Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-19T00:40:32.188Z Has data issue: false hasContentIssue false

Petrological implications of element redistribution during metamorphism: insights from meta-granite of the South Delhi Fold Belt, Rajasthan, India

Published online by Cambridge University Press:  23 February 2022

Upama Dutta
Affiliation:
Department of Applied Geology, Indian Institute of Technology (ISM) Dhanbad, Dhanbad-826004, India
Ayan Kumar Sarkar
Affiliation:
Department of Geological Sciences, Jadavpur University, Kolkata-700032, India
Sadhana M. Chatterjee*
Affiliation:
Department of Geological Sciences, Jadavpur University, Kolkata-700032, India
Anirban Manna
Affiliation:
Department of Geological Sciences, Jadavpur University, Kolkata-700032, India
Alip Roy
Affiliation:
Department of Geological Sciences, Jadavpur University, Kolkata-700032, India
Subhrajyoti Das
Affiliation:
Department of Geological Sciences, Jadavpur University, Kolkata-700032, India Present address: J. K. College, Purulia-723101, West Bengal, India
*
Author for correspondence: Sadhana M. Chatterjee, Email: smcjugeo@gmail.com

Abstract

Meta-granites of the South Delhi Fold Belt, northwestern India, contain spectacular reaction textures formed during the metamorphic replacement of primary minerals. Textural relationships imply that amphibole was replaced sequentially in two stages. Epidote + titanite + quartz symplectite formed syn-tectonically on amphibole grain boundaries/fractures, followed by post-deformational growth of euhedral garnet overprinting amphibole grains. Besides occurring as symplectite grown during deformation, titanite in this rock also developed as a post-tectonic corona around magnetite. Parent magnetite contains exsolutions of ilmenite and/or ultrafine lamellae of Ti-rich oxide (Ti-oxd). Textures involving coronal titanite suggest their formation through a magnetite + ilmenite(/Ti-oxd) + plagioclase → titanite reaction. Compositional attributes and the calculation of the gain versus loss of components during the reaction suggest that the Mn2+ for garnet (XSpss = 0.23–0.29) that grew replacing amphibole was supplied by ilmenite (Mn2+ is 0.118–0.128 apfu) as it disintegrated to form coronal titanite. The redistribution of components between the metamorphic reaction sites connects the texturally unrelated domains and suggests that these zones were in chemical equilibrium during metamorphism. We estimated the PT conditions of metamorphism for these post-tectonic assemblages as ∼650–700 °C from pseudosection modelling and conventional thermometry. Zircon data from this study suggest that the granitic rock crystallized at 988.8 ± 8.8 Ma. We propose that the metamorphic phases replaced the primary minerals during the mid Neoproterozoic tectonic activity reported from this terrane. The syn-tectonic symplectitic assemblage formed as the temperature increased during prograde metamorphism, and the post-tectonic minerals developed at peak conditions following the cessation of deformation.

Type
Original Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adak, V and Dutta, U (2021) Genesis of coronae and implications of an early Neoproterozoic thermal event: a case study from SE Chotanagpur Granite Gneissic Complex, India. Geological Magazine 158, 199218. doi: 10.1017/S0016756820000357.CrossRefGoogle Scholar
Amcoff, Ö and Figueiredo, BR (1990) Mechanisms of retrograde changes in oxide minerals from the Proterozoic Serrote da Laje deposit, northeastern Brazil. Mineralium Deposita 25, 313–22.CrossRefGoogle Scholar
Austrheim, H (1987) Eclogitization of lower crustal granulites by fluid migration through shear zones. Earth and Planetary Science Letters 81, 221–32.CrossRefGoogle Scholar
Banerjee, M, Dutta, U, Anand, R and Atlas, ZD (2019) Insights on the process of two-stage coronae formation at olivine-plagioclase contact in mafic dyke from Palghat Cauvery Shear Zone, southern India. Mineralogy and Petrology 113, 625–49.CrossRefGoogle Scholar
Barnicoat, AC and Cartwright, I (1995) Focused fluid flow during subduction oxygen isotope data from high-pressure ophiolites of the Western Alps. Earth and Planetary Science Letters 132, 5361.CrossRefGoogle Scholar
Bebout, G and Barton, M (1993) Metasomatism during subduction: products and possible paths in the Catalina Schist, California. Chemical Geology 108, 6192. doi: 10.1016/0009-2541(93)90318-D.CrossRefGoogle Scholar
Bhowmik, SK, Bernhardt, HJ and Dasgupta, S (2010) Grenvillian age high-pressure upper amphibolite-granulite metamorphism in the Aravalli-Delhi Mobile Belt, Northwestern India: new evidence from monazite chemical age and its implication. Precambrian Research 178, 168–84.CrossRefGoogle Scholar
Brodie, KH and Rutter, EH (1987) The role of transiently fine-grained reaction products in syntectonic metamorphism: natural and experimental examples. Canadian Journal of Earth Sciences 24, 556–64.CrossRefGoogle Scholar
Brodie, KH (1995) The development of orientated symplectites during deformation. Journal of Metamorphic Geology 13, 499508.CrossRefGoogle Scholar
Buddington, AF and Lindsley, DH (1964) Iron-titanium oxide minerals and synthetic equivalents. Journal of Petrology 5, 310–57.CrossRefGoogle Scholar
Buick, IS, Clark, C, Rubatto, D, Hermann, J, Pandit, M and Hand, M (2010) Constraints on the Proterozoic evolution of the Aravalli–Delhi Orogenic belt (NW India) from monazite geochronology and mineral trace element geochemistry. Lithos 120, 511–28.CrossRefGoogle Scholar
Carmichael, DM (1969) On the mechanism of prograde metamorphic reactions in quartz-bearing pelitic rocks. Contributions to Mineralogy and Petrology 20, 244–67.CrossRefGoogle Scholar
Cartwright, I and Buick, IS (2000) Stable isotope constraints on the mechanism of fluid flow during contact metamorphism around the Marulan Batholith, NSW, Australia. Journal of Geochemical Exploration 69, 291–5. doi: 10.1016/S0375-6742(00)00027-3.Google Scholar
Centrella, S (2019) The granulite- to eclogite- and amphibolite-facies transition: a volume and mass transfer study in the Lindås Nappe, Bergen Arcs, west Norway. In Metamorphic Geology: Microscale to Mountain Belts (eds Ferrero, S, Lanari, P, Goncalves, P and Grosch, EG), pp. 241–63. Geological Society of London, Special Publication no. 478.Google Scholar
Centrella, S, Austrheim, H and Putnis, A (2015) Coupled mass transfer through a fluid phase and volume preservation during the hydration of granulite: an example from the Bergen Arcs, Norway. Lithos 236–237, 245–55. doi: 10.1016/j.lithos.2015.09.010.CrossRefGoogle Scholar
Centrella, S, Austrheim, H and Putnis, A (2016) Mass transfer and trace element redistribution during hydration of granulites in the Bergen Arcs, Norway. Lithos 262, 110. doi: 10.1016/j.lithos.2016.06.019.CrossRefGoogle Scholar
Centrella, S, Beaudoin, NE, Derluyn, H, Motte, G, Hoareau, G, Lanari, P, Piccoli, F, Pecheyran, C and Callot, JP (2020) Micro-scale chemical and physical patterns in an interface of hydrothermal dolomitization reveals the governing transport mechanisms in nature: case of the Layens anticline, Pyrenees, France. Journal of the International Association of Sedimentologists 68, 834–54.Google Scholar
Centrella, S, Putnis, A, Lanari, P and Austrheim, H (2018) Textural and chemical evolution of pyroxene during hydration and deformation: a consequence of retrograde metamorphism Lithos 296–299, 245–64.CrossRefGoogle Scholar
Chatterjee, SM, Roy Choudhury, M and Das, S (2016) Reaction enhanced channelized fluid-flux along mid crustal shear zone: an example from Mesoproterozoic Phulad Shear Zone, Rajasthan, India. Journal of Earth System Science 125, 1321–8.CrossRefGoogle Scholar
Chatterjee, SM, Roy Choudhury, M, Das, S and Roy, A (2017) Significance and dynamics of the Neoproterozoic (810 Ma) Phulad Shear Zone, Rajasthan, NW India. Tectonics 36, 1432–54. doi: 10.1002/2017TC004554.CrossRefGoogle Scholar
Chatterjee, SM, Sarkar, AK, Roy, A and Manna, A (2020) Mid-Neoproterozoic tectonics of northwestern India: evidence of stitching pluton along 810 Ma Phulad Shear Zone. Tectonics 39, e2019TC005902. doi: 10.1029/2019TC005902.CrossRefGoogle Scholar
Chowdhury, P, Talukdar, M, Sengupta, P, Sanyal, S and Mukhopadhyay, D (2013) Controls of PT path and element mobility on the formation of corundum pseudomorphs in Paleoproterozoic high-pressure anorthosite from Sittampundi, Tamil Nadu, India. American Mineralogist 98, 1725–37.CrossRefGoogle Scholar
Connolly, JAD (2005) Computation of phase equilibria by linear programming: a tool for geodynamic modeling and its application to subduction zone decarbonation. Earth and Planetary Science Letters 236, 524–41 (updated in 2018).CrossRefGoogle Scholar
Dalziel, IWD (1991) Pacific margins of Laurentia and East Antarctica–Australia as conjugate rift pair: evidence and implications for an Eocambrian supercontinent. Geology 19, 598601.2.3.CO;2>CrossRefGoogle Scholar
de Meer, S, Spiers, CJ and Nakashima, S (2005) Structure and diffusive properties of fluid-filled grain boundaries: an in-situ study using infrared (micro) spectroscopy. Earth and Planetary Science Letters 232, 403–14. doi: 10.1016/j.epsl.2004.12.030.CrossRefGoogle Scholar
Deb, M, Thorpe, RI, Kristic, D, Corfu, F and Davis, DW (2001) Zircon U–Pb and galena Pb isotope evidence for an approximate 1.0 Ga terrane constituting the western margin of the Aravalli-Delhi orogenic belt, northwestern India. Precambrian Research 108, 195213.CrossRefGoogle Scholar
Deer, WA, Howie, RA and Zussman, J (2013) An Introduction to the Rock Forming Minerals. 3rd Ed. London: The Mineralogical Society, pp. 404–5.CrossRefGoogle Scholar
Dharma Rao, CV, Santosh, M, Kim, SW and Li, S (2013) Arc magmatism in the Delhi Fold Belt: SHRIMP U–Pb zircon ages of granitoids and implications for Neoproterozoic convergent margin tectonics in NW India. Journal of Asian Earth Science 78, 8399.CrossRefGoogle Scholar
Droop, GTR (1987) A general equation for estimating Fe3+ concentrations in ferromagnesian silicates and oxides from microprobe analyses, using stoichiometric criteria. Mineralogical Magazine 51, 431–5.CrossRefGoogle Scholar
Fusseis, F, Regenauer-Lieb, K, Liu, J, Hough, RM and De Carlo, F (2009) Creep cavitation can establish a dynamic granular fluid pump in ductile shear zones. Nature 459, 974–7, doi: 10.1038/nature08051.CrossRefGoogle ScholarPubMed
Ghosh, SK, Hazra, S and Sengupta, S (1999) Planar, non-planar and refolded sheath folds in the Phulad Shear Zone, Rajasthan. Journal of Geological Society of India 21, 1715–29.Google Scholar
Golani, PR, Reddy, AB and Bhattacharjee, J (1998) The Phulad Shear Zone in Central Rajasthan and its tectonostratigraphic implications. In The Indian Precambrian (ed. Paliwal, BS), pp. 272278. Jodhpur: Scientific Publishers (India).Google Scholar
Green, ECR, White, RW, Diener, JFA, Powell, R, Holland, TJB and Palin, RM (2016) Activity–composition relations for the calculation of partial melting equilibria in metabasic rocks. Journal of Metamorphic Geology 34, 845–69. doi: 10.1111/jmg.12211.CrossRefGoogle Scholar
Gregory, LC, Meert, JG, Bingen, B, Pandit, MK and Torsvik, TH (2009) Paleomagnetism and geochronology of the Malani igneous suite, Northwest India: implications for the configuration of Rodinia and the assembly of Gondwana. Precambrian Research 170, 1326.CrossRefGoogle Scholar
Gupta, SN, Arora, YK, Mathur, RK, Iqbaluddin, BP, Prasad, B, Sahai, TN and Sharma, SB (1980) Lithostratigraphic Map of Aravalli Region, Southern Rajasthan and North-eastern Gujrat. Hyderabad: Geological Survey of India.Google Scholar
Gupta, SN, Arora, YK, Mathur, RK, Iqbaluddin, BP, Prasad, B, Sahai, TN and Sharma, SB (1997) The Precambrian Geology of the Aravalli Region, Southern Rajasthan and Northeastern Gujarat. Geological Survey of India Memoir 123, 262 pp.Google Scholar
Gupta, P, Mukhopadhyay, K, Fareeduddin, RM and Reddy, MS (1995) Stratigraphy and structure of Delhi Supergroup of rocks in central part of Aravalli Range. Records of the Geological Survey of India 120, 1226.Google Scholar
Hawthorne, FC, Oberti, R, Harlow, GE, Maresch, WV, Martin, RF, Schumacher, JC and Welch, M (2012) IMA Report – Nomenclature of the amphibole supergroup. American Mineralogist 97, 2031–48.CrossRefGoogle Scholar
Heron, AM (1953) Geology of Central Rajputana. Geological Survey of India Memoir 79, 339 pp.Google Scholar
Hoffman, PF (1991) Did the breakout of Laurentia turn Gondwanaland inside out? Science 252, 1409–12.CrossRefGoogle ScholarPubMed
Holland, T and Blundy, J (1994) Non-ideal interactions in calcic amphiboles and their bearing on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology 116, 433–47.CrossRefGoogle Scholar
Holland, T and Powell, R (2003) Activity-composition relations for phases in petrological calculations: an asymmetric multicomponent formulation. Contributions to Mineralogy and Petrology 145, 492501. doi: 10.1007/s00410-003-0464-z CrossRefGoogle Scholar
Holland, TJB and Powell, R (2011) An improved and extended internally consistent thermodynamic dataset for phases of petrological interest, involving a new equation of state for solids. Journal of Metamorphic Geology 29, 333–83. doi: 10.1111/j.1525-1314.2010.00923.x.CrossRefGoogle Scholar
Hu, R, Li, F, Yu, H and Yang, J (2019) Application of ImageJ in the rock thin section image analysis: the separation and quantitative calculation of crystal-glass two phases. Journal of Earth Sciences and Environmental Studies 4, 609–16.Google Scholar
Hughes, JM, Bloodaxe, ES, Hanchar, JM and Foord, EE (1997) Incorporation of rare earth elements in titanite: stabilization of the A2/a dimorph by creation of antiphase boundaries. American Mineralogist 82, 512–6.CrossRefGoogle Scholar
Jamtveit, B, Austrheim, H and Malthe-Sørensen, A (2000) Accelerated hydration of the Earth’s deep crust induced by stress perturbations. Nature 408, 75–9.CrossRefGoogle ScholarPubMed
Konrad-Schmolke, M, Zack, T, O’Brien, PJ and Barth, M (2011) Fluid migration above a subducted slab — thermodynamic and trace element modelling of fluid–rock interaction in partially overprinted eclogite-facies rocks (Sesia Zone, Western Alps). Earth and Planetary Science Letters 311, 287–98. doi: 10.1016/j.epsl.2011.09.025.CrossRefGoogle Scholar
Kröner, A and Cordani, UG (2003) African, southern Indian and South American cratons were not part of the Rodinia supercontinent: evidence from field relationships and geochronology. Tectonophysics 375, 325–52.CrossRefGoogle Scholar
Lang, HM and Gilotti, JA (2001) Plagioclase replacement textures in partially eclogitised gabbros from the Sanddal mafic-ultramafic complex, Greenland Caledonides. Journal of Metamorphic Geology 19, 497517.CrossRefGoogle Scholar
Lang, HM and Rice, JM (1985) Regression modelling of metamorphic reactions in metapelites, Snow Peak, northern Idaho. Journal of Petrology 26, 857–87.CrossRefGoogle Scholar
Lang, HM, Watcher, AJ, Peterson, VL and Ryan, JG (2004) Coexisting clinopyroxene/spinel and amphibole/spinel symplectites in metatroctolite from the Buck Creek ultramafic body, North Carolina Blue Ridge. American Mineralogist 89, 2030.CrossRefGoogle Scholar
Larikova, TL and Zaraisky, GP (2009) Experimental modelling of corona textures. Journal of Metamorphic Geology 27, 139–51.CrossRefGoogle Scholar
Li, ZX, Bogdanova, SV, Collins, AS, Davidson, A, De Waele, B, Ernst, RE, Fitzsimons, ICW, Fuck, RA, Gladkochub, DP, Jacobs, J, Karlstrom, KE, Lu, S, Natapov, LM, Pease, V, Pisarevsky, SA, Thrane, K and Vernikovsky, V (2008) Assembly, configuration, and break-up history of Rodinia: a synthesis. Precambrian Research 160, 179210.CrossRefGoogle Scholar
Li, ZX, Zhang, L and Powell, CM (1996) Positions of the East Asian cratons in the Neoproterozoic supercontinent Rodinia. Australian Journal of Earth Science 43, 593604.CrossRefGoogle Scholar
Ludwig, K (2008) User’s Manual for Isoplot 3.60: A Geochronological Toolkit for Microsoft Excel. Berkeley Geochronology Center, Special Publication no. 4, 77 pp.Google Scholar
Malone, SJ, Meert, JG, Banerjee, DM, Pandit, MK, Tamrat, E, Kamenova, GD, Pradhana, VR and Sohl, LE (2008) Paleomagnetism and detrital zircon geochronology of the Upper Vindhyan sequence, Son Valley and Rajasthan, India: a ca. 1000 Ma closure age for the Purana Basins? Precambrian Research 164, 137–59.CrossRefGoogle Scholar
Mengel, F and Rivers, T (1991) Decompression reactions and PT conditions in high-grade rocks, Northern Labrador: PTt paths from individual samples and implications for Early Proterozoic tectonic evolution. Journal of Petrology 32, 139167. doi: 10.1093/petrology/32.1.139.CrossRefGoogle Scholar
Mezger, K and Cosca, MA (1999) The thermal history of the Eastern Ghats Belt (India) as revealed by U–Pb and 40Ar/39Ar dating of metamorphic and magmatic minerals: implications for the SWEAT correlation. Precambrian Research 94, 251–71.CrossRefGoogle Scholar
Milke, R, Neusser, G, Kolzer, K and Wunder, B (2013) Very little water is necessary to make a dry solid silicate system wet. Geology 41, 247–50.CrossRefGoogle Scholar
Moore, J, Beinlich, A, Austrheim, H and Putnis, A (2019) Stress orientation-dependent reactions during metamorphism. Geology 47, 151–4.CrossRefGoogle Scholar
Moores, EM (1991) Southwest U.S.–East Antarctic (SWEAT) connection: a hypothesis. Geology 19, 425–8.2.3.CO;2>CrossRefGoogle Scholar
Mukherjee, PK, Singhal, S, Adlakha, V, Rai, SK, Dutt, S, Kharya, A and Gupta, AK (2017) In situ U–Pb zircon micro-geochronology of MCT zone rocks in the Lesser Himalaya using LA-MC-ICPMS technique. Current Science 112, 802–10. doi: 10. 18520/cs/v112/i04/802-810.CrossRefGoogle Scholar
Nasipuri, P, Bhattacharya, A and Das, S (2009) Metamorphic reactions in dry and aluminous granulites: a Perple_X PT pseudosection analysis of the influence of effective reaction volume. Contributions to Mineralogy and Petrology 157, 301–11.CrossRefGoogle Scholar
Oliver, NHS and Bons, PD (2001) Mechanisms of fluid flow and fluid-rock interaction in fossil metamorphic-hydrothermal systems inferred from vein-wallrock patterns, geometry, and microstructure. Geofluids 1, 137–63.CrossRefGoogle Scholar
Onge, MR and Ijewliw, OJ (1996) Mineral corona formation during high-P retrogression of granulitic rocks, Ungava Orogen, Canada. Journal of Petrology 37, 553–82. doi: 10.1093/petrology/37.3.553.CrossRefGoogle Scholar
Pandit, MK, Carter, LM, Ashwal, LD, Tucker, RD, Torsvik, TH, Jamtveit, B and Bhushan, SK (2003) Age, petrogenesis and significance of 1 Ga granitoids and related rocks from the Sendra area, Aravalli Craton, NW India. Journal of Asian Earth Science 22, 363–81.CrossRefGoogle Scholar
Paton, C, Hellstrom, J, Paul, B, Woodhead, J and Hergt, J (2011) Iolite: freeware for the visualisation and processing of mass spectrometric data. Journal of Analytical Atomic Spectrometry 26, 2508–18. doi: 10.1039/c1ja10172b.CrossRefGoogle Scholar
Pisarevsky, SA, Wingate, MTD, Powell, CM, Johnson, S and Evans, DAD (2003) Models of Rodinia assembly and fragmentation. In Proterozoic East Gondwana: Supercontinent Assembly and Breakup (eds Yoshida, M, Windley, BF and Dasgupta, S), pp. 3555. Geological Society of London, Special Publication no. 206.Google Scholar
Powell, CM and Pisarevsky, SA (2002) Late Neoproterozoic assembly of East Gondwana. Geology 30, 36.2.0.CO;2>CrossRefGoogle Scholar
Pownceby, MI, Wall, VJ and O’Neill, HStC (1991) An experimental study of the effect of Ca upon garnet-ilmenite Fe-Mn exchange equilibria. American Mineralogist 76, 1580–8.Google Scholar
Price, GD (1980) Exsolution microstructures in titanomagnetite and their magnetic significance. Physics of Earth and Planetary Interiors 23, 212.CrossRefGoogle Scholar
Price, GD and Putnis, A (1979) Oxidation phenomena in pleonaste bearing titanomagnetites. Contributions to Mineralogy and Petrology 69, 355–9.CrossRefGoogle Scholar
Putnis, A (2002) Mineral replacement reactions: from macroscopic observations to microscopic mechanisms. Mineralogical Magazine 66, 689708.CrossRefGoogle Scholar
Putnis, A (2009) Mineral replacement reactions. Reviews in Mineralogy and Geochemistry 70, 87124.CrossRefGoogle Scholar
Putnis, A and Austrheim, H (2010) Fluid-induced processes: metasomatism and metamorphism. Geofluids 10, 254–69, doi: 10.1111/j.1468-8123.2010.00285.x.Google Scholar
Radhakrishna, T and Mathew, J (1996) Late Precambrian (850–800 Ma) palaeomagnetic pole for the south Indian shield from the Harohalli alkaline dykes: geotectonic implications for Gondwana reconstructions. Precambrian Research 80, 7787.CrossRefGoogle Scholar
Ravna, EK (2000) Distribution of Fe2+ and Mg between coexisting garnet and hornblende in synthetic and natural systems: an empirical calibration of the garnet–hornblende Fe–Mg geothermometer. Lithos 53, 265–77.CrossRefGoogle Scholar
Roy Choudhury, M, Das, S, Chatterjee, SM and Sengupta, S (2016) Deformation of footwall rock of Phulad Shear Zone, Rajasthan: evidence of transpressional shear zone. Journal of Earth System Science 125, 1033–40.CrossRefGoogle Scholar
Roy, AB and Jakhar, SR (2002) Geology of Rajasthan (Northwest India) Precambrian to Recent. Jodhpur: Scientific Publishers (India), 421 pp.Google Scholar
Rubie, DC (1986) The catalysis of mineral reactions by water and restrictions on the presence of aqueous fluid during metamorphism. Mineralogical Magazine 50, 399415.CrossRefGoogle Scholar
Rutter, EH and Brodie, KH (1995) Mechanistic interactions between deformation and metamorphism. Geological Journal 30, 227–40.CrossRefGoogle Scholar
Sengupta, P, Bhui, UK, Braun, I, Dutta, U and Mukhopadhyay, D (2009) Chemical substitutions, paragenetic relations and physical conditions of högbomite in Sittampundi layered anorthosite complex, south India. American Mineralogist 94, 1520–34.Google Scholar
Sengupta, S and Chatterjee, SM (2016) Microstructural variation in quartzofeldspathic mylonites and vorticity analysis using rotating porphyroclasts in the Phulad Shear Zone, Rajasthan, India. In Ductile Shear Zones: From Micro- to Macro-Scales (eds Mukherjee, S and Mulchrone, K), pp. 128–40. London: Wiley Blackwell.Google Scholar
Simpson, C and Wintsch, RP (1989) Evidence for deformation-induced K-feldspar replacement by myrmekite. Journal of Metamorphic Geology 7, 261–75.CrossRefGoogle Scholar
Singh, YK, Waele, BD, Karmakar, S, Sarkar, S and Biswal, TK (2010) Tectonic setting of the Balaram-Kui-Surpagla-Kengora granulites of the South Delhi Terrane of the Aravalli Mobile Belt, NW India and its implication on correlation with the East African Orogen in the Gondwana assembly. Precambrian Research 183, 669–88.CrossRefGoogle Scholar
Sinha-Roy, S (1988) Proterozoic Wilson Cycles in Rajasthan. In Precambrian of the Aravalli Mountain, Rajasthan, India (ed. Roy, AB), pp. 95107. Geological Survey of India Memoir 7.Google Scholar
Sláma, J, Kosler, J, Condon, DJ, Crowley, JL, Gerdes, A, Hanchar, JM, Horstwood, MSA, Morris, GA, Nasdala, L, Norberg, N, Schaltegger, U, Schoene, B, Tubrett, MN and Whitehouse, MJ (2008) Plešovice zircon—a new natural reference material for U–Pb and Hf isotopic micro-analysis. Chemical Geology 249, 135. doi: 10.1016/j.chemgeo.2007.11. 005.CrossRefGoogle Scholar
Spruzeniece, L, Piazolo, S, Daczko, NR, Kilburn, MR and Putnis, A (2016) Symplectite formation in the presence of a reactive fluid: insights from hydrothermal experiments. Journal of Metamorphic Geology 35, 281–99.CrossRefGoogle Scholar
Svahnberg, H and Piazolo, S (2013) Interaction of chemical and physical processes during deformation at fluid-present conditions: a case study from an anorthosite–leucogabbro deformed at amphibolite facies conditions. Contributions to Mineralogy and Petrology 165, 543–62.CrossRefGoogle Scholar
Tan, P, Breivik, AJ, Trønnes, RG, Mjelde, R and Azuma, R (2016) Crustal structure and origin of the Eggvin Bank west of Jan Mayen, NE Atlantic. Journal of Geophysical Research: Solid Earth 122, 4362. doi: 10.1002/2016JB013495.CrossRefGoogle Scholar
Tiwari, SK and Biswal, TK (2019) Dynamics, EPMA Th–U total Pb monazite geochronology and tectonic implications of deformational fabric in the lower-middle crustal rocks: a case study of Ambaji granulite, NW India. Tectonics 38, 2232–54.CrossRefGoogle Scholar
Tobisch, OT, Collerson, KD, Bhattacharya, TU and Mukhopadhyay, D (1994) Structural relationship and Sm–Nd isotope systematics of polymetamorphic granitic gneisses and granitic rocks from central Rajasthan, India – implications for the evolution of the Aravalli craton. Precambrian Research 65, 319–39.CrossRefGoogle Scholar
Torres-Roldan, RL, Garcia-Casco, A and García-Sánchez, PA (2000) CSpace: an integrated workplace for the graphical and algebraic analysis of phase assemblages on 32-bit Wintel platforms. Computer Geoscience 26, 779–93. doi: 10.1016/S0098-3004(00)00006-6.CrossRefGoogle Scholar
Torsvik, TH, Ashwal, LD, Tucker, RD and Eide, EA (2001a) Neoproterozoic geochronology and palaeogeography of the Seychelles microcontinent: the India link. Precambrian Research 110, 4759.CrossRefGoogle Scholar
Torsvik, TH, Carter, LM, Ashwal, LD, Bhushan, SK, Pandit, MK and Jamtveit, B (2001b) Rodinia refined or obscured: palaeomagnetism of the Malani igneous suite (NW India). Precambrian Research 108, 319–33.CrossRefGoogle Scholar
Torsvik, TH, Smethurst, MA, Meert, JG, Van der Voo, R, McKerrow, WS, Brasier, MD, Sturt, BA and Walderhaug, HJ (1996) Continental break-up and collision in the Neoproterozoic and Palaeozoic – a tale of Baltica and Laurentia. Earth-Science Reviews 40, 229–58.CrossRefGoogle Scholar
Turnock, AC and Eugster, HP (1962) Fe–Al oxides: phase relationships below 1,000°C. Journal of Petrology 3, 533–65. doi: 10.1093/petrology/3.3.533.CrossRefGoogle Scholar
Vernon, RH (1991) Questions about myrmekite in deformed rocks. Journal of Structural Geology 13, 979–85. doi: 10.1016/0191-8141(91)90050-S.CrossRefGoogle Scholar
Vincent, EA, Wright, JB, Chevallier, R and Mathieu, S (1957) Heating experiments on some natural titaniferous magnetites. Mineralogical Magazine 31, 624–55.CrossRefGoogle Scholar
Wayte, GJ, Worden, RH, Rubie, RC and Droop, GTR (1989) A TEM study of disequilibrium plagioclase breakdown at high pressure: the role of the infiltrating fluid. Contributions to Mineralogy and Petrology 101, 426–37.CrossRefGoogle Scholar
Weil, AB, Van der Voo, R, Mac Niocaill, C and Meert, JG (1998) The Proterozoic supercontinent Rodinia: paleomagnetically derived reconstructions for 1100 to 800 Ma. Earth and Planetary Science Letters 154, 1324.CrossRefGoogle Scholar
White, RW, Powell, R and Clarke, GL (2002) The interpretation of reaction textures in Fe-rich metapelitic granulites of the Musgrave Block, central Australia: constraints from mineral equilibria calculations in the system K2O–FeO–MgO–Al2O3–SiO2– H2O–TiO2–Fe2O3 . Journal of Metamorphic Geology 20, 4155.CrossRefGoogle Scholar
White, RW, Powell, R, Holland, TJB, Johnson, TE and Green, ECR (2014a) New mineral activity–composition relations for thermodynamic calculations in metapelitic systems. Journal of Metamorphic Geology 32, 261–86.CrossRefGoogle Scholar
White, RW, Powell, R, Holland, TJB and Worley, BA (2000) The effect of TiO2 and Fe2O3 on metapelitic assemblages at greenschist and amphibolite facies conditions: mineral equilibria calculations in the system K2O–FeO–MgO–Al2O3–SiO2–H2O–TiO2–Fe2O3 . Journal of Metamorphic Geology 18, 497511.CrossRefGoogle Scholar
White, RW, Powell, R and Johnson, TE (2014b) The effect of Mn on mineral stability in metapelites revisited: new ax relations for manganese-bearing minerals. Journal of Metamorphic Geology 32, 261–86. doi: 10.1111/jmg.12095.CrossRefGoogle Scholar
Whitney, DL and Evans, BW (2010) Abbreviations for names of rock-forming minerals. American Mineralogist 95, 185–7.CrossRefGoogle Scholar
Wiedenbeck, M, Allé, P, Corfu, F, Griffin, WL, Meier, M, Oberli, F, Von Quadt, A, Roddick, JC and Spiegel, W (1995) Three natural zircon standards for U–Th–Pb, Lu–Hf, trace element and REE analyses. Geostandards Newsletter 19, 123.CrossRefGoogle Scholar
Zhao, JH, Pandit, MK, Wang, W and Xia, XP (2018) Neoproterozoic tectonothermal evolution of NW India: evidence from geochemistry and geochronology of granitoids. Lithos 316–317, 330–46.CrossRefGoogle Scholar