1 Introduction

Electron micro probe analyser (EPMA) monazite dating has been extensively used for metamorphic and igneous rocks from the last few decades (Suzuki and Adachi 1991, 1994; Montel et al. 1994, 1996; Cocherie et al. 1997). An electron microprobe of thorium (Th)–uranium (U)–lead (Pb) monazite dating principal was applied to obtain the age of granulite rocks from Daltonganj. Monazite mineral is a phosphate of light weight rare-earth element [(LREE) PO4], with abundant U, Th and little Pb (Parrish 1990). The rapid accumulation of radiogenic lead (*Pb), to a required level, is possible, which can be analysed with an electron probe (Montel et al. 1996). Monazite dating with the help of an electron microprobe has been performed by several researchers based on the abundance of Th, U and Pb (Suzuki and Adachi 1991, 1994; Montel et al. 1996; Braun et al. 1998; Williams et al. 1999). The EPMA-monazite dating technique is a reliable technique to identify the recorded history of polymetamorphic events (Rosa-Costa et al. 2008; Karmakar et al. 2011; Prabhakar 2013; Bhowmik et al. 2014). Petrogenetic considerations and the textural relations of mineral phases in the granulite of the Chhotanagpur granite–gneiss complex (CGGC) suggest that M1 (1700–1600 Ma) and M4 (1000–900 Ma) are the prograde metamorphic events, which are separated by M2 and M3 (1100–1400 Ma) retrograde metamorphic events (Maji et al. 2008), and they are of younger age (830–600 Ma), the cause of which remains unexplored. Furthermore, Sanyal and Sengupta (2012) documented the four metamorphic stages (M1–M4), where the M1 metamorphic stage occurred around 1870 Ma and successively followed by the M2 metamorphic phase between 1660 and 1270 Ma, the M3 phase was recorded between 1200 and 930 Ma, followed by an M4 event at 870–780 Ma. The detail of the four metamorphic stages along with their age and technique of dating by various authors in CGGC is given in table 1.

Table 1 Four metamorphic (M1–M4) stages with geochronology from different localities of the CGGC (modified after Sanyal and Sengupta 2012).

Chatterjee et al. (2010) correlated the Precambrian eastern Indian terrain with the Sausar mobile belt in central India and the Shillong–Meghalaya granulite belt from the north-eastern part of India which can be used as an essential tool for the India–Australia–Antarctica correlation. The breakdown of the Columbia supercontinent was initiated during the Mesoproterozoic era which leads to the reconstruction of the Rodinia supercontinent between 1200 and 900 Ma (Dalziel 1991; Hoffman 1991; Moores 1991; Li et al. 2008; Bhowmik et al. 2010). Greater India is shown to have been in contact with the Antarctica–Australia Peri–Rodinian margin (Li et al. 2008) supported by the petrological and geochronological setting of the Eastern Ghats mobile belt (EGMB) of India and the Rayner province of Antarctica (Dalziel 1991; Yoshida et al. 1992; Shaw et al. 1997; Mezger and Cosca 1999; Dasgupta and Sengupta 2003). 1000–900 Ma age was preserved in older high-pressure metamorphic rocks of CGGC which is assumed to be the extension of central Indian tectonic zone (CITZ), and act as a suture zone when the SIB, NIB and MC were joined together during the Stenian–Tonian orogenesis (Bhowmik et al. 2010). Polymetamorphic events were preserved in the SIB, CITZ and CGGC with similarity with respect to petrological and geochronological properties (Chatterjee et al. 2008; Maji et al. 2008; Sanyal and Sengupta 2012). This study deals with the electron microprobe monazite dating of the garnet–hypersthene–gedrite–cordierite gneiss and the garnet–gedrite–cordierite–biotite gneiss from Daltonganj (Palamau) in the CGGC. We have documented the evidence of the tectono-metamorphic event from the north-western part of CGGC with the help of petrological, geothermobarometry and EPMA-monazite dating techniques. We have also striven to unravel the geodynamic implications of metamorphism in the granulite facies rocks of CGGC and their correlation to global-scale Grenvillian orogenesis suturing in the peninsular part of India with the Rodinia assembly.

2 Geological background

The area under investigation (latitude 23°54′50″–23°58′30″N; longitude 84°2′–84°06′30″E) belongs to the western part of CGGC. It covers a vast area of about 100,000 km2 and extends in the east–west from the provinces of Jharkhand, Orissa, Chhattisgarh, Madhya Pradesh and West Bengal and the south-eastern part of Uttar Pradesh. The CGGC includes multiple generations of mafic intrusives, namely meta-dolerite and norites to gabbros in which the corona texture is often demarcated at several places, especially in Purulia, Dumka and Daltonganj. This gneissic complex shows composite character consisting mainly of granitoid gneisses, migmatites and massive granites with enclaves of metasedimentary and meta-igneous rocks and intrusive basic and intermediate rocks (Ghose 1983, 1992; Chatterjee and Ghose 2011; Sanyal and Sengupta 2012; Yadav et al. 2016; Dwivedi et al. 2019 and references therein). The medium- to high-grade metamorphic terrain of CGGC mainly contains amphibolite to granulite facies rocks, which is sandwiched between the medium- to low-grade mobile belts. Here, the first one is located in the southern margin of CGGC and known as the North Singhbhum mobile belt, and it consists of the volcanic intrusive as well as sedimentary sequences (Saha 1994). The other mobile belt is located on the northern margin, and it extends from the north-western part of CGGC, which is called the Mahakoshal mobile belt. It preserves the metasediments, granitoids and mafic–ultramafic rocks (Roy and Devarajan 2000). The CGGC has a broad history of two high-grade metamorphic events during the Mesoproterozoic and Grenvillian ages, and both these events recorded in the entire metamorphosed rock lie within CGGC (Pandey et al. 1986a, b; Ray Barman and Bishui 1994; Chatterjee et al. 2008; Maji et al. 2008; Sanyal and Sengupta 2012; Mukherjee et al. 2017). The north-western part of CGGC (Daltonganj) consists of granulite facies rocks, charnockite and migmatitic–tonalite–granodiorite–granite gneisses (Rode 1948; Dwivedi et al. 2019). The porphyritic granitic magmatism (~1660 ± 17 Ma) was reported from the western as well as the northern part of CGGC (Chatterjee and Ghose 2011; Saikia et al. 2017). The NE part of the CGGC is dominated by charnockitic gneisses as country rocks, and it was metamorphosed to amphibole–biotite gneiss. The 1447 ± 11 Ma age was obtained by U–Pb zircon dating from the protolith of the charnockitic gneisses and further high-grade metamorphism was recorded at 943 Ma from the migmatitic charnockitic gneiss under pressure–temperature (PT) condition ~9 kbar and 780–800°C (Mukherjee et al. 2017, 2018). Magmatic events are recorded from all parts of the CGGC during the Neoproterozoic era, which includes (a) partial melting and intrusive grey granite at 1005 ± 51 Ma and pink granite intrusion at 815 ± 47 Ma (Singh and Krishna 2009) and (b) granite intrusion emplaced at 975 Ma from Daltonganj of the western CGGC (Chatterjee and Ghose 2011). The geological map (figure 1a) represents the study area (Daltonganj) with geochronological age distribution at different locations within the CGGC, and the enlarged geological map of the study area is shown in figure 1(b).

Figure 1
figure 1

(a) Geological map of the CGGC (modified after Acharyya 2003; Maji et al. 2008; Chatterjee and Ghose 2011). Abbreviations: ADMB: Aravalli–Delhi mobile belt, BC: Bastar craton, BuC: Bundelkhand craton, CGGC: Chhotanagpur granite gneiss complex, CITZ: Central Indian tectonic zone, D: Dudhi, DM: Dumka, EGB: Eastern Ghats belt, KC: Karnataka craton, MGB: Makrohar granulite belt, NSMB: North Singhbhum mobile belt, R: Rihand–Renusagar area, SC: Singhbhum craton and SMGC: Shillong–Meghalaya gneissic complex. Here, the distribution of different geochronological (Grenvillian) ages in different locations, with our analysed monazite dating age is 1424–972 Ma from the Daltonganj (CGGC) India. (b) Enlarged geological map of the area around the south-west of Daltonganj, Palamau (Jharkhand) district, India.

3 Analytical techniques

The analytical work was performed using an EPMA on a CAMECA SX five instrument at the DST-SERB National Facility, Department of Geology (CAS), Institute of Science, BHU. The thin polished section was coated with a 20 nm thin layer of carbon for electron probe microanalyses using a LEICA-EM ACE200 carbon coater instrument. The EPMA instrument CAMECA SX Five was operated with SX Five software at an accelerated voltage of 15 kV and a current of 200 nA with a LaB6 source in the electron gun for electron beam generation, which is based on a new analytical protocol for the U–Th–Pb chemical dating of monazite (Pandey et al. 2019). Andradite is used as a natural silicate mineral to verify crystal positions by using an internal standard (SP2-LiF, SP3-LPET, SP4-LTAP and SP5-PET) with suitable wavelength dispersive spectrometers (SP#) using the CAMECA SX-Five instrument. The following X-ray lines were used in the analyses: F–Kα, Na–Kα, Mg–Kα, Al–Kα, Si–Kα, P–Kα, Cl–Kα, K–Kα, Ca–Kα, Ti–Kα, Cr–Kα, Mn–Kα, Fe–Kα, Ni–Kα and Ba–Lα. Natural mineral standards: fluorite, halite, periclase, corundum, wollastonite, apatite, orthoclase, rutile, chromite, rhodonite, hematite and barite; Ni pure metal standard was supplied by CAMECA-AMETEK which was used for routine calibration and quantification. Quantification of rare-earth element (REE) analysis in monazite mineral phases and U, Th and Y elemental X-ray mapping of monazite grains were obtained at an accelerating voltage of 20 kV, and currents of a beam are 200 nA, at 0.5 μm/pixel spatial resolution. The following X-ray lines were used in the analyses: Y–Lα, La–Lα, Ce–Lα, Pr–Lα, Nd–Lα, Sm–Lα, Eu–Lα, Th–Mα and U–Mα. All REE analysis was carried out on a LiF crystal attached with SP2 and Pb, Th and U were analysed with an LPET crystal connected with the SP3 spectrometer in a CAMECA-SX five EPMA instrument. Synthetic glass standards of all REE (La to U) supplied by CAMECA-AMETEK were used for routine calibration and quantification.

Scanning electron microscope (SEM) analysis was performed at the DST-SERB National Facility, Department of Geology (CAS), Institute of Science, BHU. The SEM instrument was operated at an accelerated voltage of 15 kV and a current of 200 nA.

4 Petrography and mineral chemistry

Different samples were collected from Datam and Mahawat Muria localities, which lie in the south-west of Daltonganj within CGGC (figure 1b). In this study, two rock samples (R-91-97 and R-91-96) were selected for the analytical purpose after the detailed petrographic study from collected rock samples. Representative mineral compositional data are presented in tables 26. The mineral abbreviations used in this study are after Whitney and Evans (2010).

Table 2 Representative electron microprobe analyses and structural formula of garnet and hypersthene.

4.1 Megascopic characters

The R-91-97 contains garnet–hypersthene–gedrite–cordierite gneisses, and R-91-96 consists of garnet–gedrite–cordierite–biotite gneisses, which are medium- to coarse-grained and exhibit gneissose texture with a resinous and greasy appearance. The garnet–gedrite–cordierite–biotite gneiss consists of garnet as one of their dominant minerals displaying light pinkish colour with the dark-coloured gedrite along with biotite flakes (figure 2a). Garnet–hypersthene–gedrite–cordierite gneiss contains the porphyroblastic garnet with medium- to small-size grains of hypersthene, gedrite, cordierite, biotite and quartz (figure 2b). The presence of garnet has displayed pinkish to reddish tinge to the rock. On the weathered surface of such gneisses, the nodules of garnet are generally seen on the rock surface in the hand-specimen. The detail petrography and reaction texture of these rocks were discussed by Dwivedi et al. (2019).

Figure 2
figure 2

(a) Field photograph of the garnet–gedrite–cordierite–biotite gneiss; (b) representative hand-specimen of the garnet–orthopyroxene–gedrite–cordierite gneiss; (c) photomicrographs showing the gedrite rimmed by garnet and cordierite, where orthopyroxene coexists with garnet; (d) photomicrographs showing the small grain of brown biotite present as inclusion in gedrite and gedrites are surrounded by a huge mass of garnet and cordierite; (e) BSE image shows some accessory minerals, i.e., monazite and ilmenite with other minerals like gedrite, biotite and quartz.

4.2 Identification of monazite

Monazite occurs as an accessory phase in the rocks of the study area. In the context of the identification of monazite grain under a petrological microscope, zircon creates some confusion with monazite. Monazite and zircon have some vital diagnostic features, and they could be distinguished by their characteristic. The zircon grains have distinct prismatic as well as euhedral shape, but monazite grains are rarely rounded and anhedral in shape (Schaltegger et al. 1999). The pleochroic haloes formed by zircons were small, but monazite produced comparatively larger pleochroic haloes in flakes of biotite and cordierite, due to an abundance of Th, i.e., 3.14–7.20 wt%. However, solely on optical properties, monazite and zircon grains cannot be easily distinguished. The back-scattered electron (BSE) images were used to identify the monazite grains from the matrix and garnet.

4.3 Textural interpretations of monazite

The analyzed monazite grains occur as inclusion within porphyroblastic garnet, cordierite biotite and matrix (figure 3). Garnet shows the compositional variation from the core to the rim in which the core (Py29.5Alm69.5Grs0.7Sps0.3) contains comparatively lower Alm and Sps and slightly higher in pyrope content in comparison with the rim (Py23.3Alm75.4Grs0.6Sps0.7) (figure 4). Monazite grain in the core of garnet (figure 3a) produced the older age in R-91-97, and many of the monazite grains embedded in the periphery area (figure 3b) generates the younger age in R-91-96. Garnet is rimmed by cordierite and shows the corona structure (figure 3a) which includes biotite and quartz as inclusion that leads to the formation of cordierite. Monazite grains are also present in the cordierite, biotite (figure 3c and d), gedrite and matrix which provide a younger age.

Figure 3
figure 3

BSE images showing the microstructural and textural settings of monazite occurrences in the granulitic gneiss of Daltonganj. (a) Monazite occurring as inclusion within porphyroblastic garnet in R-91-97. (b) Monazite grain occurring as inclusion within the periphery area of garnet in R-91-97. (c) Monazite present as inclusion in the cordierite, Crd are later surrounded by garnet in R-91-96. (d) Monazite occurring as inclusion within biotite in R-91-96.

Figure 4
figure 4

(a) BSE image of garnet porphyroblast with inclusions of biotite and quartz; (be) these images represent the X-ray mapping of Fe, Mg, Mn and Ca in garnet porphyroblast; (f) XAlm, XPy, XGrs and XSps variation along the garnet porphyroblast from rim to rim.

4.4 Microscopic characters and mineral chemistry

Garnet occurs as coarse xenoblast and poikiloblast and contains gedrite, cordierite, biotite and quartz as an inclusion with some other heavy minerals i.e., monazite, ilmenite, magnetite, zircon, etc. Garnet grains are highly fractured and partially rimmed by gedrite and cordierite. The small flake of biotite occurs as inclusion along with quartz within the garnet (figure 2c). Garnet shows solid solution dominantly between almandine (65.4–74.3 mol%) and pyrope (24.3–32.8 mol%) with minor amounts of grossular (0.7–1.2 mol%) and spessartine (0.6–0.7 mol%). The XMg of garnet lies between 0.25 and 0.27 (table 2).

The BSE image shows the inclusion of biotite and quartz in garnet (figure 4a), and the elemental X-ray map of garnet reveals the enrichment of Fe and Mg elements (figure 4b and c) and depletion of Mn and Ca elements (figure 4d and e). Garnet shows the dominance of the almandine in which XAlm varies from 0.68 to 0.77. The length of the porphyroblastic garnet is 890 μm, and the rim–core–rim distribution of almandine and pyrope garnet is graphically represented (figure 4f). Here almandine shows a high-composition peak at the rim and a lower one at the core area. Pyrope represents a higher occurrence at the core and lower occurrence at the rim portion (XPy = 0.22–0.34). The Fe content of garnet increases and the Mg content decreases at the rim of the garnet porphyroblast due to locally resorbed by gedrite + cordierite minerals or by retrogressive biotite. This situation indicates the lowering of the pressure and the temperature at the rim of garnet porphyroblast as compared to the core portion.

Hypersthene is idioblastic to xenoblastic in nature and varies from medium to coarse (figure 2c). It shows strong pleochroism from pink to bluish-green in colour. Hypersthene is partially rimmed by garnet and gedrite. The XMg of hypersthene ranges between 0.52 and 0.58 (table 2).

Xenoblast and coarse aggregates of cordierite wrap the garnet and gedrite (figure 2d). Corroded cordierite is completely rimmed by garnet which provides evidence of the prograde metamorphic condition. Cordierite shows some alteration along the grain boundaries and the fractured zone. Cordierite includes magnetite, quartz, monazite, etc. as inclusions. Cordierite compositions are magnesian, and XMg ranges from 0.74 to 0.77 (table 3).

Table 3 Representative electron microprobe analyses and structural formula of gedrite and cordierite.

Gedrite is coarse-grained and has idioblastic prisms in thin sections. It is commonly associated with biotite to define foliation in rocks (figure 2d). It contains inclusions of cordierite, biotite, quartz, etc. It shows parallel extinction and pleochroism in which the colour varies from yellowish green to greenish brown. Gedrite contains a trail of biotite as inclusion which suggests the appearance of gedrite in the rock due to the breakdown of biotite + quartz. Gedrite consists of AlIV and AlVI components in sufficient quantities 0.64–1.29 and 0.47–0.78 pfu, respectively, and the XMg ranges from 0.54 to 0.58 (table 3).

Biotite occurs in the form of small flakes and as individual laths within the hypersthene (figure 2c). It occurs as inclusion in gedrite and partially wrapped by cordierite (figure 2d). The TiO2 content of biotite varies from 1.62 to 1.99 wt%, and its compositions are magnesian with the XMg ranges from 0.71 to 0.73 (table 4).

Table 4 Representative electron microprobe analyses and structural formula of biotite.

Ilmenite occurs as elongated and prismatic grains, which are very fine-grained and present as inclusion in gedrite, biotite, garnet and other mineral grains (figure 2e). Compositionally, they are magnesian–mangaon ilmenites (MgO: 0.049–0.734 wt% and MnO: 0.221–0.412 wt%) (table 5).

Table 5 Representative electron microprobe analyses and structural formula of ilmenite.

Monazite grains occur as accessory minerals in the matrix and are ubiquitous in garnet–hypersthene–gedrite–cordierite gneisses as well as garnet–gedrite–cordierite–biotite gneisses; large grains of monazite (60–80 micron) are commonly found in high-grade metamorphic rocks (Montel et al. 1996). The monazite grains contain 3.14–7.20 wt% of thorium oxide (ThO2), 0.28–1.52 wt% of uranium oxide (UO2) and 0.26–0.44 wt% of lead oxide (PbO) in the sample R-91-97. The sample R-91-96 contains 3.53–6.18 wt% of ThO2, 0.26–1.45 wt% of UO2 and 0.26–0.46 wt% of (PbO). The normalised cations based on the four oxygen basis are presented in table 6 and supplementary table 1.

Table 6 Representative electron microprobe analyses and structural formula of monazite.

5 Monazite geochronology

5.1 Sample description and U–Th–Pb systematics

R-91-97 and R-91-96 samples have been chosen for microprobe dating after systematic EPMA-BSE imaging. Monazite grains are of diameter (70–80 μm) and compositionally homogeneous (figure 5a–h). The monazite grains show the homogeneous compositional domain which is demarcated by the BSE image and X-ray elemental mapping. The X-ray elemental maps of some selective monazite grains P43 of R-91-97 and P46 of R-91-96 are shown in figure 5(a–h). Both P43 and P46 monazite grains occur as inclusions in the garnet porphyroblast and are relatively poor in yttrium (Y) elemental composition at the outer rim margin in comparison with the core (figure 5d and h). The yttrium (Y) partitioning in the monazite is directly linked to the growth or consumption by the peritectic garnet (Spear and Pyle 2010; Bhowmik et al. 2014). But, U and Th X-ray elemental maps are showing homogeneous composition in both the grains (figure 5b, c, f and g). U and Th with Pb were found to occur together in both huttonite and brabantite types of substitution. Monazite shows the compositional variation between Th (+Ca and Si) and Y (+HREE) and it reflects the different substitutions: brabantite substitution (Th4+ + Ca2+ = 2REE3+; Rose 1980) and huttonite substitution (Th4+ + Si4+ = LREE3+ + P5+; Pabst and Hutton 1951). The variation of the brabantite vs. huttonite exchange operation is presented in the plot of Th + U + Si vs. REE + Y + P (figure 6). All monazite grains contained negligible SiO2. However, they contain sufficient amounts of Ca which quantify the brabantite (Th/U + Ca ↔ 2REE) substitution. Here, the brabantite substitution is dominant in monazites.

Figure 5
figure 5

Grain-P43 of the R-91-97 sample, (a) BSE image. (b) and (c) X-ray elemental maps documenting the homogeneous pattern of Th and U elements in monazite. (d) X-ray map shows the zoning pattern at the outer part in monazite; wherein grain-P46 of the R-91-96 sample, (e) BSE image, (f) and (g) X-ray elemental maps documenting the homogeneous pattern of Th and U elements in the monazite. (h) X-ray map shows the zoning pattern at the outer part as well as the core of monazite.

Figure 6
figure 6

The bivariate plot shows the variation in the composition of monazite of two different granulite samples from Daltonganj (CGGC). Both rock types are enriched in the brabantite substitution vector as marked in the diagram.

5.2 Electron microprobe dating

Electron microprobe dating can be used as an efficient investigation tool for finding the age of metamorphism and deformational history (Williams et al. 1999). Here, EPMA monazite geochronology was conducted to find out the age, and to establish the evolutionary history of the granulite of Daltonganj.

Monazite grains vary in shape from anhedral to subhedral or rounded and size from the smaller grain (10–30 μm) to larger grain (60–80 μm). It occurs as inclusion within garnet and the matrix. Estimation of age and uncertainties are compared from different monazite grains, and monazite growth events were interpreted which recorded from the Daltonganj area of CGGC. A total of 39 EPMA ages were obtained from 39 monazite mineral grains of the two different granulite samples. The Th–U–Pb values from the different monazite grains of R-91-97 and R-91-96 samples are given in table 7. EPMA dating generates two age domains, and the calculated monazite ages range from 1348 ± 47 to 1482 ± 49 Ma and 896 ± 49 to 1050 ± 63 Ma in R-91-97, and vary from 1322 ± 64 to 1494 ± 65 Ma and 926 ± 58 to 1019 ± 59 Ma in R-91-96 (figure 7). The weighted average age distribution and probability density plot was obtained by using the ISOPLOT program (Ludwig 2011) which is depicted in figure 8a–h. The analysis of sample R-91-97 produced age population at 1424 ± 64 Ma (figure 8a and b) and 972 ± 28 Ma, with 95% confidence (figure 8c and d). The sample R-91-96 yielded age population at 1390 ± 56 Ma (figure 8e and f) and 962 ± 159 Ma, with 95% confidence (figure 8g and h). The electron microprobe dating of monazite grains has generated the two-age domain from both rocks, i.e., garnet–hypersthene–gedrite–cordierite gneiss and garnet–gedrite–cordierite–biotite gneiss, which lies around the Mesoproterozoic and Grenville orogeny age.

Table 7 EPMA dating age of monazite crystals of granulites from the Daltonganj (Palamau) area.
Figure 7
figure 7

Represents the backscattered images (BSE-SEM) of different monazite grains from two rock samples.

Figure 8
figure 8

(a) and (c) Weighted-average ages (b) and (d) probability–density ages of two distinct age domains from the R-91-97 rock sample and (e) and (g) weighted-average ages and (f) and (h) probability–density ages of two distinct age domains from the R-91-96 rock sample with 2σ uncertainty, different numbers of point analysis and MSWD (mean square of weighted deviates) for monazite from the Daltonganj area of the CGGC, plotted with the ISOPLOT program (Ludwig 2011).

6 PT condition of metamorphism

The PT conditions were estimated from the heterogeneous compositions of garnet and cordierite from garnet–hypersthene–gedrite–cordierite gneiss. The garnet–cordierite Mg–Fe exchange geothermometers and garnet–cordierite–sillimanite–quartz geobarometers were used to estimate the PT conditions and their results are summarised in tables 8 and 9. The maximum and minimum temperatures obtained at 7 kbar pressure were 788° and 656°C, respectively. At 700°C, the corresponding pressure varies from 4.80 to 7.34 kbar. By using the THERMOCALC v-3.21 thermodynamic modelling program by Holland and Powell (1998), the PTavg was estimated for garnet–hypersthene–cordierite–biotite gneiss, the estimated average temperature and pressure (PTavg) were 792°C and 7.35 kbar. The calculated pressure and temperature have corresponded to initial heating and compression until achieving a peak metamorphic condition, i.e., prograde metamorphism until the peak.

Table 8 Pressure and temperature estimates of the garnet–hypersthene–gedrite–cordierite–gneiss of the study area through conventional geothermobarometers and internally consistent data sets.
Table 9 Result of internally consistent geothermobarometry with THERMOCALC v-3.21 (Holland and Powell 1998).

7 Discussion

7.1 The timing of metamorphic events

The CGGC has a multiplex metamorphic history, based on pre-existing geological information; the CGGC has been divided into four phases of metamorphic events (M1–M4). The M1 metamorphic event is recorded at ~1870 Ma from granulite enclaves emplaced in felsic gneiss; furthermore, the M2 metamorphic event was dated between 1628 and 1270 Ma, where felsic magma intrusion occurred and further metamorphosed to form the migmatitic felsic gneiss. The M3 phase is a high-grade metamorphic event that occurred during 1200–930 Ma, followed by the M4 event (870–780 Ma) with the emplacement of the mafic dyke (Sanyal and Sengupta 2012 and references therein). The geochronological studies of various researchers from different localities of the CGGC, metamorphic phases (M1–M4), dating techniques and nature of rocks are compiled in table 1.

The PT conditions during the M1 metamorphic stage of the granulite facies rocks are difficult to derive due to their complex history of successive metamorphic events. Prograde granulite facies metamorphism was reported in the enclave suite (750–850°C/4–6 kbar), and this event was reported at >1500 Ma age by Maji et al. (2008). The PT condition and the petrographic reaction texture relations are preserved within the rocks of the CGGC, which suggest the two prominent metamorphic events M2 and M3 that correspond to Mesoproterozoic and Grenvillian orogeny age (Sanyal and Sengupta 2012). These events represent the two different episodes of progressive metamorphism, which is separated by retrogressive metamorphic events (Maji et al. 2008), but the representative age is not distinguished. U–Pb zircon dating reveals that the age of intrusive A-type felsic magma (protolith of charnockite) is at 1447 + 11 Ma (Mukherjee et al. 2017), also during 1470–1450 Ma age, the emplacement of ferroan granitoids was reported in the north-eastern part of the CGGC (Mukherjee et al. 2018). A-type granitoid magmatism and fragmentation of the Columbia supercontinent are recorded during the Mesoproterozoic era (Hoffman 1989; Frost and Frost 2011). The monazite age (1424 ± 64 Ma) is revealed as the oldest age of the garnet–hypersthene–gedrite–cordierite gneiss; it is signified as the age of the gneissic protolith. The EPMA monazite ages 972 and 962 Ma (Grenville Orogeny) represent the high-grade granulite facies event, which is recorded in the Daltonganj. Petrographical features show that recrystallisation of the amphibole-rich magmatic rock to garnet-hypersthene-bearing gneiss by the following reaction during the M3 event:

$$\text{Gedrite} + \text{quartz} = \text{Garnet} + \text{hypersthene} + \text{cordierite} + \text{albite} + \text{H}_2{\text{O}}.$$

The appearance of the hypersthene indicates that the low PT condition has changed into a high PT condition of granulite facies. The PT condition and reaction texture are interpreted from the mineral assemblage, which shows that gedrite and quartz are consumed to produce the garnet + hypersthene + cordierite mineral phases through the prograde metamorphism. Similar PT condition and Grenvillian orogenic age (975 ± 67 Ma) were obtained by Chatterjee and Ghose (2011) from Chianki village of the Daltonganj area present in the north-west of the CGGC.

7.2 Implications for the supercontinental history

The age of formation, amalgamation and reconstruction of central and eastern Indian terrain generates essential information regarding the palaeogeographic condition of supercontinents. Rogers and Santosh (2002) proposed that the Columbia supercontinent amalgamation initiated ~1900–1800 Ma and achieved their highest packing strength at 1600–1500 Ma and started to rift after 1500 Ma. During this rifting period a lot of magmatic processes were obtained, viz., crystallisation of anorthosite around 1550 Ma (Chatterjee et al. 2008), khondalite emplaced in the quartzo-feldspathic matrix around 1510 Ma (Sanyal et al. 2007), as well as charnockite gneiss emplacement during 1457 ± 63 Ma (Ray Barman et al. 1994). The development of Rodinia started from the Grenvillian orogenic age ~1100–900 Ma, and drifting was started after 750 Ma. The number of configurations and models of the Rodinia supercontinent have been proposed by different scientists, including Dalziel (1991), Hoffman (1991), Rogers (1996), Meert (2001), Wingate et al. (2002) and Li et al. (2008). The age of fragmentation from 1200 to 800 Ma was interpreted within the southern Indian granulite blocks (Yoshida et al. 2003) at some locations. Ghose (1983) and Banerji (1991) mentioned orogenic phases in the CGGC, named as the Chhotanagpur orogeny (1600–1500 Ma) and the Satpura orogeny (900–850 Ma). However, the CGGC of eastern India shows a shred of evidence of the Grenvillian-orogeny age at 1100–900 Ma which is strongly preserved, and it postulates that the Grenvillian-orogeny suture was very near the CGGC of India. In the previously proposed models, Greater India was emplaced along the western side of East Antarctica and the SW part of Australia to produce a substantial accretionary mass of western Rodinia (Dalziel 1991; Hoffman 1991; Moores 1991). They suggested that India was assembled with the Rodinia supercontinent through the continent–continent collision between 1000 and 900 Ma along the Eastern Ghats mobile belt (EGMB) and CGGC of the Indian subcontinent which corresponds to East Antarctica’s Rayner Province. Li et al. (2008) explained the palaeo-latitudinal position between Greater India and the Australian land mass at ~770–750 Ma age due to the drifting of the Indian plate away from the Australia–East Antarctica continental plate by ca. 755 Ma. The transpressional movement of the Indian and Australian continental plates may explain the 1100–1000 Ma metamorphic events investigated from the Pinjarra orogen (Bruguier et al. 1999; Fitzsimons 2003). The (~1424 Ma) older age reveals the age of emplacement of felsic magmatism similar to the rocks of the other area of the CGGC that has been mentioned in table 1. This rock was assumed to be the protolith of granulitic gneiss, which was formed by the high-grade metamorphism under granulite facies conditions (‘Grenville-age’ orogenesis, 1000–900 Ma) during M3 in Daltonganj, presumably during the assembly of Rodinia. The 1000 Ma high-grade metamorphism gives evidence of tectono-metamorphic episodes in the CGGC, CITZ and EGMB of India (figure 9).

Figure 9
figure 9

Cartographic picture showing the Rodinia assembly and position of India at ~1000 Ma (modified after Li et al. 2008).

8 Conclusion

The CGGC represents a complex metamorphic history, where Mesoproterozoic metamorphism was overlain by high-grade metamorphism of Grenvillian-orogeny (1100–900 Ma). The CGGC terrain mainly contains high-grade amphibolite facies to granulite facies rocks, which lies between two medium- to low-grade mobile belts. The north-western part of the CGGC depicts that the protolith of granulitic gneiss had been emplaced around 1424 Ma age and subsequently transformed by a high-grade metamorphic event at 972 Ma. High-grade metamorphism at 972–962 Ma from the Daltonganj area of the north-western CGGC suggests connecting link between the Satpura mobile belts of CITZ in Grenvillian-orogeny. The PT condition calculated in the present study perhaps corresponds to initial heating and compression until achieving a peak metamorphic condition. The prograde metamorphic condition in the Daltonganj area of the CGGC at the M3 metamorphic event has reported 7.35 kbar/792°C in the Grenvillian age. Thus, Grenvillian metamorphism was well documented in the CGGC, where Greater India was part of the Rodinia supercontinent.