115 1.8 Ga, the Aravalli Ocean basin closed, metamorphism and deformation of the Aravalli sed-

116 imentary rocks occurred and the Rakhabdev mafic-ultramafic rocks were emplaced (Kaur

117 et al., 2007a,b, 2009). Bhowmik et al. (2012) then posit a slab roll-back event that creates

118 basinal space for sedimentary rocks in the Sandmata back-arc basin during the 1.8-1.72 Ga

119 interval (Kaur et al., 2011a). In their scenario, slab breakoff at 1.72 Ga produces basaltic

120 underplating and granulite-facies metamorphism in the Sandmata Complex followed by the

121 eventual development of the Delhi basin during the 1.7-1.0 Ga interval. Closure of the Delhi

122 Basin at 1.0 Ga coincided with a major phase of deformation throughout India (Deb et al.,

123 2001; Leelanandam et al. (2006). In the Aravalli region, evidence for the 1.0 Ga orogeny is

124 widespread and includes emplacement of the Phulad ophiolite, 1.0 Ga granitoid emplace-

125 ment, cessation of sedimentation in several of the larger “Purana” basins (Gupta et a;.,

126 1997; McKenzie et al., 2001; Malone et al., 2008; Turner et al., 2013; Patranabis-Deb et al.,

127 2007).

128 The processes associated with the development of the Purana-I Aravalli basinal sed-

129 iments are related to the formation of the Columbia (Rogers and Santosh, 2002; Meert,

130 2012) supercontinent. The opening of the Delhi basin is poorly constrained to the 1.7-1.0

131 Ga interval, but we posit that the development of accommodation space for these basins

132 was related to the breakup of the Columbia supercontinent during the Mesoproterozoic 6

133 and thus represent Purana-II Proterozoic basins. Closure of the Delhi basin, along with

134 several other Purana-II basins occurred during the assembly phase of Rodinia.

135 The post-Delhi scenario is marked by outpouring of felsic lavas and granitic intru-

136 sions, commonly referred to as the Malani Igneous Suite (MIS). The MIS is predominately

137 felsic volcanic-plutonic suite of rocks exposed over an area of 54,000 km2 in northwestern

138 India (Figure 2; Bhushan et al., 2000; Torsvik et al., 2001a; Gregory et al., 2008). Contact

139 relationships with the underlying basement cannot be observed in all regions due to exten-

140 sive Quaternary-aged sand cover; however, Malani felsic dykes intrude the ~830 Ma pre-

141 Malani granitic gneisses and in the Barmer region (Pandit et al., 1999;

142 Pradhan et al., 2010). The initial phase of igneous activity associated with the MIS was

143 characterized by major felsic and minor mafic flows and was followed by the emplacement

144 of peraluminous and peralkaline granitic bodies. Volumetrically minor felsic and mafic

145 dykes represent the final phase of Malani activity. Age constraints for the MIS range from

146 ~750 Ma to ~770 Ma (Torsvik et al., 2001a,b; Gregory et al., 2008; van Lente et al., 2009;

147 Meert et al., in press).

148 Malani rocks form the basement for a sedimentary sequence named the Marwar Su-

149 pergroup (Figs. 3,4,5) is composed of a two-kilometer thick section of un-metamorphosed

150 and relatively un-deformed succession of sandstones, shales, carbonates, and evaporites.

151 Lithostratigraphically, the Marwar Supergroup is subdivided into three groups; the lower-

152 most Jodhpur Group, the middle Bilara Group, and the uppermost Nagaur Group (Khan,

153 1973; Pareek, 1984). The oldest unit of the Jodphur Group (Pokaran Boulder Bed or Sonia

154 sandstone) unconformably overlies the Neoproterozoic Malani Igneous Suite. The Jodphur

155 Group is a fluvio-marine succession, of cross-bedded, white to reddish sandstone and ma-

156 roon shale. The overlying Bilara Group consists of dolomite and stromatalitic limestone

157 that conformably overlie the Jodphur Group (Khilnani, 1968; Barman, 1987). In a some-

158 what confusing nomenclature, the Hanseran Evaporite Group is included within the Bilara

159 Group, but as noted by Mazumder and Strauss (2006) these are coeval facies variants with-

160 in the basin. The so-called Hanseran Group is characterized by seven evaporitic cycles of

161 dolomite, magnesite, anhydrite, halite, polyhalite, and clay bands (Dasgupta, 1996; Kumar,

162 1999; Mazumder and Strauss, 2006). Unconformably overlying the Bilara Group is a se-

163 quence of fine to coarse grained, cross-bedded, reddish brown, sandstone and siltstone of 7

164 the Nagaur Group (Pareek, 1984; Pandey and Bahadur, 2009). The youngest sedimentary

165 unit of the Nagaur Group, the Tunklian sandstone, is unconformably overlain by the Permo-

166 Carboniferous Bap boulder bed (Pareek, 1984).

167 The Marwar Supergroup was historically classified as Neoproterozoic in age based

168 upon the relatively un-deformed stratigraphy and the absence of index fossils within the

169 sequence. Assuming the Neoproterozoic Snowball Earth event was globally distributed, the

170 absence of glacial deposits within the Marwar Supergroup suggests (but does not prove) a

171 post-Marinoan age of deposition (i.e. <635 Ma). Ediacaran fossils collected from the

172 Jodphur group, including Arumberia, Beltanelliformis, Aspidella, and Hiemalora, support a

173 late-Neoproterozoic age assignment (<570 Ma; Kumar and Pandey, 2008, 2009, 2010;

174 Raghav et al., 2005; Kumar, 2012). Fossils in the Nagaur sandstone including Cruziana

175 nabatacica, Rusophycus and Dimorphichnus are ichnogenera that first appeared in the Cam-

176 brian and indicate that the deposition in the Marwar Supergroup continued into the Cam-

177 brian. It should also be noted that these ichnogenera are also found in much younger rocks

178 (as young as Permian), so the absolute range for the Marwar Supergroup is reliably con-

179 strained only as Ediacaran-Carboniferous by stratigraphy and fossils. Davis et al. (2014)

180 offer some paleomagnetic support that the Nagaur is not much younger than Cambrian.

181 Based on these observations, the Marwar Basin is therefore considered as a Purana-III de-

182 velopment.

183 Bundelkhand CratonThe Bundelkhand craton, to the east of the Aravalli-Delhi fold

184 belt, is a relatively less studied region (Fig. 6) Sharma and Rahman (2000) divided the

185 Bundelkhand craton into three distinct units: (1) Archaean-aged granite-greenstone and

186 gneiss belts that is sometimes referred to as the Enclave Suite; (2) Relatively undeformed

187 granitoid plutons and large quartz reefs know as the Granite Suite and; (3) Mafic dyke

188 swarms and other smaller scale intrusions known as the Intrusive suite.

189 The Enclave Suite is composed of intensely deformed basement rocks; predominant-

190 ly, schists, gneisses, banded iron formations (BIF’S), mafic volcanic rocks and quartzites.

191 These basement rocks are intruded by the Bundelkhand Igneous Complex that makes up

192 the bulk of exposed rocks within Bundelkhand Craton (Goodwin, 1991; Basu, 2007a).

193 Three generations of gneisses are thought to have formed at 3.2-3.3 Ga, 2.7 Ga and 2.5 Ga,

194 respectively as indicated by 207Pb/206Pb isotopic data (Mondal et al., 1997; Gopalan et al., 8

195 1990). The latter 2.5 Ga age is also considered as the stabilization age of the Bundelkhand

196 craton as it overlaps with the ages of undeformed granitoid plutons in the Bundelkhand

197 and BGC-Aravalli cratons. The Bundelkhand granite is dated to 2492 ± 10 Ma (Mondal et al.,

198 2002) and the Berach Granite to 2530 ± 3.6 Ma using U-Pb isotopic dating (R. D. Tucker,

199 personal communication, see also Wiedenbeck et al., 1996). Pradhan et al. (2012) dated

200 xenocrystic in mafic dykes at 2.7 Ga and 3.2 Ga consistent with the observations of

201 Mondal et al. (1997).

202 The Bundelkhand craton in the Central Indian shield is also characterized by various

203 Proterozoic extrusive and intrusive events. The NE-SW trending quartz reefs are the most

204 spectacular feature in the Bundelkhand granitic massif (Basu, 1986). The majority of these

205 quartz reefs are concentrated in the area bounded by Jhansi on the NW, Supa to the NE,

206 Khajuraho on the SE and Tikamgarh to the SW (Fig. 6). These giant quartz reefs and veins

207 along the brittle ductile shear zones and fault planes mark extensive hydrothermal fluid ac-

208 tivity following the crystallization of the granite plutons. The quartz reefs and associated

209 hydrothermal activity are argued to have taken place in three phases based on the K – Ar

210 geochronology: (1) 1480 ± 35 to 1660 ± 40 Ma, (2) 1790 ± 40 to 1850 ± 35 Ma and (3)

211 1930 ± 40 to 2010 ± 80 Ma (Pati et al., 1997). The broad age ranges reported here indicate

212 a need for more robust dating of these intrusive events.

213 Numerous mafic dykes intrude the Bundelkhand Igneous Complex. Rao (2004) sug-

214 gested that most of these mafic dykes were emplaced in two phases, one at 2.15 Ga and the

215 second at 2.0 Ga, based on their 40Ar/39Ar isotopic analyses. An earlier study by Sarkar

216 (1997) reported two distinct K-Ar age clusters (~1800 Ma and 1560 Ma) on mafic dykes in-

217 truding the Bundelkhand Province. Pradhan et al. (2012) provided U-Pb ages for two of the

218 suites. The older NW-SE trending dykes are dated to ~2.0 Ga and the younger cross-

219 cutting dykes to 1.1 Ga. Pradhan et al. (2012) also noted a third suite of dykes were pre-

220 sent based on distinct paleomagnetic directions, but were unable to provide age con-

221 straints on those dykes.

222 It is unclear if the BGC region and Bundelkhand cratons share a complete common

223 history. The ages of gneissic (TTG) rocks and late Archaean granitoids in both the BGC and

224 Bundelkhand cratons span the same broad age range of 3.3-2.5 Ga. Perhaps the most im- 9

225 portant link between the two cratonic nuclei is that both are overlain by the ~1.85 Ga

226 Hindoli Group. On that basis, we posit that the BGC-Bundelkhand sectors were a single

227 block by at least ~1.9 Ga and perhaps earlier.

228 Overlying the Bundelkhand craton is the Vindhyan Basin (figs. 1,5 and 6). The

229 Vindhyan basin is comprised of distinct sequences called the Lower and Upper Vindhyan

230 Supergroups. Lower Vindhyan sediments (Semri Group) blanket the Hindoli Group and

231 thus are younger than 1.9 Ga (Saxena and Pandit, 2012; Deb et al., 2002). Geochronologic

232 studies of the Lower Vindhyan indicate a depositional history beginning as early as 1.8 Ga

233 and ending sometime post 1.6 Ga (Rasmussen et al., 2002; Ray et al., 2002, 2003; Sarangi et

234 al., 2004; Malone et al., 2008). This would constrain Lower Vindhyan sedimentation to the

235 Purana-I basinal sequences.

236 Age constraints on the Upper Vindhyan Supergroup (Kaimur, Bhander and Rewa

237 Groups) are controversial (Azmi et al., 2008; Malone et al., 2008; Gregory et al., 2006;

238 McKenzie et al., 2011; Turner et al., 2013), but the onset of sedimentation is constrained to

239 be older than 1073 Ma by intrusive relationships between the Majhgawan kimberlite into

240 the Upper Vindhyan Kaimur sandstone (Gregory et al., 2006; fig. 5).

241 The younger limit for the age of the Upper Vindhyan is based on several observa-

242 tions. There is a lack of Neoproterozoic zircons in the Upper Vindhyan rocks in spite of the

243 fact that numerous Neoproterozoic source regions are known from throughout northern

244 India (including the Himalayas; Malone et al., 2008; McKenzie et al., 2011; Turner et al., in

245 press). Paleomagnetic data from ~1000-1100 Ma igneous units in India are identical to di-

246 rections observed in the Bhander and Rewa Groups (Gregory et al., 2006; Malone et al.,

247 2008; Pradhan et al., 2012; Venkateshwarlu and Rao, 2013). Paleomagnetic directions ob-

248 served in the Marwar Supergroup are distinct from the Upper Vindhyan Bhander-Rewa

249 Groups (Davis et al., in review). We contend that the sedimentary rocks of the Upper

250 Vindhyan Supergroup developed during the Purana-II basinal stage and the basin was

251 closed during the final stages of Rodinia formation (Pradhan et al., 2012; Turner et al., in

252 press; Malone et al., 2008). The alternative view (espoused by Azmi et al., 2008) is that the

253 Upper Vindhyan is correlative with the Marwar Supergroup and that both are Purana-III

254 basins.

255 Bijawar-Sonrai and Gwalior Basins 10

256 Three isolated and narrow basins are also located on the basement rocks of the

257 Bundelkhand craton. The Bijawar-Sonrai basins contain similar lithologies and are located

258 along the southern margin of the craton west of the town of Panna (fig. 6). The Bijawar

259 Group of sediments consists of two subgroups; the lower Mali Subgroup (conglomerate,

260 mafic flows and sills, sandstones and dolomites) and the upper Gangau subgroup contain-

261 ing phosphorites and sandstone. The total thickness is thought to be about 1000 m. The

262 Gwalior basin is about 25 kilometers wide and 80 kilometers long situated near the town of

263 Gwalior (fig. 6). Lithologies in the Gwalior Basin are similar to those in the Bijawars includ-

264 ing mafic flows and sills. The only available age constraints for the Bijawar and Gwalior ba-

265 sins are derived from Paleoproterozoic Rb-Sr and Sm-Nd ages on mafic rocks around

266 ~1800-1900 Ma (Crawford and Compston, 1970; Haldar and Ghosh 2000; Pandey et al.,

267 2012). A Paleoproterozoic age for the Bijawars is consistent with observations that show

268 an unconformable onlap of Lower Vindhyan sediments onto the Bijawar sediments

269 (Banerjee et al., 1982).

270 Singhbhum Craton

271 The Singhbhum Craton (also called the Singhbhum-Orissa craton; Figs. 1 & 7) lies in

272 the eastern part of India and borders the Mahanadi graben to the west, the Central Indian

273 Tectonic Zone (CITZ), and the Indo-Gangetic plain. It is bordered to the north by the

274 Chhotanagpur granite-gneiss terrain (CGGC). The CGGC is thought to be an extension of the

275 Central Indian Tectonic Zone (CITZ). The craton is subdivided into several different as-

276 semblages including the Older Metamorphic Group (OMG), the Older Metamorphic Tonalite

277 Gneisses (OMTG) the Singhbhum granite and the Iron Ore Group (IOG). The relationships

278 between the various units tend to be obscured by the scattered nature of the outcrop and

279 the paucity of reliable age data (see Mazumder et al., 2012). In our discussion we view the

280 Iron Ore Group (IOG) and the Older Metamorphic Group (OMG & OMTG) as broadly co-

281 genetic and the oldest units on the craton. Older detrital zircons found within these

282 gneissic units hint at an older basement that has been reworked. 11

283 The Iron Ore Group (IOG) is a greenstone-gneiss sequence (Eriksson et al., 2006;

284 Mondal et al., 2007). The entire IOG occurs as a supracrustal suite composed of three fold

285 belts: the Jamda-Koira, the Gorumahishani-Badampahar, and the Tomka-Daitari (Mondal et

286 al., 2007). It is divided into an Older and a Younger section, with similar compositions but

287 differing ages.

288 The Older IOG (southern sequence) is comprised of clastic sedimentary rocks

289 formed in a shallow marine setting along with syn-depositional volcanic rocks that togeth-

290 er suggest large scale rifting (Eriksson et al., 2006) or an arc-forearc sequence

291 (Mukhopadhyay et al., 2012). The Older IOG formed prior to the intrusion of the

292 Singhbhum Granite and was thought to have an age range between 3.3 and 3.1 Ga, based

293 solely on associations to nearby rocks and available ages for related rocks (Mondal et al.,

294 2007; Eriksson et al., 2006). A geochronologic study of the IOG (Mukhopadhyay et al.,

295 2008) yielded an age of 3.51 Ga for dacitic lavas. This crystallization age confirms that the

296 IOG formed just prior to the earliest Singhbhum (SGB-A) granites. Detrital zircons found in

297 the OMTG suite (see below) may have been derived from the Older IOG (Acharyya et al.,

298 2010).

299 The Older Metamorphic Group (OMG) and the Older Metamorphic Tonalite Group

300 (OMTG) outcrop as irregularly distributed enclaves within the Singhbhum granitoids. They

301 include micaceous schists, quartzites, calc-silicate, para-amphibolites, ortho-amphibolites

302 and tonalite-trondhjemite gneisses. The OMG and the OMTG are commonly described as

303 separate elements within the Singhbhum nucleus, but sparse geochronological data from

304 both the OMG and the OMTG rocks ranges in age from 3.2-3.5 Ga (Mondal et al., 2007; Misra

305 et al., 1999; Acharyya et al., 2010). Detrital and xenocrystic zircons from both groups range

306 from 3.5-3.8 Ga (Saha, 1994; Naqvi and Rogers, 1987; Misra et al., 1999). The younger,

307 Mesoarchean ages are thought to reflect metamorphic events within the belt (3.2-3.3 Ga).

308 Acharyya et al. (2010) noted that the older end of the age range in the OMG and OMTG

309 overlap with those in the lower-grade rocks of the Iron Ore Group. They argued that the

310 OMG and OMTG are metamorphosed equivalents to the rocks in the Older IOG and we

311 agree with that conclusion. Therefore, the Older IOG, the OMG and the OMTG represent the

312 oldest suites of outcrop in the Singhbhum craton. 12

313 The OMG is intruded by the approximately 10,000 km2 Singhbhum Granitic complex

314 (fig. 7 and 8). The complex includes a dozen separate domal magmatic bodies. Whether

315 these plutons were emplaced in a single magmatic event or several remains a subject of

316 debate; however older data combined with recent studies suggest polyphase emplacement

317 of granitoids spanning nearly 500 million years of the Archaean (~3.5-3.0 Ga; Naqvi & Rog-

318 ers, 1987; Acharrya et al., 2010). These are sometimes referred to as Singhbhum A (SGA-

319 older group) and Singhbhum B (SGB-younger group) granitoids (Saha, 1994;

320 Mukhopadhyay, 2001; Acharyya et al, 2010)

321 The Singhbhum Granite (SG) complex includes two different types of granite. One

322 set of granites displays HREE depletion and is dated at 3.3 Ga (Mondal et al., 2007). Other

323 varieties of granites produce a fractionated LREE pattern and flat HREE and are dated at

324 ~3.1 Ga (Mondal et al., 2007). Misra et al. (1999) report an age of 3.33 Ga for the

325 Singhbhum ‘phase II” granites and ages of 3.08 Ga and 3.09 Ga for the Mayurbhanj granite.

326 Reddy et al. (2008) obtained SHRIMP U-Pb and Pb-Pb ages from the “Singhbhum” granite

327 with a discordant upper intercept age of 3.30 Ga and a more robust 207Pb-206Pb age for the

328 most concordant zircons of 3.29 Ga. The Sushina nepheline syenite body yielded the

329 youngest (SHRIMP) ages from the SG complex of ~0.9 Ga (Reddy et al., 2008). Acharyya et

330 al. (2008) reported U-Pb ages between 3.53 and 3.45 Ga from an early phase of Singhbhum

331 granite intrusion that appears to be coeval with the dacitic lavas within the Iron Ore Group.

332 The geochronological data from the Singhbhum granites therefore favors a multi-stage em-

333 placement. The oldest intrusions of granites at ~3.45-3.5 Ga were followed by a secondary

334 emplacement at 3.3 Ga, tertiary intrusions at 3.1 Ga and perhaps a young suite of granites

335 dated at ~0.9 Ga. More robust data on each of the granitic intrusions along with detailed

336 geologic mapping of their intrusive relationships will improve our understanding of the

337 development of the Singhbhum craton.

338 The Younger Iron Ore Group formed after the Phase I Singhbhum Granite intrusions

339 has a suggested depositional age > 2.55 and < 3.0 Ga. It is comprised of shallow or shelfal

340 marine along with greenstone and banded iron formation (Eriksson et al., 2006). An older

341 age for this sequence is possible given the 3.1 Ga age for the Mayurbanj granite that in-

342 trudes the Simlipal Basinal Group overlying the younger Iron Ore Group. The exact rela- 13

343 tionship between the younger Iron Ore Group and other volcano-sedimentary sequences in

344 the Singhbhum craton require more robust geochronological data.

345 Within the Singhbhum cratonic region there are a number of volcano-sedimentary

346 ‘basins’ that are difficult to correlate due to poor age control and scattered outcrop. We de-

347 scribe the current status of these regions below with the understanding that a single good

348 age may drastically alter our interpretation.

349

350 Simlipal/Dhanjori/Singhbhum Supracrustals

351

352 The Dhanjori basin (fig 7) rests unconformably over granitoid rocks of the

353 Singhbhum granite complex and is thought to be (along with the Simlipal Basin), the first

354 volcano-sedimentary sequence deposited after intrusion of the Singhbhum-B granites. The

355 Dhanjori basin contains terrestrial-fluvial deposits that are overlain by mafic-ultramafic

356 volcanic and volcaniclastic rocks (Bhattacharya and Mahapatra, 2008; Erickson et al., 2006;

357 Mazumder, 2005). Based on field observations and limited geochronological data,

358 Acharyya et al. (2008) proposed a late Archean/Earliest Paleoproterozoic age (~2.5-2.4 Ga)

359 for the Dhanjori sequence. In contrast, Mishra and Johnson (2005) argue for a much older

360 (>2.86 Ga) age for the bulk of the Dhanjori rocks on the basis of a Pb-Pb age for volcanic

361 rocks in the upper part of the basin. It should be noted that the significance of this age was

362 disputed by Roy and Sarkar (2006).

363 The relatively undeformed Simlipal Basin is located in the eastern part of the

364 Singhbhum craton. The basinal volcano-sedimentary sequence sits unconformably atop

365 the Iron Ore Group and Singbhum Phase II granites. It is intruded by the ~3.1 Ga

366 Mayurbanj Granite and gabbro. Assuming the field relationships and ages are correctly de-

367 termined, then deposition in the Simlipal basin spanned the interval from ~3.3-3.1 Ga

368 (Misra, 2006) although other authors consider both the Simlipal and Dhanjori suite of rocks

369 as Paleoproterozoic in age (Mazumder et al., 2012; Ramakrishnan and Vaidyanadhan,

370 2008).

371 The relationship between the members of the Singhbhum Group (Chaibasa and

372 Dhalbhum Formations) along with the Simlipal and Dhanjori Formations is equally enig-

373 matic (Saha et al., 1988; Sarkar and Saha, 1983; Gupta and Basu, 2000; Mazumder et al., 14

374 2012; Misra, 2006; Mazumder and van Loon, 2012). The Singhbhum Group consists of a

375 lower Chaibasa Formation dominated by schists along with amphibolites, quartzites and

376 tuffs and an upper Dhalbhum Formation containing phyllites, quartzites and mafic sills. In

377 some stratigraphic models, the Chaibasa conformably overlies the Dhanjori Formation and

378 has an uncomformable relation with the overlying Dhalbhum Formation (Gupta and Basu,

379 2000; Mazumder and van Loon, 2012) whereas in others the Dhanjori overlies the

380 Singhbhum Group (Saha et al., 1988). Misra (2006) considers the Singhbhum Group, the

381 Dhanjori Formation and the Simplipal Basinal sediments as cotemporaneous and of

382 Archean age whereas Mazumder et al. (2012) consider the sequences to be of

383 Paleoproterozoic age.

384 The Dalma meta-volcanic sequence conformably overlies the Dhalbhum Formation

385 and is dominated by mafic-ultramafic volcanic rocks. Age constraints on the Dalma se-

386 quence are poor. Misra and Johnson (2005) obtained whole rock Rb-Sr ages between 2.4-

387 2.5 Ga that they considered to reflect the timing of metamorphism. These authors consid-

388 ered the Dalma sequence to be roughly age-equivalent (~2.8 Ga) to the Dhanjori volcanics.

389 Roy and Sarkar (2006) argued that the Rb-Sr system appeared to be highly disturbed and

390 attached no significance to those ages.

391 The Chandil Formation is yet another enigmatic volcano-sedimentary sequence that

392 is geographically situated between the Chhotanagpur granite gneiss complex to the north

393 and the Dalma sequence to the south. Erickson et al. (2006) envisioned a fluvial-aeolian to

394 shallow marine setting for this sequence. As with most other supracrustal sequences on

395 the Singhbhum craton, the age of the Chandil Formation is poorly constrained. Tuffs within

396 the volcanic sequence are dated to ~1.5 Ga (Sengupta et al., 2000) whereas granites intrud-

397 ing the Chandil are dated to 1.6 Ga (Mazumder et al., 2005). Ages from the Chandil For-

398 mation (reported in abstract only) appear to be more robust U-Pb ages of ~1.63 Ga

399 (Mazumder and van Loon, 2012; Nelson et al., 2007; Reddy et al., 2009).

400 Controversy abounds on the age of the supracrustal sequences overlying the crystal-

401 line basement rocks of the Singhbhum craton. If the new ages (~1.63 Ga) on the Chandil

402 felsic volcanic rocks cited above are correct and stratigraphic continuity exists between the

403 Chandil Formation, Dalma volcanics and the Singhbhum Group, then all are likely 15

404 Paleoproterozoic in age (Mazumder and van Loon, 2012) rather than Neoarchean as envi-

405 sioned by Misra (2006).

406 In the absence of more robust geochronological data, it appears that most of the me-

407 ta-sedimentary sequences in the Singhbhum region are either Archean in age or part of

408 Purana-I basinal development within India.

409 Dyke swarms cut across much of the Singbhum craton and the largest suite of these

410 mafic to intermediate dykes is collectively known as the “Newer Dolerites” (Bose, 2008).

411 The Newer Dolerites are subdivided into at least two distinct generations, related by cross-

412 cutting relationships and distinct geochemical signatures. Emplacement ages are poorly

413 constrained, ranging from 1600 to 950 Ma, based mainly on K-Ar dating (Naqvi & Rogers,

414 1987; Bose, 2008; Srivastava et al., 2000). Bose (2008; and sources therein) suggests three

415 distinct pulses of magmatism, based mainly on available K-Ar data, at 2100 ± 100 Ma, 1500

416 ± 100 Ma, and 1100 ± 200 Ma. These ages should be viewed with caution until more robust

417 U-Pb ages become available.

418 The dykes vary from a few meters to 700 m in thickness and can extend for several

419 kilometers. They predominantly strike NNE-SSW or NNW-SSE. The dolerites exhibit a vari-

420 ety of textures including fine, medium and coarse-grained varieties (Mir et al., 2012).

421 The Chhotanagpur Granite Gneiss Complex (CGGC) forms the northern boundary of

422 the Singhbhum craton and is separated from it by the Singhbhum Shear Zone (SSZ; Figure

423 xx). The CGGC and the SSZ form a northeasterly extension of the Central Indian Tectonic

424 Zone (CITZ; Figures 1 & 6). The CGGC is composed of gneisses, granites and numerous

425 metasedimentary enclaves (Sharma, 2010). There have been numerous geochronological

426 studies on rocks within the CGGC, but relatively few U-Pb ages. The oldest rocks date to 2.3

427 Ga with the bulk of ages ranging between 1.6 and 0.9 Ga (see Misra, 2006 for a review).

428 Sharma (2010) argues that the CGGC represents a separate crustal block that accret-

429 ed to the Singhbhum craton during the Proterozoic. There are any number of alternative

430 models for the development of the CGGC and SSZ (See Misra, 2006). At the present time, it

431 is impossible to distinguish between the many options and it may be that the CGGC repre-

432 sents the margin of the Singbhum craton that was caught up in collisional orogenesis dur-

433 ing the suturing of the northern Indian cratons with the southern cratons along the CITZ

434 between 1.6-1.5 Ga. 16

435

436 The Kolhan Group

437

438 In the southern part of the Singhbhum craton, there is a minor supracrustal suite

439 known as the Kolhan Group (figures 6 & 7). The age of the Kolhan Group is unconstrained,

440 but Mukhopadhyay et al. (2006) argued that sedimentation likely began at about 1.1 Ga and

441 thus it would correspond to the Purana-II basinal sequences. The Kolhan Group formed in

442 an intracratonic basin with a westward slope and was subsequently deformed into a syn-

443 clinal structure. Elongate domes and basins and dome-in-dome structures dominate the

444 eastern part of the basin (Naqvi & Rogers, 1987). The Kolhan Group is subdivided into

445 three different formations; the Mungra sandstone (25 m thick), the Jinkphani limestone (80

446 m thick) and the Jetia shale (1000 meters). As a whole, the Kolhan Group is a transgressive

447 feature that is interpreted as having formed in a rift setting that is perhaps related to the

448 fragmentation of the Rodinia supercontinent (Bandopadhyay and Sengupta, 2004;

449 Mukhopadhyay et al., 2006).

450 Bastar Craton

451 The Bastar (a.k.a Bhandara or Central Indian) Craton in central India (Figure 8) is

452 bordered by the Godavari rift (to the south); by the Mahandi Rift (to the northeast); by the

453 Central Indian Tectonic Zone (to the north); by the Eastern Ghats mobile belt (to the east)

454 and the Deccan traps (to the west). The craton can be subdivided into distinct lithotectonic

455 assemblages. These include a suite of basement rocks collectively known as the “Gneissic

456 Complex” which includes the Amgaon and Sukma gneisses. The second dominant suite in-

457 cludes granitoid bodies of different ages that are intrusive into the “Gneissic Complex”.

458 The basement “Gneissic Complex” is dominated by tonalite-trondhjemite gneisses

459 and granites with ages between 2.5-2.6 Ga that are interpreted to reflect a major interval of

460 crustal accretion (Ramakrishnan & Vaidyanadhan, 2008; Santosh et al., 2004; Sarkar et al.,

461 1981; Sarkar et al., 1990; Sarkar et al., 1993). The oldest ages reported from the basement

462 rocks were derived from a tonalite sample with a U-Pb upper intercept age of ~3.6 Ga 17

463 (Ghosh, 2004), a ~3.6 Ga age from a granitic sample (Rajesh et al., 2009) and a 3.51 Ga age

464 from a xenocrystic within a gneiss (Sarkar et al., 1993). The former two ages of ~3.6

465 Ga represent the oldest rocks discovered so far within Peninsular India although there are

466 some older xenocrystic zircons from the Singhbhum craton (see above).

467

468 Supracrustal Sequences

469

470 The Bastar Craton contains at least three major supracrustal/volcanic sequences of

471 rocks (Fig. 8), the Dongargarh, the Sakoli, and the Sausar suites along with numerous scat-

472 tered enclaves. There are very few age constraints on any of these units and most are arbi-

473 trarily assigned to the Paleoproterozoic. Further subdivisions and supracrustal sequences

474 are described in the Bastar craton, but a lack of geochronological data makes it extremely

475 difficult to discern if they are truly distinct units or merely poorly correlated across the

476 craton. Thus we adopt a more simplistic subdivision for the purposes of this introduction.

477 Dongargarh Supergroup

478 The Dongargarh Supergroup extends from the Chhattisgarh basin in the east to the

479 Sakoli in the west and is composed of three smaller groups of rocks, the Amagaon,

480 Nandgaon, and Khairagarh groups (Figure 8).

481 The Amagaon granites and gneisses are presumed to have formed during the

482 Amagaon Orogeny at ca. 2.3 Ga. The group consists mainly of gneisses with secondary

483 schists and quartzites (Naqvi & Rogers, 1987).

484 The Nandgaon Group contains two volcanic suites (the Bijli and Pitepani suites) that

485 are dominated by rhyolites with secondary dacites, andesites, and basalts (Neogi et al.,

486 1996). The Bijli rhyolite was dated using Rb-Sr techniques at 2180 ± 25 Ma and 2503 ± 35

487 Ma (Sarkar et al., 1981; Krishnamurthy et al., 1988) and has localized inclusions of

488 Amagaon granite (Naqvi & Rogers, 1987). The Dongarhgarh volcanic rocks have Rb-Sr ages

489 of 2465 ± 22 Ma and 2270 ± 90 Ma (Sarkar et al., 1981; Krishnamurthy et al., 1988).

490 Chakraborty and Sensarma (2008) largely dismiss the inconsistency within the Rb-Sr data

491 and argue, on the basis of correlation with well-dated units in the Singhbhum craton, that 18

492 the Nandgoan Group was developed ~2.5 Ga. We view all these age estimates as tentative

493 until more precise U-Pb ages are acquired.

494 The Khairagarh Group unconformably overlies the Nandgaon and consists of shales,

495 sandstones, and igneous rocks. The basal formations are divided into a conformable se-

496 quence the basal Shale, the Bortalao formation, and an intra-trappean shale (Naqvi & Rog-

497 ers, 1987). Immediately overlying the Khairagarh Group are the Sitagota and Mangikhuta

498 volcanic suites separated by the Karutola sandstone (Neogi et al., 1996). The four volcanic

499 suites within the Dongargarh Group erupted periodically between ca. 2462 and 1367 Ma

500 (Neogi et al., 1996), but this age range is poorly constrained.

501 Sakoli Group

502 The Sakoli Group consists of low-grade metamorphic rocks of undetermined age in

503 a large synclinorium. The Sakoli Group is a volcano-sedimentary deposit comprised of

504 (youngest to oldest) slates and phyllites, bimodal volcanic suite and schists, metabasalts

505 and cherts and conglomerates and Banded Iron Formations (BIF’s; Bandyopadhyay et al.,

506 1990). Two stages of deformation are thought to have occurred, creating a sequence of

507 overfolded bedding and a period of progressive metamorphism followed by retrogression

508 (Naqvi & Rogers, 1987). Unconformably overlying the Sakoli Group are the Permo-Triassic

509 Gondwana Supergroup and the Late Cretaceous Deccan basalts.

510 The age of the Sakoli Group is not known. Rb-Sr ages on metavolcanics and tuffs

511 yield ages of 1295 ± 40 and 922 ± 33 Ma but the significance of these ages is difficult to in-

512 terpret (Bandopadhyay et al., 1990).

513 Sausar Group

514 The Sausar Group of metasediments and manganese-bearing ores were once

515 thought to be the oldest formations in central India. The Sausar polymetamorphic belt that

516 contains the sediments is part of the larger Central Indian Tectonic Zone (CITZ) and is ap-

517 proximately 300 kilometers in length and 70 kilometers in width (Naqvi and Rogers, 1987;

518 Figure 8). Detailed geochronologic studies are lacking within this belt; however, Roy et al.

519 (2006) argued that the main phase of metamorphism (-grade) took place be-

520 tween 800-900 Ma (Rb-Sr and Sm-Nd ages). They also noted that the Sausar Belt was

521 bounded on the north and south by granulite belts of different ages. The southern granulite

522 belt hosts a charnockite that yielded a Sm-Nd isochron age of 2672 ± 54 Ma. A mafic granu- 19

523 lite within the southern belt yielded a Sm-Nd age of 1403 ± 99 Ma. The northern granulite

524 yielded a Sm-Nd age of 1112 ± 77 Ma. The granulites in the north and south also yield Rb-

525 Sr isochron ages in the range of 800-900 Ma.

526 Bhowmik et al. (2009) suggested that the pre-1.0 Ga Indian landmass consisted of at

527 least three micro-continental blocks, the North Indian block, the South Indian Block and the

528 Marwar block that were united by ~1.0 Ga (fig. 1; Meert et al., 2010). Peak and retrograde

529 stages of metamorphism are recorded in schists from the central domain of the Central In-

530 dian Sausar Mobile Belt at 1062 ± 13 Ma and 993 ± 19 Ma monazite ages (Bhowmik et al.

531 2012). The Aravalli/Delhi region is also characterized by granitic intrusions with ages of

532 ~1.0 -1.1 Ga (Deb et al. 2001; Biju-Sekhar et al. 2003; Buick et al. 2006; Just et al. 2011). If

533 correct, then the Sausar sedimentary sequence should be considered part of the Purana-II

534 basins of Peninsular India.

535 In contrast Stein et al. (2004) argue that the juxtaposition between the northern and

536 southern Indian cratonic nuclei along the CITZ took place during the earliest

537 Paleoproterozoic based on Re-Os ages from within the Sausar Belt (Malanjkhand granitoid

538 batholith). They report a Re-Os age of 2490 ± 2 Ma for the granitoid that is nearly identical

539 to U-Pb zircon ages of 2478 ± 9 Ma and 2477 ± 10 Ma for the same unit (Ranigrahi et al.,

540 2002). Cu-Mo-Ag mineralization ages associated with the intrusions ranged from 2446-

541 2475 Ma (Stein et al., 2004). Stein et al. (2004) note that the region underwent significant

542 ~1100-1000 Ma reworking, but the main assembly of cratons occurred during the latest

543 Archaen to earliest Paleoproterozoic (~2.5 Ga) along the Sausar Belt (e.g. CITZ).

544 Mafic Dyke Swarms

545 The Bastar Craton is intruded by numerous mafic dyke swarms, spanning an area of

546 at least 17,000 km3, that cross cut the various granitoids and supracrustal rocks of the re-

547 gion (French et al., 2008). The swarms are given regional names, but many may belong to

548 the same intrusive episode. These include the Gidam-Tongpal swarm, the

549 Bhanupratappur-Keskal swarm, the Narainpur-Kondagaon swarm and the Bijapur-Sukma

550 swarm (Ramachandra et al., 1995). A majority of the dykes in the southern Bastar craton

551 trend NW-SE, paralleling the Godavari rift, and these dykes are thought to have exploited 20

552 preexisting faults. The northern dykes are oblique to the Mahanadi rift in a NNW-SSE di-

553 rection (French et al., 2008).

554 Geochronologic constraints on many of the swarms are poor although recent work

555 suggests a major episode of igneous activity and dyke intrusion around 1.9 Ga (French et

556 al., 2008). The Paleoproterozoic dyke swarms are dated using U-Pb baddeleyite/zircon

557 techniques at 1891.1 ± 0.1 Ma and 1883 ± 1.4 Ma and include boninite-norite and sub-

558 alkaline mafic dykes, most of which display some degree of metamorphism (Srivastava,

559 2008; Srivastava et al., 2004; French et al., 2008). French et al. (2008) and Srivastava et al.

560 (2008 and 2004) interpret the Precambrian dyke swarms as remnants of a large igneous

561 province. French et al. (2008) noted that this activity is coeval with mafic magmatism in

562 both the Superior craton of North America and along the northern margin of the Kaapvaal

563 craton although they did not link the regions together paleogeographically and instead ar-

564 gued for a mantle upwelling on a global scale. In contrast, Srivastava and Singh (2003)

565 linked the dykes to Laurentia and Antarctica in a “Columbia-type” paleogeography.

566 The younger dyke swarms represent the youngest igneous events in the Bastar

567 Craton and mainly include metagabbros and metadolerites (Subba Rao et al., 2008).

568 Hussain et al. (2008) postulated that these dykes were derived from subduction constitu-

569 ents that were altered in the mantle lithosphere. A subduction related genesis is consistent

570 with the increased incompatible lithophile elements seen in the geochemical analysis

571 (Subba Rao et al., 2008), but this does not preclude different genetic models for the young-

572 er dykes. Recent geochronological evidence from the more felsic end-members of these

573 dykes (Lakhna Swarm) yielded a SHRIMP U-Pb age of 1450 Ma (Ratre et al., 2010) and U-

574 Pb zircon age of 1466 Ma for a rhyolitic dyke (Pisarevsky et al., 2013).

575 Sedimentary Basins

576 The Bastar Craton contains two large Purana-II basins, the Chhattisgarh Basin and

577 the Indravati Basin along with four minor basins with poorer age constraints (fig 8 & 9). It

578 is likely that these were all part of a single large basin and the outcrops are now isolated via

579 differential erosion.

580 The 36,000 km2 Chhattisgarh Basin is comprised of a ~1500 meter thick sedimen-

581 tary layer (the Chhattisgarh Supergroup) of conglomerates, orthoquartzites, sandstones, 21

582 shales, limestones, cherts, and dolomites (Fig 9; Naqvi & Rogers, 1987). The sedimentary

583 sequence has been divided into a basal Chandarpur series and an upper Raipur series

584 (Naqvi & Rogers, 1987; Patranabis-Deb et al., 2007). The Chandarpur Group consists of a

585 shale-dominated sequence containing conglomerate and coarse arkose sandstone formed

586 as coalescing fan-fan delta deposits, storm-dominated shelf deposits, and high-energy

587 shoreface deposits. The Raipur Group, however, underwent outer shelf, slope and basin

588 deposition and consists of a limestone-shale dominated sequence (Chaudhuri et al., 2002).

589 In the eastern part of the basin lies the “Purana” succession. The Purana contains a

590 proximal conglomerate-shale-sandstone assemblage and a distal limestone-shale assem-

591 blage. The conglomerate-shale-sandstone assemblage unconformably overlies the base-

592 ment and is thought to correspond to the Chandarpur group. The limestone-shale assem-

593 blage, on the other hand, is thought to correspond to the Raipur series (Deb, 2004).

594 The age of the Chhattisgarh Supergroup is becoming better established. Previous

595 thoughts extended deposition within the Chhattisgarh basin to as young as 500 Ma (Naqvi,

596 2005). However, rhyolitic tuffs near the top of the Chhattisgarh sequence (the Sukhda and

597 Sapos Tuffs) yielded ages of 1011 ± 19 Ma and 990 ± 23 Ma (Sukhda tuff) and 1020 ± 15 Ma

598 (Sapos tuff) using U-Pb SHRIMP techniques on magmatic zircons (Patranabis-Deb et al,

599 2007). This led the authors of that paper to conclude that the Purana basins may be up to

600 500 Ma older than the ‘consensus’ agreement. A tuff from the basal part of the Chhattis-

601 garh (Singhora Group) is dated to ~1500 Ma.

602

603 Indravati, Khariar, Pakhal and Sukma Basins

604

605 The 9000 square-kilometer Indravati Basin consists of shales, dolomites, sand-

606 stones, quartz arenites, limestones, and conglomerates showing little to no deformation or

607 metamorphism (figure 9). The sediments are thought to have a shallow marine or lagoonal

608 depositional environment (Maheshwari et al., 2005). The basin is lithologically similar to

609 the Chhattisgarh and it is postulated that at one point the two were connected and later

610 eroded into what are now discrete basins (Naqvi & Rogers, 1987). The sandstone member

611 is correlated with the Chopardih Formation of the Chhattisgarh Basin and was dated to 22

612 700-750 Ma using K-Ar on glauconite (Kruezer, 1977). More recently, Mukherjee et al.

613 (2012) dated a tuff layer near the top of the Indravati sequence (Jagdalpur Formation) to

614 ~1000 Ma strengthening the connection to the Chhattisgarh Basin.

615 Less is known about the stratigraphy and age of the remaining smaller basins

616 (Khariar, Pakhal, Sukma; see fig 8). Das et al. (2010) dated monazite grains found in a tuff

617 layer near the base of the Khariar Basin and obtained a cluster of ages at 1455 ± 47 Ma.

618 Dykes of the Lakhna swarm (1450-1460 Ma) intrude the Bastar craton and Ratre et al.

619 (2010) argue that these dykes are eroded by the overlying Khariar sediments and thus the

620 onset of sedimentation must be younger than ~1460 Ma (Purana-II). Das et al. (2011) ar-

621 gued that this lower part of the Khariar Basin is correlative to the Singhora Group of the

622 Chhattisgarh Basin. We agree with this conclusion, but note that there is disagreement re-

623 garding the closure age of these basins (Ratre et al., 2011).

624

625 Dharwar Craton

626 Eastern Dharwar Domain

627 The Dharwar craton is split into eastern and western domains. The western bound-

628 ary of the Eastern Dharwar domain (EDD) is poorly defined and is constrained to a 200 km

629 wide lithologic transitional zone from the peninsular gneisses of the Western Dharwar

630 craton to the Closepet Granite (fig. 10). The Closepet granite is a good approximation of the

631 western boundary (Ramakrishnan & Vaidyanadhan, 2008). The EDD is bounded to the

632 north by the Deccan traps and the Bastar craton, to the east by the Eastern Ghats mobile

633 belt, and to the south by the Southern Granulite terrane (Balakrishnan et al., 1999). Major

634 lithotectonic units include the Dharwar Batholith (dominantly granitic), greenstone belts,

635 intrusive volcanics, and middle Proterozoic to more recent sedimentary basins (Fig. 10;

636 Naqvi and Rogers, 1987; Ramakrishnan & Vaidyanadhan, 2008). There are a number of

637 sedimentary basins resting on the basement of the Eastern Dharwar domain including the

638 Cuddapah, Pranhita-Godavari and Kurnool (figures 9 and 10).

639 Greenstone Belts

640 Greenstone and belts of the EDD are concentrated in the West. These N-S

641 trending belts diminish in the east and are ultimately covered by the Cuddapah Basin 23

642 (Ramakrishnan and Vaidyanadhan, 2008). Metamorphism is limited to

643 greenschist/amphibolite facies (Chadwick et al., 2000).

644 A number of reliable age determinations have been published in recent years for

645 some of the larger belts (Rogers et al., 2007; Anand, 2007; Sarma et al., 2008; Chardon et

646 al., 2002; Jayananda et al., 2012).

647 Age determinations from the ~40 kilometer wide Kolar Schist Belt range from 3.1 to

648 about 2.5 Ga (Balakrishnan, 1990; Krogstad et al., 1991; Jayananda et al., 2000, 2013; Char-

649 don, 2002; Jayananda et al., 2013). The bulk of the ages cluster around 2.5-2.6 Ga and

650 Jayananda et al (2013) suggest a three-phase evolution of the belt with greenstone em-

651 placement around 3.1 Ga, felsic and granitic intrusions around 2.7 Ga and late-phase felsic

652 magmatism around 2.5 Ga.

653 The Sandur schist belt, located to the north of the Closepet granite, is characterized

654 by greenschist-facies metamorphism with amphibolite grade rocks occurring at the mar-

655 gins (Fig 10; Naqvi and Rogers, 1987). Granites within the Sandur Schist Belt were dated

656 using SHRIMP U-Pb at 2600-2500 Ma (Ramakrishnan and Vaidyanadhan, 2008). Rhyolites

657 from the Sandur greenstone belt yield a SHRIMP zircon U-Pb age of 2658 ± 14 Ma (Nutman

658 et al., 1996) and Naqvi et al. (2002) report Sm-Nd ages of 2706 ± 84 Ma for basalts and

659 komatiites.

660 The Ramagiri Penakacherla Sirigeri and Hundgund Belts (RPSH) represent two

661 discontinuus schist belts in the EDD. The RPSH belts are intruded by a series of granites

662 and gneisses that provide minimum age constraints for the metamorphic protoliths of >

663 ~2500 Ma. Basalts from the Ramigiri greenstone belt are dated to 2746 ± 64 Ma (Pb-Pb;

664 Zachariah et al., 1995) and appear to be coeval with those of the nearby Sandur greenstone.

665 Jayananda et al. (2013) summarized geochronological data from this region and ages range

666 from ~2470 Ma to 2707 Ma. The older ages (~2700 Ma) are derived from pyroclastic and

667 metabasalts whereas gneiss and granitoid ages from Chenna and Central Ramigiri belts

668 date to between 2550-2650 Ma (Balakrishan et al., 1999; Zachariah et al. (1995).

669 The Kolar-Kadiri-Jonnagiri-Hutti (KKJH) superbelt is located in the southern portion

670 of the EDD and is a discontinuous band of linear belts (Fig. 10). The southern portion of the

671 superbelt grades into a characteristic charnockitic terrain, while the north end (Kadiri belt) 24

672 disappears beneath the Cuddapah basin. Felsic volcanics in the Kadiri belt were recently

673 dated to 2556 Ma (Jayananda et al., 2013). The Kolar region contains mostly amphibolite

674 grade metamorphic rocks. As with the other greenstone belts in the region, the KKJH is in-

675 truded by various felsic dykes that provide minimum age constraints. Pb-Pb isochron data

676 provide an upper estimate at 2700 Ma for the protolith. This age is consistent with SHRIMP

677 U-Pb zircon analysis of granites and gneisses located in the belt (Ramakrishnan and

678 Vaidyanadhan, 2008). A second SHRIMP U-Pb zircon age of ~2550 Ma was found in vari-

679 ous intrusions within the KKJH which provides a younger limit for the superbelt (Rogers et

680 al., 2007). Ages in the Hutti Belt (granitoids, rhyolites, felsic volcanics cluster around 2550

681 Ma (Vasudev et al., 2000; Rogers et al., 2007; Anand, 2007; Sarma et al., 2008; Jayananda et

682 al., 2013).

683 The Velligallu-Raichur-Gadwal (VRG) belt is located to the south of, beneath, and

684 north of the Cuddapah basin (Fig. 10). The group is split to the south of the basin and

685 emerges to the north as a single unit before diverging again. The southern portion is divid-

686 ed by granite and is composed of metabasalts (amphibolites). The northern portion con-

687 tains pillowed metabasalts and boninites that are typically formed during the early stages

688 of subduction. Jayananda et al. (2012) report a SIMS U-Pb age of 2697 Ma for felsic volcan-

689 ic rocks in the Veligallu belt.

690 Maiban et al. (2011) obtained a number of 207Pb/206Pb ages from metasedimentary

691 rocks in the southern part of the Dharwar craton near . Older ‘cores’ yielded ages

692 in the range from 3.3-3.5 Ga (along with younger 2.5-2.9 Ga) ages.

693 In summary, age constraints on the greenstone belts in the Eastern Dharwar domain

694 are known from only a few locations and all appear to be Neo-Archaean in age as compared

695 to those in the Western Dharwar domain (WDD) described below. The relationships and

696 stratigraphy of the gneissic rocks in the region are difficult to discern mainly due to the

697 dismembered nature of the outcrop and the limited geochronology.

698

699 Dharwar Batholith

700

701 The Dharwar Batholith is a term first used by Chadwick et al. (2000) to describe a

702 series of parallel plutonic belts (fig. 10). Peninsular Gneisses refer to the majority of the fel- 25

703 sic basement rocks within the Eastern Dharwar domain; however, the EDD rocks are com-

704 positionally different than the WDD gneisses, more granitic than gneissic, and hence the

705 new terminology of Dharwar batholith is more appropriate for these plutonic belts

706 (Ramakrishnan and Vaidyanadhan, 2008). Age constraints from the WDD Peninsular

707 gneisses suggest an early Archaean age, whereas the granitic gneisses of the Dharwar

708 Batholith are of late Archaean age. The plutonic belts are approximately 15-25 km wide,

709 hundreds of km long and separated by greenstone belts (described above). They trend NW

710 to SE except for in the south where the trends become predominately north-south. The

711 belts are mostly mixtures of juvenile multipulse granites and diorites, (Chadwick et al.,

712 2000; Ramakrishnan and Vaidyanadhan, 2008). Geochronologic information for this unit

713 comes from SHRIMP U-Pb zircon measurements that constrain the emplacement of the

714 Dharwar Batholith to between 2700-2500 Ma (Nutman et al., 1996; Nutman and Ehlers,

715 1998; Friend and Nutman, 1991; Krogstad et al., 1995). Ages for granitic units appear to

716 decrease from west to east; however, gneissic protolith ages of > 2900 Ma are inferred

717 from inherited zircons within younger dykes near Harohalli intruding the gneissic rocks

718 (Pradhan et al., 2008).

719

720 Closepet Granite

721

722 The Closepet Granite is located on the western margin of the EDC and is a linear

723 batholith that trends ~North-South. The granite is 400 km long and approximately 20-30

724 km wide with sheared margins. Recent studies suggest that the similar convexity of adja-

725 cent schist belts and granitic plutons may indicate that the Closepet is a 'stitching pluton'

726 formed during the suturing of the Eastern and Western Dharwar cratons (fig. 10;

727 Ramakrishnan and Vaidyanadhan, 2008). The exposed rock is divided into northern and

728 southern components by a part of the Sandur Schist belt; however, both sections appear to

729 be lithologically similar at the outcrop level (Naqvi and Rogers, 1987). The Closepet Gran-

730 ite is dated to 2513 ± 5 Ma (Friend and Nutman, 1991) and appears to be part of a wide-

731 spread NeoArchaean phase of plutonism (Mojzsis et al., 2003; Maibam et al, 2011;

732 Jayananda et al., 2013) in both the eastern and western Dharwar domains that we consider

733 to mark the stabilization age for the WDD and EDD. 26

734

735 Post-Cratonization Intrusive Events

736

737 The majority of intrusive events of the Eastern Dharwar domain (EDD) are repre-

738 sented by mafic dykes, kimberlites and lamproites. Many of the clusters occur around the

739 Cuddapah basin and have three main trends: NW-SE, E-W, and NE-SW. These trends are

740 associated with various paleostress orientations during the Proterozoic to Late Cretaceous

741 (Srivastava and Shah, 2008). Most of the dykes disappear beneath the Cuddapah basin, in-

742 dicating that intrusion of the host granitic-gneiss took place before the basin developed.

743 These dykes all formed after the migmatitic activity of the host granitoids and are virtually

744 free of any effects of metamorphism and deformation (Chakrabarti et al., 2004). Five major

745 dyke clusters of the EDC, described below, include: (1) Hyderabad, (2) Mahbubnagar, (3)

746 Harohalli/Bangalore, (4) Anantapur and (5) Tirupati/Chitoor (Fig. 10).

747 The Hyderabad cluster is located to the north of the Cuddapah basin (Fig. 10).

748 Widely spaced NNE-SSW to N-S trending dykes traverse ENE-WSW and WNW-ESE oriented

749 dykes. The majority of the dykes present are doleritic in composition (Murthy, 1995).

750 Whole rock K-Ar ages of local dykes indicate emplacement between 1471 ± 54 Ma and

751 1335 ± 49 Ma (Mallikarjuna et al., 1995), but these well may reflect a younger isotopic dis-

752 turbance. Recent geochronological studies in the area demonstrate that some of the E-W

753 trending dykes are around 2.4 Ga (Kumar et al., 2012a) or N-S trending dykes around 2.2

754 Ga (Kumar et al., 2012b).

755 Located to the NW of the Cuddapah basin (Fig. 10), the Mahabubnagar dyke swarm

756 intrudes local granitic gneisses with Rb-Sr ages of 2.5-2.4 Ga and 2.2-2.1 Ga. The mafic

757 dykes are predominantly gabbroic; however, dolerite and metapyroxenite are also present.

758 They are oriented NW-SE and can be up to 50 km long and average 5-30 m wide. Chilled

759 margins are common with coarse aphyric or -rich interiors. Pooled regression

760 results from Sm-Nd analysis gives an emplacement age of 2173 ± 64 Ma (Pandey et al.,

761 1997). These results are duplicated by French et al. (2004), who obtained ages of ~2180

762 Ma using U-Pb techniques on near-by dykes. In light of the Sm-Nd and U-Pb ages for the

763 dykes, it appears that the 2.2-2.1 Ga Rb-Sr ages cited above for the gneisses in the region

764 may reflect disturbance due to dyke intrusion. 27

765 The Harohalli/Bangalore swarm is located between the southwestern portion of the

766 Cuddapah basin and the southeastern limb of the Closepet granite (Fig. 10). The dyke clus-

767 ter is split into an older group made up of dolerites, trending E-W (Bangalore dyke swarm),

768 and a younger group of alkaline dykes that trend approximately N-S (Harohalli alkaline

769 dykes; Pradhan et al., 2008). The Bangalore dyke swarm provided robust U-Pb ages of

770 2365.5 ± 1.1 Ma and 2370 ± 1 Ma (French et al., 2004; Halls et al., 2007). Initial Rb-Sr

771 whole rock measurements of the Harohalli alkaline dykes constrained ages to 850-800 Ma

772 (Ikramuddin and Steuber, 1976; Anil-Kumar et al., 1989). However, U-Pb ages of 1192 ± 10

773 Ma produced by Pradhan et al. (2008) on the alkaline dykes challenge these earlier esti-

774 mates. E-W, N-S and NE-SW trending dykes are also found nearby in the Tiptur-Hassan re-

775 gion. These dykes are assigned to the 2.4 Ga cluster (Halls et al., 2007; Belica et al, in

776 press), the 2.2 Ga swarm (Kumar et al., 2012b) or to the 1.9 Ga swarm (Belica et al., in

777 press).

778 Just west of the Cuddapah basin is the Anantapur dyke swarm (Fig. 10). This cluster

779 is less studied than other areas; however, some poorly constrained ages are available. The

780 NE-SW and ENE-WSW oriented dykes of the Anantapur swarm are dated using K-Ar meas-

781 urements and are poorly constrained between 1900-1700 Ma and 1500-1350 Ma, respec-

782 tively (Mallikarjuna et al., 1995; Murthy et al., 1987). Several detailed paleomagnetic stud-

783 ies on dykes from this region indicates that the dykes are part of several larger swarms

784 including the 2.4 Ga and 1.9 Ga swarms seen elsewhere in the Dharwar and Bastar cratons

785 (Halls et al., 2007; Piispa et al., 2011; Meert et al., 2011; Belica et al., in press). Additional

786 considerations based on cross-cutting relationships between dykes in this area suggest that

787 a ~2.0-2.1 Ga dyke swarm is also present in the region (Belica et al., in press).

788 To the south of the Cuddapah Basin is a very dense cluster of mostly E-W trending

789 dykes (with subordinate NW-SE trends). These are the so-called Tirupati (or Chitoor) clus-

790 ter. There are K-Ar and 40Ar/39Ar age determinations on dykes in the Tirupati/Chitoor

791 swarm. The E-W trending dykes have K-Ar ages of 1073 and 1349 Ma and one 40Ar/39Ar

792 total fusion age of 1333 ± 4 Ma (Mallikarjuna et al., 1995); however, it is likely that these

793 ages reflect some disturbance in the K-Ar system rather than crystallization ages. NW-SE

794 trending dykes have K-Ar ages of 935 and 1280 Ma (Mallikarjuna et al., 1995), but French 28

795 et al., (20xx) show that at least one of these dykes is ~2.2 Ga. Disturbed 40Ar/39Ar ages by

796 Goutham et al. (2011) yielded ages of 1200 Ma and 800 Ma, but the paleomagnetic direc-

797 tions in that study are consistent with Neoproterozoic overprints seen elsewhere in the

798 Dharwar craton or with well-dated 1.9 Ga paleomagnetic poles (Halls et al., 2007; Pradhan

799 et al., 2008; Belica et al., in press).

800 Kimberlites and lamproites are found in relative abundance in four areas within the

801 EDC. Concentrations can be found distributed around the Cuddapah Basin (Kumar et al.,

802 2007; Fig. 10). They are characteristically potassic volcanic rocks that are sometimes dia-

803 mondiferous. The main areas of kimberlite-lamproite intrusions are known as the

804 Wajrakarur, Narayanpet, Krishna and Nallamalai fields. Each of these fields contains mul-

805 tiple pipes. There are excellent age constraints on many of these fields. The Wajrakur field

806 is probably the best dated of the four. Rb-Sr ages on the Wajrakur field form a tight cluster

807 between 1091-1102 Ma and a recent U-Pb age on perovskite is 1124 +5/-3 Ma (Kumar et

808 al., 2007). A newly discovered cluster at Sidanpalli (north of Wajrakur) yielded an Rb-Sr

809 whole-rock mineral isochron age of 1093 ± 4 Ma (Kumar et al., 2007). Miller and Har-

810 graves (1994) report a U-Pb perovskite age of 1079 Ma for the Mulgiripalli pipe, but analyt-

811 ical details were not provided. Rb-Sr ages on kimberlites from the Kotakonda and Mudalbid

812 kimberlite intrusions yielded ages of 1084 ± 14 and 1098 ± 12 Ma (Kumar et al., 2001). It

813 should be noted that there are 40Ar/39Ar ages from the Kotakonda kimberlite that are much

814 older. Chalapathi-Rao et al. (1999) obtained plateau ages of 1401 ± 5 Ma for a phlogopite

815 separate from Kotakonda and 1417 ± 8 Ma from a lamproite at Chelima. The discrepancy

816 in the Rb-Sr and 40Ar/39Ar ages from Kotakonda were recently addressed by Gopalan and

817 Kumar (2008) who applied K-Ca dating to samples from the Kotakonda swarm and ob-

818 tained ages of 1068 ± 19 Ma. Gopalan and Kumar (2008) argue that the 40Ar/39Ar results of

819 Chalapathi Rao et al. (1999) are affected by excess argon and the Kotakonda field is ~1100

820 Ma. It is unclear if the lamproite in Chelima represents an older suite of lamproitic intru-

821 sion. It is possible that the kimberlitic intrusions into the Dharwar craton all occurred

822 within a relatively narrow time frame from ~1050-1100 Ma. It should be noted that many

823 other kimberlites around the globe were emplaced during this same interval of time includ-

824 ing elsewhere in India (Majhgawan, Madhya Pradesh for example). 29

825

826 Sedimentary Basins in the Dharwar Craton

827

828 Cuddapah Basin

829

830 The Cuddapah basin, located in the eastern portion of the EDD, is one of the better

831 studied basins in India (Figs. 9, 10 and 11). It covers an area of approximately 44,500 km2

832 and the convex western margin spans nearly 440 km. The eastern margin of the basin is

833 represented by a thrust fault while all other boundaries are part of the ‘Epi-Archaean Un-

834 conformity’ (a non-conformity associated with undisturbed contact to older Archaean

835 rocks). The sediments and minor volcanics of the basin are estimated to be approximately

836 12 km thick and made up of two distinct stratigraphic groups (fig. 11). The Cuddapah Su-

837 pergroup is the older unit and is present throughout the basin. The Kurnool Group was de-

838 posited unconformably over the Cuddapah rocks and is concentrated in the western por-

839 tion of the basin. The basin is surrounded by granitic gneisses, dykes, and sills, all of which

840 terminate at the basin boundary and appear to have formed before deposition. The young-

841 est igneous activity in the basin is the kimberlite and lamproite field located near the basin

842 center (Fig. 10; Chakrabarti et al., 2006).

843 Two competing hypotheses for the initiation of basinal subsidence and deposition

844 were forwarded. Chatterjee and Bhattacharji (2001) propose that the basin was formed

845 due to a mantle induced thermal trigger. Evidence for this comes from the presence of a

846 large subsurface mafic body in the southwestern portion of the basin that provided episod-

847 ic magmatism to form the abundant dykes and lava flows in and around the basin. The ag-

848 es for this magmatic event cluster around 1.9-2.1 Ga (ages of dykes, sills and volcanics near

849 the basal part of the Cuddapah Supergroup. A second hypothesis suggests that deep basin

850 margin faults played a major role in controlling the evolution of the basin (Chaudhuri et al.,

851 2002). Evidence for these marginal faults comes from seismic studies and Bouguer anoma-

852 ly interpretations.

853 Lower limits for the onset of basin formation (assuming a thermal origin) can be in-

854 ferred by ages of a mafic dyke on the southwest border of the craton and the Pullivendla sill

855 along the western margin of the basin. Chatterjee and Bhattacharji (2001) report a 40Ar- 30

856 3940 age of 1879 ± 5 Ma for the mafic dyke that is coeval with the 1885.4 ± 3.1 Ma U-Pb age

857 on the Pulivendla Sill by French et al. (2008). Unpublished paleomagnetic data from the

858 ~1.9 Ga Bastar dykes are identical to the Cuddapah traps volcanics, Cuddapah Basin sedi-

859 ments and the Pullivendla sill (Clark, 1982; Belica et al., in press). The available vidence

860 suggests a thermal pulse of ~1.9 Ga for the initiation of Purana-I basin formation in the

861 Cuddapah (Cuddapah Supergroup sediments).

862 Sedimentation in the Cuddapah Basin was discontinuous and numerous unconform-

863 ities exist within the Cuddapah Supergroup. A major unconformity separates the Cuddapah

864 Supergroup from the overlying Kurnool Group, but there are also important unconformi-

865 ties present within the Cuddapah Supergroup as well. An angular unconformity exists be-

866 tween the Chitravati Group and the overlying Nallamalai Group (figure 11). The Nallamalai

867 Group and overlying Srisailam quartzite may represent Purana-II basinal development alt-

868 hough there no age controls to confirm this hypothesis.

869 Age constraints on the Kurnool Supergroup are lacking, but Goutham et al. (2006)

870 correlate the Kurnool Group sediments with those in the Upper Vindhyan and assign all to

871 the Neoproterozoic; however, such a correlation is based more on tradition rather than on

872 strong correlative evidence and radiometric dating. The Kurnool Group would therefore be

873 one of the Purana-III basins in India.

874

875 Pranhita-Godavari Basin

876

877 The Pranhita-Godavari (P-G) basin is made up of two NW-SE trending basins sand-

878 wiched between the Dharwar and Bastar cratons (figures 1, 9 and 10). It is one of several

879 Purana basins formed (at least partially) on the Dharwar craton. The Cuddapah (see

880 above) lies to the south of the P-G basin and the Bhima (discussed below) basin lies to the

881 southwest. The Paleozoic-Mesozoic aged Gondwana sediments lay between the eastern

882 and western portions of the P-G basin (Chaudhuri, 2003; Ramakrishnan and Vaidyanadhan,

883 2008). The sedimentary sequence within the basin consists of a series of unconformity-

884 bounded packages reaching an aggregate thickness of ~6000 m. The rocks are mildly de-

885 formed and weakly metamorphosed. 31

886 There are numerous stratigraphic interpretations (and names) for the P-G sequence,

887 but we present the version favoured by Chaudhuri (2003) and Conrad et al. (2011). Ac-

888 cording to these classification schemes, the basinal sediments are collectively referred to as

889 the Godavari Supergroup and contain three unconformity-bounded groups (from oldest to

890 youngest) known as the Pakhal Group, the Penganga (or Albaka) Group and the Sullavai

891 Group (fig 9).

892 In the southwestern basin, the basal Pakhal Group is composed of two subunits

893 called the Mallampalli and Mulug Subgroups. The Mallampalli Subgroup is predominately

894 limestone and quartz arenite whereas the Mulug subgroup contains a basal conglomerate

895 followed by a carbonate-rich shelfal sequence. Unconformably overlying the Pakhal Group

896 is the Penganga (Albaka) Group composed of mature sandstones and shales and the

897 Chanda limestone. The uppermost Sullavai Group is floored by a conglomerate and is com-

898 posed of aeolian sandstones (Fig 9; Conrad et al., 2011).

899 The northeastern basin contains Mulug Subgroup (Pakhal Group), the Penganga

900 (Albaka) Group and the Sullavai Group sediments. The Proterozoic sedimentary sequence

901 is unconformably overlain by the Paleozoic-Mesozoic aged Gondwana Supergroup.

902 Age constraints are limited within the basin. Chaudhuri (2003) gives a range be-

903 tween 1330-790 Ma for the sequence. Conrad et al. (2011) provided 40Ar/39Ar ages for

904 diagenetic glauconite at various levels within the P-G basin. The Pakhal Group (Purana-1

905 Basin) yielded ages between 1565-1686 Ma and a single age determination of 1180 Ma was

906 obtained from the overlying Penganga Group. Given the lacunae present in the sequence

907 and the correlations made between the Prahnita-Godavari Basin (Dharwar craton) and the

908 nearby Chhattisgarh, Indravati and Khariar Basins (Bastar Craton), we conclude that all

909 three main basinal sequences (Purana-I, II and III) may be represented in this region of In-

910 dia (fig 9).

911

912 Bhima-Kaladgi Basins

913

914 The Bhima Basin is located between the northern margin of the EDD and the Deccan

915 Trap flows (Figs. 9 & 10). The basin is much smaller than the Cuddapah and covers 5,200

916 km2 with the longest portion having an axis of 160 km (NE-SW). The southern portion of 32

917 the basin is bounded by an unconformity with the underlying granitic gneisses while the E-

918 W and NW-SE borders are bounded by faults. The full extent of the basin is unknown due

919 to the Deccan Trap covering the basin in the north. The Bhima group is predominantly

920 composed of limestones; however, sandstone and conglomerate sediments exist between

921 the basement and the upper sequence limestones. The oldest age for the formation of the

922 Bhima basin is constrained by the underlying granitic gneisses to ~2500 Ma (Sastry et al.,

923 1999). It is currently under debate as to whether the basin formed during the Meso

924 (Purana II) or Neoproterozoic (Purana-III; Malone et al., 2008; Patranabis-Deb et al., 2007).

925

926 Western Dharwar Domain

927

928 The Western Dharwar Domain (WDD) is located in south-west India (Fig. 12). It is

929 bounded to the east by the Eastern Dharwar domain, to the west by the Arabian Sea, and to

930 the south by a transition into the so-called “Southern Granulite terrane.” The remaining

931 boundary to the north is buried under younger sediments and the Cretaceous Deccan

932 Traps. The division between the Western and Eastern Dharwar domains is based on the

933 nature and abundance of greenstones, as well as the age of surrounding basement and de-

934 gree of regional metamorphism (Rollinson et al., 1981).

935 The Archaean tonalitic-trondhjemitic-granodioritic (TTG) gneisses are found

936 throughout the Western Dharwar domain, dated at 3.3 to 3.4 Ga via whole rock Rb-Sr and

937 Pb-Pb methods (Pitchamuthu and Srinivasan, 1984; Bhaskar Rao et al., 1991; Naha et al.,

938 1991). U-Pb zircon ages ranging from 3.5 – 3.6 Ga have also been published. Three genera-

939 tions of volcanic-sedimentary greenstone granite sequences are present in the WDC: the

940 3.1 – 3.3 Ga Group, the 2.6 – 2.9 Ga Dharwar Supergroup (Radhakrishna and

941 Vaidyanadhan, 1997) and 2.5 – 2.6 Ga calc-alkaline to high potassic granitoids, the largest

942 of which is the Closepet Granite (Jayananda et al., 2008). The Dharwar supracrustal rocks

943 uncomformably overlie widespread gneiss- of the Complex

944 (3.0 – 3.3 Ga) that encloses the Sargur schist belts (Naqvi and Rogers, 1987).

945 The WDC shows an increase in regional metamorphic grade from greenschist and

946 amphibolite facies in the north and granulite facies in the south. The metamorphic grade

947 increase corresponds to a paleopressure increase from 3 – 4 kbar in the amphibolite facies 33

948 to as much as 9–10 kbar (35 km paleodepth) in the highest-grade granulite-transition zone

949 along the southern margin of the craton (Mojzsis et al., 2003). A nearly continuous cross

950 section of Late Archaean crust that has been tectonically upturned and channeled by ero-

951 sion is exposed in the WDC.

952 The Sargur Group

953 The Sargur Group greenstone belts display well-preserved volcano-sedimentary se-

954 quences. Generally these comprise of ultramafic to mafic volcanic rocks (komatiitic to

955 tholeiitic sources) that shows an up-section transition to felsic volcanic rocks, often inter-

956 preted to be related to a calc-alkaline source (e.g. Naqvi, 1981; Charan et al., 1988; Srikantia

957 and Bose, 1985; Srikantia and Venkataramana, 1989; Srikantia and Rao, 1990; Venkata

958 Dasu et al., 1991; Devapriyan et al., 1994; Subba Rao and Naqvi, 1999; Paranthaman, 2005).

959 These include the Ghattihosahalli, the J.C. Pura, the Bansandra area, the Kalyadi area, and

960 the Nuggihalli belt (Jayananda et al., 2008).

961 The Sargur Group developed from several distinct geodynamic processes across a

962 span of millions of years. Detrital zircons from the schist yield a Pb-Pb evaporation age of

963 ~ 3.3 Ga. SHRIMP U-Pb analysis yielded ages between 3.1 – 3.3 Ga, with some analyses

964 yielding a 3.6 Ga age inherited from the protolith. Sm-Nd model ages of ~ 3.1 Ga were cal-

965 culated from the ultramafic units. Rb-Sr dating on anorthosite fell into this range as well,

966 resulting in a 3.1 Ga age for the unit. When taken together, this geochronologic dataset may

967 constrain the age of the Sargur group to ~3.1 Ga. The older ages present in these analyses

968 are likely inherited from the basement material, and may represent the older limit of the

969 group. The Sargur unit appears to have formed in a subduction setting, likely derived from

970 the melting of oceanic slab materials (Martin, 1986). Komatiites found in the Sargur Group

971 are interpreted by Jayananda et al. (2008) to be related to plume events, and may have

972 originally been elements of oceanic plateaus. These accreted oceanic plateaus then served

973 as a base for further subduction related processes, represented by the series of mafic to fel-

974 sic volcanic units emplaced over and intruded the ultramafic plateau sequences (Jayananda

975 et al., 2008).

976 The komatiitic-tholeiitic volcanism observed in the WDC is part of a larger scale

977 process that led to the growth of the proto-craton. The 3.35 Ga volcanism appears to have

978 been pene-contemporaneous with the formation of the TTG basement, and provided host- 34

979 ing for the intrusion of the TTG protoliths. The melting events that lead to the ultra-mafic

980 volcanism occurred over a range of depths and co-existed with mantle peridotite; however,

981 evidence for the presence of garnet in the residue is unclear (Jayananda et al., 2008). Trace

982 element and Nd isotope data rule out the assimilation of continental materials into the

983 . Instead, the komatiite show the characteristic geochemical evidence of a

984 depleted mantle source (Boyet and Carlson, 2005). Mantle depletion at 3.35 Ga is poten-

985 tially significant. It would suggest that the earlier extraction of enriched materials from the

986 mantle had depleted the upper mantle prior to 3.35 Ga.

987

988 The Dharwar Supergroup

989 The Dharwar Supergroup is exposed in two large schist belts that have been divided

990 into two sub-sections, the Bababudan Group and the Chitradurga Group. The Bababudan

991 Group is spread over a 300 km long and 100 – 150 km wide area, and is made up of the

992 Babadudan schist belt, Western Ghats belt, and the Shimoga schist belt. The Bababudan

993 schist belt covers an area of approximately 2500 km2. The base of this unit is represented

994 by the Kartikere conglomerate that discontinuously extends along the southern margin of

995 the belt for ~ 40 km. This unit grades into a quartzite. The detrital zircon population from

996 the quartzite suggests that the sediments were mainly derived from the

997 . The overlying formations typically consist of metabasalts with intercalated

998 metasedimentary units, with occasional gabbroic sills, minor BIF, and phyllites. These are

999 thought to represent a variety of terrestrial environments, ranging from braided fluvial sys-

1000 tems to sub-aerial lava flows. The Western Ghats Belt is a large schist belt about 2200 km2

1001 in extent, and about 150 km by 15 km in dimension. The stratigraphy closely resembals

1002 the Babaudan belt; however, a major group of basalts, felsic volcanics, and pyroclastic units

1003 is also seen in the upper levels. The Shimoga schist belt is a large (25,000 km2) NW trend-

1004 ing belt separated from the previous two by outcropping TTG basement gneiss. The con-

1005 tact between these basement gneisses and the schist belt is observed as a zone of high

1006 grade metamorphism, often with kyanite and garnet phases present. Granitioid intrusions

1007 are also present in the north of the belt.

1008 Proterozoic Dyke Swarms 35

1009 Mafic dyke swarms varying in orientation and composition, intrude many areas of

1010 the WDC. Murthy et al. (1987) noted that the dykes are prevalent north of latitude 13°N

1011 and east of longitude 78°E, but that the dykes trend out towards latitude 12°N and are

1012 nearly gone south of latitude 11°N. All of the dykes post-date migmatitic activity in the host

1013 granitoids and are thus free of overprints of deformation and metamorphism.

1014 There are three main dyke swarms of Proterozoic age in the Western Dharwar

1015 Craton known as the (1) Hassan-Tiptur dykes; (2) dykes and (3) “Dharwar” dykes

1016 (Radhakrishna and Mathew, 1993).

1017 The Hassan-Tiptur dyke swarm contains two suites of dykes an older amphibolite

1018 and epidioritic swarm and younger and more widespread doleritic dykes. Age constraints

1019 are lacking on both suites of dykes. The Mysore dykes trend E-W and form a dense swarm

1020 near the town of Mysore. Recent geochronological and paleomagnetic studies on these

1021 dyke swarms indicates that three-phases of dyke intrusions are present in the swarms

1022 (~2.4 Ga, ~2.2 Ga and 1.9 Ga;

1023 Proterozoic Sedimentary Basins

1024 The E-W trending Kaladgi-Badami basin is the only significant Proterozoic

1025 intracratonic basin of the Western Darwar craton located along the northern edge of the

1026 craton (Fig. 12). This basin formed on TTG gneisses and greenstones of Archaean age. The

1027 Kaladgi Supergroup preserves the record of sedimentation in the basin, and consists of

1028 sandstones, mudstones and carbonates. The textural and mineralogical maturity of this ba-

1029 sin increased over time, indicating that the regional relief surrounding the basin declined

1030 over time, with the clastic sediments being derived from the local gneiss and greenstone

1031 rock (Dey et al., 2009). An angular unconformity between the two constituent groups (The

1032 lower Bagalkot and overlying Badami) suggests a period of uplift in the basins history

1033 (Jayaprakash et al., 1987). Deformation in Bagalkot group is significant, whereas the upper

1034 group only exhibited mild deformation (Kale and I'hansalkar, 1991).

1035

1036 Granitic Intrusions

1037

1038 Late to post-tectonic Dharwar potassic granite plutons (~ 2.5-2.6 Ga) that are as-

1039 sumed to reflect crustal reworking in WDC, occur as isolated intrusions cutting across the 36

1040 foliation and banding of the Peninsular Genisses (~ 3.0 Ga; Jayananda et al., 2006). In many

1041 cases, these plutons occur as distinct types either separately or as parts of larger composite

1042 intrusions, likely related to the generation of melts at differing depths within the crust

1043 (Sylvester, 1994). Several classes of TTG’s are present as well, broadly split into classical

1044 TTG and transitional TTG that formed 500 Ma later. These transitional TTG’s are believed

1045 to be lower crustal derived melts, and share the garnet residue signal of the high K gran-

1046 ites; however, this similarity may also indicate a mixing between these two melts

1047 (Jayananda et al., 2006). There is still uncertainty as to the role the late potassic granites

1048 played in the cratonization of the WDC. Jayananda et al. (2006) suggest that they may be

1049 related either to a thermal event prior to the termination of craton stabilization, or that

1050 they actually represent part of a longer term (~100 Ma) stabilization. Age data from Taylor

1051 et al. (1984) for the various intrusions range from 3080 ± 110 Ma (Rb-Sr) and 3175 ± 45

1052 Ma (Pb-Pb Isochron) for the Chikmagalur Granite to 2605 ± 18 Ma (Pb-Pb isochron). Much

1053 of the data is based on older, whole rock isotopic work.

1054 The Chitradurga Granite is an elongate, lenticular body of late to post-tectonic gran-

1055 ite, about 60 km long and 15 km wide. The granite is clearly intrusive into the Jogimaradi

1056 lavas of the Bababudan Group as well as into the TTG basement. The Chitradurga is

1057 granite grading into granodiorite and quartz monzonite. Chadwick et al. (2007) dated the

1058 granite using Pb-Pb and Rb-Sr isochrons yielding an age of ~ 2.6 Ga, as well as SIMS U-Pb

1059 zircon age of ~ 2610 Ma. The Jampalnaikankote Granite is a ~2.6 Ga (Rb-Sr) roughly oval

1060 shaped pluton that intrudes the Chitradurga schist belt. The Arsikere and Banavara Gran-

1061 ites are thought to be from a single pluton that is connected at depth. The Arsikere granitic

1062 batholith is approximately 75 km2 and oval in shape. The intrusion is primarily a potassic

1063 biotite granite that yielded a Rb-Sr age of ~ 2.6 Ga, and a SIMS U-Pb zircon age of ~ 2615

1064 Ma. The Chamundi Granite is another potassic pluton, with associated radial and parallel

1065 dykes, that intrudes the peninsular gneiss. The granite has been dated via Rb-Sr at ~ 800

1066 Ma.

1067 Summary

1068 Peninsular India is an amalgam of Archean nuclei that were sutured together by at

1069 least mid-Proterozoic time (~1.6 Ga) or perhaps by the end of the Archean. Following sta-

1070 bilization of the individual blocks, a series of Proterozoic to early Paleozoic sedimentary 37

1071 basins opened on the basins. These basinal sequences are most commonly known as the

1072 “Purana” basins and new age constraints on sedimentation are broadly consistent with tec-

1073 tonic events related to the assembly (or dispersal) of the supercontinents Columbia,

1074 Rodinia and Gondwana.

1075 1076 Acknowledgements: This research was made possible through grants by the US National Science 1077 Foundation (to JGM) EAR04-09101 and EAR09-10888. The authors also wish to thank the many students and 1078 colleagues who participated in various aspects of the work described in this paper. 1079 1080 1081

1082

1083 References 1084 1085 Acharyya, S.K., Gupta, A., Orihashi, Y., 2008. U-Pb zircon dates (LA-ICP-MS) of some felsic magmatic rocks from the basal parts of 1086 the Dhanjori basin and their stratigraphic implication, East Singhbhum, India. IAGR Conference Series 5, 151-152. 1087 Acharyya, S.K., Gupta, A., Orihashi, Y., 2010. New U-Pb zircon ages from Paleo-Mesoarchean TTG gneisses of the Singhbhum craton, 1088 Eastern India, Geochemical Journal, 44, 81-88. 1089 Anil-Kumar, Sivaraman, T.V., Bhaskararao, Y.J. and Gopalan, K., 1989. Rb-Sr ages of two dyke swarms from the Dharwar craton, Kar- 1090 nataka (abstract). Internationa Symposium on the Structure and Dynamics of the Indian Lithosphere NGR 1:3. 1091 Azmi, R.J., Joshi, D., Tewari, B.N., Joshi, M.N., Srivastava, S.S., 2008. A synoptic view on the current discordant geo-and 1092 biochronological ages of the Vindhyan Supergroup, central India, Journal of Himalayan Geology, 29, 177-191. 1093 Azmi, R.J., Joshi, D., Tewari, B.N., Joshi, M.N., Srivastava, S.S., 2008. A synoptic view on the current discordant geo- and 1094 biochronological ages of the Vindhyan Supergroup, central India, Journal of Himalayan Geology, 29, 177-191Bhowmik, S.K. and 1095 Dasgupta, S., 2012. Tectonothermal evolution of the Banded Gneiss Complex in central Rajasthan, NW India: Present status and cor- 1096 relation, Journal of Asian Earth Sciences, 49, 339-348. 1097 Balakrishnan, S., 1990. Late Archean suture zone at Kolar schist belt; isotopic and geochemical evidences. Group discussion on Suture 1098 zones, young and old; Seminar on Himalayan geology, Abstracts, p. 23. 1099 Balakrishnan, S., Hanson, G.N. & Rajamani, V., 1999. U-Pb Ages for Zircon and Titanite from the Ramagiri Area, Southern India: Evi- 1100 dence for Accretionary Origin of the Eastern Dharwar Craton during the Late Archean. Journal of Geology 107, 69-86. 1101 Bandyopadhyay B.K., Roy A. and Huin A.K., 1990. Structure and tectonics of a part of the central Indian shield. Geological Society of 1102 India Memoir 31, 433–467. 1103 Bandopadhyay, P. C. & Sengupta, S., 2004. The Paleoproterozoic supracrustal Kolhan Group in Singhbhum Craton, India and the Indo- 1104 African supercontinent. Gondwana Research, Japan (JPN) 7, 1228-1235. 1105 Banerjee, D.M., Khan, M.W.Y., Srivastava, N. and Saigal, G.C., 1982. Precambrian phosphorites in the Bijawar rocks of Hirapur-Bassia 1106 area, Sagar District, Madhya Pradesh, India. Mineralium Deposita, 17, 349-362. 1107 Basu, A. K., 1986. Geology of the Bundelkhand Granite massif, Central India. Records of the Geological Survey of India 117, 61-124. 1108 Basu, A.R., Sharma, M., Premo, W.R., 1996. U-Pb age of an Older Metamorphic Group schist: earliest terrain of the eastern Indian 1109 Craton. In: Saha, A.K. (Ed.), Recent Researches in Geology and Geophysics of the Precambrians, RRG vol. 16, Hindustan Publishing 1110 Corporation New Delhi, India, pp. 93–102. 1111 Barman, G., 1987. Stratigraphical position of the Marwar Supergroup in the light of stromatolite study. Geological Survey of India Spe- 1112 cial Publication #11, 72-80. 1113 Belica, M.E., Piispa, E.J., Meert, J.G., Pesonen, L.J., Plado, J., Pandit, M.K., Kamenov, G.D., 2014. Paleoproterozoic mafic dyke swarms 1114 from the Dharwar craton, India-constraining paleomagnetic poles from 2.37-1.88 Ga and rethinking the Columbia supercontinent, 1115 Precambrian Research, in press. 1116 Bhaskar Rao, Y.J., Naha, K., Srinivasan, R., Gopalan, K., 1991. Geology, geochemistry and geochronology of the Archaean Peninsular 1117 Gneiss around Gorur, Hassan District, , India. Indian Academy of Science (Earth and Planetary Sciences) Proceedings 100, 1118 399–412. 1119 Bhattacharya, H. N. & Mahapatra, S. 2008. Evolution of the Proterozoic rift margin sediments – North Singhbhum Mobile Belt, Jhar- 1120 khand–Orissa, India. Precambrian Research, 162, 302–316. 1121 Bhowmik, S. K., Saha, L., Dasgupta, S., and Fukuoka, M., 2009. Metamorphic phase relations in orthopyroxene-bearing granitoids: im- 1122 plication for high-pressure metamorphism and prograde melting in the continental crust. Journal of Metamorphic Geology 27, 295- 1123 315. 38

1124 Bhowmik, S. K., Wilde, S. A., Bhandari, A., Pal, T., and Pant, N. C., 2012, Growth of the Greater Indian Landmass and its assembly in 1125 Rodinia: Geochronological evidence from the Central Indian Tectonic Zone: Gondwana Research, 22, 54-72. 1126 Bickford, M. E., Basu, A., Mukherjee, A., Hietpas, J., Schieber, J., Patranabis-Deb, S., Ray, R. K., Guhey, R., Bhattacharya, P., and 1127 Dhang, P. C., 2011a, New U-Pb SHRIMP Zircon Ages of the Dhamda Tuff in the Mesoproterozoic Chhattisgarh Basin, Peninsular 1128 India: Stratigraphic Implications and Significance of a 1-Ga Thermal-Magmatic Event: Journal of Geology, v. 119, no. 5. 1129 Bickford, M. E., Basu, A., Patranabis-Deb, S., Dhang, P. C., and Schieber, J., 2011b, Depositional History of the Chhattisgarh Basin, 1130 Central India: Constraints from New SHRIMP Zircon Ages: Journal of Geology, v. 119, no. 1. 1131 Biju Sekhar, S., Yokoyama, K., Pandit, M.K., Okudaira, T. and Santosh, M., 2003. Paleoproterozoic magmatism in Delhi Fold Belt, NW 1132 India and its significance: evidence from EPMA chemical ages of zircons. Journal of Asian Earth Sciences, 22, 189-207. 1133 Bose, M.K., 2008. Petrology and geochemistry of Proterozoic ‘Newer Dolerite’ and associated ultramafic dykes within Singhbhum gran- 1134 ite pluton, eastern India. In: Srivastava, R.K. (ed) Indian Dykes: Geochemistry, Geophysics, and Geochronology. Narosa Publishing 1135 House Pvt. Ltd., New Delhi, India, 413-446. 1136 Boyet, M., Carlson, R.W., 2005. 143Nd evidence for early (>4.53 Ga) global differentiation of silicate Earth. Science 309, 577–581. 1137 Buick, I.S., Allen, C., Pandit, M., Rubatto, D., Herman, J., 2006. The Proterozoic magmatic and metamorphic history of the banded 1138 gneissic complex, central Rajasthan, India: LA-ICP-MS U–Pb zircon constraints. Precambrian Research 151, 119–142. 1139 Buick, I.S., Clark, C., Rubatto, D., Hermann, J., Pandit, M., Hand, M., 2010. Constraints on the Proterozoic evolution of the Aravalli- 1140 Delhi Orogenic belt (NW India) from monazite geochronology and mineral trace element geochemistry. Lithos 120, 511–528. 1141 Bhushan S. K., 2000. Malani rhyolites – a review. Gondwana Research 3, 65-77. 1142 Chadwick, B., Vasudev, V.N. & Hegde, G.V., 2000. The Dharwar craton, southern India, interpreted as the result of Late Archaean 1143 oblique convergence. Precambrian Research 99, 91-111. 1144 Chadwick, B., Vasudev, V., Hegde, G.V., Nutman, A.P., 2007. Structure and SHRIMP U/Pb zircon ages of granites adjacent to the 1145 Chitradurga schist belt: implications for Neoarchean convergence in the Dharwar craton, southern India, Journal of the Geological 1146 Society of India, Precambrian Research, 69, 5-24. 1147 Chakraborty, T. and Sensarma, S., 2008. Shallow marine and coastal eolian quartz arenites in the Neoarchean-Palaeoproterozoic 1148 Karutola Formation, Dongargarh Volcano-sedimentary succession, central India. Precambrian Research, 162, 284-301. 1149 Chalapathi-Rao, N.V., Miller, J.A., Gibson, S.A., Pyle, D.M., Madhavan, V., 1999. Precise 40Ar/39Ar dating of the Kotakanda kimberlite 1150 and Chelima lamproite, India: implication as to the timing of mafic dyke swarm activity in the Eastern Dharwar craton. Journal of the 1151 Geological Society of India 53, 425-433. 1152 Charan, S.N., Naqvi, S.M., Ramesh, S.L., 1988. Geology and geochemistry of spinifex-textured peridotitic komatiite from Mayasandra 1153 Schist Belt. Journal of the Geological Society of India 32, 343–350. 1154 Chardon, D., Peucat, J-J., Jayananda, M., Choukroune, P., Fanning, C.M., 2002. Archean granite–greenstone tectonics at Kolar (South 1155 India): interplay of diapirism and bulk inhomogenous shortening during juvenile magmatic accretion. Tectonics 21, 1016, 1156 http://dx.doi.org/10.1029/2001TC901032. 1157 Chatterjee, P., De, S, Ranaivoson, M., Mazumder, R., Arima, M., 2013. A review of the ~1600 Ma sedimentation, volcanism and 1158 tectonothermal events in the Singhbhum craton, Eastern India, Geoscience Frontiers, 4, 277-287. 1159 Chaudhuri, A.K. Saha, D., Deb, G.K., Deb, S.P., Mukherjee, M. K. & Ghosh, G., 2002. The Purana basins of southern cratonic province 1160 of India; a case for Mesoproterozoic fossil rifts. Gondwana Research 5, 23-33. 1161 Chaudhuri, A.K., 2003. Stratigraphy and paleogeography of the Godavari Supergroup in the south-central Pranhita-Godavari Valley, 1162 . Journal of Asian Earth Sciences 21, 595-611. 1163 Chatterjee, N. & Bhattacharji, S., 2001. Origin of the felsic and basaltic dikes and flows in the Rajula-Palitana-Sihor area of the Deccan 1164 Traps, Saurashtra, India; a geochemical and geochronological study. International Geology Review 43, 1094-1116. 1165 Clark, D.A., 1982. Preliminary paleomagnetic results from the Cuddapah traps of Andhra Pradesh, Monograph-2, On Evolution of the 1166 intracratonic Cuddapah Basin. HPG, Hyderabad, India, pp. 47–51. 1167 Conrad, J.E., Hein, J.R., Chaudhuri, A.K., Patranabis-Deb, S., Mukhopadhyay, J., Deb, G.K., Beukes, N., 2011. Constraints on the de- 1168 velopment of Proterozoic basins in Central India from 40Ar/39Ar analysis of authigenic glauconite minerals, Geological Society of 1169 America Bulletin, 123, 158-167. 1170 Crawford, A.R. and Compston, W., 1970. The age of the Vindhyan System of Peninsular India, Journal Geological Society of London, 1171 125, 351-371. 1172 Das, K., Yokoyama, K., Chakraborty, P.P., Sarkar, A., 2009. Basal Tuffs and Contemporaneity of the Chattisgarh and Khariar Basins 1173 Based on New Dates and Geochemistry. Journal of Geology, 117, 88-102. 1174 Das Gupta, S.P., 1996. Marwar Supergroup evaporates, Rajasthan. Geological Society of India Memoir, 36, 49-58. 1175 Das Gupta, S.K., Bulgauda, S.S., 1994. An overview of the geology and hydrocarbon occurrence in western part of Bikaner-Nagaur ba- 1176 sin. Indian Jour. Petrol. Geol., 3, 1-17. 1177 Davis, J.K., Meert, J.G., Pandit, M.K., 2014. Paleomagnetic analysis of the Marwar Supergroup, Rajasthan, India and proposed 1178 interbasinal correlations, Journal of Asian Earth Sciences, in press. 1179 Deb, M., 1999. Metallic mineral deposits of Rajasthan. In: P. Kataria (Ed.) Proc. Sem. on Geology of Rajasthan: Status and Perspective. 1180 Geology Dept. MLS University, Udaipur, India, pp. 213–237. 1181 Deb, M., Thorpe, R.I., Krstic, D., Corfu, F. and Davis, D.W., 2001. Zircon U-Pb and galena Pb isotopic evidence for an approximate 1.0 1182 Ga terrane constituting the western margin of the Aravalli-Delhi orogenic belt, northwestern India. Precambrian Research 108, 195- 1183 213. 1184 Deb, M., Thorpe, R., Krstic, D., 2002. Hindoli Group of rocks in the Eastern Fringe of the Aravalli-Delhi Orogenic belt- Archean sec- 1185 ondary greenstone belt or Proterozoic supracrustals? Gondwana Research, 5, 879-883. 39

1186 Deb, M. and Thorpe, R.I., 2004. Geochronological constraints in the Precambrian geology of Rajasthan and their metallogenic implica- 1187 tions in: Deb et al. (eds) Sediment-hosted lead-zinc sulphide deposits; attributes and models of some major deposits in India, Austral- 1188 ia and Canada, Narosa Publishing House, New Delhi, India, 246-263. 1189 Devapriyan, G.V., Anantharamu, T.R., Vidyadharan, K.T., Raghu Nandan, K.R., 1994. Spinifex textured peridotitic komatiite from 1190 Honnabetta area, Nagamangala schist belt, Karnataka. Journal of the Geological Society of India 44, 483–493. 1191 de Wall, H., Pandit, M.K., Chauhan, N.K., 2012. Paleosol at the Archean – Proterozoic contact in Udaipur. Precam Research 216-219, 1192 120 -131. 1193 Dey, S., Rai, A.K., Chaki, A., 2008. Palaeoweathering, composition and tectonics of provenance of the Proterozoic intracratonic 1194 Kaladgi-Badami basin, Karnataka, southern India: Evidence from sandstone petrography and geochemistry. Journal of Asian Earth 1195 Sciences doi: 10.1016/j.jseaes.2008.10.003 1196 Dharma Rao, C.V., Santosh, M., Purohit, R., Wang, J., Jiang, X., Kusky, T., 2011a. LA- ICP-MS U–Pb zircon age constraints on the 1197 Paleoproterozoic and Neoarchean history of the Sandmata Complex in Rajasthan within the NW Indian Plate. Journal of Asian Earth 1198 Sciences 42, 286–305. 1199 Dharma Rao, C.V., Santosh, M., Wu, Y., 2011b. Mesoproterozoic ophiolite mélange from the SE periphery of Indian plate: U–Pb zircon 1200 ages and tectonic implications. Gondwana Research 19, 384–401. 1201 Eriksson, P. G., Mazumder, R., Catuneanu, O., Bumby, A. J., Ilondo, B. O., 2006. Precambrian continental freeboard and geological evo- 1202 lution: A time perspective. Earth-Science Reviews 79, 165-204. 1203 Fareeduddin, Kröner, A., 1998. Single zircon age constraints on the evolution of Rajasthan granulite. In: Paliwal, B.S. (Ed.), The Indian 1204 Precambrian. Scientific Publishers, India, pp. 547–556. 1205 French, J.E., Heaman, L.M., Chacko, T. and Rivard, B., 2004. Global mafic magmatism and continental breakup at 2.2 Ga.: Evidence 1206 from the Dharwar craton, India. Geological Society of America Abstracts with Program, 36, 340. 1207 French, J.E., Heaman, L.M., Chacko, T., Srivastava, R.K., 2008. 1891–1883 a southern Bastar craton-Cuddapah mafic igneous events, 1208 India: a newly recognized large igneous province. Precambrian Research 160, 308–322. 1209 French, J.E. and Heaman, L.M., 2010. Precise U-Pb dating of Paleoprotoerozoic mafic dyke swarms of the Dharwar craton, India: Impli- 1210 cations for the existence of the Neoarchean supercraton Sclavia. Precambrian Research 183, 416-441. 1211 Friend, C.R.L., Nutman, A.P., 1991. SHRIMP U–Pb Geochronology of the Closepet Granite and Peninsular Gneiss, Karnataka, South 1212 India. Journal of the Geological Society of India 38, 357–368. 1213 Ghosh, J.G., 2004. 3.56 Ga tonalite in the central part of the Bastar Craton, India: oldest Indian date. Journal of Asian Earth Sciences, 23, 1214 359-364. 1215 Goodwin, A.M., 1991. Precambrian Geology: The Dynamic Evolution of the Continental Crust. Academic Press, London, 666 pp. 1216 Goutham, M.R., Raghubabu, K., Prasad, C.V.R.K., Subbarao, K.V. and Damodar Reddy, V., 2006. A Neoproterozoic Geomagnetic Field 1217 Reversal frpm the Kurnool Group, India: Implications for Stratigraphic Correlation and Formation of Gondwana. Journal of the Geo- 1218 logical Society of India, 67, 221-233. 1219 Gopalan K., Kumar A., 2008. Phlogopite K-Ca dating of Narayanpet kimberlites, south India: Implications to the discordance between 1220 their Rb-Sr and Ar/Ar ages. Precambrian Research 167, 377-382. 1221 Gopalan, K., MacDougall, J.D., Roy, A.B. and Murali, A.K., 1990. Sm-Nd evidence for 3.3 Ga old rocks in Rajasthan, northwestern In- 1222 dia. Precambrian Research, 48, 287-297. 1223 Goutham, M.R., Subbarao, K.V., Prasad, C.V.R.K., Piper, J.D.A., Miggins, D.P., 2011. Proterozoic mafic dykes from the southern mar- 1224 gin of the Cuddapah Basin, India: Part 2- Paleomagnetism and Ar-Ar geochronology, in: Srivastava et al. (eds) Dyke Swarms: Keys 1225 for Geodynamic Interpretation, Springer Press, Heidelberg, 73-93. 1226 Gregory, L.C., Meert, J.G., Pradhan, V., Pandit, M.K., Tamrat, E., Malone, S.J., 2006. A paleomagnetic and geochronologic study of the 1227 Majhgawan Kimberlite, India: Implications for the age of the Vindhyan Supergroup. Precambrian Research 149, 65-75. 1228 Gregory, L.C., Meert, J.G., Bingen, B.H. Pandit, M.K. and Torsvik, T.H., 2008. Paleomagnetic and geochronologic study of Malani 1229 Ingeous suite, NW India: implications for the configuration of Rodinia and the assembly of Gondwana, Precambrian Research, 170, 1230 13-26. 1231 Gupta, B.C., 1934. The geology of central Mewar, Memoirs of the Geological Society of India, 65, 107-168. 1232 Gupta, A. & Basu, A. 2000. North Singhbhum Proterozoic mobile belt Eastern India—a review. In: National Seminar Commemorating 1233 Dr M. S. Krishnan’s Birth centenary (Proceedings Volume). Geological Survey of India, Special Publications, 55, 195–226. 1234 1235 Gupta SN, Arora YK, Mathur RK, Iqballuddin, Prasad B, Sahai TN, Sharma SB, 1997, The Precambrian Geology of the Aravalli Re- 1236 gion, Southern Rajasthan and Northeastern Gujarat. Mem Geol Surv India 123: 262 pp. 1237 Haldar, D. and Ghosh, R.N., 2000. Eruption of Bijawar lava: An example of Precambrian volcanicity under stable cratonic conditions. 1238 Geological Survey of India Special Publication 57, 151-170. 1239 Halls, H.C., Kumar, A., Srinivasan, R. & Hamilton, M.A., 2007. Paleomagnetism and U/Pb geochronology of easterly trending dykes in 1240 the Dharwar Craton, India; clouding, radiating dyke swarms and the position of India at 2.37 Ga. Precambrian Research 155, 1241 47-68. 1242 Heron, A.M., 1953. Geology of central Rajputana, Memoirs of the Geological Society of India, 79, 339 pp. 1243 Hokada, T., Horie, K., Satish-Kumar, M., Ueno, Y., Nasheeth, A., Mishima, K., Shiriashi, K., 2013. An appraisal of Archaean 1244 supracrustal sequences in the Chitradurga Schist belt, Western Dharwar craton, southern India, Precambrian Research, 227, 99-119. 1245 Ikramuddin, M. & Stueber, A.M, 1976. Rb-Sr ages of Precambrian dolerite and alkaline dikes, Southeast Mysore State, India. Lithos 9, 1246 235-241. 1247 Iyenger, S. V. P. & Murthy, Y. G. K. 1982. The evolution of the Archaean Proterozoic crust in parts of Bihar and Orissa, eastern India. 1248 Geological Survey of India Record, 112, 1–5. 40

1249 Jayananda, M., Moyen, J.F., Martin, H., Peucat, J.J., Auvray, B., Mahabalesawar, B. 2000. Late Archaean (2550–2520 Ma) juvenile 1250 magmatism in the Eastern Dharwar craton, southern India: constraints from geochronology, Nd–Sr isotopes and whole rock geo- 1251 chemistry. Precambrian Research 99, 225–254. 1252 Jayananda, M., Chardon, D., Peucat, J-J, Capdevila, R., Martin, H., 2006. 2.61 Ga potassic granites and crustal re-working, western 1253 Dharwar craton (India): tectonic, geochronologic and geochemical constraints, Precambrian Research, 150, 1-26. 1254 Jayananda, M., Kano, T., Peucat, J.-J and Channabasappa, S., 2008. 3.35 Ga komatiite volcanism in the western Dharwar craton, south- 1255 ern India: Constraints from Nd isotopes and whole-rock geochemistry. Precambrian Research 162,160-179. 1256 Jayananda, M., Peucat, J-J., Chardon, D., Krishna-Rao, B., Fanning, C.M., Corfu, F., 2013. Neoarchean greenstone volcanism and con- 1257 tinental growth, Dharwar craton, southern India: Constraints from SIMS U-Pb zircon geochronology and Nd isotopes, Precambrian 1258 Research, 227, 55-76. 1259 Just, J., Schulz, B., de Wall, H., Jourdan, F., Pandit, M.K. 2011 Monazite CHIME/EPMA dating of the Erinpura granitoid deformation: 1260 implications for neoproterozic tectonothermal evolution of NW India. Gondwana Res., 19, 402-412. 1261 Kale, V.S., I’hansalkar, V.G., 1991. Purana basins of peninsular India: a review. Basin Research 3, 1-36. 1262 Kaur, P., Chaudhri, N., Okrusch, M., and Koepke, J., 2006, Palaeoproterozoic A-type felsic magmatism in the Khetri Copper Belt, Raja- 1263 sthan, northwestern India: petrologic and tectonic implications: Mineralogy and Petrology, 87, 81-122. 1264 Kaur, P., Chaudhri, N., Raczek, I., Kroener, A., and Hofmann, A. W., 2007a. Geochemistry, zircon ages and whole-rock Nd isotopic sys- 1265 tematics for Palaeoproterozoic A-type granitoids in the northern part of the Delhi belt, Rajasthan, NW India: implications for late 1266 Palaeoproterozoic crustal evolution of the Aravalli craton: Geological Magazine, 144, 361-378. 1267 Kaur, P., Chaudhri, N., Raczek, I., Kroener, A., Okrusch, M., and Hofmann, A. W., 2007b, Records of A-type and I-type plutonism from 1268 the northern Aravalli craton, NW India: Age, petrogenesis and regional tectonic implications: Geochimica Et Cosmochimica Acta, 1269 71, 15. 1270 Kaur, P., Chaudhri, N., Raczek, I., Kroener, A., and Hofmann, A. W., 2009, Record of 1.82 Ga Andean-type continental arc magmatism 1271 in NE Rajasthan, India: Insights from zircon and Sm-Nd ages, combined with Nd-Sr isotope geochemistry: Gondwana Research, 16, 1272 56-71. 1273 Kaur, P., Chaudhri, N., Raczek, I., Kroener, A., Hofmann, A. W., and Okrusch, M., 2011a, Zircon ages of late Palaeoproterozoic (ca. 1274 1.72-1.70 Ga) extension-related granitoids in NE Rajasthan, India: Regional and tectonic significance: Gondwana Research, 19, 4, 1275 1040-1053. 1276 Kaur, P., Zeh, A., Chaudhri, N., Gerdes, A., and Okrusch, M., 2011b, Archaean to Palaeoproterozoic crustal evolution of the Aravalli 1277 mountain range, NW India, and its hinterland: The U-Pb and Hf isotope record of detrital zircon: Precambrian Research, 187, 155- 1278 164. 1279 Khilnani, B.V., 1968. Stromatolites of Vindhyan limestone, from Bilara, district Jodhpur, Rajasthan. Quarterly Journal of the Geological 1280 Mineralogical and Meteorological Society of India, 40, 1-6. 1281 Khan, E.A., 1973. Discovery of fossil brachiopod of Cambrian affinity from Trans-Aravalli Vindhyan sequence, Jodhpur, Rajasthan. 1282 Seminar on recent advances in the geology of Rajasthan and Gujarat, Abstract 74-75. 1283 Khan, M.S., Smith, T.E., Raza, M., Huang, J., 2005. Geology, Geochemistry and tectonic significance of mafic-ultramafic rocks of 1284 Mesoproterozoic Phulad ophiolite suite of South Delhi Fold Belt, Gondwana Research, 8, 553-566. 1285 Krishnamurthy P., Chaki A., Pandey B.K., Chimote J.S. and Singh S.N., 1988. Geochronology of the granite–rhyolite suites of the 1286 Dongargarh supergroup central India. Proceedings of the Fourth National Symposium on Mass Spectrometry, pp. EPS-2/1–EPS-2/3. 1287 Krogstad, E.J., Hanson, G.N., Rajamani, V. 1991. U–Pb ages of zircon and sphene for two gneiss terranes adjacent to the Kolar Schist 1288 Belt, South India: evidence for separate crustal evolution histories. Journal of Geology 99, 801–816. 1289 Krogstad, E.J., Hanson, G.N. & Rajamani, V., 1995. Sources of continental magmatism adjacent to the late Archean Kolar suture zone, 1290 South India; distinct isotopic and elemental signatures of two late Archean magmatic series. Contributions to Mineralogy and Petrol- 1291 ogy 122, 159-173. 1292 Kruezer, H., Harre, W., Kursten, M., Schinitzer, W.A., Murthi, K.S. and Shrivastava N.K., 1977. K/Ar dates of two glauconites from the 1293 Chanderpur-series (Chhattisgarh/India): on the stratigraphic status of the Late Precambrian Basins in Central India. Geol. Jb., 28, 23- 1294 36. 1295 Kumar, A., Hamilton, M.A., Halls, H., 2012a. A Paleoproterozoic giant radiating dyke swarm in the Dharwar Craton, southern India. 1296 Geochem. Geophys. Geosyst., 13, doi:10.1029/2011GC003926. 1297 Kumar, A., Nagaraju, E., Besse, J., Rao, Y.J., Bhaskar, 2012b. New age, geochemical and paleomagnetic data on a 2.21 Ga dyke swarm 1298 from south India: Constraints on Paleoproterozoic reconstruction, Precambrian Research, 220–221, 123-138. 1299 Kumar, S., 2012. Stratigraphy and correlation of the Neoproterozoic deposits of central and western India: an overview, Geological So- 1300 ciety of London Special Publication, 366, 75-90. 1301 Kumar, S., and Pandey, S. K., 2008. Discovery of trilobite tracefossils from the Nagaur Sandstone, the Marwar Supergroup, Dulmera ar- 1302 ea, Bikaner District, Rajasthan. Current Science 94, 1081-1084. 1303 Kumar, S., and Pandey, S. K. 2009. Note on the occurrence of Arumberia Banksi and associated fossils from the Jodhpur Sandstone, 1304 Marwar Supergroup, Western Rajasthan. Journal of the Palaeontological Society of India 54, 41-48. 1305 Kumar, S., and Pandey, S.K., 2010. Trace fossils from the Nagaur Sandstone, Marwar Supergroup, Dulmera area, Bikaner district, Raja- 1306 sthan, India. Journal of Asian Earth Sciences 38, 77-85. 1307 Kumar, V., 1999. Eocambrian sedimentation in Nagaur–Ganganagar Evaporite Basin, Rajasthan. Journal of the Indian Association of 1308 Sedimentology, 18, 201-210. 1309 Leelanamdam, C., Burke, K., Ashwal, L.D., Webb, S.J., 2006. Proterozoic mountain building in Peninsular India: an analysis based 1310 primarily on alkaline rock distribution, Geological Magazine, 143, 195-212. 41

1311 Mahmoud, M.Y., Mitra, A.K., Dhar, R., Sarkar, S., Mandal, N., 2008. Repeated emplacement of syntectonic in Precambrian 1312 granite gneisses: Indication of pulsating brittle-ductile rheological transitions. In: Srivastava, R.K. (ed) Indian Dykes: Geochemistry, 1313 Geophysics, and Geochronology. Narosa Publishing House Pvt. Ltd., New Delhi, India, 495-510. 1314 Maibam, B., Goswami, J.N., Srinivasan, R., 2011. Pb-Pb zircon ages of Archean metasediments and gneisses from the Dharwar craton, 1315 southern India: Implications for the antiquity of the Eastern Dharwar craton, Journal of Earth System Science, 120, 643-661. 1316 Mallikarjuna, R.J., Bhattacharji, S., Rao, M.N. & Hermes, O.D., 1995. 40Ar- 39Ar ages and geochemical characteristics of dolerite dykes 1317 around the Proterozoic Cuddapah Basin, South India. Memoir - Geological Society of India 33, 307-328. 1318 Malone, S.J., Meert, J.G., Banerjee, D.M., Pandit, M.K., Tamrat, E., Kamenov, G.D., Pradhan, V.R., Sohl, L.E., 2008. Paleomagnetism 1319 and detrital zircon geochronology of the Upper Vindhyan sequence, Son Valley and Rajasthan, India: A ca. 1000 Ma closure age for 1320 the Purana basins? Precambrian Research 164, 137-159. 1321 Malviya, V.P., Arima, M., Pati, J.K., Kaneko, Y., 2004. First report of metamorphosed basaltic pillow lava from central part of the 1322 Bundelkhand craton, India: An island arc setting of possible Late Archaean age, Gondwana Research, 7, 1338-1340. 1323 Malviya, V.P., Arima, M., Pati, J.K., Kaneko, Y. 2006. Petrology and geochemistry of metamorphosed basaltic pillow lava and basaltic 1324 komatiite in Mauranipur area: subduction related volcanism in Archaean Bundelkhand craton, Central India, Journal of Mineralogical 1325 and Petrological Sciences, 101, 199-217. 1326 Martin, H., 1986. The effects of steeper Archaean geothermal gradients on geochemistry of subduction zone magmas. Geology 14, 753– 1327 756. 1328 Mazumder, A. and Bhattacharya, S.K., 2004. Stable isotopic study of late Neoproterozoic-Early Cambrian (?) sediments from Nagaur- 1329 Ganganagar Basin, western India; possible signatures of global and regional C-isotopic events, Geochemical Journal, 38, 163-175. 1330 Mazumder, A. and Strauss, H., 2006. Sulfur and Strontium isotopic compositions of carbonate and evaporate rocks from the late 1331 Neoproterozoic-early Cambrian Bilara Group: Constraints on intrabasinal correlation and global sulfur cycle., Precambrian Research, 1332 149, 217-230. 1333 Mazumder, R., 2005. Proterozoic sedimentation and volcanism in the Singhbhum crustal province, India and their implications; Sedi- 1334 mentary systems and sequence stratigraphy related to Precambrian sea level change; a special issue dedicated to Pradip K. Bose. Sed- 1335 imentary Geology 176, 167-193. 1336 Mazumder, R. and van Loon, A.J., 2012. Depositional history of diamictites from the Late Paleoproterozoic Dalma Formation (E India), 1337 Sedimentary Geology, 251–252, 49-57. 1338 Mazumder, R., Van Loon, A.J., Mallik, L., Reddy, S.M., Altermann, W., Eriksson, P.G., De, S., 2012. Mesoarchean-Paleoproterozoic 1339 stratigraphic record of the Singhbhum crustal province, eastern India: A Synthesis, Geological Society of London Special Publication 1340 365, 31-49. 1341 McKenzie, N. R., Hughes, N. C., Myrow, P. M., Xiao, S., and Sharma, M., 2011, Correlation of Precambrian-Cambrian sedimentary 1342 successions across northern India and the utility of isotopic signatures of Himalayan lithotectonic zones: Earth and Planetary Science 1343 Letters, v. 312, no. 3-4. 1344 Meert, J.G. and Lieberman, B.S., 2008. The Neoproterozoic assembly of Gondwana and its relationship to the Ediacaran-Cambrian Ra- 1345 diation, Gondwana Research, 14, 5-21. 1346 Meert, J.G., Pandit, M.K. and Kamenov, G.D., 2013. Further geochronological and paleomagnetic constraints on Malani (and pre- 1347 Malani) magmatism in NW India, Tectonophysics, in review. 1348 Meert, J.G., 2012. What’s in a name? The Columbia (Palaeopangea/Nuna) Supercontinent, Gondwana Research, 21, 987-993. 1349 Meert, J.G., Pandit, M.K., Pradhan, V.R., Kamenov, G.D., 2011. Preliminary report on the paleomagnetism of 1.88 Ga dykes from the 1350 Bastar and Dharwar cratons, Gondwana Research, 20, 335-343. 1351 Meert, J.G., Pandit, M.K., Pradhan, V.R., Banks, J.C., Sirianni, R., Stroud, M., Newstead, B., Gifford, J., 2010. The Precambrian tectonic 1352 evolution of India: A 3.0 billion year odyssey, J. Asian Earth Sci., 39, 483-515. 1353 Miller, K.C., Hargraves, R.B., 1994. Paleomagnetism of some Indian kimberlites and lamporites. Precambrian Research, 69, 259–267. 1354 Mir, A.R., Alvi, S.H., Balaram, V., 2012. Geochemistry of mafic dykes in the Singhbhum-Orissa craton: implications for subduction- 1355 related metasomatism of the mantle beneath the craton, International Geology Review, 52, 79-94. 1356 Misra, S., 2006. Precambrian chronostratigraphic growth of Singhbhum-Orissa craton, Eastern Indian Shield: An alternative model, 1357 Journal of the Geological Society of India, 67, 356-378. 1358 Misra, S., Deomurari, M. P.; Wiedenbeck, M.; Goswami, J. N.; Ray, S.; and Saha, A. K. 1999. 207Pb/206Pb zircon age ages and the evolu- 1359 tion of the Singhbhum craton, eastern India: an ion microprobe study. Precambrian Research, 93:139–151. 1360 Mishra, S., and Johnson, P.T., 2005, Geochronological constraints on evolution of the Singhbhum Mobile belt and associated basic 1361 volcanics of Eastern Indian Shield. Gondwana Research, 8, 129–142. 1362 Mojzsis, S.J., Devaraju, T.C. and Newton, R.C., 2003. Ion microprobe U-Pb determinations on zircon from the late Archaean granulite 1363 facies transition zone of southern India. Journal of Geology, 111, 407–425. 1364 Mondal, M. E. A., Deomurari, M. P., Goswami, J. N., Rahman, A. and Sharma, K. K., 1997. 207Pb/206Pb zircon ages of samples from 1365 Bundelkhand massif, Central India; Abstract in International Conference on Isotopes in Solar System, 80-81. 1366 Mondal, M.E.A., Goswami, J.N., Deomurari, M.P., Sharma, K.K., 2002. Ion microprobe 207Pb/206Pb ages of zircons from the 1367 Bundelkhand Massif, northern India: implications for crustal evolution of the Bundelkhand-Aravalli supercontinent. Precambrian 1368 Research 117, 85-100. 1369 Mondal, S. K., Frei, R., Ripley, E. M. 2007. Os isotope systematics of mesoarchean chromitite-PGE deposits in the Singhbhum Craton 1370 (India): Implications for the evolution of lithospheric mantle. Chemical Geology 244, 391-408. 1371 Mukherjee, A., Bickford, M.E., Hietpas, J., Schieber, J., Basu, A., 2012, Implications of a Newly Dated ca. 1000-Ma Rhyolitic Tuff in 1372 the Indravati Basin, Bastar Craton, India. Journal of Geology, 120, 477-485. 1373 Mukhopadhyay, D. (2001) The Archaean nucleus of Singhbhum: The present state of knowledge. Gondwana Res. 4, 307– 318. 42

1374 Mukhopadhyay, J., Ghosh, G., Nandi, A. K., Chaudhuri, A. K., Gutzmer, J., de Kock, M., 2006. Depositional setting of the Kolhan 1375 group; its implications for the development of a meso to neoproterozoic deep-water basin on the South Indian craton. South African 1376 Journal of Geology, 109, 183-192. 1377 Mukhopadhyay, J., Beukes, N.J., Armstrong, R.A., Zimmermann, U., Ghosh, G., Medda, R.A., 2008. Dating the oldest greenstone in In- 1378 dia: A 3.51 Ga precise U-Pb zircon age for dacitic lavas of the Southern Iron Ore Group, Singhbhum Craton, J. Geology, 116, 449- 1379 461. 1380 Mukhopadhyay, J., Ghosh, G., Zimmerman, U., Guho, S., Mukehrjee, T., 2012. A 3.51 Ga bimodal volcanics-BI-ultramafic succession 1381 from Singhbhum craton: implications for Paleoarchean geodynamic processes from the oldest greenstone succession of the Indian 1382 subcontinent, Geological Journal, 47, 284-311. 1383 Murthy, Y.G.K., Baburao, V., Guptasarma, D., Rao, J.M., and Rao, M.N., 1987. Tectonic, petrochemical and geophysical studies of 1384 mafic dyke swarms around the Proterozoic Cuddapah basin, south India; In: Mafic Dyke Swarms, H.C. Halls and W.F. Fahrig (Eds.), 1385 Geological Association of Canada Special Paper 34, 303-316. 1386 Naha, K., Srinivasan, R. and Jayaram, S., 1991. Sedimentological, structural and migmatitic history of the Archaean Dharwar tectonic 1387 province, southern India. Proceedings of the Indian Academy of Science (Earth Planetary Science), 100, 413-433. 1388 Naqvi, S.M., 1981. The oldest supracrustals of the Dharwar craton, India. Journal of the Geological Society of India 22, 458–469. 1389 Naqvi, S.M. 2005. Geology and the evolution of the India Plate. New Delhi, Capital, 450 p. 1390 Naqvi, S.M., Rogers, J. J.W., 1987. Precambrian . Oxford University Press Inc., 223. 1391 Naqvi, S.M., Manikyamba, C., Gnaneshwar, R., Subba Rao, T., Ram Mohan, D.V., Srinavasa Sarma, D., 2002. Geochemical and isotop- 1392 ic constraints of Neoarchean fossil plume for evolution of volcanic rocks of Sandur greenstone belt, India. Journal of the Geological 1393 Society of India, 47, 153-164. 1394 Neogi, S., Miura, H., Hariya, Y., 1996. Geochemistry of the Dongargarh volcanic rocks, Central India: implications for the Precambrian 1395 mantle. Precambrian Research, 76, 77-91. 1396 Nelson, D.R., Bhattacharyay, H.N., Mishra, S., Dasgupta, N., Altermann, W., 2007. New SHRIMP U–Pb zircon dates from the 1397 Singhbhum craton, Jharkhand–Orissa region, India. In: Banerjee, S. (Ed.), Abstracts International Conference on Precambrian Sedi- 1398 mentation & Tectonics & 2nd Global Precambrian Sedimentation Syndicate Meeting. Indian Institute of Technology, Bombay, p. 47. 1399 Nutman AP, Ehlers K., 1998. Evidence for multiple Palaeoproterozoic thermal events and magmatism adjacent to the Broken Hill Pb- 1400 Zn-Ag orebody, Australia. Precambrian Research, 90, 203-238. 1401 Nutman, A.P., Chadwick, B., Krishna-Rao, B., Vasudev, V.N., 1996. SHRIMP U-Pb zircon ages of acid volcanic rocks in the 1402 Chitradurga and Sandur Groups and granites adjacent to Sandur Schist Belt. Journal of the Geological Society of India, 47, 153-161. 1403 Pandey, D.K. and Bahadur, T., 2009. A review of the stratigraphy of Marwar Supergroup of west-central Rajasthan, Journal of the Geo- 1404 logical Society of India, 73, 747-758. 1405 Pandey, B.K., Gupta, J.N., Sarma, K.J. & Sastry, C.A., 1997. Sm–Nd, Pb–Pb and Rb–Sr geochronology and petrogenesis of the mafic 1406 dyke swarm of Mahbubnagar, South India: implications for Paleoproterozoic crustal evolution of the Eastern Dharwar Craton. Pre- 1407 cambrian Research, 84, 181–196. 1408 Pandey, U.K., Sastry, D.V.L.N., Pandey, B.K., Roy, M., Rawat, T.P.S., Ranjan, R. Shrivastava, V.K., 2012. Geochronological (Rb-Sr 1409 and Sm-Nd) studies on intrusive gabbros and dolerite dykes from parts of northern and central Indian cratons: implications for the age 1410 and onset of sedimentation in the Bijawar and Chhattisgarh basins and uranium mineralogy, Journal Geological Society of India, 79, 1411 30-40. 1412 Pandit, M. K., Shekhawat, L. S., Ferreira, V.P., Sial, A. N., Bohra, S. K., 1999. Trondhjemite and Granodiorite Assemblages from West 1413 of Barmer: Probable Basement for Malani Magmatism in Western India. Journal of Geological Society of India, 53, 89-96. 1414 Pandit, M.K., de Wall, H. and Chauhan, N.K. 2008 Paleosol at the Archean–Proterozoic contact in NW India revisited: Evidence for ox- 1415 idizing conditions during paleo-weathering? Journal of Earth System Science, 117, 201–209. 1416 Pandit, M. K., Sial, A.N., Jamrani, S.S. and Ferreira, V.P. (2001). Carbon isotopic profile across the Bilara Group rocks Trans-Aravalli 1417 Marwar Supergroup in western India: implications for Neoproterozoic – Cambrian transition. Gondwana Research 4, 387-394. 1418 Paranthaman, S., 2005. Geology and Geochemistry of Archaean Ghattihosahalli mafic-ultramafic complex, Chitradurga, Karnataka. 1419 Journal of the Geological Society of India, 66, 653–657. 1420 Pareek, H. S. 1984 Pre-Quaternary geology and mineral resources of northwestern Rajasthan. Mem Geol Surv India 115, 99p. 1421 Pati, J.K., Raju, S., Mamgain, V.D., Ravi Shanker, 1997. Gold mineralization in parts of Bundelkhand Granitoid Complex (BGC). Jour- 1422 nal Geological Society of India 50, 601-606. 1423 Patranabis-Deb, S., Bickford, M.E., Hill, B., Chaudhari, A.K., Basu, A., 2007. SHRIMP ages of zircon in the uppermost tuff in 1424 Chattisgarh Basin in central India require up to 500 Ma adjustments in Indian Proterozoic stratigraphy. Journal of Geology, 115, 407– 1425 416. 1426 Piispa, E. J., Smirnov, A. V., Pesonen, L. J., Lingadevaru, M., Anantha Murthy, K. S., & Devaraju, T. C., 2011. An integrated study of 1427 the paleomagnetism and geochemistry of Proterozoic dykes, Dharwar craton, southern India, in: Srivastava et al. (eds) Dyke Swarms: 1428 Keys for Geodynamic Interpretation, Springer Press, Heidelberg, 33-46. 1429 Pisarevsky, S.A., Biswal, T.K., Wang, X., de Waele, B., Ernst, R., Soderlund, U., Tait, J.A., Ratre, K., Singh, Y.K., Cleve, M., 2012. 1430 Paleomagnetic, geochronological and geochemical study of Mesoproterozoic Lakhna dykes, India: implications for the late 1431 Mesoproterozoic supercontinent, Lithos, doi:10.1016/j.lithos.2012.07.015. 1432 Pitchamuthu, C.S. and Srinivasan, R., 1984. The Dharwar Craton. Perspective Report Series, 7, Indian National Science Academy, 3-34. 1433 Pradhan, V.R., Pandit, M.K., Meert, J.G., 2008. A cautionary note on the age of the paleomagnetic pole obtained from the Harohalli dyke 1434 swarms, Dharwar craton, southern India. In: Srivastava, et al. (Ed.), Indian Dykes. Narosa Publishing House, New Delhi, India, 339– 1435 352. 43

1436 Pradhan, V.R., Meert, J.G., Pandit, M.K, Kamenov, G., Gregory, L.C. and Malone, S.J., 2010. India’s changing place in global Protero- 1437 zoic reconstructions: New geochronologic constraints on key paleomagnetic poles from the Dharwar and Aravalli/Bundelkhand 1438 cratons, Journal of Geodynamics, 50, 224-242. 1439 Pradhan, V.R., Meert, J.G., Pandit, M.K., Kamenov, G.D. and Mondal, E.A., 2012. Tectonic evolution of the Precambrian Bundelkhand 1440 craton, central India: Insights from paleomagnetic and geochronological studies on the mafic dyke swarms, Precambrian Research, 1441 198-199, 51-76. 1442 Radhakrishna, B.P. and Vaidyanadhan, R., 1997. Geology of Karnataka. Geological Society of India, 353 p. 1443 Radhakrishna, T. and Mathew, J., 1993. Proterozoic palaeomagmatism of the south Indian shield and tectonic constraints. Memoirs of 1444 the Geological Society of India, 26, 321–336. 1445 Raghav, K.S., De, C., Jain, R.L., 2005. The first record of Vendian Medusoids and trace fossil- bearing algal matgrounds from the basal 1446 part of the Marwar SuperGroup of Rajasthan, India. Indian Mineralogy, 59, 22–30. 1447 Rajesh, H.M., Mukhopadhyay, J., Beukes, N.J., Gutzmer, J., Belyanin, G.A., Armstrong, R.A. 2009. Evidence for an early Archaean 1448 granite from Bastar craton, India. Journal of the Geological Society of London, 166, 193–196 1449 Ramachandra H.M., Mishra V.P. and Deshmukh S.S., 1995. Mafic dykes in the Bastar Precambrian: study of the Bhanupratappur-Keskal 1450 mafic dyke swarm. In: T.C. Devaraju, Editor, Mafic Dyke Swarms of Peninsular, Memoirs of the Geological Society of India, 33, 1451 183–207. 1452 Ramakrishnan, M., and Vaidyanadhan, R., 2008. Geology of India:Volume 1. Geological Society of India, 994 pp. 1453 Ramakrishnan, M., 1990. Crustal Development in Southern Bastar Central Indian Craton, Geological Society of India Special Publica- 1454 tion 28, 44-66. 1455 Rasmussen, B., Bose, P.K., Sakar, S., Banerjee, S., Fletcher, I.R., McNaughton, N.J., 2002. 1.6 Ga U-Pb zircon age for the Chorhat 1456 Sandstone, Lower Vindhyan, India: Possible implications for the early evolution of animals. Geology, 20, 103-106. 1457 Ratre, K., de Waele, B., Biswal, T.K., Sinha, S., 2010. SHRIMP geochronology for the 1450 Ma Lakhna dyke swarm: Its implication for 1458 the presence of Eoarchaean crust in the Bastar Craton and 1450-517 Ma depositional age for Purana basin (Khariar), Eastern Indian 1459 Peninsula, Journal of Asian Earth Sciences, 30, 565-577. 1460 Rao, J.M., 2004. The wide-spread 2 Ga dyke activity in the Indian shield-evidences from Bundelkhand mafic dyke swarm, Central India 1461 and their tectonic implications. Gondwana Research 7, 1219-1228. 1462 Ray, J.S., Martin, M.W., Veizer, J., Bowring, S.A., 2002. U-Pb Zircon dating and Sr isotope systematic of the Vindhyan SuperGroup, 1463 India. Geology, 30, 131-134. 1464 Ray, J.S., Veizer, J., Davis, W.J., 2003. C, O, Sr and Pb isotope systematics of carbonate sequences of the Vindhyan Supergroup, India: 1465 Age, diagenesis, correlations, and implications for global events. Precambrian Research, 121, 103-140. 1466 Reddy, S., Clark, C. & Mazumder, R., 2008. Was there are pre-Nuna India-Antarctic connection? Evidence from SHRIMP U-Pb zircon 1467 dating, 33rd IGC Conference Abstracts 33, #1344159. 1468 Reddy, S., Clark, C., Mazumder, R., 2009. Temporal constraints on the evolution of the Singhbhum crustal province from U–Pb 1469 SHRIMP data. In: Saha, D., Mazumder, R. (Eds.), International Conference on Paleoproterozoic Supercontinents and Global Evolu- 1470 tion, UNESCO-IGCP 509, Abstract Volume: International Association for Gondwana Research Conference Series, 9, pp. 17–18. 1471 Rogers, J.J.W. and Santosh, M., 2002. Configuration of Columbia, a Mesoproterozoic supercontinent. Gondwana Research, 5, 5-22. 1472 Rogers, JJW and Santosh, M 2003. Supercontinents in Earth history. Gondwana Research, 6, 357-368. 1473 Rollinson, H.R., Windley, B.F., Ramakrishnan, M., 1981. Contrasting high and intermediate pressures of metamorphism in the Archaean 1474 Sargur Schists of southern India. Contributions to Mineralogy and Petrology, 76, 420–429. 1475 Roy, A.B. and Jakhar, S.R. (2002). Geology of Rajasthan (NW India) – Precambrian to Recent. Scientific Publishers (India), Jodhpur, 1476 421 pp. 1477 Roy, A.B., Kroner, A., 1996. Single zircon evaporation ages constraining the growth of the Archean Aravalli Craton, northwestern Indi- 1478 an shield. Geological Magazine, 133, 333–342. 1479 Roy, A., Sarkar, A., 2006. Discussion on paper by Misra and Johnson Gondwana Res. 8, 129–142 1480 Roy, A. B., Kröner, A., Bhattacharya, P. K. & Rathore, S., 2005. Metamorphic evolution and zircon geochronology of early Proterozoic 1481 granulites in the Aravalli Mountains of northwestern India. Geological Magazine, 142, 287–302. 1482 Roy, A.B., Kagami, H., Yoshida, M., Roy, A., Bandyopadhyay, B. K., Chattopadhyay, A., 2006. Rb/Sr and Sm/Nd dating of different 1483 metamorphic events from the Sausar mobile belt, central India; implications for Proterozoic crustal evolution. Journal of Asian Earth 1484 Sciences, 26, 61-76. 1485 Roy, A.B., Kroner, A., Rathore, S., Laul, V., Purohit, R., 2012. Tectono-metamorphic and geochronologic studies from Sandmata Com- 1486 plex, northwest Indian shield: Implication on exhumation of late-Paleoproterozoic granulites in and Archaean-early Paleoproterozoic 1487 granite-gneiss terrane, Journal of the Geological Society of India, 79, 323-334. 1488 Saha, A.K. 1994. Crustal evolution of Singhbhum-North Orissa, Eastern India, Geological Survey of India 27, 314 pp. 1489 Saha, A. K., Ray, S. L. & Sarkar, S. N. 1988. Early history of the Earth: evidence from the Eastern Indian Shield. In: Mukhopadhyay, D. 1490 (ed.) Precambrian of the Eastern Indian Shield. Memoirs of the Geological Society of India, 8, 13–37 1491 Santosh, M., Yokoyama, K., Acharya, S.K. 2004. Geochronology of tec- tonic evolution of Karimnagar–Bhopalpatnam granulite belts, 1492 Central India. Gondwana Research 7, 501–518. 1493 Sastry, R.G., Pujari, P.R. & Lal, S., 1999. Basement structure below the Indo-Gangetic alluvial plains (IGP), western U.P., based on 1494 gravity and magnetic data. Gondwana Research Group Memoir, 6, 321-327. 1495 Sarangi, S., Gopalan, K., Kumar, S., 2004. Pb-Pb age of earliest megascopic, eukaryotic alga bearing Rhotas formation, Vindhyan Su- 1496 perGroup, India: implications for Precambrian atmospheric oxygen evolution. Precambrian Research, 121, 107-121. 1497 Sarkar, S.N., Gopalan, K., Trivedi, J.R. 1981. New data on the geochronology of Precambrian of Bhandara–Durg. Central India. Indian 1498 Journal of Earth Sciences 8(2), 131–151. 44

1499 Sarkar, A., Sarkar, G., Paul, D.K., Mitra, N.D. 1990. Precambrian geochronology of central Indian shield: a review. Geological Survey 1500 of India, Special Publication 28, 453–482. 1501 Sarkar, G., Corfu, F., Paul, D.K., Mc Naughton, N.J., Gupta, S.N., Bishui, P.K. 1993. Early Archaean crust in Bastar Craton, Central In- 1502 dia: a geochemical and isotopic study. Precambrian Research 62, 127–137. 1503 Sarkar, A., 1997. Geochronology of Proterozoic Mafic dykes from the Bundelkhand craton, Central India. Abstract— International Con- 1504 ference on Isotopes in Solar System, November 11–14, 98-99. 1505 Sarkar, S.N. and Saha, A.K., 1983. Structure and tectonics of the Singhbhum-Orissa Iron Ore craton, eastern India Structure and Tecton- 1506 ics of the Precambrian Rocks. Recent Researches in Geology, 10, Hindusthan Pub., New Delhi, 1–25. 1507 Sarma, D.S., Fletcher, I.R., Rasmussen, B., McNaughton, N.J., Ram Mohan, M., Groves, D.I., 2008. The timing of gold mineralization 1508 of Hutti gold deposit, Dharwar craton, South India, Economic Geology, 103, 1715-1727. 1509 Saxena, A. and Pandit, M.K., 2012. Geochemistry of Hindoli Group metasediments, SE Aravalli craton, NW India: Implications for 1510 paleoweathering and provenance, J. Geol. Soc India, 79, 267-278. 1511 Sengupta, S., Sarkar, G., Roy, A. K. G., Bhaduri, S. K., Gupta, S. N., & Mandal, A. 2000. Geochemistry and Rb-Sr geochronology of ac- 1512 id tuffs from the northern fringe of the singhbhum craton and their significance in the Precambrian evolution. Indian Minerals, 54, 1513 43-56. 1514 Sharma, K. K. and Rahman A., 2000. The Early Archaean-Paleoproterozoic crustal growth of the Bundelkhand craton, northern Indian 1515 shield; In: M. Deb (Ed.): Crustal evolution and Metallogeny in the northwestern Indian Shield. Narosa Publishing House, New Delhi, 1516 51-72. 1517 Sharma, R. 2010. Cratons and Fold Belts of India, Springer Lecture Notes in Earth Sciences, 127, 304 pp. 1518 Sinha Roy, S., Malhotra, G., Mohanty, M.K., 1998. Geology of Rajasthan, Geological Society of India, Bangalore, 278 pp. 1519 Subba Rao, D.V., Naqvi, S.M., 1999. Archaean Komatiites from the older schist belt of Kalyadi in Western Dharwar Craton, Karnataka. 1520 Journal of the Geological Society of India, 53, 347–354 1521 Srikantia, S.V., Bose, S.S., 1985. Archaean Komatiites from Banasandra area of Kibbanahalli arm of Chitradurga Supracrustal belt in 1522 Karnataka. Journal of the Geological Society of India, 26, 407–417. 1523 Srikantia, S.V.,Venkataramana, P., 1989. The Archaean komatiites of Nagamangala supracrustal belt, Karnataka. Journal of the Geologi- 1524 cal Society of India, 33, 210–214. 1525 Srikantia, S.V., Rao, M.S., 1990. Unusual concentric structure in komatiite of Kibbanahalli Arm of Chitradurga supracrustal belt near 1526 Banasandra, Karnataka. Journal of the Geological Society of India 36, 424–429. 1527 Srivastava, R. K., Singh, R. K., Verma, S. P., 2000. Juxtaposition of India and Antarctica during the Precambrian: Inferences from geo- 1528 chemistry of mafic dykes. Gondwana Research, 3, 227-234. 1529 Srivastava, R.K. and Singh, R.K., 2003. The Paleoproterozoic dolerite dyke swarm of the southern Bastar craton, Central East India: Ev- 1530 idence for the Columbia supercontinent. Memoirs of the Geological Society of India, 52, 163-177. 1531 Srivastava, R. K., Singh, R. K., Verma, S. P., 2004. Neoarchaean mafic volcanic rocks from the southern Bastar greenstone belt, Central 1532 India: petrological and tectonic significance. Precambrian Research, 131, 305-322. 1533 Srivastava, D.C. Shah, J., 2008. Limitations of conjugate fractures and scope of reactivated faults in palaeostress analysis. Journal of the 1534 Geological Society of India, 71, 303-312. 1535 Srivastava, R.K., Gautam, G.C., 2008. Precambrian mafic dyke swarms from the southern Bastar Central India craton: Present and future 1536 perspectives. In: Srivastava, R.K. (ed) Indian Dykes: Geochemistry, Geophysics, and Geochronology. Narosa Publishing House Pvt. 1537 Ltd., New Delhi, India (IND), 367-376. 1538 Stein, H.J., Hannah, J.L., Zimmerman, A., Markey, R.J., Sarkar, S.C. and Pal, A.B. (2004). A 2.5 Ga porphyry Cu-Mo-Au deposit at 1539 Malanjkhand, central India: implications for Late Archaean continental assembly. Precambrian Res., vol. 134, pp. 189–226. 1540 Subba Rao, D.V., Sridhar, D.N., Balaram, V., Nagaraju, K., Gnaneshwara Rao, T., Keshavakrishna, A., Singh, U. P., 2008. Proterozoic 1541 mafic-ultramafic dyke swarms in the vicinity of Chattisgarh-Khariar-Singhora basins in northern Bastar craton, Central India. In: 1542 Srivastava, R.K. (ed) Indian Dykes: Geochemistry, Geophysics, and Geochronology. Narosa Publishing House Pvt. Ltd., New Delhi, 1543 India (IND), 377-396. 1544 Sylvester, P.J., 1994. Archean granite plutons. In: Condie, C. (Ed.), Archaean Crustal Evolution. Elsevier, Amsterdam, 261–314. 1545 Taylor, P.N., Chadwick, B., Moorbath, S., Ramakrishnan, M., Viswanatha, M.N., 1984. Petrography, chemistry and isotopic ages of Pen- 1546 insular Gneiss, Dharwar acid volcanic rocks and the Chitradurga granite with special reference to the late Archaean evolution of the 1547 Karnataka craton. Precambrian Research, 23, 349–375. 1548 Trendall, A.F., de Laeter, J.R., Nelson, D.R., Mukhopadhyay, D., 1997a. A precise U-Pb age for the base of Mulaingiri formation 1549 (Babbudan Group) Dharwar Supergroup of the Karnataka craton, Journal of the Geological Society of India, 50, 161-170. 1550 Trendall, A.F., de Laeter, J.R., Nelson, D.R., Bhaskar Rao, Y.J., 1997b. Further zircon U-Pb age for the Daginkatte formation, Dharwar 1551 Supergroup, Karnataka craton. Journal of the Geological Society of India, 55, 529-540. 1552 Torsvik, T.H., Carter, L., Ashwal, L.D., Bhushan, S.K., Pandit, M.K., Jamtveit, B., 2001a. Rodinia refined or obscured: Paleomagnetism 1553 of the Malani Igneous Suite (NW India). Precambrian Research, 108, 319-333. 1554 Torsvik T.H., Ashwal, L.D., Tucker, R.D., Eide, E.A., 2001b. Neoproterozoic geochronology and palaeogeography of the Seychelles 1555 microcontinent: the India link. Precambrian Res. 110, 47-59. 1556 Turner, C.C., Meert, J.G., Pandit, M.K., Kamenov, G.D., 2013. A detrital zircon transect across the Son Valley sector of the Vindhyan 1557 Basin, India: Implications for basin evolution and paleogeography, Precambrian Research, in press. 1558 van Lente, B; Ashwal, L.D., Pandit, M.K., Bowring, S.A, Torsvik, T.H., 2009. Neoproterozoic hydrothermally-altered basaltic rocks 1559 from Rajasthan, northwest India; implications for late Precambrian tectonic evolution of the Aravalli Craton, Precambrian Research, 1560 170, 202-222. 1561 Venkatadasu, S.P., Ramakrishnan, M., Mahabaleswar, B., 1991. Sargur–Dharwar relationships around the komatiite rich 1562 Jayachamarajapura greenstone belt, Karnataka. Journal of the Geological Society of India, 38, 577–592. 45

1563 Venkateshwarlu, M. and Chalapathi Rao, N.V., 2013. New palaeomagnetic and rock magnetic results on Mesoproterozoic kimberlites 1564 from the Eastern Dharwar craton, southern India: Towards constraining India’s position in Rodinia. Precambrian Research 224, 588- 1565 596. 1566 Wiedenbeck, M., Goswami, J.N., 1994. An ion-probe single zircon 207Pb/206Pb age from the Mewar Gneiss at Jhamarkotra, Rajasthan. 1567 Geochimica et Cosmochimica Acta, 58, 2135–2141. 1568 Wiedenbeck, M., Goswami, J.N., Roy, A.B., 1996a. Stabilisation of the Aravalli craton of the north-western India at 2.5 Ga: an ion- 1569 microprobe zircon study. Chemical Geology, 129, 325–340. 1570 Wiedenbeck, M., Goswami, J.N., Roy, A.B., 1996b. An ion microprobe study of single zircons from the Amet granite, Rajasthan. Jour- 1571 nal of the Geological Society of India, 48, 127–137. 1572 Zachariah, J.K., Hanson, G.N., Rajamani, V., 1995. Postcrystallization disturbance in the neodymium and lead isotope systems of 1573 metabasalts from the Ramigiri schist belt, southern India. Geochimica et Cosmochimica Acta, 59, 3189-3203. 1574 1575 1576 1577 Figure Legends 1578 1579 Figure 1: Generalized tectonic map of Indian subcontinent including Purana Basins, fold belts and cratonic regions. 1580 Fold belts; AFB = Aravalli Fold Belt, DFB = Delhi Fold Belt, EGMB = Eastern Ghat Mobile Belt, SMB=Satpura 1581 Mobile Belt, NSL = Narmada Son lineament, CIS = Central Indian Suture and PCSZ = Palghat-Cauvery Shear 1582 Zone. Purana Basins; VB=Vindhyan Basin; PG=Prahnita-Godavari Basin; Ch.B=Chhattisgarh Basin; 1583 CuB=Cuddapah Basin; KBB=Kaladgi-Bhima Basin; MB=Marwar Basin; IB=Indravati Basin. Other Elements; 1584 EDD=Eastern Dharwar Domain; WDD=Western Dharwar Domain; MR=Mahandi Rift; R=Rajhmahal Traps; 1585 CG=Closepet Granite. SIB=South Indian Block; NIB=North Indian Block. Modified from Meert et al., 2010. 1586 1587 Figure 2: The supercontinents of (a) Columbia (Meert, 2012; Rogers and Santosh, 2002); (b) Rodinia (Li et al., 1588 2008) and (c) Gondwana (Meert and Lieberman, 2008). 1589 1590 Figure 3: Generalized geological map of the Aravalli region after Roy and Jakhar (2002). BGC=Banded Gneiss 1591 Complex. 1592 1593 Figure 4: Map of the Marwar Supergroup sedimentary sequence in Rajasthan (modified from Pareek, 1981). 1594 1595 Figure 5: Stratigraphic sections for the Marwar Supergroup and the Upper and Lower Vindhyan Supergroups. Pre- 1596 viously conjectured relationships between the Marwar Supergroup (Rajasthan) and the Upper Vindhyan Super- 1597 group (Rajasthan and Son Valley Sectors) are indicated by dashed lines. 1598 1599 Figure 6: Generalized Precambrian geological map of the Bundelkhand craton of north-central India including the 1600 extent of Deccan traps in the southeastern area. The Vindhyan basin partially encircles the granitic basement. 1601 Three smaller Paleoproterozoic basins are shown including the Sonrai, Gwalior and Bijawar (after Malviya et al., 1602 2004, 2006) 1603 1604 Figure 7: Singhbhum Craton Map after Iyengar and Murthy (1982), Misra (2006) and Meert et al. (2010). Abbrevi- 1605 ations IOG=Iron Ore Group; SBG-1,2,3= Singhbhum granite, OMG=Older Metamorphic Group; OMTG=Older 1606 Metamorphic Tonalite Gneiss; MG=Mayurbanj granite; NSO=North Singhbhum Orogen; SSZ=Singhbhum Shear 1607 Zone; PLG=Pala Lahara Gneiss. 1608 1609 Figure 8: Summary of the cratonic elements in the Singhbhum craton. Abbreviations used SG=Singhbhum granite; 1610 MG=Mayurbanj granite; (dz)=detrital zircon ages. 1611 1612 Figure 9: Map of the Bastar craton showing the Proterozoic sedimentary basins (Indravati, Khariar, Pakhal, Sukma 1613 and Chhattisgarh along with the major cratonic elements (modified from Meert et al., 2010). 1614 1615 Figure 10: Correlation of sedimentary basins in the Dharwar and Bastar cratons (modified from Conrad et al., 2011) 1616 excluding the Cuddapah basin. 1617 1618 Figure 11: Map of the Eastern Dharwar domain. Regions in polygons represent major dyke swarms within the 1619 EDD. Abbreviations used: H&B=Harohalli and Bangalore dyke swarm; T=Tirupati dyke swarm; A=Anantapur 1620 dyke swarm; M=Mahabubnagar dyke swarm; H=Hyderabad dyke swarm; RPSH=Ramagiri-Penakacherla- 46

1621 Sirigeri-Hundgund Belt; VRG=Velligalu-Raichur-Gadwal Belt; KKJH=Kolar-Kadiri-Jonnaguri-Hutti Belt (mod- 1622 ified from Meert et al., 2010). 1623 1624 Figure 12: Cuddapah basin stratigraphic nomenclature includes (from base to top of section), the Papaghni Group; 1625 Chitravati Group; Nallamalai Group; Kurnool Group (after Ramakrishnan and Vaidyanadhan, 2008). 1626 1627 Figure 13: Sketch map of the western Dharwar craton showing major lithologic boundaries after Naqvi and Rogers 1628 (1987) and Ramakrishnan and Vaidyanadhan (2008). 1629 1630 Peninsular India

68 72 76 80 84 88 92 96 36 N

H 200 0 200 400 32 I Km

M Delhi A S L A A S A MB NIB Y ARAVALLI VB BUNDELKHAND DFB VB VB NSL R 24

B

F

A C I T Z Kolkata NSL SMB ChBChB CIS M 20 R GBPG BASTAR SINGHBHUM Mumbai A IB R Hyderabad EGMB LEGEND A KBB 16 B SIB Himalayas

I CuB: CuB Deccan Traps NA S DHARWAR WDD CG EGMB R Rajmahal Traps EDD O F Purana Basins Chennai GAL 12 E PCSZ Closepet Granite A

Southern BAY Granulites B E N 8 + 1000 km + +

+ N. Australia

+ + + + W. Australia

+ S. Australia + + + Gondwana after Gray et al., 2007 +

+ + + Shield Indian East

+ Shield Antarctic Arabian- Nubian +Shiel d SL + Mad

+ + + + + + + + + + + + Craton Meta- craton Ka lahari

+ Sahara + + + + +

Congo + Craton Australia + + a + +

+ + RP + di SF + In Mad a st c + a ti + + + +

E Siberia + rc + + a West Craton t African + n + + A + Craton +

Amazonian + Baltica + + Kal Laurentia East Gondwana Congo West Gondwana SAM Subduction Zone East African Brasiliano-Damar a Kuungan Orogen Neoproterozoi c Orogen Orogen Mesozoic - Tertiary Palaeozoic Orogen Orogen Palaeozoic-Mesozoic Orogen Precambrian Shield Orogen + + + + + + (c) + 60 N 60 S 60 WAfr 30 N 30 30 S 30 Equator Columbia ~1.5 Ga (a) (b) 74OE 76OE Delhi Jaipur INDIA

Jaipur 27ON

Ajmer Jodhpur P

Sendra

Western Margin Fault Banas Lineament O 26 N

O ‛ 26 00 N Mangalwar Delwara lineament P KaligumanSand Mata Lineament Complex (BGC II)

+ Great Boundary Fault P Bundi

25ON Chittaurgarh P=Phulad Ophiolite P Mt. Abu Udaipur BGC I

Q Alluvium

Deccan Traps K/T Marwar Supergroup PC/C

Malani Igneous Suite Post Delhi Granites Vindhyan Supergroup Delhi Supergroup PROTEROZOIC Aravalli Supergroup

+ 0 50km Older Granites Banded Gneissic (a) Complex (BGC) ARCHAEAN 72° 73° 74° N

INDIA 30° Suratgara

0 88 km 29°

Bikaner 28°

Nagaur Pokaran 27°

Jodhpur 26°

72° 73° 74°

Cenozoic Nagaur Bilara Jodhpur Malani Basement Marwar Supergroup Marwar Supergroup Vindhyan Supergroup

Proposed Correlations

Permo- Bap cgl. U. Bhander s.s Carboniferous Tunklian s.s. Trilobita(?) Ediacara (?) Sirbu sh.

L. Bhander s.s Priapulids Nagaur s.s. Bhander l.s. Bhander Gp.

Ganugarh Sh.

Pondlo dol.

Gotan l.s. Rewa s.s. Supergroup Upper Vindhyan

Dhanapa dol.

Chuaria Girbhakar s.s. Ediacara Jhiri Sh.

Sonia s.s. Kaimur s.s. Kaimur Gp.Kaimur Gp. Rewa

Jhodphur Gp. Gp. Bilara Gp. Nagaur Majhgawan Pokaran congl. Kimberlite 1073 Ma Malani Chuaria Rhotas Gp. Rhyolite Kheinjua Gp. 750-800 Ma (U-Pb) ~1600 Ma (U-Pb) Deonar Gp.

Mizapur Gp. Supergroup Lower Vindhyan Lower ~1900 Ma (U-Pb) Hindoli Gp. 78 E 79 E 80 E 81 E Kanpur

Gwalior basin 0 km 50

26 N Indo-Gangetic Alluvial plains N

Banda Jhansi

Mahoba Mauranipur 25 N Babina Bundelkhand Craton

Panna Lalitpur Bijawar basin

Sonrai basin

Deccan Traps 24 N

Legend Marginal basins Alluvium Mafic dyke Deccan Traps Bundelkhand tectonic zone Vindhyan Supergroup Giant quartz vein 85 30’ E 86 E 86 30’ E

Chhotanagpur

NSO Tamar 23 N

Dhanjori Jamshedpur SSZ

NSO OMG Chaibasa SBG-1 Dalma Manoharpur IOG Jamda IOG Jorapokhar

MG Kolhan

SBG-1 SBG-3 OMG Baripada OMTG 22 N Koira Simlipal

IOG Karanja

SBG-2

OMTG

MG PLG Niligiri SBG-3 IOG

10k m IOG

North GONDWANAS 21 N

Legend Gondwana Singhbhum Singhbhum Mayurbanj OlderM etamorphic N.S inghbhum Supergroup Granite-3 Granite-1 Granites Group Orogen Kolhan Singhbhum Maic Iron-Ore OlderM etamorphic Mafic Supergroup Granite-2 Volcanics Group TonaliteG roup Volcanics U-Pb Pb-Pb Kolhan Group Ages

Chandil Fm.

1.63 Ga

Dalma Volcanics

Dhalbhum Fm.

Chaibasa Fm.

2.86 Ga Dhanjori Fm.& Simlipal Basinal MG 3.10 Ga Sequence

Younger Iron Ore Group

SG

3.51 Ga SG 3.53 Ga Older Metamorphic Group

Older Iron Ore Group PALEOARCHEAN MESOARCHEAN or PALEOPROTEROZOIC NEOARCHEAN

EOARCHEAN (?) 3.5-3.8 Ga (dz) 78° 80° 82° 84°

M N aha 0 100 km n Belt adi 23° obile R ura M ift Satp Singhbhum Craton

Chhattisgarh Basin

Raipur 21° Dongarhgarh Belt

lt Khariar Basin Be ile b o M

ats h G n r

ste S a Proterozoic A Jagdalpur E Basins Indravati Basin 19° Pakhal Basin Bijapur Supracrustal Prahnita-Godavari Deccan Traps Rocks Unclassified Granulite Belt Granite and Dharwar Sukma Basin Gneiss Craton Gondwanas Dongargarh Sausar Granite Belt Dharwar Craton Bastar Craton

Pranhita-Godavari Valley Chattisgarh Basin Indravati Basin Khariar Basin Western Belt Eastern Belt (Eastern Subbasin) Gondwana Supergroup ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~

Sullavai Group Nandeli Sh Kharsiya Group Sullavai Sarnadih SS Group ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ Zircon SHRIMP; Tuff Churtela Sh 1007 ± 7 Ma Tuff 975-1000 Ma Tuff U-Pb Satnala Sh Saradih Lst Jagdalpur Fm Gunderdehi Sh Raipur Sarangarh Group Chanda Lst Kanger Fm * PC-20/2; approx. 1180 Ma Lst. Bijepur Sh

Brown Sh p

ka Kansapathar Fm u

Pranhita SS o Gomarda Fm

r * Chandarpur

G Alba

Penganga Group

Nalla Gutta SS Lohardi Fm Group Cherakur Fm ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ ? ~~~~~~~~ + + ? Rajaram Upper Sandstone Lst + + Singhora ? ? Group ~~~~~~~~ Middle Ramgundam SS + + Mulug Subgroup UPG-1; 1565 ± 6 Ma Tiratgarh Fm Shale * Somanpalli Tuff EPMA; 1500 Ma Group BJ-1; 1620 ± 6 Ma + ~~~~~~~~ Tuff Jakaram Congl * + + 1455 ± 47 Ma ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ + + + + + shale + + + ~~~~~~~~ Lower + BASTAR + +++

PURANA-I PURANA-II -III PURANA + LPG-1 BASTAR + Sandstone Pakhal Group Pakhal Pandikunta * * LPG-3; 1686 ± 6 Ma + ~~~~~~~~ Lst + ++++Lakhna ++++Dykes

Mallampalli Bolapalli Fm ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ Subgroup + 1460 Ma + + + + + + + + + + + BASTAR DHA R WAR 74 76 78 80 82

Godavari Basin 20 Bastar Craton

Deccan Basalts

Bhima 18 Basin H Kaladgi M Basin VRG

ts R a 16 P h lt S G H e rn B te Cuddapah ile Western Dharwar s Sa b A a o Domain Basin E M

VRG T 14 H&B KKJH

N 200 km

Southern Granulites 12

LegendEasternDharwarDomain

Dharwar Batholith Kimberlite/Lamproite

Greenstone Belts Mafic dyke swarms

Closepet Granite

Proterozoic Basins Gulcheru Qtzite Vempalle Formation Papaghni Cuddapah Basin Group Stratigraphy ~2.4 Gaand~2.1 Ga dykes Chitravati Tadpatri Formation Gandikota Qtzite Group 1885 +/-3Ma Qtzite Pullivendla

Pullampet Nagari Qtzite Formation Nallamai Group Quartzite Conglomerate Tuff/Shale Crystalline Rocks Precambrian Nallamai

Bairenkonda Qtzite Cumbum Formation Group Sandstone Arksoic Ma Basalt/ Shale Calcareous Shale f ic Sill Kurnool Dolomite Shale w/ Limestone Dolomite Limestone Flaggy Group Unconformity Baganpalle Koilkuntla Panem Auk Sh. Ma Narji Formation Qtzite Nandyal Qtzite Breccia Chert/Chert Shale/Dolomite Calcareous f ic Dykes WesternDharwarDomain

75 76 X 100 km X Deccan Traps X X X X X X X X X X X 16 X X

Eastern Dharwar Domain Cenozoic

X X Deccan Traps

14 Proterozoic Basin

Closepet & Related Granites

Dharwar Supergroup - Greenschist-Facies

Sargur Schist Belts

12 Peninsular TTG Gneisses

3,000-m.y. Intrusive Rocks

SouthernGranulite Gneiss-Granulite Transition Zone Terrane