Hostname: page-component-7bb8b95d7b-s9k8s Total loading time: 0 Render date: 2024-09-30T01:33:03.802Z Has data issue: false hasContentIssue false

Geochemical constraints on the petrogenesis of mafic rocks (metadolerites) from the Proterozoic Shillong Basin, Northeast India: implications for growth of the Greater Indian Landmass

Published online by Cambridge University Press:  27 April 2023

Pallabi Basumatary
Affiliation:
Department of Geology, Cotton University, Guwahati, Assam, India
Ashima Saikia
Affiliation:
Department of Geology, University of Delhi, Delhi, India
Tribujjal Prakash
Affiliation:
Department of Geology, Cotton University, Guwahati, Assam, India
Bibhuti Gogoi*
Affiliation:
Department of Geology, Cotton University, Guwahati, Assam, India
*
Corresponding author: Bibhuti Gogoi, Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

The Paleo-Mesoproterozoic Shillong Basin of the Assam-Meghalaya Gneissic Complex is exposed in parts of Northeast India. The studied metadolerites are from the volcano-sedimentary sequence of Shillong Basin from the Borjuri area in the Mikir Massif. This episode of mafic magmatism can be correlated with the Columbia supercontinent formation and bears significance to its reconstruction. The present work discusses the field, petrography and geochemical characteristics of the metadolerites, which occur in close association with the quartzites of the Shillong Group of rocks (metasedimentary rocks of the Shillong Basin). Our data show distinctive characteristics of subduction-related magmatism exhibiting high LREE/HREE, large ion lithophile element/high field strength element ratios and pronounced negative Nb anomaly. Elemental ratios such as Zr/Ba (0.21–0.46), La/Nb (1.23–2.32) and Ba/Nb (30.08–56.90) point to a fluid-enriched lithospheric mantle source in a subduction regime. Metadolerites plot in the field of ‘back-arc basin basalts’ in tectonic discrimination diagrams reinforcing a subduction zone tectonic setting. The mafic rocks correspond to a 6–10 % partial melting of a mantle source incorporating spinel+garnet lherzolite. The metamorphic P-T of the metadolerites estimated from plagioclase-hornblende geothermobarometer (7–8 kbar, 664 °C) is indicative of amphibolite facies metamorphism in a medium P-T zone. Based on the comparative analysis of field observation, petrography, geochemistry and geological ages given by previous workers, we infer that the Shillong Basin represents a back-arc rift region and is the eastern continuation of the Bathani volcano-sedimentary sequence of the Chotanagpur Granite Gneiss Complex marking continuation of the Central Indian Tectonic Zone to the Mikir Massif.

Type
Original Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1. Introduction

The genesis of mafic magma in both interplate and intraplate areas is a key concern in igneous petrology. Mafic magmas are generated as a result of (i) rise of local geothermal gradient by physical convection, (ii) subsequent lowering of solidus by influx of volatile constituents or by (iii) localized spike in geothermal gradient (frictional heating) due to tectonic conditions (Wyllie, Reference Wyllie1988). Mafic magmas also undergo fractional crystallization and crustal contamination, which facilitates change in their chemical composition (Rollinson, Reference Rollinson1993; Pe-piper et al. Reference Pe-piper, Matarangas, Reynolds and Chatterjee2003). Variations in the trace element and isotopic abundances in the mafic magmas point to geochemically distinct sources in the mantle (Wyllie, Reference Wyllie1988). The presence or absence of incompatible/mobile trace elements and high/low isotopic signatures in the melt can define a source as enriched or depleted (Pearce & Norry, Reference Pearce and Norry1979; Anderson, Reference Anderson1995). Melts from the depleted reservoir are mostly prominent at spreading ridges. On the other hand, it is suggested that enriched magmas are most evident in infant subduction zones, new rifting regions and intraplate environments away from spreading ridges (Anderson, Reference Anderson1995). Therefore, geochemical studies of mafic igneous rocks can provide insights into the processes of generation of mafic magmas from mantle, magma evolution and related tectonic setting.

The Indian subcontinent is endowed with mafic magmatic bodies emplaced over a wide range of time intervals from the Archaean belts of south India to the Cretaceous-Tertiary Deccan basalt sequences (Watts & Cox, Reference Watts and Cox1989; Ahmad & Tarney, Reference Ahmad and Tarney1991). Imprints of mafic magmatism are observed in all the Precambrian cratonic blocks of the Indian shield (Ray et al. Reference Ray, Saha, Koeberl, Thoni, Ganguly and Hazra2013). Numerous mobile belts separate these cratonic blocks. Among the prominent mobile belts, the Central Indian Tectonic Zone (CITZ) is an E-W striking belt, extending eastward through the Chotanagpur Granite Gneiss Complex (CGGC) to the Assam-Meghalaya Gneissic Complex (AMGC) (Acharyya, Reference Acharyya2001). The AMGC of northeastern India comprises a variety of rock types including migmatitic gneisses, metasedimentary rocks, mafic magmatic rocks and several granitoid plutons. It is considered that thermal disturbances in the upper mantle initiated rifting within the AMGC resulting in the formation of a basin now known as the Shillong Basin. Subsequent sedimentation within this basin resulted in the formation of the Shillong Group of metasedimentary rocks (Bhattacharjee & Rahman, Reference Bhattacharjee and Rahman1985). A group of mafic intrusives (i.e. metadolerites) are found within the Shillong Group of rocks. These intrusives are considered evidence of subduction-related mafic magmatism in the AMGC (Ray et al. Reference Ray, Saha, Koeberl, Thoni, Ganguly and Hazra2013). Stratigraphically, they have been classified as the Khasi greenstones (Mazumder, Reference Mazumder1986; Srinivasan et al. Reference Srinivasan, Sen and Bandopadhyay1996).

Studies on mafic rocks from different parts of the CITZ have been carried out by various researchers (Ghosh et al. Reference Ghosh, Fallick, Paul and Potts2005; Kumar & Ahmad, Reference Kumar and Ahmad2007; Ghose & Chatterjee, Reference Ghose, Chatterjee, Srivastava, Sivaji and Chalapathi Rao2008; Srivastava & Ahmad, Reference Srivastava and Ahmad2008; Ahmad & Wanjari, Reference Ahmad and Wanjari2009; Srivastava, Reference Srivastava2012; Alam et al. Reference Alam, Choudhary, Mouri and Ahmad2017; Ahmad et al. Reference Ahmad, Paul, Negi, Akhtar, Gogoi and Saikia2021). However, work on its easternmost extension, that is, the AMGC, is limited. Even though some amount of work has been done on the Proterozoic mafic rocks (i.e. Khasi greenstone) of the Khasi Hills in Meghalaya (Hazra & Ray, Reference Hazra and Ray2009; Mallikharjuna et al. Reference Mallikharjuna Rao, Poornachandra Rao and Sarma2009; Ray et al. Reference Ray, Saha, Koeberl, Thoni, Ganguly and Hazra2013), very few reports exist on the mafic rocks of the Mikir Massif in Assam. This paper focuses on the Proterozoic mafic rocks (metadolerites) of the Mikir Massif. Petrological studies on the basis of field relationships, petrography, mineral compositional analyses with electron probe micro-analyser and geochemistry with major and trace element data are being carried out to give a detailed overview of the genesis of these rocks and their tectonic setting. Inferences of the study will elucidate the formation of the Proterozoic Shillong Basin.

2. Geological setting

The CITZ is an ENE–WSW-trending Proterozoic mobile belt which divides the Greater Indian Landmass into two crustal provinces, viz. Northern Indian Block and Southern Indian Block (Acharyya, Reference Acharyya2003). The Mahakoshal Mobile Belt (MMB) occupies the northern region of the CITZ, while the southern belt of the CITZ consists of the Sausar Mobile Belt, the CGGC and the AMGC. The Bathani volcano-sedimentary sequence (BVSs) is a volcano-sedimentary suite located in the northern fringe of the CGGC and is the eastern continuation of the MMB. Bora et al. (Reference Bora, Kumar, Yi, Kim and Lee2013) attained a U–Pb SHRIMP zircon 206Pb/238U age of 1753 ± 9 Ma, and Yadav et al. (Reference Yadav, Ahmad, Kaulina and Bayanova2015) reported a U–Pb ID-TIMS zircon age of 1873 ± 13 Ma from the calc-alkaline granitoids of MMB. Chatterjee & Ghose (Reference Chatterjee and Ghose2011) obtained a 1697±17 Ma monazite crystallization age from the porphyritic granites of northern CGGC close to BVSs. Saikia et al. (Reference Saikia, Gogoi, Kaulina, Lialina, Bayanova and Ahmad2017, Reference Saikia, Gogoi, Ahmad, Kumar, Kaulina and Bayanova2019) reported a U–Pb ID-TIMS ages of ca. 1700–1600 Ma and a Rb-Sr whole rock isochron age of 1664 ± 130 Ma for the granites of BVSs.

The AMGC is the lone illustration of the Proterozoic gneissic basement rocks in Northeast India. This area is tectonically disconnected from mainland India by the Garo-Rajmahal depression (Eremenco et al. Reference Ermenco, Negi, Nasianov, Seregin, Despande, Sengupta, Talukdar, Sastri, Sokaluv, Pavbukov, Dutta and Raju1969). The AMGC is also flanked by two subduction zones, viz. (i) the Indian Plate subducted beneath the Burmese Plate to its southeast and (ii) Indian Plate subducted beneath the Eurasian Plate to its north. This seismically and geodynamically active area is suggested to be the eastern extension of the Chotanagpur Granite Gneissic Complex (Desikachar, Reference Desikachar1974). This sector is restricted dominantly in the states of Meghalaya and Assam in Northeast India. The Mikir Massif is separated from the Meghalaya part of AMGC by the NW-SE-trending Kopili lineament (Fig. 1). The NE-SW-trending Mesoproterozoic Shillong Basin occupies the northwestern part of Mikir Massif.

Figure 1. Tectonic map of India showing the location of Mikir Massif (MM) along with the Archean cratons and Proterozoic mobile belts (modified after Chatterjee & Ghose, Reference Chatterjee and Ghose2011). The study area is shown within the black square. Abbreviations: ADMB, Aravalli Delhi Mobile Belt; AMGC, Assam-Meghalaya Gneissic Complex; BC, Bastar Craton; BuC, Bundelkhand Craton; BVSs, Bathani volcano-sedimentary sequence; CGGC, Chotanagpur Granite Gneiss Complex; CITZ, Central Indian Tectonic Zone; DC, Dharwar Craton; EGB, Eastern Ghats Belt; KL, Kopili Lineament; MMB, Mahakoshal Mobile Belt, MM, Mikir Massif; NSMB: North Singhbhum Mobile Belt; SB, Satpura Belt; SC, Singhbhum Craton; SONA, Son-Narmada lineaments.

Lithologically, the AMGC comprises basement gneissic complex, amphibolites, migmatites, calc-silicate rocks, granulites, Shillong Group of metasedimentary rocks and Mesozoic-Tertiary igneous and sedimentary rocks (Nandy, Reference Nandy2001; Srivastava & Sinha, Reference Srivastava and Sinha2004). The Shillong Group of rocks is divided into two formations, viz. the Lower Metapelitic Formation and the Upper Quartzitic Formation. The Shillong Group of rocks unconformably overlies the basement gneissic complex (Nandy, Reference Nandy2001; Srivastava & Sinha, Reference Srivastava and Sinha2004). Several discordant granitoid plutons (500–700 Ma) are found intruded within the basement gneissic complex as well as the Shillong Group of rocks (Ghosh et al. Reference Ghosh, Bhalla, Paul, Sarkar, Bishui and Gupta1991; Kumar, Reference Kumar1990). Moreover, Gogoi et al. (Reference Gogoi, Majumdar, Cottle and Dutta2019) reported an LA-inductively coupled plasma mass spectrometry (ICP-MS) U–Pb zircon age of 1644 ± 33 Ma, 1599 ± 17 Ma and 1550 ± 25 Ma from the porphyritic granites of Shillong Basin in the Mikir Massif. Metadolerites occur as concordant and discordant bodies within the Shillong Group (Fig. 2) and are referred to as Khasi greenstone (Mazumder, Reference Mazumder1986). These Khasi greenstones had been subjected to low-grade metamorphism (Srinivasan et al. Reference Srinivasan, Sen and Bandopadhyay1996).

Figure 2. Regional geological map of Mikir Massif showing the study area within the black box (modified after Dhurandhar et al. Reference Dhurandhar, Pandey and Raminaidu2019).

The major lithological components of the Mikir Massif consist of amphibolites, sillimanite-bearing gneisses, granulites, banded iron formation and granite gneisses (Ghosh et al. Reference Ghosh, Fallick, Paul and Potts2005; Gogoi et al. Reference Gogoi, Majumdar, Cottle and Dutta2019). The Shillong Group of rocks in Mikir Massif is illustrated by low-grade schists such as quartz sericite, garnetiferous and ferruginous schists, quartzite, conglomerate and cross-bedded arenites (Dhurandhar et al. Reference Dhurandhar, Pandey and Raminaidu2019), although the Mikir Massif is similar in lithology to the Meghalaya part of AMGC but it differs in the metamorphic grade. A generalized stratigraphic succession of the Mikir Massif is presented in Table 1. The present study is carried out near the Borjuri village of Assam in Northeast India. Here, metadolerites related to the Khasi greenstone are found to be in contact with quartzites of the Shillong Group. These mafic bodies are exposed at a number of places within the Shillong Group of rocks. However, due to very thick vegetation, it is difficult to undertake field investigations in our study area. The present study has been carried out at a large quarry site near the Borjuri village spreading across an area of about 2 sq km (Fig. 2).

Table 1. A generalized stratigraphic succession of Mikir Massif (after Awati et al. Reference Awati, Harikrishnan, Sinha, Gupta and Gupta1995; Majumdar & Gogoi, Reference Majumdar and Gogoi2020)

?No absolute age is available for the metadolerites of the Shillong Basin. The age of these rocks are estimated from palaeomagnetic investigations and age of other rock associations from the Shillong Basin.

3. Field relationships

The Proterozoic mafic intrusives occur in the northwestern domain of the Mikir Massif near the Borjuri village of Nagaon district, Assam. These mafic rocks are basically melanocratic metadolerites (Fig. 3a). The metadolerites share sharp contacts with the quartzitic rocks of the Shillong Group (Fig. 3b-d). The metadolerites show variation in grain size. The rocks are relatively finer towards contact with the quartzites. The metadolerites and the metasedimentary quartzite sequence in the study area have a regional trend of NE-SW and a regional dip towards SE. The area is highly jointed and consists of two sets of prominent joints. The dominant regional joints are systematic, closely spaced joints striking N45°E. The regional joints are crisscrossed by another set of joints at a higher angle striking N30°W. A set of conjugate joints are also seen sporadically in the area (Fig. 3e, f).

Figure 3. Field photographs showing (a) a metadolerite hand specimen. (b–d) Quartzites of the Shillong Group share sharp contacts with the metadolerites or Khasi greenstones. (e, f) Closely spaced vertical joints crisscrossed by another set of joints at a higher angle (shown by yellow lines) and a set of conjugate joints (shown by white lines) were observed in metadolerites.

4. Analytical methods

Mineral chemical analyses of the metadolerites were performed on polished thin-sections using a CAMECA SX 100 electron microprobe at Electron Microprobe Analyser Laboratory, Geological Survey of India, Faridabad (India). The analyses were carried out using an acceleration voltage of 15 kV and beam current of 10 nA with a beam diameter of ca. 1 μm. Standards used include periclase for Mg, wollastonite for Si and Ca, albite for Na, rhodonite for Mn, orthoclase for K, corundum for Al, apatite for P, hematite for Fe, metallic chromium and titanium for Cr and Ti, halite for Cl, metallic zinc for Zn, fluorite for F and barite for Ba. The error in major element concentrations is <1%, whereas the error on trace elements varied between 3 and 5%. For all the elements, Kα lines were analysed with a peak counting time of 10 s and a background counting time of 5 s. The corrections applied to the raw data were that of Pouchou and Pichoir (PAP) procedure (Pouchou & Pichoir, Reference Pouchou, Pichoir, Brown and Packwood1987).

The major oxides, trace elements and rare earth element (REE) analyses were carried out at the Wadia Institute of Himalayan Geology, Dehradun (India). Major oxides and selected trace elemental analyses were performed using a Bruker S8 Tiger Sequential X-ray Spectrometer with Rh excitation source, following the processes explained by Saini et al. (Reference Saini, Mukherjee, Rathi and Khanna2000). The operating conditions for major oxides were no filter, vacuum path and 20/40 kV; for trace elements, they were no filter, vacuum path and 55/60 kV. For major and minor oxides, the overall accuracy in relative standard deviation (RSD) percentage is <5%; for trace elements, it is <12%. The average precision is better than 2% (Saini et al. Reference Saini, Mukherjee, Rathi, Khanna and Purohit2007). The ICP-MS analyses of the REEs were done using the equipment PerkinElmer SCIEX quadrapole type ICP-MS, ELAN DRC-e. For REE analysis, sample solutions were introduced into the argon plasma by means of a peristaltic pump and a cross flow nebulizer. Sample digestion and preparation of solutions were carried out following the procedures of Balaram et al. (Reference Balaram, Saxena, Manikyamba and Ramesh1990). To minimize matrix effect, United States Geological Survey (BHVO-1, AGV-1) and  Geological Survey of Japan (JB1-a, JP-1) samples were used as rock standards. RSD for most of the samples is better than 10%.

5. Results

5.a. Petrography

Petrographically, the metadolerites can be divided into medium-grained and coarse-grained varieties. However, mineralogically they are similar and dominantly consist of amphibole, biotite, plagioclase and minor amounts of chlorite, apatite, titanite and Fe-oxides. Amphibole, plagioclase and biotite were analysed by an electron microprobe to determine their compositions. Representative mineral chemical analyses of these minerals are presented in Supplementary Tables S1, S2 and S3.

The amphiboles are basically secondary phases after pyroxene and are medium to coarse-grained having an elongated prismatic habit and some occur as anhedral masses as well (Fig. 4a). On the amphibole classification diagram (Hawthorne et al. Reference Hawthorne, Oberti, Harlow, Maresch, Martin, Schumacher and Welch2012), the amphiboles plot in the field of magnesio-hornblende and pargasite (Supplementary Fig. S1). Hence, the amphiboles can be classified as calcic-amphiboles. The Ca content of the amphiboles ranges from 1.57 to 1.95 apfu (atom per formula unit), Aliv varies from 1.49 to 1.72 apfu, (Na+K)A from 0.42 to 0.62 apfu and Mg# ranges from 0.37 to 0.42. The biotite grains are brown-coloured tabular crystals showing mottled birds-eye extinction (Fig. 4b, c). On the biotite classification diagram (Tischendorf et al. Reference Tischendorf, Gottesmann, Förster and Trumbull1997), the biotites are classified as Fe-biotites (Supplementary Fig. S2). The Si value of biotite ranges from 6.28 to 7.10 apfu, total Al abundance ranges from 1.35 to 1.50 apfu and Fe+Mg values ranges from 2.45 to 2.70 apfu. On the FeO+MnO–MgO–10TiO2 ternary diagram (Nachit et al. Reference Nachit, Ibhi and Ohoud2005), the analysed biotite crystals distinctly plot within the re-equilibrated biotite field (Supplementary Fig. S3), which suggests that biotite crystals in the studied rocks are secondary in nature. The plagioclase crystals generally occur as medium-grained, lamellar-twinned lath-shaped crystals. The analysed plagioclase grains are of oligoclase and andesine composition (An21 to An32) (Supplementary Fig. S4). Minor acicular apatite grains and anhedral Fe-oxides occur as discrete mineral phases in the section (Fig. 4d).

Figure 4. Representative photomicrographs of the metadolerites illustrating the mineral assemblages (a) plane-polarized light (PPL) images of secondary amphibole, (b, c) amphibole, biotite and Fe-oxide, (d) amphibole and Fe-oxide and (e, f) cross-polarized light (CPL) images of plagioclase laths embedded within secondary amphibole. Mineral abbreviations: Am, amphibole; Bt, biotite; Pl, plagioclase.

Both the coarse-grained and medium-grained metadolerites show inequigranular, holocrystalline texture. Primarily, plagioclase laths are embedded within amphibole grains, which resemble an ophitic texture (Fig. 4e, f). There is a possibility that the amphiboles present are secondary minerals formed by the transformation from pre-existing pyroxene grains. This suggests that the mafic rocks have been influenced by metamorphism leading to the formation of secondary amphiboles.

5.b. Geochemistry

The major and trace element distributions of the Borjuri metadolerites are presented in Table 2. In general, for the geochemical classification of subvolcanic rocks for un-metamorphosed samples, the total alkali versus silica diagram is used (La Bas et al. Reference Le Bas, Le Maitre, Streckeisen and Zanettin1986). However, the studied mafic rocks have undergone deformation and metamorphism. Hence, the immobile trace elements such as REEs and high field strength elements (HFSEs) are given more importance. The HFSE ratio of parent rocks is similar to that of the altered rocks (Finlow-Bates & Stumpfl, Reference Finlow-Bates and Stumpfl1981; MacLean & Barrett, Reference MacLean and Barrett1993). We have therefore adopted a classification diagram based on immobile elements, which is generally used for altered rocks. In the Zr/Ti versus Nb/Y discrimination diagram after Pearce (Reference Pearce and Wyman1996), all the studied samples plot in the field of basalt (Fig. 5a). The mafic rocks show tholeiitic trend in the (Na2O+K2O)-FeOT-MgO or AFM (alkalis-iron-magnesium) plot (Fig. 5b; Irvine & Baragar, Reference Irvine and Baragar1971). We have drawn a comparison of our data with known Indian occurrences of Proterozoic mafic rocks from the BVSs of CGGC: pillow basalts and dolerites after Ahmad et al. (Reference Ahmad, Paul, Negi, Akhtar, Gogoi and Saikia2021) and dolerites after Gogoi (Reference Gogoi2022).

Table 2. Major oxides (wt %), trace and rare earth elements (ppm) composition of the metadolerites of Borjuri from Mikir Massif part of Assam-Meghalaya Gneissic Complex

LOI, loss on ignition; BDL, below detection limit.

Figure 5. (a) Geochemical classification diagram based on immobile trace elements (after Pearce, Reference Pearce and Wyman1996) and (b) AFM diagram displaying tholeiitic nature of magma for the metadolerites of Borjuri (after Irvine & Baragar, Reference Irvine and Baragar1971). Data used for comparison are from Proterozoic mafic rocks of Bathani volcano-sedimentary sequence: Ahmad et al. (Reference Ahmad, Paul, Negi, Akhtar, Gogoi and Saikia2021) and Gogoi (Reference Gogoi2022).

The primitive mantle-normalized multi-element diagram is shown in Figure 6a. The diagram exhibits distinct peaks at large ion lithophile elements (LILE) Rb, K and Pb, and noteworthy troughs at HFSEs Nb, Zr and Yb. The elevated concentrations of LILE in the rocks can be attributed to dehydration of the subducted slab and transportation of the LILE-rich fluids to the overlying mantle wedge (Saunders et al. Reference Saunders, Tarney and Weaver1980; Wilson & Davidson, Reference Wilson and Davidson1984). The negative aberrations depicted by Nb and Yb can be ascribed to the subduction zone magmatic regime (Sheraton et al. Reference Sheraton, Black, McCulloch and Oliver1990; Zhao et al. Reference Zhao, Shiraishi, Ellis and Sheraton1995). Chondrite-normalized REE diagram of the studied metadolerites displays elevated patterns for LREE ((La/Yb)N=3.49–4.97; (Gd/Yb)N=1.98–2.30)) and nearly flat patterns for HREE (Fig. 6b). Most of the samples exhibit positive europium anomaly (Eu/Eu*=0.99–1.13) indicating accumulation of plagioclase. The REEs displaying sub-parallel trends suggest that the rocks are cogenetic in nature. The trace and REE patterns of mafic rocks from the Shillong Basin are similar to those of mafic rocks from the BVSs, with the samples from the former displaying slightly enriched character (Fig. 6a, b).

Figure 6. (a) Primitive mantle-normalized multi-element diagram (normalizing values after Sun & McDonough, Reference Sun and McDonough1989) and (b) chondrite-normalized rare earth element patterns (normalizing values after Boyton, Reference Boynton1984) for the metadolerite samples of Borjuri. For comparison purpose, trace element data in yellow and green patterns are plotted from Ahmad et al. (Reference Ahmad, Paul, Negi, Akhtar, Gogoi and Saikia2021) and Gogoi (Reference Gogoi2022) respectively.

5.c. Pressure and temperature estimates

For geothermobarometric studies, six pairs of mineral chemical data of plagioclase and amphibole were selected from the metadolerites. Various researchers have proposed several methods (Blundy & Holland, Reference Blundy and Holland1990; Holland & Blundy, Reference Holland and Blundy1994; Anderson & Smith, Reference Anderson and Smith1995; Anderson, Reference Anderson1996) for geothermobarometry calculations. Amphiboles are crucial minerals for geothermobarometry as they are combinedly affected by parameters such as pressure, temperature and oxygen fugacity. From experimental and field studies regarding the depth of emplacement of plutons, it is seen that the Al content of amphibole is considered a function of pressure and temperature (Hollister et al. Reference Hollister, Grissom, Peters, Stowell and Sisson1987; Thomas & Ernst, Reference Thomas and Ernst1990; Blundy & Holland, Reference Blundy and Holland1990; Schmidt, Reference Schmidt1992; Anderson & Smith, Reference Anderson and Smith1995).

The metamorphic P-T conditions were estimated using the amphibole-plagioclase geothermobarometry. To calculate pressure, we have used the formulation of Al-in-hornblende barometer by Anderson & Smith (Reference Anderson and Smith1995). The errors associated with the Al-in-hornblende barometer are ± 0·6 kbar. The pressure obtained for the samples varies between 7 and 8 kbar (Supplementary Table S4). Holland & Blundy (Reference Holland and Blundy1994) proposed two geothermometers. One applies to quartz-bearing igneous rocks (edenite + 4quartz = tremolite + albite), while the other is for both quartz-bearing and quartz-free igneous rocks (edenite + albite = richterite + anorthite). For our samples, compositions of amphibole along with coexisting plagioclase were used for temperature calculations. The calculation was carried out with the help of amphibole-plagioclase geothermometer (edenite-richterite thermometer) after Holland & Blundy (Reference Holland and Blundy1994). The errors associated with the thermometer are ± 40°C. The temperature of the samples ranges from 633°C to 703°C with an average of 664°C (Supplementary Table S4). In the calcic amphibole discrimination diagram (Supplementary Fig. S5) by Fleet & Barnett (Reference Fleet and Barnett1978), the amphiboles distinctly plot in the low-pressure metamorphic field. In the temperature–pressure diagram, the estimated temperature and pressure of the metadolerites plot in the amphibolite facies and medium P-T zone (Supplementary Fig. S6).

6. Discussion

6.a. Post-crystallization alteration and crustal contamination

The field as well as petrographic observations shows post-crystallization alteration effects in the metadolerites of Borjuri area. Amphibole, biotite and chlorite observed under a microscope are secondary minerals that have formed from the replacement of primary Fe-Mg minerals, such as pyroxene. Isochemical and volume-to-volume replacement are responsible for the replacement of original pyroxene by amphiboles which do not bring about any changes in the original igneous textures (Redman & Keays, Reference Redman and Keays1985). The occurrence of such secondary metamorphic minerals helps to ascertain the degree of alteration. Other than the petrographic study, the extent of post-magmatic alteration processes can also be evaluated with the help of Harker variation diagrams. A series of binary variation diagrams (Fig. 7) of selected major and trace elements have been plotted against SiO2, and the observations show positive, negative and scattered trends. The elements CaO, Na2O, P2O5 and Sr show positive trends in a linear or non-linear way, which is suggestive of the accumulation of plagioclase and apatite in the evolving magma. However, with an increase in SiO2, the elements Fe2O3, TiO2, K2O, Rb, Cr and Y all show negative trends. This is indicative of the fractionation of amphibole, biotite, clinopyroxene, titanite and Fe-Ti oxides in the evolving magma. The positive and negative trends displayed by these major and minor elements suggest near primary magmatic characteristics. On the other hand, the elements MgO and Al2O3 show plots of scattered nature, which decipher post-crystallization alteration effects or the post-magmatic mobility of these elements.

Figure 7. Bivariate diagrams of (a–h) major elements versus SiO2 and (i–l) selected trace elements versus SiO2.

Mantle-derived mafic magmas are prone to crustal assimilation during ascent or when emplaced in crustal magma chambers (Mohr, Reference Mohr, Halls and Fahrig1987; Pe-piper et al. Reference Pe-piper, Matarangas, Reynolds and Chatterjee2003). It is therefore crucial to assess the effects of crustal contamination on the chemical composition of the primary mafic magma. Elemental ratios such as Ba/Nb, Ti/Y, La/Yb, Ba/Zr, La/Sm, Zr/Y, Nb/La, Nb/Th and Nb/Y are frequently used to evaluate crustal assimilation in mafic rocks (Rollinson, Reference Rollinson1993). Trace-element ratio plots for the Borjuri metadolerites are presented in Figure 8a–d. Plots show that the studied metadolerites plot away from the mantle array, which likely suggests an influence of crustal assimilation (Fig. 8a, b). The Nb/La versus SiO2 and Nb/Th versus Nb/La diagrams also suggest a crustal contamination in the samples (Fig. 8c, d).

Figure 8. (a) Ti/Y versus Ba/Nb, (b) Ti/Y versus Ba/Zr, (c) Nb/La versus SiO2 and (d) Nb/La versus Nb/Th variation diagrams of the mafic rocks of Borjuri.

6.b. Magmatic evolution

Mafic rocks derived from the lithospheric mantle have higher LILE and LREE element abundances and lower HFSE than those generated by the asthenospheric mantle melting. As a result of these variations in elemental abundances between the asthenospheric and lithospheric mantle, an arrangement of HFSE–LREE ratios is useful for understanding the source of mantle-derived rocks. The Zr/Ba values are being routinely used to differentiate between asthenospheric (Zr/Ba > 0.5) and lithospheric sources (Zr/Ba = 0.3–0.5) (Ormerod et al. Reference Ormerod, Hawkesworth, Leeman and Menzies1988; Menzies et al. Reference Menzies, Kyle, Jones and Ingram1991). The Zr/Ba range of the studied samples varies from 0.21 to 0.46 indicating the involvement of the lithospheric mantle. Moreover, the elemental ratio of La/Nb of the Borjuri metadolerites ranges from 1.23 to 2.32. This is comparable to the lithospheric mantle where La/Nb > 1 (Fitton et al. Reference Fitton, James, Kempton, Ormerod and Leeman1988; Thompson & Morrison, Reference Thompson and Morrison1998). The ratio La/Nb also therefore suggests an origin from the lithospheric mantle for the metadolerites. As HFSEs like Nb are depleted in the lithospheric mantle with respect to the LREEs, low Nb/La implies a lithospheric mantle origin and higher Nb/La suggests an OIB-like asthenospheric mantle source. The plot of (La/Yb) versus (Nb/La) (Smith et al. Reference Smith, Sánchez, Walker and Wang1999) further confirms the origin of the metadolerites from lithospheric and mixed lithospheric–asthenospheric mantle sources (Fig. 9a).

Figure 9. (a) La/Yb versus Nb/La (Smith et al. Reference Smith, Sánchez, Walker and Wang1999) plot showing lithospheric and mixed lithospheric mantle source for Borjuri mafics. (b) Th/Sm versus Th/Ce plot of the Borjuri mafics shows the input of subducted sediment. (c) Nb/Y versus Rb/Y and (d) Nb/Zr versus Th/Zr, indicating fluid-related enrichment in Borjuri mafics.

The trace elemental signatures such as enrichment of LILE with respect to HFSE and negative Nb anomaly are all aspects of subduction-related mantle metasomatism. Generally, the enrichment of mantle source region can be ascribed to metasomatism by (1) fluxing of slab-derived fluids obtained from the dehydration of oceanic crust (Hawkesworth et al. Reference Hawkesworth, Gallagher, Hergt and McDermott1993; Turner et al. Reference Turner, Hawkesworth, Rogers and King1997) and (2) addition of hydrous melts of subducted sediments (Peacock et al. Reference Peacock, Rushmer and Thompson1994; Hawkesworth et al. Reference Hawkesworth, Turner, McDermott, Peate and Van Calsteren1997) or of the MORB segment of the slab (Peacock et al. Reference Peacock, Rushmer and Thompson1994; Stern & Kilian, Reference Stern and Kilian1996). The distribution and variability of elements such as REE, Th and Ba provide valuable constraints for identifying the role of slab-derived fluids and slab-derived melts in mantle metasomatism (Spandler & Pirard, Reference Spandler and Pirard2013; Guo et al. Reference Guo, Li, Fan, Li, Zhao, Huang and Xu2015). In the Th/Sm versus Th/Ce plot (Fig. 9b), the metadolerites illustrate a sediment-input trend indicative of sediment-related melt enrichment. The vertical and horizontal trends shown by the ratio plots of Rb/Y versus Nb/Y (Fig. 9c) and Nb/Zr versus Th/Zr (Fig. 9d), respectively, indicate fluid-related enrichment of the mafic magma. Source enrichment by fluids from the subducted slab is also supported by high Ba/Nb (> 28; Fitton et al. Reference Fitton, James, Kempton, Ormerod and Leeman1988). Ba/Nb ratio of the studied metadolerites ranges from 30.08 to 56.90. Fluids obtained from the subducted slab will modify the overlying mantle wedge geochemically. It will also lower the peridotite solidus to generate magmas (Stolper & Newman, Reference Stolper and Newman1994; Grove et al. Reference Grove, Parman, Bowring, Price and Baker2002). Trace elemental geochemical attributes of the metadolerites of Borjuri strongly suggest derivation from a fluid-enriched lithospheric mantle source in a subduction regime.

A petrogenetic model based on Ni versus Zr was used for evaluating the lherzolite mantle composition. It is evident from the Ni versus Zr plot (Condie et al. Reference Condie, Bobrow, Card, Halls and Fahrig1987) that the metadolerites are products of a primary magma generated by 6%–10% melting of a lherzolite mantle source (Fig. 10a). REE ratios and their abundances (e.g. Sm/Yb and La/Yb) are extensively used to determine the source of mafic magmas and the degree/variation of mantle melting (McKenzie & O’Nions, Reference McKenzie and O’Nions1991; Su et al. Reference Su, Zheng, Griffin, Zhao, Tang, Ma and Lin2012). In the Sm/Yb versus La/Yb diagram, the studied samples plot below or near the spinel + garnet lherzolite (50:50) melting curve with primitive mantle as the initial composition (Fig. 10b). Here, the Sm/Yb values are lower than that of the garnet lherzolite melting curve and higher than that of the spinel lherzolite melting curve. Therefore, it can be inferred that the parental mafic magma for metadolerites was obtained from a mantle source incorporating spinel + garnet lherzolite. Moreover, the Sm/Yb versus La/Yb plot indicates an approximately 6%–10% partial melting of mantle lherzolite was necessary for the generation of parental magma of the metadolerites. Thus, from the trace element ratio plots it is evident that the parental magma of the metadolerites from Borjuri area was produced by 6%–10% melting of the lherzolite mantle source.

Figure 10. Plots for the metadolerites of Borjuri (a) Ni versus Zr petrogenetic model for a lherzolite mantle composition. Curve 1 signifies batch melting at 1500 °C (1 atm. equivalent) with percentages along the line representing degree of partial melting. Curves 2, 3 and 4 demonstrate olivine fractionation with elimination of olivine marked in 5% increments (after Condie et al. Reference Condie, Bobrow, Card, Halls and Fahrig1987). (b) La/Yb versus Sm/Yb diagram where the Borjuri metadolerites cluster near the spinel+garnet lherzolite curve (after Liu et al. Reference Liu, Ma, Zhang, Xiong, Huang and Jiang2014). Abbreviation: PM, primitive mantle.

6.c. Geotectonic framework

Rocks of basaltic composition are ubiquitous in every tectonic environment. They are geochemically sensitive to the changes in the geotectonic framework. Thus, the geochemical signature of basaltic rocks is commonly used to discriminate tectonic settings. The metadolerites of Borjuri area show high LREE/HREE and LILE/HFSE ratios, which are distinguishing traits of magmas produced in subduction zones (Hawkesworth et al. Reference Hawkesworth, Gallagher, Hergt and McDermott1993; Gorton & Schandl, Reference Gorton and Schandl2000). In the Th/Yb versus Nb/Y tectonic discrimination diagram (Pearce, Reference Pearce2008), composition of the metadolerites clusters near E-MORB composition (Fig. 11a) displaying resemblance to back-arc basalts and dolerites of the BVSs (Ahmad et al. Reference Ahmad, Paul, Negi, Akhtar, Gogoi and Saikia2021; Gogoi, Reference Gogoi2022). Furthermore, fluid immobile elements such as Zr, Ti, V and Y, which are independent of secondary processes (e.g. weathering and metamorphism), help in evaluating the tectonic setting of altered basic rocks. In the Ti/100-Zr-Yx3 ternary diagram (after Pearce & Cann, Reference Pearce and Cann1973), the metadolerites display mixed signatures of mid-oceanic ridge basalt, island arc tholeiites and calc-alkali basalts (Fig. 11b). The observed hybrid mixture between mid-oceanic ridge basalt and island arc tholeiites for the studied mafic rocks can be associated with back-arc basin basalts (Pearce & Peate, Reference Pearce and Peate1995; Pearce & Stern, Reference Pearce, Stern, Christie, Fisher, Lee and Givens2006). In the Ti/Zr versus Zr (Fig. 11c) and V/Ti versus Zr (Fig. 11d) discrimination diagrams after Woodhead et al. (Reference Woodhead, Eggins and Gamble1993), the mafic rocks distinctly plot in the field of back-arc basin. Thus, it can be inferred that the parental magmas responsible for the generation of the metadolerite intrusives were produced in a back-arc extensional basin associated with subduction dynamics. Our data show similar geochemical characteristics to the known occurrences of mafic rocks from the BVSs.

Figure 11. Tectonic discrimination diagrams for the Borjuri metadolerites. (a) Nb/Yb versus Th/Yb plot after Pearce (Reference Pearce2008). (b) Ti/100-Zr-Yx3 ternary diagram after Pearce & Cann (Reference Pearce and Cann1973). (c) Ti/Zr versus Zr after Woodhead et al. (Reference Woodhead, Eggins and Gamble1993). (d) Zr versus V/Ti after Woodhead et al. (Reference Woodhead, Eggins and Gamble1993). Field of Andean back-arc basalt is taken from Espanon et al. (Reference Espanon, Chivas, Kinsley and Dosseto2014). Abbreviations: BABB, Back-arc basin basalt; CAB, Calc-alkali basalts; IAB, Island-arc basalt; IAT, Island-arc tholeiites; MORB, Mid-oceanic ridge basalt; WPB, Within plate basalt. Symbols as in Figure 5.

6.d. Geodynamic implications

The Shillong Basin is an integral unit of AMGC that developed in the east-central part of the Meghalaya plateau and continued up to the Mikir Massif. This NE-SW interior basin is believed to have originated due to a thermal event followed by subsidence in a tensional stress regime (Nandy, Reference Nandy2001). The initiation of sedimentation in the basin commenced as early as 1900 Ma and continued till 1400 Ma (zircon dates of quartzites by Bidyananda & Deomurari, Reference Bidyananda and Deomurari2007). The metasedimentary province of Shillong Basin is intruded by several mafic and felsic dykes and sills. These mafic units, which have been subjected to amphibolite-facies metamorphism, occur in close association with the quartzites of the Shillong Group of rocks in the Mikir Massif (Fig. 3b-d). In this study, we attempt to give the tectonic framework responsible for the formation of the Shillong Basin. The geochemical signatures of the metadolerites presented in this work indicate a subduction zone setting for the generation of these rocks induced by fluid flux melting of the lithospheric mantle in a back-arc region. This implicates that the Shillong Basin represents a back-arc rift basin.

Recorded geochronological evidence suggests that AMGC has undergone Columbian, Grenvillian and Pan-African magmatic events (Yin et al. Reference Yin, Dubey, Webb, Kelty, Grove, Gehrels and Burgess2010; Chatterjee et al. Reference Chatterjee, Bhattacharya, Duarah and Mazumdar2011; Kumar et al. Reference Kumar, Rino, Hayasaka, Kimura, Raju, Terada and Pathak2017; Gogoi et al. Reference Gogoi, Majumdar, Cottle and Dutta2019). Shreds of evidence of such events from AMGC establish its involvement in the amalgamation and break-up of Columbia, Rodinia and Pangea. The Paleo-Mesoproterozoic felsic magmatism of the AMGC (Bidyananda & Deomurari, Reference Bidyananda and Deomurari2007; Yin et al. Reference Yin, Dubey, Webb, Kelty, Grove, Gehrels and Burgess2010; Kumar et al. Reference Kumar, Rino, Hayasaka, Kimura, Raju, Terada and Pathak2017; Gogoi et al. Reference Gogoi, Majumdar, Cottle and Dutta2019; Doley et al. Reference Doley, Bhagabaty, Sarma, Singh and Zou2022) is comparable to the Aravalli Belt of northwest India (Kaur et al. Reference Kaur, Zeh, Chaudhri, Gerdes and Okrusch2013), Eastern Ghats Mobile Belt of southeast India (Dobmeier et al. Reference Dobmeier, Lütke, Hammerschmidt and Mezger2006), MMB of central India (Bora et al. Reference Bora, Kumar, Yi, Kim and Lee2013) and CGGC (Ray Barman et al. Reference Ray Barman, Bishui and Sarkar1990; Ray Barman & Bishui, Reference Ray Barman and Bishui1994; Acharyya, Reference Acharyya2003; Hossain et al. Reference Hossain, Tsunogae, Rajesh, Chen and Arakawa2007; Saikia et al. Reference Saikia, Gogoi, Kaulina, Lialina, Bayanova and Ahmad2017; Gogoi, Reference Gogoi2022). On drawing a comparison between the geochemical trends shown by the metadolerites of Shillong Basin and mafic rocks of BVSs, it is clear that the similarity is very significant. Moreover, in the present study it is noted from field observations that the joints present in the metadolerites show a trend of NE-SW and a dip towards SE which is directly comparable to the regional trend of CITZ (ENE-WSW). From a comparative analysis of our geochemical data, similar lithological and structural frameworks and associated geological dates given by previous authors, we envisage that the Shillong Basin is the eastern continuation of the BVSs and MMB of the CITZ (Fig. 12).

Figure 12. Map of Indian subcontinent displaying the ENE-WSW-trending Central Indian Tectonic Zone (CITZ), Chotanagpur Granite Gneiss Complex (CGGC) and Assam-Meghalaya Gneissic Complex (AMGC) (modified after Saikia et al. Reference Saikia, Gogoi, Kaulina, Lialina, Bayanova and Ahmad2017). This continuous mobile belt demarcates the boundary between the North Indian Block (NIB) from the South Indian Block (SIB) (Acharyya Reference Acharyya2003; Bhandari et al. Reference Bhandari, Pant, Bhowmik and Goswami2011; Chatterjee & Ghose, Reference Chatterjee and Ghose2011; Rekha et al. Reference Rekha, Upadhyay, Bhattacharya, Kooijman, Goon, Mahato and Pant2011; Bhowmik et al. Reference Bhowmik, Wilde, Bhandari, Pal and Pant2012). The NIB comprises of the Bundelkhand Craton (BuC), whereas the SIB is a union of the Dharwar Craton (DC), the Singhbhum Craton (SC) and the Bastar Craton (BC). The Aravalli Delhi Mobile Belt (ADMB) and the Eastern Ghats Belt (EGB) are also shown. Two volcano-sedimentary sequences of CITZ, namely Mahakoshal Mobile Belt (MMB) and its eastern continuation Bathani volcano-sedimentary sequence (BVSs), are shown. The present study projects the Shillong Basin of AMGC as the eastern continuation of the BVSs. The ages reported from the MMB are taken from: 1) Sarkar et al. (Reference Sarkar, Boda, Kundu, Mamgain and Shankar1998); 2) Bora et al. (Reference Bora, Kumar, Yi, Kim and Lee2013); 3) Yadav et al. (Reference Yadav, Ahmad, Kaulina and Bayanova2015). The ages reported from the BVSs are taken from: 1) Chatterjee & Ghose (Reference Chatterjee and Ghose2011); 2) Saikia et al. (Reference Saikia, Gogoi, Kaulina, Lialina, Bayanova and Ahmad2017, Reference Saikia, Gogoi, Ahmad, Kumar, Kaulina and Bayanova2019). The ages reported from the Shillong Basin are taken from: 1) Gogoi et al. (Reference Gogoi, Majumdar, Cottle and Dutta2019). The ages reported from the CITZ are taken from: 1) Bhandari et al. (Reference Bhandari, Pant, Bhowmik and Goswami2011); 2) Bhowmik et al. (Reference Bhowmik, Wilde and Bhandari2011). The ages reported from the CGGC are taken from: 1) Pandey et al. (Reference Pandey, Gupta and Lall1986); 2) Chatterjee et al. (Reference Chatterjee, Crowley and Ghose2008). The ages reported from the AMGC are taken from: 1) Chatterjee et al. (Reference Chatterjee, Mazumdar, Bhattacharya and Saikia2007); 2) Yin et al. (Reference Yin, Dubey, Webb, Kelty, Grove, Gehrels and Burgess2010).

In various hypothesized models, the Greater Indian Landmass has been an essential element in the build-up of the Columbia supercontinent (Zhao et al. Reference Zhao, Cawood, Wilde and Sun2002; Hou et al. Reference Hou, Santosh, Qian, Lister and Li2008; Pesonen et al. Reference Pesonen, Mertanen and Veikkolainen2012; Zhang et al. Reference Zhang, Li, Evans, Wu, Li and Dong2012; Pisarevsky et al. Reference Pisarevsky, Biswal, Wang, DeWaele, Ernst, Södurlund, Tait, Ratre, Singh and Cleve2013; Bhowmik, Reference Bhowmik2019; Chattopadhyay et al. Reference Chattopadhyay, Bhowmik and Roy2020; Sequeira et al. Reference Sequeira, Bhattacharya and Bell2022). On the basis of previous literatures, it can be deduced that Columbia supercontinent assembled during 1900–1600 Ma and fragmented at around 1500–1200 Ma (Evans, Reference Evans2013). An even wider time span for tectonic history of the Columbia supercontinent, that is, ca 2100–1800 Ma for its amalgamation, 1800–1600 Ma for intracontinental rifting and 1300–1200 Ma for break-up, have been considered (Saha et al. Reference Saha, Bhowmik, Bose and Sajeev2016; Wang et al. Reference Wang, He, Carranza and Cui2019; Meert & Santosh, Reference Meert and Santosh2022). A large number of literatures are available regarding the position of Indian subcontinent within the Columbia supercontinent. India was placed near Laurentia and the North China Craton by Hou et al. (Reference Hou, Santosh, Qian, Lister and Li2008), while Zhang et al. (Reference Zhang, Li, Evans, Wu, Li and Dong2012) have placed it adjacent to Australia, and Pisarevsky et al. (Reference Pisarevsky, Biswal, Wang, DeWaele, Ernst, Södurlund, Tait, Ratre, Singh and Cleve2013) have positioned India next to the Sarmatia margin of Baltica. Thus, India’s position in Columbia reconstructions has largely remained dubious due to inadequate geological and geochronological data. During Columbia formation, the northern and southern Indian blocks merged to form a single continental block, that is, the Greater Indian Landmass. Thus, India was not a single continental block until the amalgamation of Columbia (Saikia et al. Reference Saikia, Gogoi, Kaulina, Lialina, Bayanova and Ahmad2017). To get a closer outlook on the assembly, configuration and breaking up of the Greater Indian Landmass in relation to Columbia supercontinent, it is necessary to look for a continuation of mobile belts similar to CITZ in other continental blocks. Identifying Paleo-Mesoproterozoic mobile belt with an associated back-arc rift basin will allow us to position the hypothesized proto-continental blocks in the hypothetical supercontinental models.

7. Conclusions

The AMGC is exemplary of a number of magmatic events which have led to many felsic and mafic intrusives within the basement gneisses as well as the Shillong Group of rocks. The studied metadolerites related to the Khasi greenstones are found as intrusive in contact with quartzites of the Borjuri area in the Mikir Massif. These metadolerites have undergone metamorphism reaching amphibolite facies conditions with estimated pressure and temperature of 7–8 kbar and 664 °C, respectively. The studied mafic rocks depict well-preserved signatures that define the geotectonic framework of a subduction zone setting, more specifically a back-arc extensional setting. Comparing geochemical data of the metadolerites and taking into consideration structural affiliations along with available geochronological dates, marked similarity is seen between the Shillong Basin of AMGC and BVSs of CGGC. Thus, we can conclude that the Shillong Basin is the eastern extension of the BVSs and also MMB of the CITZ.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756823000237

Acknowledgements

Bibhuti Gogoi acknowledges the DST-SERB grant vide Project No. CRG/2020/002635. Pallabi Basumatary acknowledges the CSIR-JRF fellowship No. 09/1236(11154)/2021-EMR-I. We would like to express our sincere gratitude to three anonymous reviewers for their insightful and constructive comments. We thank Dr. Sarah Sherlock for editorial handling of the manuscript. Useful suggestions by Dr. Amiya Baruah are gratefully acknowledged.

Competing interests

The authors declare that in this present work there is no conflict of interests.

References

Acharyya, SK (2001) Geodynamic setting of the Central Indian Tectonic Zone in central, eastern and north eastern India. Geological Survey of India, Special Publication 64, 1735.Google Scholar
Acharyya, SK (2003) The nature of Mesoproterozoic central Indian tectonic zone with exhumed and reworked older granulites. Gondwana Research 6, 197214.CrossRefGoogle Scholar
Ahmad, M and Wanjari, N (2009) Volcano-Sedimentary Sequence in the Munger-Rajgir Metasedimentary Belt, Gaya District, Bihar. Indian Journal of Geoscience 63, 351–60.Google Scholar
Ahmad, M, Paul, AQ, Negi, P, Akhtar, S, Gogoi, B, and Saikia, A (2021) Mafic rocks with back-arc E-MORB affinity from the Chotanagpur Granite Gneiss Complex of India: relicts of a Proterozoic Ophiolite suite. Geological Magazine 158, 1527–42.CrossRefGoogle Scholar
Ahmad, T and Tarney, J (1991) Geochemistry and petrogenesis of Garhwal volcanics: implications for evolution of the north Indian lithosphere. Precambrian Research 50, 6988.CrossRefGoogle Scholar
Alam, M, Choudhary, AK, Mouri, H and Ahmad, T (2017) Geochemical characterization and petrogenesis of mafic granulites from the Central Indian Tectonic Zone (CITZ). Geological Society, London, Special Publications 449, 207–29.CrossRefGoogle Scholar
Anderson, D (1995) Lithosphere, asthenosphere, and perisphere. Reviews of Geophysics 33, 125–49.CrossRefGoogle Scholar
Anderson, JL (1996) Status of thermobarometry in granitic batholiths. Transactions of Royal Society Edinburgh, Earth Sciences 87, 125–38.CrossRefGoogle Scholar
Anderson, JL and Smith, DR (1995) The effects of temperature and fO 2 on the Al-in hornblende barometer. American Mineralogist 8, 549–59.CrossRefGoogle Scholar
Awati, AB, Harikrishnan, T, Sinha, RM, Gupta, KR and Gupta, RK (1995) The Proterozoic Shillong basin of North East India: conceptual model for hosting unconformity related uranium mineralization. Exploration and Research for Atomic Minerals 8, 141–54.Google Scholar
Balaram, V, Saxena, VK, Manikyamba, C and Ramesh, SL (1990) Determination of rare earth elements in Japanese rock standards by inductively coupled plasma mass spectrometry. Atomic Spectroscopy 11, 1923.Google Scholar
Bhandari, A, Pant, NC, Bhowmik, SK and Goswami, S (2011) _1.6 Ga ultrahigh-temperature granulite metamorphism in the Central Indian Tectonic Zone: insights from metamorphic reaction history, geothermobarometry and monazite chemical ages. Geological Journal 46, 198216.CrossRefGoogle Scholar
Bhattacharjee, CC and Rahman, S (1985) Structure and lithostratigraphy of the Shillong Group of rocks in East Khasi hills of Meghalaya. Bulletin of Geological Mining and Metallurgy Society of India 53, 90–9.Google Scholar
Bhowmik, SK (2019) The current status of orogenesis in the Central Indian Tectonic Zone: a view from its Southern Margin. Geological Journal 54, 2912–34.CrossRefGoogle Scholar
Bhowmik, SK, Wilde, SA and Bhandari, A (2011) Zircon U–Pb/Lu–Hf and monazite chemical dating of the tirodi biotite gneiss: implication for Latest Paleoproterozoic to Early Mesoproterozoic Orogenesis in the Central Indian Tectonic Zone. Geological Journal 46, 574–96.CrossRefGoogle Scholar
Bhowmik, SK, Wilde, SA, Bhandari, A, Pal, T and Pant, NC (2012) Growth of the Greater Indian landmass and its assembly in Rodinia: geochronological evidence from the Central Indian Tectonic Zone. Gondwana Research 22, 5472.CrossRefGoogle Scholar
Bidyananda, M and Deomurari, MP (2007) Geochronological constraints on the evolution of Meghalaya massif, northeastern India: an ion microprobe study. Current Science 93, 1620–3.Google Scholar
Blundy, JD and Holland, TJB (1990) Calcic amphibole equilibria and a new amphibole plagioclase geothermometer. Contributions to Mineralogy and Petrology 104, 208–24.CrossRefGoogle Scholar
Bora, S, Kumar, S, Yi, K, Kim, N and Lee, TH (2013) Geochemistry and U-Pb SHRIMP zircon chronology of granitoids and microgranular enclaves from Jhirgadandi Pluton of Mahakoshal Belt, Central India Tectonic Zone, India. Journal of Asian Earth Sciences 70–71, 99114.CrossRefGoogle Scholar
Boynton, WV (1984) Cosmochemistry of the rare earth elements: meteorite studies. In Developments in geochemistry, Elsevier 2, 63–114.CrossRefGoogle Scholar
Chatterjee, N and Ghose, NC (2011) Extensive Early Neoproterozoic high-grade metamorphism in North Chotanagpur Gneissic Complex of the Central Indian Tectonic Zone. Gondwana Research 20, 362–79.CrossRefGoogle Scholar
Chatterjee, N, Bhattacharya, A, Duarah, BP and Mazumdar, AC (2011) Late Cambrian reworking of Paleo-Mesoproterozoic granulites in Shillong-Meghalaya gneissic complex (Northeast India): evidence from PT pseudosection analysis and monazite chronology and implications for East Gondwana assembly. The Journal of Geology 119, 311–30.CrossRefGoogle Scholar
Chatterjee, N, Crowley, JI and Ghose, NC (2008) Geochronology of the 1.55 Ga Bengal anorthosite and Grenvillian metamorphism in the Chotanagpur Gneissic complex, eastern India. Precambrian Research 161, 303–16.CrossRefGoogle Scholar
Chatterjee, N, Mazumdar, AC, Bhattacharya, A and Saikia, RR (2007) Mesoproterozoic granulites of the Shillong-Meghalaya Plateau: evidence of westward continuation of the Prydz Bay Pan-African suture into Northeastern India. Precambrian Research 152, 126.CrossRefGoogle Scholar
Chattopadhyay, A, Bhowmik, SK and Roy, A (2020) Tectonothermal evolution of the Central Indian Tectonic Zone and its implications for Proterozoic supercontinent assembly: the current status. Episodes Journal of International Geoscience 43, 132–44.Google Scholar
Condie, KC, Bobrow, DJ and Card, KD (1987) Geochemistry of Precambrian mafic dikes from the southern Superior Province. In Mafic dike swarms (eds Halls, HC and Fahrig, WF), pp. 95108. Geological Association of Canada Special Paper 34.Google Scholar
Desikachar, OSV (1974) A review of the tectonic and geological history of eastern India in terms of ‘plate tectonics’ theory. Journal of the Geological Society of India 33, 137–49.Google Scholar
Dhurandhar, AP, Pandey, UK and Raminaidu, C (2019) Petrochemistry and Sr, Nd, Pb isotopic characteristics of basic Dykes of Mikir Hills, Assam. Journal of the Geological Society of India 94, 559–72.CrossRefGoogle Scholar
Dobmeier, C, Lütke, S, Hammerschmidt, K, and Mezger, K (2006) Emplacement and deformation of the Vinukonda meta-granite (Eastern Ghats, India)—Implications for the geological evolution of peninsular India and for Rodinia reconstructions. Precambrian Research 146, 165178.CrossRefGoogle Scholar
Doley, D, Bhagabaty, B, Sarma, G, Singh, AK and Zou, X (2022) Geochemistry of Late Palaeoproterozoic (1.69 Ga) A-type Mayong granitoids in Shillong Plateau, north-east India: implication for anorogenic magmatism during Columbia Supercontinent cycle. Geological Journal 57, 662–80.CrossRefGoogle Scholar
Ermenco, NA, Negi, BS, Nasianov, MV, Seregin, AM, Despande, BG, Sengupta, SN, Talukdar, SN, Sastri, VV, Sokaluv, IP, Pavbukov, AT, Dutta, AK and Raju, ATR (1969) Tectonic map of India–Principles of preparation. Bulletin of ONGC 6, 1111.Google Scholar
Espanon, VR, Chivas, AR, Kinsley, LPJ and Dosseto, A (2014) Geochemical variations in the Quaternary Andean back-arc volcanism, southern Mendoza, Argentina. Lithos 208–209, 251–64.CrossRefGoogle Scholar
Evans, DAD (2013) Reconstructing pre-Pangean supercontinents. Geological Society of America Bulletin 125, 1735–51.CrossRefGoogle Scholar
Finlow-Bates, T and Stumpfl, EF (1981) The behaviour of so-called immobile elements in hydrothermally altered rocks associated with volcanogenic submarine-exhalative ore deposits. Mineralium Deposita 16, 319–28.CrossRefGoogle Scholar
Fitton, JG, James, D, Kempton, PD, Ormerod, DS and Leeman, WP (1988) The role of lithospheric mantle in the generation of late Cenozoic basic magmas in the western United States. Journal of Petrology 1, 331–49.CrossRefGoogle Scholar
Fleet, ME and Barnett, RL (1978) ivAl/viAl partitioning in calciferous amphiboles from the Frood Mine Sudbury, Ontario. Canadian Mineralogist 16, 527–32.Google Scholar
Ghose, NC and Chatterjee, N (2008) Petrology, tectonic setting and source of dykes and related magmatic bodies in the Chotanagpur Gneissic Complex, Eastern India. Indian Dykes: Geochemistry, Geophysics and Geochronology (eds Srivastava, RK, Sivaji, C and Chalapathi Rao, NV), pp. 471493. New Delhi, India: Narosa Publ. House Pvt. Ltd. Google Scholar
Ghosh, S, Bhalla, JK, Paul, DK, Sarkar, A, Bishui, PK and Gupta, SN (1991) Geochronology and geochemistry of granite plutons from East Khasi Hills, Meghalaya. Journal of the Geological Society of India 37, 331–42.Google Scholar
Ghosh, S, Fallick, AE, Paul, DK and Potts, PJ (2005) Geochemistry and origin of Neoproterozoic granitoids of Meghalaya, Northeast India: implications for linkage with amalgamation of Gondwana supercontinent. Gondwana Research 8, 421–32.CrossRefGoogle Scholar
Gogoi, A, Majumdar, D, Cottle, J and Dutta, P (2019) Geochronology and geochemistry of Mesoproterozoic porphyry granitoids in the northern Karbi Hills, NE India: implications for early tectonic evolution of the Karbi Massif. Journal of Asian Earth Sciences 79, 6579.CrossRefGoogle Scholar
Gogoi, B (2022) Late Paleoproterozoic bimodal magmatic rocks in the Nimchak Granite Pluton of the Bathani volcano-sedimentary sequence, Eastern India: implications for the Columbia supercontinent formation with respect to the Indian landmass. Periodico di Mineralogia 91, 120.Google Scholar
Gorton, MP and Schandl, ES (2000) From continents to island arcs: a geochemical index of tectonic setting for arc-related and within-plate felsic to intermediate volcanic rocks. Canadian Mineralogist 38, 1065–73.CrossRefGoogle Scholar
Grove, T, Parman, S, Bowring, S, Price, R and Baker, M (2002) The role of an H2O-rich fluid component in the generation of primitive basaltic andesites and andesites from the Mt. Shasta region, N California. Contributions to Mineralogy and Petrology 142, 375–96.CrossRefGoogle Scholar
Guo, F, Li, H, Fan, W, Li, J, Zhao, L, Huang, M and Xu, WL (2015) Early Jurassic subduction of the Paleo-Pacific Ocean in NE. China: petrologic and geochemical evidence from the Tumen mafic intrusive complex. Lithos 224, 4060.Google Scholar
Hawkesworth, CJ, Gallagher, K, Hergt, JM and McDermott, F (1993) Mantle slab contributions in arc magmas. Annual Review of Earth and Planetary Sciences 21, 175204.CrossRefGoogle Scholar
Hawkesworth, CJ, Turner, SP, McDermott, F, Peate, DW and Van Calsteren, P (1997) U-Th isotopes in arc magmas: implications for element transfer from the subducted crust. Science 276, 551–5.CrossRefGoogle ScholarPubMed
Hawthorne, FC, Oberti, R, Harlow, GE, Maresch, WV, Martin, RF, Schumacher, JC and Welch, MD (2012) Nomenclature of the amphibole group. American Mineralogist 97, 2031–48.CrossRefGoogle Scholar
Hazra, S and Ray, J (2009) Nature of mafic magmatism in ‘Khasi Greenstone’ around Laitlyngkot, East Khasi Hills, Meghalaya, northeastern India. Indian Journal of Geology 79, 4758.Google Scholar
Holland, TJB and Blundy, JD (1994) Non-ideal interactions in calcic amphiboles and their bearing on amphibole–plagioclase thermometry. Contributions to Mineralogy and Petrology 116, 433–47.CrossRefGoogle Scholar
Hollister, LS, Grissom, GC, Peters, EK, Stowell, HH and Sisson, VB (1987) Confirmation of the empirical correlation of Al in hornblende with pressure of solidification of calc-alkaline plutons. American Mineralogist 72, 231–9.Google Scholar
Hossain, I, Tsunogae, T, Rajesh, HM, Chen, B and Arakawa, Y (2007) Palaeoproterozoic U–Pb SHRIMP zircon age from basement rocks in Bangladesh: a possible remnant of the Columbia Supercontinent. Comptes Rendus Geoscience 339, 979–86.CrossRefGoogle Scholar
Hou, G, Santosh, M, Qian, X, Lister, GS and Li, J (2008) Configuration of the Late Paleoproterozoic super continent Columbia: insights from radiating mafic dyke swarms. Gondwana Research 14, 395409.CrossRefGoogle Scholar
Irvine, TN and Baragar, WRA (1971) A guide to the geochemical classification of the common volcanic rocks. Canadian Journal of Earth Sciences 8, 523–48.CrossRefGoogle Scholar
Kaur, P, Zeh, A, Chaudhri, N, Gerdes, A and Okrusch, M (2013) Nature of magmatism and sedimentation at a Columbia active margin: insights from combined U–Pb and Lu–Hf isotope data of detrital zircons from NW India. Gondwana Research 23, 1040–52.CrossRefGoogle Scholar
Kumar, A and Ahmad, T (2007) Geochemistry of mafic dykes in part of Chotanagpur gneissic complex: petrogenetic and tectonic implications. Geochemical Journal 41, 173–86.CrossRefGoogle Scholar
Kumar, S (1990) Petrochemistry and geochronology of pink granite from Sonsak, East Garo Hills, Meghalaya. Journal of Geological Society of India 35, 3945.Google Scholar
Kumar, S, Rino, V, Hayasaka, Y, Kimura, K, Raju, S, Terada, K, and Pathak, M (2017) Contribution of Columbia and Gondwana Supercontinent assembly- and growth-related magmatism in the evolution of the Meghalaya Plateau and the Mikir Hills, Northeast India: constraints from U-Pb SHRIMP zircon geochronology and geochemistry. Lithos 277, 356–75.CrossRefGoogle Scholar
Le Bas, MJ, Le Maitre, RW, Streckeisen, A and Zanettin, B (1986) A chemical classification of volcanic rocks based on the total alkalies-silica diagram. Journal of Petrology 27, 745–50.CrossRefGoogle Scholar
Liu, B, Ma, CQ, Zhang, JY, Xiong, FH, Huang, J and Jiang, HA (2014) 40Ar–39Ar age and geochemistry of subduction-related mafic dikes in northern Tibet, China: petrogenesis and tectonic implications. International Geology Review 56, 5773.CrossRefGoogle Scholar
MacLean, WH and Barrett, TJ (1993) Lithogeochemical techniques using immobile elements. Journal of Geochemical Exploration 48, 109–33.CrossRefGoogle Scholar
Majumdar, D and Gogoi, A (2020) Prospectivity of Mesoproterozoic magmatism in the northern Karbi Hills, NE India for porphyry copper mineralization. Ore Geology Reviews 120, 103467.CrossRefGoogle Scholar
Mallikharjuna Rao, J, Poornachandra Rao, GVS and Sarma, KP (2009) Precambrian mafic magmatism of Shillong Plateau, Meghalaya and their evolutionary history. Journal of the Geological Society of India 73, 143–52.CrossRefGoogle Scholar
Mazumder, SK (1986) The Precambrian framework of the Khasi Hills, Meghalaya. Record Geological Survey of India 117, 159.Google Scholar
McKenzie, D and O’Nions, RK (1991) Partial melt distributions from inversion of rare earth element. Journal of Petrology 32, 1021–91.CrossRefGoogle Scholar
Meert, JG and Santosh, M (2022) The Columbia supercontinent: retrospective, status, and a statistical assessment of paleomagnetic poles used in reconstructions. Gondwana Research 110, 143–64.CrossRefGoogle Scholar
Menzies, MA, Kyle, PR, Jones, M and Ingram, G (1991) Enriched and depleted source components for tholeiitic and alkaline lavas from Zuni-Bandera, New Mexico: inferences about intraplate processes and stratified lithosphere. Journal of Geophysical Research: Solid Earth 96, 13645–71.CrossRefGoogle Scholar
Miyashiro, A (1973) Paired and unpaired metamorphic belts. Tectonophysics 17, 241–54.CrossRefGoogle Scholar
Miyashiro, A (1994) Metamorphic Petrology. Boca Raton, FL: CRC Press.Google Scholar
Mohr, PA (1987) Crustal contamination in mafic sheets: a summary. In Mafic Dyke Swarms (eds Halls, HC and Fahrig, WC), pp. 7580. St John’s: Geological Association of Canada, Special Publication no. 34.Google Scholar
Nachit, H, Ibhi, A and Ohoud, MB (2005) Discrimination between primary magmatic biotites, reequilibrated biotites and neoformed biotites. Comptes Rendus Geoscience 337, 1415–20.CrossRefGoogle Scholar
Nandy, DR (2001) Geodynamics of northeastern India and the adjoining region. ACB Publications, 111–130.Google Scholar
Ormerod, DS, Hawkesworth, CJ, Leeman, WP and Menzies, MA (1988) The identification of subduction-related and within-plate components in basalts from the western U.S.A. Nature 333, 349–53.CrossRefGoogle Scholar
Pandey, BK, Gupta, JN and Lall, Y (1986) Whole-rock and mineral Rb–Sr isochron ages for the granites of Bihar Mica Belt of Hazaribagh, Bihar, India. Indian Journal of Earth Sciences 13, 157–62.Google Scholar
Peacock, SM, Rushmer, T and Thompson, AB (1994) Partial melting of subducting oceanic crust. Earth and planetary science letters 121, 227–44.CrossRefGoogle Scholar
Pearce, JA (1996) A users guide to basalt discrimination diagrams. In Trace Element Geochemistry of Volcanic Rocks: Applications for Massive Sulphide Exploration (ed. Wyman, DA), pp. 79113. St. John’s, Canada: Geological Association of Canada, Short Course Notes no. 12.Google Scholar
Pearce, JA (2008) Geochemical fingerprinting of oceanic basalts with applications to the ophiolites classification and the search for Archean oceanic crust. Lithos 100, 1448.CrossRefGoogle Scholar
Pearce, JA and Cann, JR (1973) Tectonic setting of basic volcanic rocks determined using trace element analyses. Earth and Planetary Science Letters 19, 290300.CrossRefGoogle Scholar
Pearce, JA and Norry, MJ (1979) Petrogenetic implications of Ti, Zr, Y, and Nb variations in volcanic rocks. Contributions to Mineralogy and Petrology 69, 3347.CrossRefGoogle Scholar
Pearce, JA and Peate, DW (1995) Tectonic implications of the composition of volcanic arc magmas. Annual Review of Earth and Planetary Sciences 23, 281–5.CrossRefGoogle Scholar
Pearce, JA and Stern, RJ (2006) Origin of back-arc basin magmas: trace element and isotope perspectives. In Back-Arc Spreading Systems; Geological, Biological, Chemical, and Physical Interactions (eds Christie, DM, Fisher, CR, Lee, SM and Givens, S), pp. 6386. Geophysical Monograph Series 166. Washington, DC: American Geophysical Union.CrossRefGoogle Scholar
Pe-piper, G, Matarangas, D, Reynolds, PH and Chatterjee, AK (2003) Shoshonites from Agios Nectarios, Lesbos, Greece: origin by mixing of felsic and mafic magma. European Journal of Mineralogy 15, 77–8.CrossRefGoogle Scholar
Pesonen, LJ, Mertanen, S and Veikkolainen, T (2012) Paleo-Mesoproterozoic supercontinents – a paleomagnetic view. Geophysica 48, 547.Google Scholar
Pisarevsky, SA, Biswal, TK, Wang, XC, DeWaele, B, Ernst, R, Södurlund, U, Tait, JA, Ratre, K, Singh, YK and Cleve, M (2013) Palaeomagnetic, geochronological and geochemical study of Mesoproterozoic Lakhna Dykes in the Bastar Craton, India: implications for the Mesoproterozoic supercontinent. Lithos 174, 125–43.CrossRefGoogle Scholar
Pouchou, JL and Pichoir, F (1987) Basic expression of “PAP” computation for quantitative EPMA. In X-ray Optics and Microanalysis (eds Brown, JD and Packwood, RH), pp. 249253. Ontario: University of Western Ontario.Google Scholar
Ray Barman, T and Bishui, PK (1994) Dating of Chotanagpur gneissic complex of eastern Indian Precambrian shield. Record Geological Survey of India 127, 2527.Google Scholar
Ray Barman, T, Bishui, PK and Sarkar, A (1990) Dating of Early Precambrian Granite-Greenstone Complex of the eastern Indian Precambrian shield with special references to Chotanagpur Granite Gneiss Complex. Records of the Geological Survey of India 123, 25–7.Google Scholar
Ray, J, Saha, A, Koeberl, C, Thoni, M, Ganguly, S and Hazra, S (2013) Geochemistry and petrogenesis of Proterozoic mafic rocks from East Khasi Hills, Shillong Plateau, Northeastern India. Precambrian Research 230, 119–37.CrossRefGoogle Scholar
Redman, BA and Keays, RR (1985) Archaean basic volcanism in the Eastern Goldfields Province, Yilgarn Block, Western Australia. Precambrian Research 30, 113–52.CrossRefGoogle Scholar
Rekha, S, Upadhyay, D, Bhattacharya, A, Kooijman, E, Goon, S, Mahato, S and Pant, NC (2011) Lithostructural and chronological constraints for tectonic restoration of Proterozoic accretion in the Eastern Indian Precambrian shield. Precambrian Research 187, 313–33.CrossRefGoogle Scholar
Rollinson, HR (1993) Using Geochemical Data: Evaluation, Presentation, Interpretation. New Jersey: Pearson Prentice Hall, 352.Google Scholar
Saha, D, Bhowmik, SK, Bose, S and Sajeev, K (2016) Proterozoic tectonics and trans-Indian mobile belts: a status report. Proceedings of the Indian National Science Academy 82, 445–60.CrossRefGoogle Scholar
Saikia, A, Gogoi, B, Ahmad, M, Kumar, R, Kaulina, T and Bayanova, T (2019) Mineral chemistry, Sr-Nd isotope geochemistry and petrogenesis of the granites of Bathani volcano-sedimentary sequence from the northern fringe of Chotanagpur Granite Gneiss Complex of eastern India. Bern: Springer Nature, Society of Earth Scientists, pp. 79120.Google Scholar
Saikia, A, Gogoi, B, Kaulina, T, Lialina, L, Bayanova, T and Ahmad, M (2017) Geochemical and U–Pb zircon age characterization of granites of the Bathani volcano-sedimentary sequence, Chotanagpur granite gneiss complex, Eastern India: vestiges of the Nuna supercontinent in the Central Indian Tectonic Zone. In Crustal Evolution of India and Antarctica: The Supercontinent Connection. Geological Society of London, Special Publication 457, 233–252.CrossRefGoogle Scholar
Saini, NK, Mukherjee, PK, Rathi, MS and Khanna, PP (2000) Evaluation of energy-dispersive x-ray fluorescence spectrometry in the rapid analysis of silicate rocks using pressed powder pellets. X-Ray Spectrometry 29, 166–72.3.0.CO;2-L>CrossRefGoogle Scholar
Saini, NK, Mukherjee, PK, Rathi, MS, Khanna, PP and Purohit, KK (2007) A proposed amphibolite reference rock sample (AMH) from Himachal Pradesh. Journal of Geological Society of India 69, 799802.Google Scholar
Sarkar, A, Boda, MS, Kundu, HK, Mamgain, VV and Shankar, R (1998) Geochronology and geochemistry of Mesoproterozoic intrusive plutonites from the eastern segment of the Mahakoshal greenstone belt, Central India. In: International Seminar on Precambrian Crust in Eastern and Central India, UNESCOIUGS- IGCP-368, Abstract Volume. Geological Survey of India, Kolkata, 82–85.Google Scholar
Saunders, AD, Tarney, J and Weaver, SD (1980) Transverse geochemical variations across the Antarctic Peninsula: implications for the genesis of calc-alkaline magmas. Earth and Planetary Science Letters 46, 344–60.CrossRefGoogle Scholar
Schmidt, MW (1992) Amphibole composition in tonalite as a function of pressure: an experimental calibration of the Al-in-hornblende barometer. Contributions to Mineralogy and Petrology 110, 304–10.CrossRefGoogle Scholar
Sequeira, N, Bhattacharya, A and Bell, E (2022) The∼ 1.4 Ga A-type granitoids in the “Chottanagpur crustal block”(India), and its relocation from Columbia to Rodinia? Geoscience Frontiers 13, 101138.CrossRefGoogle Scholar
Sheraton, JW, Black, LP, McCulloch, MT and Oliver, RL (1990) Age and origin of a compositionally varied mafic dyke swarm in the Bunger Hills, East Antarctica. Chemical Geology 85, 215–46.CrossRefGoogle Scholar
Smith, EI, Sánchez, A, Walker, JD and Wang, K (1999) Geochemistry of mafic magmas in the Hurricane volcanic field, Utah: implications for small- and large-scale chemical variability of the lithospheric mantle. The Journal of Geology 107, 433–48.CrossRefGoogle Scholar
Spandler, C and Pirard, C (2013) Element recycling from subducting slabs to arc crust: a review. Lithos 170–171, 208–23.CrossRefGoogle Scholar
Srinivasan, P, Sen, S and Bandopadhyay, PC (1996) Study of variation of Palaeocene Eocene sediments in the shield areas of Shillong Plateau. Records of Geological Survey of India 129, 77–8.Google Scholar
Srivastava, RK (2012) Petrological and geochemical studies of paleoproterozoic mafic dykes from the Chitrangi Region, Mahakoshal Supracrustal Belt, Central Indian Tectonic Zone: petrogenetic and tectonic significance. Journal of the Geological Society of India 80, 369–81.CrossRefGoogle Scholar
Srivastava, RK and Ahmad, T (2008) Precambrian mafic magmatism in the Indian shield: an introduction. Journal of the Geological Society of India 72, 914.Google Scholar
Srivastava, RK and Sinha, AK (2004) Geochemistry and Petrogenesis of early Cretaceous sub alkaline mafic dykes from Swangkre–Rongmil, East Garo Hills, Shillong Plateau, NE India. Journal of Earth System Science 113, 683–97.CrossRefGoogle Scholar
Stern, CR and Kilian, R (1996) Role of the subducted slab, mantle wedge and continental crust in the generation of adakites from the Andean Austral Volcanic Zone. Contributions to Mineralogy and Petrology 123, 263–81.CrossRefGoogle Scholar
Stolper, E and Newman, S (1994) The role of water in the petrogenesis of Mariana trough magmas. Earth and Planetary Science Letters 121, 293325.CrossRefGoogle Scholar
Su, YP, Zheng, JP, Griffin, WL, Zhao, JH, Tang, HY, Ma, Q and Lin, XY (2012) Geochemistry and geochronology of Carboniferous volcanic rocks in the eastern Junggar terrane, NW China: implication for a tectonic transition. Gondwana Research 22, 1009–29.CrossRefGoogle Scholar
Sun, SS and McDonough, WF (1989) Chemical and isotopic systematics of oceanic basalts: implications for mantle composition and processes. Geological Society, London, Special Publications 42, 313–45.CrossRefGoogle Scholar
Thomas, WM and Ernst, WG (1990) The aluminum content of hornblende in calc-alkaline granitic rocks: a mineralogic barometer calibrated experimentally to 12 kbars. In Fluid–mineral interactions: a tribute to HP Eugster (eds RJ Spencer and IM Chou), pp. 59–63. Geochemical society of London Special Publication 2.Google Scholar
Thompson, RN and Morrison, MA (1998) Asthenospheric and lower-lithospheric mantle contributions to continental extensional magmatism: an example from the British Tertiary Province. Chemical Geology 68, 115.CrossRefGoogle Scholar
Tischendorf, G, Gottesmann, B, Förster, HJ and Trumbull, RB (1997) On Li-bearing micas: estimating Li from electron microprobe analyses and improved diagram for graphical representation. Mineralogical Magazine 61, 809–34.CrossRefGoogle Scholar
Turner, S, Hawkesworth, C, Rogers, N and King, P (1997) U-Th isotope disequilibria and ocean island basalt generation in the Azores. Chemical Geology 139, 145–64.CrossRefGoogle Scholar
Wang, C, He, X, Carranza, EJM and Cui, C (2019) Paleoproterozoic volcanic rocks in the southern margin of the North China Craton, central China: implications for the Columbia supercontinent. Geoscience Frontiers 10, 1543–60.CrossRefGoogle Scholar
Watts, AB and Cox, KG (1989) The Deccan Traps: an interpretation on terms of progressive lithospheric flexure in response to a migrating lead. Earth and Planetary Science Letters 93, 8597.CrossRefGoogle Scholar
Wilson, M and Davidson, JP (1984) The relative roles of crust and upper mantle in the generation of oceanic island-arc magmas. Philosophical Transactions of Royal Society of London A 310, 661–74.Google Scholar
Woodhead, J, Eggins, S and Gamble, J (1993) High-field strength and transition element systematics in island arc and back-arc basin basalts: evidence for multiphase melt extraction and a depleted mantle wedge. Earth and Planetary Science Letters 114, 491504.CrossRefGoogle Scholar
Wyllie, PJ (1988) Magma genesis, plate tectonics, and chemical differentiation of the Earth. Reviews of Geophysics 26, 370404.CrossRefGoogle Scholar
Yadav, B, Ahmad, T, Kaulina, T and Bayanova, T (2015) Geochemistry and petrogenesis of Paleo-Proterozoic granitoids from Mahakoshal Supracrustal Belt (MSB), CITZ. EGU General Assembly Conference Abstracts 17, 5256.Google Scholar
Yin, A, Dubey, CS, Webb, AAG, Kelty, TK, Grove, M, Gehrels, GE and Burgess, WP (2010) Geologic correlation of the Himalayan orogen and Indian craton: part 1. Structural geology, U-Pb zircon geochronology, and tectonic evolution of the Shillong Plateau and its neighboring regions in NE India. GSA Bulletin 122, 336–59.CrossRefGoogle Scholar
Zhang, S, Li, ZX, Evans, DAD, Wu, H, Li, H and Dong, J (2012) Pre-Rodinia supercontinent Nuna shaping up: a global synthesis with new paleomagnetic results from North China. Earth and Planetary Science Letters 353–354, 145–55.CrossRefGoogle Scholar
Zhao, G, Cawood, PA, Wilde, SA and Sun, M (2002) Review of global 2.1-1.8 Ga orogens: implications for a pre-Rodinia supercontinent. Earth-Science Reviews 59, 125–62.CrossRefGoogle Scholar
Zhao, JX, Shiraishi, K, Ellis, DJ and Sheraton, JW (1995) Geochemical and isotopic studies of syenites from the Yamato Mountains, East Antarctica: implications for the origin of syenitic magmas. Geochimica et Cosmochimica Acta 59, 1363–82.CrossRefGoogle Scholar
Figure 0

Figure 1. Tectonic map of India showing the location of Mikir Massif (MM) along with the Archean cratons and Proterozoic mobile belts (modified after Chatterjee & Ghose, 2011). The study area is shown within the black square. Abbreviations: ADMB, Aravalli Delhi Mobile Belt; AMGC, Assam-Meghalaya Gneissic Complex; BC, Bastar Craton; BuC, Bundelkhand Craton; BVSs, Bathani volcano-sedimentary sequence; CGGC, Chotanagpur Granite Gneiss Complex; CITZ, Central Indian Tectonic Zone; DC, Dharwar Craton; EGB, Eastern Ghats Belt; KL, Kopili Lineament; MMB, Mahakoshal Mobile Belt, MM, Mikir Massif; NSMB: North Singhbhum Mobile Belt; SB, Satpura Belt; SC, Singhbhum Craton; SONA, Son-Narmada lineaments.

Figure 1

Figure 2. Regional geological map of Mikir Massif showing the study area within the black box (modified after Dhurandhar et al. 2019).

Figure 2

Table 1. A generalized stratigraphic succession of Mikir Massif (after Awati et al. 1995; Majumdar & Gogoi, 2020)

Figure 3

Figure 3. Field photographs showing (a) a metadolerite hand specimen. (b–d) Quartzites of the Shillong Group share sharp contacts with the metadolerites or Khasi greenstones. (e, f) Closely spaced vertical joints crisscrossed by another set of joints at a higher angle (shown by yellow lines) and a set of conjugate joints (shown by white lines) were observed in metadolerites.

Figure 4

Figure 4. Representative photomicrographs of the metadolerites illustrating the mineral assemblages (a) plane-polarized light (PPL) images of secondary amphibole, (b, c) amphibole, biotite and Fe-oxide, (d) amphibole and Fe-oxide and (e, f) cross-polarized light (CPL) images of plagioclase laths embedded within secondary amphibole. Mineral abbreviations: Am, amphibole; Bt, biotite; Pl, plagioclase.

Figure 5

Table 2. Major oxides (wt %), trace and rare earth elements (ppm) composition of the metadolerites of Borjuri from Mikir Massif part of Assam-Meghalaya Gneissic Complex

Figure 6

Figure 5. (a) Geochemical classification diagram based on immobile trace elements (after Pearce, 1996) and (b) AFM diagram displaying tholeiitic nature of magma for the metadolerites of Borjuri (after Irvine & Baragar, 1971). Data used for comparison are from Proterozoic mafic rocks of Bathani volcano-sedimentary sequence: Ahmad et al. (2021) and Gogoi (2022).

Figure 7

Figure 6. (a) Primitive mantle-normalized multi-element diagram (normalizing values after Sun & McDonough, 1989) and (b) chondrite-normalized rare earth element patterns (normalizing values after Boyton, 1984) for the metadolerite samples of Borjuri. For comparison purpose, trace element data in yellow and green patterns are plotted from Ahmad et al. (2021) and Gogoi (2022) respectively.

Figure 8

Figure 7. Bivariate diagrams of (a–h) major elements versus SiO2 and (i–l) selected trace elements versus SiO2.

Figure 9

Figure 8. (a) Ti/Y versus Ba/Nb, (b) Ti/Y versus Ba/Zr, (c) Nb/La versus SiO2 and (d) Nb/La versus Nb/Th variation diagrams of the mafic rocks of Borjuri.

Figure 10

Figure 9. (a) La/Yb versus Nb/La (Smith et al. 1999) plot showing lithospheric and mixed lithospheric mantle source for Borjuri mafics. (b) Th/Sm versus Th/Ce plot of the Borjuri mafics shows the input of subducted sediment. (c) Nb/Y versus Rb/Y and (d) Nb/Zr versus Th/Zr, indicating fluid-related enrichment in Borjuri mafics.

Figure 11

Figure 10. Plots for the metadolerites of Borjuri (a) Ni versus Zr petrogenetic model for a lherzolite mantle composition. Curve 1 signifies batch melting at 1500 °C (1 atm. equivalent) with percentages along the line representing degree of partial melting. Curves 2, 3 and 4 demonstrate olivine fractionation with elimination of olivine marked in 5% increments (after Condie et al. 1987). (b) La/Yb versus Sm/Yb diagram where the Borjuri metadolerites cluster near the spinel+garnet lherzolite curve (after Liu et al. 2014). Abbreviation: PM, primitive mantle.

Figure 12

Figure 11. Tectonic discrimination diagrams for the Borjuri metadolerites. (a) Nb/Yb versus Th/Yb plot after Pearce (2008). (b) Ti/100-Zr-Yx3 ternary diagram after Pearce & Cann (1973). (c) Ti/Zr versus Zr after Woodhead et al. (1993). (d) Zr versus V/Ti after Woodhead et al. (1993). Field of Andean back-arc basalt is taken from Espanon et al. (2014). Abbreviations: BABB, Back-arc basin basalt; CAB, Calc-alkali basalts; IAB, Island-arc basalt; IAT, Island-arc tholeiites; MORB, Mid-oceanic ridge basalt; WPB, Within plate basalt. Symbols as in Figure 5.

Figure 13

Figure 12. Map of Indian subcontinent displaying the ENE-WSW-trending Central Indian Tectonic Zone (CITZ), Chotanagpur Granite Gneiss Complex (CGGC) and Assam-Meghalaya Gneissic Complex (AMGC) (modified after Saikia et al. 2017). This continuous mobile belt demarcates the boundary between the North Indian Block (NIB) from the South Indian Block (SIB) (Acharyya 2003; Bhandari et al. 2011; Chatterjee & Ghose, 2011; Rekha et al. 2011; Bhowmik et al. 2012). The NIB comprises of the Bundelkhand Craton (BuC), whereas the SIB is a union of the Dharwar Craton (DC), the Singhbhum Craton (SC) and the Bastar Craton (BC). The Aravalli Delhi Mobile Belt (ADMB) and the Eastern Ghats Belt (EGB) are also shown. Two volcano-sedimentary sequences of CITZ, namely Mahakoshal Mobile Belt (MMB) and its eastern continuation Bathani volcano-sedimentary sequence (BVSs), are shown. The present study projects the Shillong Basin of AMGC as the eastern continuation of the BVSs. The ages reported from the MMB are taken from: 1) Sarkar et al. (1998); 2) Bora et al. (2013); 3) Yadav et al. (2015). The ages reported from the BVSs are taken from: 1) Chatterjee & Ghose (2011); 2) Saikia et al. (2017, 2019). The ages reported from the Shillong Basin are taken from: 1) Gogoi et al. (2019). The ages reported from the CITZ are taken from: 1) Bhandari et al. (2011); 2) Bhowmik et al. (2011). The ages reported from the CGGC are taken from: 1) Pandey et al. (1986); 2) Chatterjee et al. (2008). The ages reported from the AMGC are taken from: 1) Chatterjee et al. (2007); 2) Yin et al. (2010).

Supplementary material: File

Basumatary et al. supplementary material

Figures S1-S6 and Tables S1-S4

Download Basumatary et al. supplementary material(File)
File 956.6 KB