You are on page 1of 15

Geoscience Frontiers 11 (2020) 597–611

H O S T E D BY Contents lists available at ScienceDirect

Geoscience Frontiers
journal homepage: www.elsevier.com/locate/gsf

Research Paper

Characterization of partial melting events in garnet-cordierite gneiss from


the Kerala Khondalite Belt, India
Nilanjana Sorcar a, Kumar Batuk Joshi a, *, Elson P. Oliveira b, J.K. Tomson a, V. Nandakumar a
a
ESSO–National Centre For Earth Science Studies, Thiruvananthapuram, Kerala, India
b
Department of Geology, Institute of Geosciences, University of Campinas, Brazil

A R T I C L E I N F O A B S T R A C T

Handling Editor: Richard M Palin Phase equilibria modelling coupled with U–Pb zircon and monazite ages of garnet–cordierite gneiss from Valli-
kodu Kottayam in the Kerala Khondalite Belt, southern India are presented here. The results suggest that the area
Keywords: attained peak P–T conditions of ~900  C at 7.5–8 kbar, followed by decompression to 3.5–5 kbar and cooling to
Kerala khondalite belt (KKB) 450–480  C, preserving signatures of the partial melting event in the field of high to ultra-high temperature
Garnet-cordierite gneiss
metamorphism. Melt reintegration models suggest that up to 35% granitic melt could have been produced during
Partial melting
metamorphism at ~950  C. The U–Pb age data from zircons (~1.0 – ~0.7 Ga) and chemical ages from monazites
Pseudosection
Monazite-zircon geochronology (~540 Ma and ~941 Ma) reflect a complex tectonometamorphic evolution of the terrain. The ~941 Ma age
reported from these monazites indicate a Tonian ultra-high temperature event, linked to juvenile magmatism/
deformation episodes reported from the Southern Granulite Terrane and associated fragments in Rodinia, which
were subsequently overprinted by the Cambrian (~540 Ma) tectonothermal episode.

1. Introduction HT–UHT metamorphism across crustal fragments, necessitating an


approach that combines the use of petrology, phase equilibrium model-
Partial melting of lower crustal rocks is an important mechanism for ling and in-situ geochronology of HT and UHT assemblages within such
crustal differentiation and granulite genesis (Vielzeuf et al., 1990). terranes. Here, this method is adopted to characterise lower crustal
Numerous natural and experimental data sets (Le Breton and Thompson, partial melting processes, under varying P–T conditions and at different
1988; Vielzeuf and Holloway, 1988; Peterson and Newton, 1989; Car- times in the evolutionary history of the Kerala Khondalite Belt (KKB),
rington and Harley, 1995; Carrington and Watt, 1995; Stevens et al., which is among the most important crustal segments of the SGT.
1997) have firmly established the significance of dehydration melting
(e.g. muscovite, biotite and hornblende melting) under varying
physico-chemical conditions within the middle and lower crust. Pre- 1.1. Geological setting
cambrian HT/UHT terranes are important domains to understand the
melting processes and scales of melt transfer. Regarded as exhumed root The SGT is a collage of Precambrian high-grade crustal blocks sepa-
zones of ancient orogenic belts, these terranes provide an excellent rated by transcrustal brittle–ductile shear/suture zones with an crustal
window into middle–lower crust evolution (Nandakumar and Harley, evolutionary history that ranges from the Late Archean to the Neo-
2019). In southern India, vast exposures of deep crustal rocks in the proterozoic (Fig. 1). The four distinct crustal domians of the SGT, from
Southern Granulite Terrane (SGT) make it a key region in the recon- north to south, are the Salem Block, the Madurai Block, the Trivandrum
struction of the Gondwana supercontinent. The exact spatial and tem- Block, and the Nagercoil Block. They retain a record of multiple
poral evolution of the distinct crustal blocks within the SGT, and their magmatic, metamorphic and deformation episodes in the history of the
correlation to similar tectonic units in the presently dispersed Gondwana SGT (Bhaskar Rao et al., 2003, 2008; Chetty et al., 2003; Collins et al.,
continental fragments, have been the subject of active debate. This is, in 2007; Clark et al., 2009; Santosh et al., 2009, 2012; Plavsa et al., 2012;
part, due to the paucity in data for correlating the precise chronology of Peucat et al., 2013; Tomson et al., 2013; Brandt et al., 2014; Collins et al.,
2014; Plavsa et al., 2015; Kumar et al., 2017 and references therein). The

* Corresponding author.
E-mail address: kr.batukjoshi@gmail.com (K.B. Joshi).
Peer-review under responsibility of China University of Geosciences (Beijing).

https://doi.org/10.1016/j.gsf.2019.05.013
Received 5 July 2018; Received in revised form 28 April 2019; Accepted 23 May 2019
Available online 19 July 2019
1674-9871/© 2019 China University of Geosciences (Beijing) and Peking University. Production and hosting by Elsevier B.V. This is an open access article under the
CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 1. A simplified geological sketch of southern


India showing (from south to north) the Southern
Granulite Terrane (SGT), the metamorphic transition
zone (TZ), and the western and eastern parts of the
Dharwar craton (WDC and EDC); modified from
Vijaya Kumar et al. (2017). Sample locations, and the
Neoarchaean and Neoproterozoic (more specifically,
Ediacaran–Early Cambrian) granulite blocks are
marked; the latter includes the Madurai Block (MB)
comprising the Northern Madurai Block (NMB),
Southern Madurai Block (SMB), Trivandrum Block
(TB), and Nagercoil Block (NB). Prominent highland
massifs include Coorg (C), Nilgiri (NG), Biligiri
Rangan (BR), MaleMahadeshwara (MM), Shevaroy
(SH), Kollimalai (Ko), Madras (MA), Kodaikanal
(KH), Varusanadu (VH), Cardamom (CH), and
Nagercoil (NH). Major shear zones include the Moyar
Shear Zone (MSZ), Bhavani Shear Zone (BSZ), Mettur
Shear Zone (MtSZ), Salem-Attur Shear Zone (SASZ),
Gangavalli Shear Zone (GSZ), Palghat–Cauvery Shear
System (PCSS), Palghat–Cauvery Shear Zone (PCSZ),
Karur–Kambam–Painavu–Trichur Shear Zone
(KKPTSZ), Suruli Shear Zone (SSZ), and Achankovil
Shear Zone (ASZ).

transcrustal shear/suture zones in the SGT can be traced across conti-


nents, which is suggestive of their imporatance in establishing spatial
relationships with other fragments of supercontinents such as Rodinia
and Gondwana (Kriegsman, 1995; Ghosh et al., 2004; Collins et al., 2007;
Santosh et al., 2009; Kr€ oner et al., 2012; Brandt et al., 2014 and refer-
ences therein).
The Trivandrum Block also known as the Kerala Khondalite Belt
(KKB; Chacko et al., 1987), is composed of granulite facies metasedi-
mentary rocks like khondalite (aluminous garnet þ cordierite þ silli-
minite gneiss), highly migmatized garnet-bearing felsic gneiss
(leptynitic), granitic orthogneiss, charnockite and localised mafic gran-
Fig. 2. Field photographs of garnet-cordierite gneiss from the Vallikodu Kot-
ulite. The KKB is separated from the Madurai Block by the Achankovil
tayam quarry showing (A) near-parallel fabric defined by gneissic layering
Shear Zone (AKSZ) to the north, and bounded by the Nagercoil Block in comprising Grt þ Crd mesosomes of 2–3 cm width (coin for scale, 2.4 cm
the south. For the KKB, metasedimentary packages and associated rocks diameter) and a quartzo-feldpsathic leucosome; and (B) cordierite overprinting
record Nd model protolith ages from the Archean to the Paleoproterozoic the foliation in a garnet-bearing leucosome.
(Harris et al., 1994; Cenki et al., 2004; Tomson et al., 2013) whereas
U–Pb protolith crystallization ages range from 2.1 Ga to 1.7 Ga (Kr€ oner 700–800  C to 6.5–7.0 kbar at 900–950  C (Chacko et al., 1987; Nan-
et al., 2012, 2015; Whitehouse et al., 2014; Harley and Nandakumar, dakumar and Harley, 2000; Cenki et al., 2002; Braun and Broker, 2004)
2016; Kumar et al., 2017 and references therein). Detrital zircon U–Pb and between 830 and 925  C at 6–9 kbar (Blereau et al., 2016). These
ages from metasedimentary gneisses in the KKB are sourced from calculations were contradicted by Tadokoro et al. (2008), who reported
Palaeoproterozoic and Neoarchaean rocks, and were deposited after ca. UHT conditions of >950  C at pressures up to 12 kbar.
1900 Ma. They are correlated with Paleoproterozoic metasedimetray For the present study, a garnet-cordierite gneiss sample (Fig. 2 A, B)
rocks from Madagascar and Sri Lanka (Collins et al., 2007, 2014), sug- was collected from the Vallikodu (V) Kottayam quarry, hereafter referred
gesting that they originated from East African rocks (Collins et al., 2003). to as V. Kottayam (N 09 130 12.4"; E 76 480 40.600 ), which is situated
Neoproterozoic to Cambrian age regional granulite-facies metamorphism within the KKB, close to the Achankovil Shear Zone.
has been reported from the KKB at ca. 570–515 Ma in zircons from
orthogneisses, paragneisses and migmatitic leucosomes, which coincides
2. Petrography
with the final stages of Gondwana amalgamation (Cenki et al., 2004;
Shabeer et al., 2005; Collins et al., 2007, 2014; Kr€
oner et al., 2012, 2015;
The dominant lithology from V. Kottayam is migmatitic garnet-
Harley and Nandakumar, 2014; Taylor et al., 2014; Whitehouse et al.,
cordierite gneiss that is characterized by the alternation of quartzo-
2014). Monazite chemical ages have been used to date the HT–UHT
feldspathic leucosomes and mesosomes (Fig. 2A) containing a combina-
metamorphism (Santosh et al., 2003, 2005, 2006a, b); P–T estimates for a
tion of garnet–cordierite–orthopyroxene–spinelss–plagioclase–K-feld-
high to ultra–high temperature (UHT) rock range from 5.5 to 7.0 kbar at
spar–quartz–ilmenite, with minor biotite making up the main

598
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

assemblage. Garnet, cordierite, orthopyroxene and spinelss are generally


present as large irregular-shaped porphyroblasts. Texturally, cordierite
occurs as porphyroblasts (Fig. 3A), inclusions within garnet porphyro-
blasts (Fig. 3B) and as rims over porphyroblastic garnet (Fig. 3C). Garnet
porphyroblasts usually engulf spinelss, sillimanite, biotite, zircon and
monazite as inclusions. Cordierite porphyroblasts also contain small in-
clusions of rounded plagioclase, alkali feldspar, opaques, and spinel-
ilmenite aggregates. Partial resorption of garnet, orthopyroxene and
cordierite is widespread, which gives rise to a variety of textures.
Plagioclase and cordierite rims are typically observed around garnet
crystals, though plagioclase corona structures may also be seen around
orthopyroxene. Locally, cordierite breaks down to fine-grained in-
tergrowths of biotite  plagioclase  K-feldspar (Fig. 3A). Textural
studies further suggest that the leucosome is essentially composed of
quartz and plagioclase, with perthitic K-feldspar and minor biotite,
cordierite and garnet (Fig. 2B). The leucosome could be overprinted by
fine-grained vermicular intergrowths of different phases like plagioclase
and quartz, appearing to corrode coarse alkali feldspar (Fig. 3D) and by
biotite-quartz symplectites surrounding cordierite (Fig. 3E). Cuspate
quartz grains in the domain of feldspar (Fig. 3F) and rounded quartz
grains within large perthitic feldspar suggest the presence of a pre-
existing melt. In migmatite leucosomes, a narrow plagioclase rim
within the quartz layer indicates evidence of melt migration (Fig. 3G).
Monazite and zircon form inclusions in porphyroblasts and are also
present in leucosomes.

3. Analytical techniques

Major elements were analysed using X-ray fluorescence (XRF) tech-


nique using a Bruker S4 Pioneer wavelength dispersive (WD) XRF in-
strument at the National Centre for Earth Science Studies (NCESS),
Trivandrum, India. The detection limit for major elements was ~0.01%
and analytical precision was always better than 1%. The precision and
accuracy of calibration curves and data reliability is available at http://
cess.res.in/groups/crustal-processes-crpgroup/laborataries/xrf-lab-2
(Ravindra Kumar and Sreejith, 2016). The XRF data of representative
samples are given in Table 1.
Fig. 3. Microscopic features of the garnet-cordierite gneissic samples showing Mineral compositions and monazite U–Th–total-Pb concentrations
textural features: (A) Cordierite porphyroblast surrounded by the fine-grained were analysed using a CAMECA SX-100 EPMA at the Department of
intergrowth among K-feldspar, plagioclase and biotite; (B) cordierite inclusion Geology and Geophysics, Indian Institute of Technology, Kharagpur,
within garnet porphyroblast which is rimmed by plagioclase; (C) garnet rimmed
India. The instrument was operated at an accelerating voltage of 15 kV,
by cordierite in association with orthopyroxene porphyroblast; (D) worm-like
15 nA beam current and 1–2 μm beam diameter. Natural silicate and
intergrowths of plagioclase, quartz and cordierite corroding alkali feldspar;
(E) biotite–quartz symplectites surrounding cordierite grain; (F) cuspate quartz oxide standards were used for calibration, and raw data were corrected
grains in the domain of feldspar; (G) narrow rim of plagioclase within the layer using the ZAF program. Representative mineral compositions are given in
of quartz. Mineral abbreviations are used after Kretz (1983): Kfs–K-feldspar; Appendices A–I.
Crd–Cordierite; Grt–Garnet; Pl–Plagioclase; Bt–Biotite; Opx–Orthopyroxene; Matrix as well as inclusion monazite grains were analysed for
Qtz–Quartz; Spl–Spinel. U–Th–total-Pb on selected spots, based on backscattered electron (BSE)
imaging and X-ray element imaging. Analytical procedures followed are

Table 1
Bulk compositions used in the construction of phase diagrams.
Oxides XRF whole rock composition (wt.%) Recalculated bulk after addition of 10% Recalculated bulk after addition of 20% Recalculated bulk after addition of 30%
leucosome-composition leucosome-composition leucosome-composition

KKB3 (Grt-Crd KKB3L (wt.%) (wt.%) (wt.%)


Gneiss) (Leucosome)

SiO2 62.85 72.26 63.71 64.42 65.02


TiO2 0.97 0.02 0.88 0.81 0.75
Al2O3 17.26 15.99 17.14 17.05 16.97
FeO 6.28 0.68 5.77 5.35 4.99
MnO 0.13 0.01 0.12 0.11 0.1
MgO 2.26 0.18 2.07 1.91 1.78
CaO 2.86 1.31 2.72 2.6 2.5
Na2O 3.78 3.31 3.74 3.7 3.67
K2O 2.65 5.93 2.95 3.2 3.41
O2 0.45 – 0.41 0.38 0.35

599
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

done using heavy liquids like bromoform and methylene iodide. Zircons
were finally handpicked under a binocular microscope.
Approximately 90–100 handpicked zircons from the heaviest non-
magnetic fractions of each sample were mounted in transparent epoxy
resin and polished to expose the interiors of the crystals. BSE and cath-
odoluminescence (CL) images were obtained prior to analysis, using a
scanning electron microscope (SEM) in order to characterise the internal
structures of individual grains. The U–Pb zircon analytical work was
performed at the Institute of Geosciences, UNICAMP. Isotope data were
acquired on an SF-ICP-MS Element XR (Thermo Scientific), coupled with
an Excite.193 (Photon Machines) laser ablation system, equipped with a
two-volume HelEx ablation cell. The acquisition protocol followed Nav-
arro et al. (2015) with details from Verma et al. (2016); 25 μm was used
as the spot size. Data were reduced off-line using the Iolite software
(version 2.5) following the method of Paton et al. (2010), which involves
the subtraction of gas blanks followed by down-hole fractionation
correction and comparison with the behaviour of the 91500 reference
zircon (Wiedenbeck et al., 1995). Peixe zircon standard (ID-TIMS age of
564  4 Ma; cf. Dickinson and Gehrels, 2003) was used to monitor the
quality of the reduction procedures; during the analytical sessions 15
Peixe zircon analyses yielded an average age of 564.4  9.5 Ma.
Representative CL images and geochronological data from zircons are
given in Fig. 4 and Appendix L respectively.

Fig. 4. Cathodoluminescence images of the zircons analysed in this study; 4. Mineral chemistry
measured points marked by hollow white circles.
The overall composition of garnet in the garnet-cordierite gneiss from
described in Pant et al. (2009) and Bhowmik et al. (2014). For the the KKB is Alm54–64Prp24–39Sps1–7Grs2–4. Garnet porphyroblasts associ-
chemical mapping of monazite, X-ray element maps were generated for ated with cordierite do not show typically show any zoning, except for
Y, Th and U at 20 kV and 100 nA, with 50 ms dwell times and a step size the few which are in contact with biotite. Zoned garnet show very faint
of 0.1 μm at a resolution of 512  512 pixels. Spot analysis of monazite diffusion zoning, with a rimward depletion of Mg and an increase in Fe
was conducted at an accelerating voltage of 20 kV, a beam current of 200 and Mn.
nA and with a beam size of 1 μm. X-ray data were calibrated against Cordierite, irrespective of their texture, are highly magnesian with
well-characterized natural and synthetic standards. The specific stan- XMg (¼ Mg/(Mg þ Fe)) between 0.76 and 0.85. But intergrowth phase
dards used for monazite analysis were: galena for Pb; UO2 for U; ThO2 for cordierites have lower XMg (XMg ¼ ~0.76).
Th; synthetic silica–aluminium glass containing 4% rare earth element Orthopyroxene associated with garnet and cordierite has XMg values
(REE) for La, Ce, Nd, Pr, Sm, Ho, Dy and Gd; apatite for P and Ca; Yttrium between 0.59 and 0.65, with low contents of Ti, Mn and Ca. Al2O3
Aluminium Garnet (YAG) for Y; corundum for Al; hematite for Fe; and contents in the orthopyroxene cores range from 5.95 wt.% to 8.17 wt.%,
Th-glass for Si. X-ray lines selected for age analysis were Y Lα, Pb Mβ, U whereas rims typically have a slightly lower Al2O3 contents (5.67–7.21
Mβ and Th Mα. Pb Mα was measured on a large pentaerythyritol (LPET) wt.%).
crystal; counting time at the peak was 300 s while the background was Spinel is hercynite rich with XMg (¼ Mg/(Mg þ Fe2þþ Zn)) ranging
measured on both sides for 150 s. For U, the U Mβ line was used to avoid from 0.32 to 0.50. The ZnO content is 1.27–4.42 wt.%, providing a
interference with the Th Mβ line; the counting time was 200 s on the peak gahnite component, i.e., XZn (¼ Zn/(Mg þ Fe2þþ Zn)) of 0.03–0.09.
and 100 s on the background. The Th Mα peak was also counted for 200 s, Opaques containing only Fe appear to be exclusively magnetite when the
with 100 s measured on the background. U and Th were measured on PET total wt.% is recalculated for Fe3þ.
crystals. The X-ray lines used to measure REE were La (Lα), Ce (Lα), Nd Ileminite composition is close to the Fe end member having Xilm (¼
(Lβ), Pr (Lβ), Sm (Lα), Ho (Lβ), Dy (Lα) and Gd (Lβ). Background positions Fe2þ/(Mg þ Fe2þþ Mn)) of 0.90–0.97. These coarse-grained ilmenite are
were selected using wavelength dispersive spectroscopy (WDS) spectra, characterized by low to moderate hematite contents (XFe3þ  100 ¼ 5–17
with care taken to avoid major interferences. The reproducibility of the mol.%). Geikelite component is close to 1–6 mol.%.
analytical data was further checked with the reference, a Moacyr The alkali feldspar in leucosomes is mostly perthitic, with a compo-
monazite from Brazil with a207Pb/235U age of 483  1 Ma (Crowley et al., sition of Or65–98Ab2–21An1–7. The overall composition of plagioclase is
2005). The age calculated during calibration was 492  13 Ma (MSWD ¼ Ab65–75An23–34Or1–3 with anorthite content independent of the textural
0.79, n ¼ 7) at the 95% confidence level. However, the age calculation position of the phase. However, the albite content is higher in plagioclase
was carried out by employing an improved method by Cocherie and inclusions and intergrowths, and lower for plagioclase in the leucosome;
Albarede (2001), and plotted by Isoplot/Ex version 3.0 (Ludwig, 2008). matrix plagioclase also has a higher albite component as compared to
The results of all the analyses are presented in Appendices J–K. those lying close to garnet or orthopyroxene.
Zircon separation from the samples were carried out at the National Biotite is mostly unzoned; those lying adjacent to garnet and the ones
Centre for Earth Science Studies, Thiruvananthapuram, Kerala, India and included within garnets have higher XMg values ranging from 0.69 to
at the Institute of Geosciences, UNICAMP, Brazil. Sample preparation 0.80 and 0.70–0.75, respectively, as compared to the ones present in the
involved crushing approximately 10 kg of the sample to approximately matrix (XMg ¼ 0.66–0.74). Biotite grains that are present as intergrown
0.25 mm grain-size, and then washing the crushed sample with de- phases are more ferric (XFe ¼ 0.41–0.42) than those in the matrix or
ionized water to remove ultra-fine particles to prevent muddying the contact biotite. TiO2 content of biotite is variable, ranging between 3.79
heavy liquid used in later stages. Magnetic minerals were separated from wt.% and 8.08 wt.%. Biotite present within the intergrowth has higher
the dried sample powder, first by a hand-held magnet and subsequently TiO2 concentration than that of contact, matrix and inclusion biotite.
using an isodynamic separator. The separation of heavy minerals was Fluorine content in the studied biotites is negligible, irrespective of its
mode of occurrence.

600
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

5. Mineral reactions paths followed by these partially molten rocks.


Locally, garnet porphyroblasts are partly resorbed and display
The garnet porphyroblast in association with cordierite contains in- embayed grain boundaries. These embayments are characterized by
clusions of biotite, sillimanite and quartz. This texture suggests the vermicular intergrowths with compositions changing gradually from the
following reaction might have occurred during the early stages of garnet boundary towards the leucosome, which could be anhedral
metamorphism (Spear et al., 1999): cordierite, cordierite-quartz (Fig. 3D) and locally, biotite-quartz sym-
plectites (Fig. 3E). These textures could be explained by the following
Bt þ Sil þ Pl þ Qtz → Grt þ Crd þ Melt þ (Kfs) (i)
melt consuming reaction (Cenki et al., 2002):
A local presence of large, subhedral orthopyroxene grains in the Grt þ Melt → Crd þ Bt þ Qtz (vii)
garnet-cordierite gneiss suggests the occurrence of the multivariate
KFMASH melting reaction (Vielzuef and Montel, 1994; Stevens et al., In addition, porphyroblastic cordierite may have also reacted with the
1997): melt forming biotite-quartz intergrowths. The melt also exhibits graphic
intergrowth involving quartz and plagioclase (Fig. 3D), which surrounds
Bt þ Qtz þ Pl → Opx þ Crd þ Melt þ (Kfs) (ii)
the perthitic K-feldspar. Hence, the last melt possibly crystallized at the
Production of K-feldspar in biotite-melting reactions depends on the solidus via (Cenki et al., 2002):
K2O/H2O ratio of biotite and melts (Carrington and Watt, 1995). Melt → Qtz þ Kfs þ Pl þ V  Bt (viii)
Such reactions are likely to be multivariant in natural rocks and
therefore, occur over a P–T interval. Other than P and T, factors as bulk Late fluid release from the crystallizing melt according to reaction
rock XMg and Al2O3 content, fO2, F and Ti in biotite, XCa of plagioclase and (viii) may possibly explain the partial chloritization of the Fe–Mg phases
aH2 O , also influence these reactions. Vielzeuf and Montel (1994) in their like Opx (Cenki et al., 2002).
experiments produced orthopyroxene þ cordierite by reaction (ii) at
830–850  C and 3 kbar. At higher pressures, spinel and garnet succes- 6. Phase diagram modelling of melting event
sively substituted for cordierite. Experimental data from Carrington and
Harley (1995), on a more magnesian pelitic bulk composition (XMg ¼ 86) The metamorphic evolution of the garnet-cordierite gneiss from the
in the system KFMASH, defined a narrow stability field for Opx þ Crd þ KKB has been modelled using calculated P–T pseudosections to under-
Bt þ Qtz þ melt at 900–920  C, where at pressures of 8 kbar and above, stand the paragenetic changes it underwent. Applying free energy
garnet appeared instead of orthopyroxene. Orthopyroxene and cordierite minimization tools in conjunction with internally consistent thermody-
were also produced by Stevens et al. (1997) in highly magnesian psam- namic databases (e.g., Perple_X from Connolly, 1990, with the database
mopelitic composition (XMg ¼ 81) at 5 kbar and 850–950  C. Garnet of Holland and Powell, 1998) to partially molten systems is challenging
appeared at 10 kbar in the same bulk composition. However, the available due to the open system behaviour of high-temperature rocks which have
experimental data collectively define an upper limit for melting reaction local domains of equilibrium. The major problem with this approach
(ii) at 7–8 kbar and at temperatures in excess of 850  C. remains the choice of bulk composition because high-grade rocks nor-
Locally, aggregates of spinel and cordierite commonly separate mally undergo significant melt-loss during or prior to peak meta-
partially resorbed garnets from the leucosome, which can be explained morphism (Brown, 2002). As a result, present rock chemistry may not
by the reaction (Cenki et al., 2002): represent pristine bulk composition (which is the composition prior to
melt loss). Here, two different bulk chemistries (pre-peak and post-peak)
Bt þ Grt þ Sil þ Pl →Spl þ Crd þ (Kfs) þ Melt (iii)
are selected to model the partial melting events in the studied granulite.
Again, orthopyroxene þ spinel þ quartz could have been stabilized by
the following reaction in a slightly Fe-rich bulk composition (Vielzeuf 6.1. Pre-peak metamorphic condition and melting
and Montel, 1994):

Bt þ Pl þ Qtz → Opx þ Spl þ (Kfs)þ Melt (iv) Pre-peak metamorphic conditions are considered by recasting the
bulk composition which underwent biotite-dehydration melting. In this
In addition to the factors mentioned earlier, this reaction is also study, a melt–reintegration approach is followed to obtain the bulk
influenced by the presence of Zn in spinelss (Hensen and Harley, 1990). composition prior to melt loss. It is a popular approach to reconstruct the
The experimental results of Vielzeuf and Montel (1994), and Stevens prograde history of melt-depleted rocks using phase equilibria modelling
et al. (1997) demonstrate that the orthopyroxene þ spinel assemblage is (White et al., 2004). Different methods of calculating and reintegrating
preferred over garnet in psammopelitic compositions at relatively low melt composition have been applied by previous studies (Bartoli, 2017
pressure (5–6 kbar) and high fO2. The melt thus produced is mostly and references therein). Bartoli (2017) used one-step and multi-step
removed from the system, which enabled the preservation of the procedures for melt-reintegration and showed that the overall topology
HT/UHT assemblage (White et al., 2011). However, we consider relic of phase diagrams is similar, irrespective of the method applied; thus,
textures like cuspate quartz grains in the domain of plagioclase (Fig. 3F), reintegrating a certain amount of melt can be sufficient to restore a
rounded quartz grains within large perthitic feldspar to be suggestive of convincing prograde history (i.e. melting conditions, melt productivity
the former presence of melt. etc.) of residual migmatites and granulites. However, in this study, the
Textural studies further suggest that large garnet grains (containing single step melt-reintegration approach is followed.
inclusions of sillimanite and quartz) show the development of cordierite/ Since the preservation of peak assemblages requires significant
plagioclase corona indicating the following reactions (Sorcar et al., removal of melt from the system, it is mandatory to add the lost melt
2014): component to the present bulk composition. As the melt composition for
a pelitic to semipelitic protolith is almost granitic, we looked for a suit-
Grt þ Sill þ Qtz → Crd (v)
able granitic rock which crystallized in the same (peak) metamorphic
Grt þ Sill þ Qtz → Pl (vi) realm of the granulite. Pre-peak bulk composition was thus obtained by
adding the adjacent leucosome composition (considered as crystallized
Similar textures were also reported by Santosh (1987), and Nanda- melt formed during the partial melting of the granulite) of the sample
kumar and Harley (2000) from a nearby area in the KKB. These two re- KKB-3L to the whole rock chemistry of the garnet-cordierite gneiss
actions involve large changes in volume and could have occurred during sample KKB3. The recalculated bulk composition was then used for
post-peak decompression, which provides added constraints on P–T constructing P–T pseudosections with the software Perple_X (Version

601
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 5. (A) T–O2 diagram indicating the stability fields of the observed assemblages with changing O2 values; (B) T–H2O diagram showing stability of the mineral
assemblages as a function of temperature and H2O-content at 7.5 kbar. Mineral abbreviations are used after Kretz (1983) as follows: Kfs–K-feldspar; Crd–Cordierite;
Grt–Garnet; Pl–Plagioclase; Bt–Biotite; Opx–Orthopyroxene; Qtz–Quartz; Splss–Spinel solid solution; Ilm–Ilmenite; Sil–Sillimanite; Ky–Kyanite;
O2–Oxygen; H2O–Water.

6.7.9) in the system Na2O–CaO–K2O–FeO–MgO–MnO–Al2O3–SiO2– water-undersaturated melt (Vielzuef and Holloway, 1988; Stevens et al.,
TiO2–H2O–O2 (NCKFMASHTMnO). Several types of activi- 1997). Again, melt segregation requires a rheological critical melt per-
ty–composition relationships for solid solutions are used on a trial basis centage (RCMP) to be achieved due to dehydration melting. Experi-
(with reference to solution_model_v.6.7.9 of Perple_X_6.7.9). The solu- mental data has suggested that the RCMP would occur at about 30 vol.%
tion model of Tajcmanova et al. (2009) was considered for biotite. This is melt (Van der Molen and Paterson, 1979). In this study, it is assumed a
a reciprocal margules solution with dependent end members. There were maximum 30 wt.% of anhydrous granitic leucosome composition can be
two independent mixing sites incorporating Mn, Fe2þ, and Mg at Site 1 added to the whole rock chemistry of the granulite to retrieve the bulk
and four components (Al, Si, Fe3þ, Ti) at Site 2. The solution models of rock composition prior to the melt loss event.
Holland and Powell (1998) were used for muscovite, garnet, and The granulite was metamorphosed under oxidized conditions, which
orthopyroxene. The orthopyroxene solution model is a reciprocal with a is evident from the coexistence of spinel–magnetite (magnetite solid
speciation-type of solution, whereas the muscovite and garnet solutions solution) and ilmenite–hematite (hematite solid solution). To constrain
models are margules-type of solutions with end member fractionation. the oxygen fugacity, the approach outlined by Korhonen et al. (2012),
The spinel solution model, after White et al. (2002), is an ideal, and Kelsey and Hand (2015) was followed in this study. Accordingly, a
margules-type of solution. For ilmenite, the macroscopic-type solution T–O2 pseudosection is constructed at 7.5 kbar (Fig. 5A) using the
model used was from White et al. (2000). In the case of plagioclase measured composition of the rock to show the stability fields of the
feldspar, the solution model of Newton and Haselton (1981) was used. observed assemblages with changing O2 values (Table 1). Fig. 5A shows
This is also a margules solution. In the case of cordierite, an ideal that the garnet–orthopyroxene–cordierite–spinel and ilmenite bearing
anhydrous/hydrous Mg/Fe HP cordierite model with the correction of assemblages stabilize with 0.3–0.6 wt.% of O2, respectively, within the
hfcrd_i stoichiometry by L. Baumgartner (on 5/6/03) was taken for temperature range 920–980  C. Therefore, a median O2 value of 0.45
pseudosection calculation. This is a reciprocal solution with an internal, wt.% was chosen to construct the P–T pseudosection (Fig. 5A).
dependent end member. For the melt, the solution model of Holland and The composition of water (H2O) present at the time of metamorphism
Powell (2001), and White et al. (2001) was taken for pseudosection cannot be chemically determined; for this, an isobaric (at 7.5 kbar, cor-
construction, in a partially molten system. This is also a margules solu- responding to the pressure of metamorphism) T–H2O pseudosection
tion, and this model uses an internal routine to compute the entropy of (Fig. 5B) was constructed using whole rock composition (Table 1)
the melt. The pre-peak condition of metamorphism was recomposed by (following the method adopted in Sorcar et al., 2014). It shows that in
adding 10%, 20% and 30% anhydrous granitic leucosome composition to order to stabilize the observed mineral assemblage in the rock (gar-
the whole rock chemistry of the granulite (Table 1). The modal percent of net–orthopyroxene–cordierite–spinel–ilmenite–K-feldspar) at the tem-
the melt was estimated from the abundance of quartz and feldspar grains perature of metamorphism, the rock should have contained at least ~1
showing melt-crystallizing textures, for which a value of 30% was taken wt.% H2O (Fig. 5B). Moreover, a maximum of 2.0 wt.% H2O can be
as the maximum following the method adopted by Ganguly et al. (2017). present under saturated conditions in a psammitic bulk composition
Although the actual amount of melt loss in events along the prograde (Holtz et al., 2001). Similarly, the isobaric (at 7.5 kbar) T–H2O diagrams
path is unknown, each melting reaction (described in Section 6) seems to are constructed using recalculated bulk compositions (after adding 10%,
have been able to produce more melt than required for it to escape, ~7 20% and 30% anhydrous granitic leucosome composition to the whole
vol.% (~6 wt.%), corresponding to the melt connectivity transition of rock chemistry of the studied granulite. For each recalculated bulk
Rosenberg and Handy (2005). Therefore, at least ~10 wt.% of anhydrous composition, the diagrams demonstrate that the rock should have con-
granitic leucosome composition has to be added to the whole rock tained at least ~1 wt.% H2O to stabilize the observed mineral assemblage
chemistry of the granulite to retrieve the bulk rock composition prior to in the rock (garnet–orthopyroxene–cordierite–spinel–ilmenite–K-
the melt loss event. On the other hand, under favourable conditions the feldspar) at the temperature of metamorphism. Here, H2O ¼ 1.00 wt.% is
biotite breakdown reaction can produce a large amount (40 wt.%) of used for pseudosection construction.

602
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 6. P–T pseudosections constructed at re-calculated bulk compositions in


the system Na2O–CaO–K2O–FeO–MgO–MnO–Al2O3–SiO2–TiO2–H2O–O2
demonstrating pre-peak to peak conditions of metamorphism. The pre-peak
condition of metamorphism is recomposed by changing the leucosome compo-
sition from (A) 10% to (B) 20% to (C) 30% of the whole rock chemistry of the
garnet-cordierite gneiss. Each pseudosection is contoured with modal abun-
dance of the melt. Regions of the diagram where stability fields of several as-
semblages are closely juxtaposed on each other are amalgamated and marked as
a single, shaded and unlabeled field for clarity of illustration. The reader is
referred to the text for details. Mineral abbreviations: Ms–Muscovite, others are
same with Fig. 5.

It is to be noted that the ‘melt-in’ curve appears close to 740  C in the


low pressure region (5–8 kbar) in all the pseudosections (Figs. 6A–C and
7) irrespective of the bulk composition used. The resulting diagrams also
show that biotite dehydration melting initiates close to 750  C, at 6–8
kbar and the biotite-out temperature is close to 880  C. This is in
agreement with the experimental data (Bohlen et al., 1983; Peterson and
Newton, 1989; Vielzeuf and Clemens, 1992; Carrington and Harley,
1995; Stevens et al., 1997; Das et al., 2001, 2003) although these ex-
periments did not consider F content in biotite and its effect on dehy-
dration melting (Skjerlie et al., 1993; Bose et al., 2005 and references
therein). It is also noted that the field consisting of peak metamorphic
assemblage found in the studied rock (Grt–Crd–Opx–Splss–Pl–Kfs–
Qtz–Ilm) is extended towards higher temperature with gradual addition
of the melt to the residuum bulk composition. From the calculated melt
modes (Fig. 6 A–C), we found that melt productivity gradually increases
with melt addition; however, up to 35% melt can be generated in the
temperature interval of 740–950  C. Extracting this melt from the system
would cause considerable change in the initial bulk composition.

6.2. Peak to post-peak metamorphic stage

The whole rock composition (Table 1) of the sample has been used for
the construction of P–T pseudosections in the oxidized system
Na2O–CaO–K2O–FeO–MgO–MnO–Al2O3–SiO2–TiO2–H2O–O2 (NCKFMA
SHTMnO). We have used the internally consistent thermodynamic
database of Holland and Powell (1998) with subsequent updates
(described in the previous section). Modes of minerals/melts of solid
solution minerals were computed with the software Werami_6.7.9, a part
of the Perple_X _6.7.9 package. The value of O2 is assumed as 0.45 wt.%
following the procedure described above.
The appearance and disappearance of key mineral phases are shown
in the computed pseudosection (Fig. 7). The ‘melt-in’ curve shows a steep
slope (almost pressure insensitive) and successive fields (with increasing
temperature) of an interval of dehydration melting of muscovite (Fig. 7)
followed by a broad interval of dehydration melting of biotite. The low-
variance field with the assemblage of Grt–Opx–Crd–Splss–Ilm–Pl–Kfs–L
closely represents the peak metamorphic assemblage of the sample. This
field has a pressure-temperature window of 6–8 kbar and 900–980  C.
The peak field is confined by the cordierite-in line for the down-
temperature, the garnet-out line for the down-pressure, and the
orthopyroxene-out line for the up-pressure. There is a rapid increase in
the amount of melt produced in the biotite-melting interval after
exhaustion of muscovite in the system (Fig. 8A), in conformity with
experimental data of Le Breton and Thompson (1988), and Clemens
(1990). A maximum of 30% melt can be produced at the highest tem-
perature in this melting interval, and the melt becomes progressively
depleted in XNa (i.e. becomes more granitic) (Fig. 8B). There is an in-
crease in the modal abundance of garnet and spinel (Fig. 8C and D) in the
biotite-melting interval, which indicates that garnet and spinel are per-
(caption on next column) itectic phases in the biotite dehydration-melting reaction (see also
textural evidence described earlier). Again, the modal variation of
cordierite and orthopyroxene (Fig. 8E and F) shows that their abundance
increases with temperature rise in the low P region in the pseudosection,
which supports their occurrence as porphyroblasts in the rock. Since

603
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 7. P–T pseudosections constructed at whole rock


composition from the studied sample in the system Na2O–
CaO–K2O–FeO–MgO–MnO–Al2O3–SiO2–TiO2–H2O–O2 to
represent peak- and post-peak conditions of meta-
morphism. Regions of the diagram where stability fields of
several assemblages are closely juxtaposed on each other
have been amalgamated and illustrated as a single, shaded
and unlabeled field for clarity of illustration. The stability
field of the peak assemblage found in the studied sample is
shown in bold black letters. Coloured lines demarcate the
regions where a particular phase is present; the colour code
is explained in the legend. Location of peak P–T point ob-
tained by individual thermobarometry is marked as star
and black dotted line shows the interpreted clockwise P–T
path followed by the rock. The reader is referred to the
figure legend for an explanation of symbols and the text for
discussion of the data. The details of the mineral abbrevi-
ations are mentioned in Fig. 6 caption.

F-absent biotite in the granulite formed during retrogression, the biotite temperatures are 885–920  C from KKB4 at an assumed pressure of 8
þ cordierite þ quartz þ plagioclase þ spinel þ ilmenite þ K-feldspar kbar. In case of sample KKB 3, also the “peak” T condition (890–960  C)
retrograde mineral assemblage defines a P–T condition of <6 kbar and following the thermometer of Nichols et al. (1992) is in close agreement
<840  C. The retrograde field is well constrained by the garnet-in line at with that obtained from KKB4. Estimates of pressure at the thermal
the upper pressure and the biotite-out line at the upper temperature. The maximum attained by these rocks (based on core analyses) lie in the
above results and the occurrence of cordierite and plagioclase coronas range of 8  0.5 kbar using the model by Ganguly et al. (1996).
suggest a clockwise P–T path. The pressure and temperature for the retrograde stage was con-
strained by biotites (without fluorine) coexisting with garnet porphyro-
7. Geothermobarometry and P–T evolution blasts in the studied samples by considering the garnet–biotite Fe–Mg
exchange thermometer and garnet–aluminosilicate–plagioclase–quartz
The results of P–T calculations using various thermometers and ba- (GASP) geobarometer of Ganguly et al. (1996). The estimated P–T con-
rometers are presented in Table 2. Based on petrographic and mineral dition of the retrograde stage (based on rim composition) lies in the range
chemical studies, the porphyroblastic garnet, orthopyroxene, cordierite, of 3.5–5 kbar and 450–480  C.
plagioclase, and quartz are considered to be in equilibrium at the peak However, the combined results of various geothermobarometric es-
stage of metamorphism. The garnet–orthopyroxene geothermometer timates indicate that the studied garnet-cordierite gneiss from the KKB
(Lee and Ganguly, 1988) and the garnet–aluminosilicate-(sillimanite reached peak temperatures of ~845–1000  C, with maximum pressure of
present as inclusion within garnet)–plagioclase–quartz geobarometer ~8 kbar. Individual thermobarometry also determined the retrograde
(Ganguly et al., 1996) are therefore used to estimate the peak meta- P–T condition clustering around 3.5–5 kbar and 450–480  C for all the
morphic P–T conditions of the KKB3 samples. The Fe–Mg exchange samples indicating significant decompression, a likely component of
thermometers formulated by Lee and Ganguly (1988), yield maximum retrogression along a clockwise P–T trajectory. This implies that the peak
temperatures (based on core analyses) of 870–1000  C at 8.0 kbar for assemblage was Qtz–Kfs–Pl–Grt–Crd–Spl–Melt–Ilm (Opx) for all the
KKB3 (with the appearance of orthopyroxene as porphyroblasts). Since KKB samples in this study. In the calculated pseudosection, this field
orthopyroxene is not present in other samples of garnet-cordierite gneiss occupies an area with a small range in temperature (920–980  C) and a
from KKB4, an estimate of peak temperature is obtained from the somewhat more extended range in pressure of 5–8 kbar. Peak P–T
spinel–sillimanite–garnet–quartz assemblage (Nichols et al., 1992). calculated using individual thermobarometry for all samples using the
Using the thermometer of Nichols et al. (1992), the estimated method outlined above lie close to the stipulated stability field.

604
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 8. Contours of modal abundances of different phases and concentrations of different elements in a selected phase in the pseudosection: (A) Melt mode; (B) XNa in
melt; (C) modal abundance of garnet; (D) modal abundance of spinelss; (E) modal abundance of cordierite; (F) modal abundance of orthopyroxene. The details of the
mineral abbreviations are mentioned in the figure caption of Fig. 6.

Finally, overall P–T trajectory for KKB samples can be obtained from garnet-cordierite gneisses. A few round- to oval-shaped monazite grains
the above combining results of pseudosection calculations (Fig. 7) and are included within porphyroblastic garnet or cordierite, while the rest
individual thermobarometry (Table 2). In this study, the prograde path of are dispersed through a plagioclase–perthite–quartz matrix or locally
the KKB rocks are poorly constrained, the first anchor point of the path is associated with Pl–Qtz–Kfs–Bt intergrowths.
the peak P–T condition, plotted as a star (Fig. 7). The plot shows that the X-ray element composition maps of U, Th and Y for selected matrix
KKB reached peak temperatures of ~900  C (in spite of some scatter) monazite grains along with their BSE images (e.g. Fig. 9A–D) shows
with a maximum pressure of ~7.5 kbar. This stage is followed by complex zoning patterns with three domains of contrasting BSE charac-
decompression, evident from the production of a corona structure on ters. X-ray element composition maps reveal the presence of truncated
cordierites over garnet. This is an eventful period in the history of these high-U and low-Th zones (Fig. 9B and C). The oldest age data (from 934
rocks, where changes in the modal abundances of garnet and melt take to 965 Ma) plot within the U-rich patchy zones in the central portion,
place. Once this stage is reached, a phase of near isobaric cooling occurs. which may indicate relicts of an older monazite core. This zone is
During this stage, some of the back reactions involving melts (e.g. reac- sequentially enveloped by thick domain of low-Th and high-Y contents
tion vii) along with the development of fine-grained intergrowth followed by a domain of the highest Th and lowest Y contents (Fig. 9D).
involving biotite and Qtz (e.g. reaction viii) may occur. Such P–T paths An outermost thin zone is characterized by moderate-Th.
are consistent with clockwise P–T path described by other workers from Spot dates of monazite grains from both the matrix and the inclusions
the Trivandrum block in the KKB (Santosh, 1986; Chacko et al., 1996; vary within the range of ca. 410 Ma to 965 Ma (Fig. 10A). Few matrix
Nandakumar and Harley, 2000; Pattison et al., 2003; Ishii et al., 2006; monazite populations show older ages, ranging from 798 Ma to 965 Ma.
Shimizu et al., 2009; Harley and Nandakumar, 2014, 2016; Johnson The older age ranging from 927 Ma to 965 Ma yields a weighted mean of
et al., 2015; Blereau et al., 2016). 941  18 Ma (95% conf., MSWD ¼ 0.23; 10 spots, Fig. 10B). Another
group of data yield spot dates ranging from 409 Ma to 618 Ma (matrix
8. Monazite chemical dating grains: 88 spots; inclusions: 52 spots). It is worth noting that the MSWD
of the weighted mean (540  3.7 Ma) for this data is greater than the
Two textural modes of monazites have been observed in the studied maximum permissible value (Wendt and Carl, 1991). Therefore, the

605
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Table 2 cannot be differentiated from morphological and chemical criteria. As a


Results of thermobarometric calculations of garnet-cordierite gneiss using result, the age is henceforth described as ca. 540 Ma, which may repre-
different models of geothermobarometers. Please see legend for explanation of sent the probable age of the event/events related to the growth of garnet
acronyms. in the granulite as evident from the occurrence of the Y-depleted outer
Sample Method Peak condition Retrograde rim of the monazite (Williams et al., 2007).
condition Compositionally, all monazites show depletion of huttonite compo-
P T ( C) P T ( C) nent (Fig. 11A); younger monazites have a wide variation in Si as well as
(kbar) (kbar) Th þ U, while those with older ages have restricted variations. Monazites
KKB 3A Grt-Opx thermometer by Lee [8] 1000 with younger ages, on the other hand, show two groups of enrichment
and Ganguly (1988) and depletion in the cheralite component, as compared to the ones with
Inclusion thermometry by [8] 960 older ages (Fig. 11B). A wide variation in Th, U, Y makes it difficult to
Nichols et al. (1992)
relate variation in monazite ages with their respective chemistry thereby
GASP by Ganguly et al. (1996) 8 [850] 5 [500]
GB-GASP by Ganguly et al. 4.6 480 suggesting that the growth of different zones might have occurred under
(1996) variable physicochemical conditions. This observation is in sharp
contrast to monazites studied from the Central Indian Tectonic Zone
KKB 3B Grt-Opx thermometer by Lee [8] 870 (CITZ), wherein a clear distinction was made between monazites
and Ganguly (1988)
growing under varying P–T conditions (Bhomik et al., 2014).
Inclusion thermometry by [8] 890
Nichols et al. (1992)
GASP by Ganguly et al. (1996) 8.4 [850] 4.5 [500] 9. U–Pb zircon geochronology

KKB 4A Inclusion thermometry by [8] 920 The analytical methods for the U–Pb zircon dating are given in the
Nichols et al. (1992) section on analytical techniques and the results for the samples are listed
GASP by Ganguly et al. (1996) 8.2 [850] 3.5 [500]
GB-GASP by Ganguly et al. 3.3 467
in the supplementary material. Fig. 4 shows CL images of zircons and
(1996) Fig. 12A and B presents geochronological results.
The studied zircons from the garnet-cordierite gneiss range from
KKB 4B Inclusion thermometry [8] 885 subhedral to anhedral grains of approximately 100–450 μm in size, with
Nichols et al. (1992) xenocrystic cores. The grains are colorless ranging from transparent to
GASP by Ganguly et al. (1996) 7.5 [850] 4 [500]
cloudy in appearance, and most of them have a visible (in CL) core-rim
GB-GASP by Ganguly et al. 3.5 450
(1996) zoning structure. Cores in some of the grains show brighter CL
response as compared to the rims, while in some the rims are brighter
GB-GASP: simultaneous calc. of Grt-Bt thermometer (Ferry and Spear, 1978) and
(Fig. 4). Zircons from this study show oscillatory, sector and patchy
Grt-Pl-Als-barometer (Koziol and Newton, 1989) modified with garnet activity
zoning with or without inherited cores. In some cases, the oscillatory
model after Ganguly et al. (1996); GASP: Grt-Als-Pl-barometer (Koziol and
Newton, 1989) modified with garnet activity model after Ganguly et al. (1996); zoning is cut-off by re-homogenised or irregular concentric zoning which
Nichols et al. (1992) and Grt-Spl barometer in FASZn system by Nichols et al. are characteristic of metamorphic zircons (Corfu et al., 2003).
(1992); [ ]: estimated P or T for calculation. A total of 79 zircon spots were analysed from the garnet-cordierite
gneiss. Nine spots were deleted owing to their high common Pb con-
tent (f206 > 1), while the rest plot on the concordia line between 400 Ma
and 1000 Ma (Fig. 12A). Sixty-eight spots with less than 20% discordance
were additionally plotted in the probability distribution diagram and
histogram shown in Fig. 11B. The ca. 840 Ma age cluster suggests the
main provenance (n ¼ 16). The youngest zircon population (n ¼ 6) has a
206
Pb/238U weighted average age of 532  16 Ma, which is the same as
the monazite age, within uncertainity. The histogram indicates that the
zircons from the studied sample have ages ranging from ca. 350–1000 Ma
with a dominance at 700–1000 Ma. A single grain with 207Pb/206U age of
ca. 2530 Ma is reported.
The studied zircon grains have a wide variation in their Th/U ratios
(0.01–1.48; Appendix L) which have no correlation with their respective
spot age. Rubatto (2002) suggested that the Th/U ratio could be an
important criterion to distinguish between zircon produced by magmatic
(Th/U > 0.1) and metamorphic (Th/U < 0.1) processes. However, Harley
et al. (2007) and Das et al. (2011) pointed out that a simple correlation
between Th/U ratio and the zircon forming processes, whether magmatic
or metamorphic, can be problematic as the enrichment and depletion of
U and Th is dependent on multiple factors. Yakymchuk et al. (2018) used
U and Th partition coefficients to model their behaviour; they suggested
that the breakdown and growth of monazite in equillibrium with zircon,
Fig. 9. Representative backscatter electron (BSE) images and X-ray element along with the timing of their growth during cooling and melt crystalli-
distribution (U, Th, and Y) maps of monazites in the studied samples; (A) BSE zation, controls the Th/U ratios in suprasolidus metamorphic zircon.
image showing complex zoning patterns with spot positions and calculated spot
ages; X-ray maps for (B) U, (C) Th and (D) Y. Color scale (right) indicates 10. Discussion
relative number of X-ray counts per pixel.

10.1. Characterization of the partial melting event


weighted mean plot is not shown, as it is just the statistical distribution of
spot ages. This age domain is wide, in range of ~120–130 Ma, which may It is essential to a develop quantitative mineral equilibria modelling
represent a single event or more than one event in a short time span that of a partially molten rock system to constrain the P–T evolution which

606
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 10. Calculated population ages from monazite grains analysed in this study; (A) probability density plot of weighted ages showing two peaks at ca. 540 Ma and
941 Ma respectively and (B) population with weighted mean age of 941  18 Ma.

Fig. 11. Binary plots showing the composition of monazite spots of different age populations; (A) Si vs. Th þ U and (B) Ca vs. Th þ U.

Fig. 12. (A) U–Pb concordia plot with all the analysed spots showing Archean inheritance and (B) histogram showing probability density of the analysed zircon grains.

helps to better our understanding on crustal differentiation and the when the amount of melt removal is uncertain, is done by constructing
evolution of orogens. U–Pb dating from accessory minerals has been pseudosections with the measured bulk composition of migmatitic gneiss
immensely successful in pinning down the time factor for metamorphic (KKB3) through the application of Gibb’s free energy minimization
and/or melting events along a P–T path. Based on evidence from technique. The pseudosections are contoured with modal abundance and
experimental and natural rock data, we argue that migmatization of compositional isopleths of the key phases. A melt reintegration approach
gneisses was triggered by several biotite dehydration-melting reactions is considered to determine the protolith composition of the study area.
at ~7.5 kbar and above 850  C (details in geothermobarometric section), Accordingly, the reintegrated composition (Table 1), which falls in the
which caused complete elimination of prograde biotite in the studied shaded region of the diagram (Fig. 13A–D) might represent an acceptable
rock. Subsequent to the formation of megacrystic orthopyroxene þ protolith composition, i.e. showing lower FeO, MgO, Al2O3, CaO, TiO2 as
garnet þ cordierite in the peak metamorphic stage, reactions (v) to (viii) well as N2O/K2O ratio compared to the measured bulk composition of the
(as mentioned in previous section) occurred during decompression and studied granulite. The analyses of the pseudosections (Figs. 6A–C and 8)
associated cooling. This is supported by the presence of plagioclase/ establish that melt productivity gradually increases with melt addition;
cordierite corona over garnet as well as the development of worm like no more than 35% granitic melt can be generated in the temperature
intergrowth among Qtz–Pl–Kfs  Bt (Fig. 2D, G). interval of 740–950  C. This estimation is consistent with the experi-
Modelling of phase equilibria in partially molten rocks in the KKB, mental result, which shows that under favourable conditions the biotite

607
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Fig. 13. Binary plots showing reintegrated compositions suggesting protolith composition of garnet-cordierite gneiss.

breakdown reaction can produce a large amount (up to 45%–50%) of metamorphic events, similar to the ones previously reported (Braun
water-undersaturated melt (Vielzuef and Holloway, 1988). Following its et al., 2007; Catlos et al., 2008; Collins et al., 2014; Taylor et al., 2014;
generation, melt extraction can result in the preservation of Harley and Nandakumar, 2016; Kumar et al., 2017; Tang et al., 2018 and
granulite-facies mineral assemblages (e.g. Fyfe, 1973; White and Powell, references therein) from the SGT, India. The younger ages, which are the
2002; Johnson et al., 2012; Palin et al., 2016a, b; Stuck and Diener, 2018) dominant and most common population are all from monazite occur-
in the study area. Cordierite gneisses from the KKB attained peak P–T rences, while the older ages come from texturally-controlled monazite
conditions of ~900  C at 7.5–8 kbar, and this stage is followed by its cores. The older ages can be attributed to relict domains related to earlier
decompression to 3.5–5 kbar, with cooling to 450–480  C. Such P–T monazite growth, resetting or recrystallisation during metamorphism or
paths are consistent with the clockwise P–T path described in cordierite magmatism (Braun and Brocker, 2004). Monazites with older age do-
gneisses from the Achankovil Zone (Santosh, 1987) which reported a mains have ca. 540 Ma low-Y rims, which are associated with the growth
moderate to rapid isothermal decompression at a maximum temperature of garnet during the high-temperature metamorphism in the Pan-African
of ca. 750  C followed by near-isobaric cooling at ca. 5 kbar. Further orogeny. Similar ages from monazites have also been reported from other
south of the KKB, Satish-Kumar and Harley (1998) proposed an isobaric localities in India (Braun et al., 1998; Braun and Broker, 2004; Clark
cooling path at a temperature range of 840–750  C for calc-silicate et al., 2015), Central Africa (Moller et al., 2000; Muhongo et al., 2001;
samples. These findings are supported by Nandakumar and Harley Hauzenberger et al., 2007; Tchato et al., 2009; Owona et al., 2011;
(2000), who concluded that the initial post-peak P–T path was dominated Tenczer et al., 2013), Madagascar (Jons and Schenk, 2008; Collins et al.,
by near-isobaric cooling until temperatures of 775–800  C and pressures 2012; Boger et al., 2015; Holder et al., 2018), Sri Lanka (Boger et al.,
of 5.5–6 kbar were reached. However, Cenki et al. (2002) suggested 2014 and references therein) and Antartica (Shiraishi et al., 2008;
higher pressure (6–7 kbar) and temperature 900–950  C, while Mor- Grantham et al., 2013; Hokada et al., 2013; Osanai et al., 2013; Boger
imoto et al. (2004) and Tadokoro et al. (2008) suggested UHT conditions et al., 2014). 950–850 Ma ages have been reported from monazites
of >950  C, at pressures up to 12 kbar. The most recent estimates based preserved in the cores of garnet porphyroblasts (Braun et al., 2007) from
on calculated pseudosections have demonstrated peak metamorphic the southern part of the Madurai Block (MB) which were interpreted as
conditions in the range of 830–925  C and 6–9 kbar followed by indications of a high-temperature metamorphic event. Our study reports
decompression (Blereau et al., 2016). Thus, our present results are in line for the first time the presence of Tonian ages (~941 Ma) in matrix
with the argument of clockwise P–T evolution for the garnet-cordierite monazites from KKB pelitic migmatites (garnet-cordierite gneiss), which
gneiss peak high temperature – moderate pressure stage (~900  C at is significant, considering the recently reported U–Pb ages from the
7.5–8 kbar) followed by decompression and cooling to 450–480  C at southern part of the MB (ca. 1.0–0.9 Ga) adjacent to the KKB (Kumar
3.5–5 kbar. These results suggest that metasedimentary units of the KKB et al., 2017) and from the KKB (ca. 1.0–0.7 Ga; this study). The ca.
reflect the burial, heating and eventual exhumation of 1.0–0.9 Ga zircon ages have been reported from charnockite orthog-
moderately-thickened crust. neisses of the MB which indicate juvenile magmatism in the SGT.
Additionally, ~0.8–0.7 Ga magmatic and metamorphic ages have also
10.2. Neoproterozoic/Cambrian tectonothermal events been reported from the SGT (Braun and Appel, 2006; Kooijman et al.,
2011; Sato et al., 2011; Teale et al., 2011; Plavsa et al., 2012) suggesting
The Fortunian (~ca. 540 Ma) and Tonian (~941 Ma) monazite ages the presence of a cryptic tectonothermal event in the region (Collins
from the garnet-cordierite gneiss from the KKB clearly bring out two et al., 2014). Based on recently available age data (Tucker et al., 2011;

608
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Kr€oner et al., 2012, 2015), the KKB and the adjoining Nagercoil Block are Archean, Neoproterozoic terrane boundary in the granulite terrane. In:
Ramakrishnan, M. (Ed.), Tectonics of the Southern Granulite Terrane, Kuppam-Palani
broadly correlated with the Anosyen–Androyen domain of southern
Geotransect, vol. 50. Geological Society of India Memoir, pp. 297–317.
Madagascar and the Highland complex of Sri Lanka. The ca. 1.0–0.9 Ga Bhaskar Rao, Y.J., Vijaya Kumar, T., Krishna, K.V.S.S., Tomson, J.K., 2008. The emerging
zircon ages have been reported from east Dronning Maud Land, pattern of crust formation and recycling history in the Precambrian Dharwar Craton
Antarctica (Jacobs et al., 2015), west and south Madagascar (the Dabo- and the Southern Granulite Terrane, Southern India: Constraints from recent
geochronological and isotopic results, vol. 74. Geological Society of India Memoir,
lava Suite; Tucker et al., 2011a, 2014) suggesting a geographical connect pp. 147–168.
between the terranes (Kumar et al., 2017 and references therein). New Bhomik, S.K., Wilde, S.A., Bhandari, A., Sarbadhikari, B., 2014. Zoned monazite and
geological age data presented in this study, together with previously zircon as monitors for the thermal history of granulite terranes: an example from the
Central Indian Tectonic Zone. J. Petrol. 55, 585–621.
reported juvenile magmatic ages from the southern part of the MB, Blereau, E., Clark, C., Taylor, R.J., Johnson, T.E., Fitzsimons, I.C.W., Santosh, M., 2016.
conclusively point towards an older tectonothermal event (Tonian) that Constraints on the timing and conditions of high-grade metamorphism, charnockite
was previously not well recorded within the KKB. This period overlaps formation and fluid–rock interaction in the Trivandrum Block, southern India.
J. Metamorph. Geol. 34, 527–549.
with the amalgamation of the supercontinent Rodinia, and juvenile Boger, S.D., Hirdes, W., Ferreira, C.A.M., Schulte, B., Jennet, T., Fanning, C.M., 2014.
magmatic ages are most widely reported from the Dabolava Suite (Tucker From passive margin to volcano-sedimentary forearc: the Tonian to Cryogenian
et al., 2014), Madagascar. The Tonian tectonothermal event in the KKB eolution of the Anosyen Domain of southeastern Madagascar. Precambrian Res. 247,
159–186.
may be a representation of the juvenile magmatism and associated Boger, S.D., Hirdes, W., Ferreira, C.A.M., Jennet, T., Dallwig, R., Fanning, C.M., 2015. The
HT/UHT metamorphism during the Rodinia amalgamation. This was 580-520 Ma Gondwana suture of Madagascar and its continuation into Antartica and
extensively overprinted by Cambrian tectonothermal episodes which Africa. Gondwana Res. 28, 1048–1060.
Bohlen, S.R., Boettcher, A.L., Wall, V.J., Clemens, J.D., 1983. Stability of phlogopite-
have been reported throughout the SGT (Santosh et al., 2009; Brandt
quartz and sanidine-quartz: a model for melting in the lower crust. Contrib. Mineral.
et al., 2014; Collins et al., 2014; Harley and Nandakumar, 2016; Kumar Petrol. 83, 270–277.
et al., 2017 and references therein). Bose, S., Das, K., Fukuoka, M., 2005. Fluorine content of biotite in granulite-grade
metapelitic assemblages and its implications for the Eastern Ghats granulites. Eur. J.
Mineral. 17, 665–674.
11. Conclusion Brandt, S., Raith, M.M., Schenk, V., Sengupta, P., Srikantappa, C., Gerdes, A., 2014.
Crustal evolution of the Southern Granulite terrane, south India: new
geochronological and geochemical data for felsic orthogneisses and granites.
The present study characterizes the evolution of the KKB by con-
Precambrian Res. 246, 91–122.
straining the phenomena of partial melting events through phase equi- Braun, I., Appel, P., 2006. U-Th-total Pb dating of monazite from orthogneisses and their
libria modelling as well as geochronological analyses of monazite and ultra-high temperature metapelitic enclaves: implications for the multistage tectonic
zircons. Various geothermobarometric estimates indicate that the area evolution of the Madurai Block, southern India. Eur. J. Mineral. 18, 415–427.
Braun, I., Broker, M., 2004. Monazite dating of granitic gneisses and leucogranites from
experienced peak temperatures of ~845–1000  C, with maximum pres- Kerala Khondalite Belt, southern India: implications for Late Proterozoic crustal
sure of ~8 kbar registering a clockwise P–T evolution. Pseudosections evolution in East Gondwana. Int. J. Earth Sci. 93, 13–22.
and melt-reintegration suggest that, with increasing pressure and tem- Braun, I., Montel, J.-M., Nicollet, C., 1998. Electron microprobe dating monazites from
high-grade gneisses and pegmatites of the Kerala Khondalite Belt, southern India.
perature, a maximum of 35% granitic melt was produced. Integration of Chem. Geol. 146, 65–85.
reaction histories constrained by textural evidence, kinetically- Braun, I., Cenki-Tok, B., Paquette, J.L., Tiepolo, M., 2007. Petrology and U–Th–Pb
constrained individual thermobarometry and pseudosection calcula- geochronology of the sapphirine–quartz-bearing metapelites from Rajapalayam,
Madurai Block, Southern India: evidence for polyphase Neoproterozoic high-grade
tions suggest a clockwise P–T path reflecting burial, heating and eventual metamorphism. Chem. Geol. 241, 129–147.
exhumation of moderately-thickened crust. Further, the chemical ages Brown, M., 2002. Prograde and retrograde processes in migmatites revisited.
from monazites and in-situ U–Pb isotopes from zircons point towards a J. Metamorph. Geol. 20, 25–40.
Carrington, D.P., Harley, S.L., 1995. Partial melting and phase relations in high-grade
complex tectonometamorphic evolution in the terrain with the clear
metapelites: an experimental petrogenetic grid in the KFMASH system. Contrib.
indication of a Tonian metamorphic event subsequently overprinted by Mineral. Petrol. 120, 270–291.
Cambrian tectonothermal episode. Carrington, D.P., Watt, G.R., 1995. A geochemical and experimental study of the role of K-
feldspar during water-undersaturated melting of metapelites. Chem. Geol. 122, 59–76.
Catlos, E.J., Dubey, C.S., Sivasubramanian, P., 2008. Monazite ages from carbonatites and
Acknowledgements high-grade assemblages along the kambam fault (southern granulite terrane, south
India). Am. Mineral. 93, 1230–1244.
Cenki, B., Kriegsman, L.M., Braun, I., 2002. Melt-producing and melt-consuming reactions
NS, KBJ, JKT, VN thank the Director, NCESS and the Secretary, MoES,
in the Achankovil cordierite gneisses, South India. J. Metamorph. Geol. 20, 543–561.
Government of India for providing funds and unconditional support to Cenki, B., Braun, I., Br€ ocker, M., 2004. Evolution of the continental crust in the Kerala
carry out this work. Authors also thank Prof. Somnath Dasgupta for in- Khondalite Belt, southernmost India: evidence from Nd isotope mapping, U–Pb and
depth discussions that helped improve the manuscript. The manuscript Rb–Sr geochronology. Precambrian Res. 134, 275–292.
Chacko, T., Ravindra Kumar, G.R., Newton, R.C., 1987. Metamorphic PT conditions of the
also benefitted from discussions with Profs. Sankar Bose, Kaushik Das Kerala (South India) Khondalite belt: a granulite facies supracrustal terrain. J. Geol.
and N. C. Pant. NS acknowledges support from Eldose and Nishanth in 95, 343–358.
the thin section and XRF laboratories at NCESS. Prof. B. Mishra and Chacko, T., Lamb, M., Farquhar, J., 1996. Ultra high temperature metamorphism in the
Kerala Khondalite Belt. In: Santosh, M., Yoshida, M. (Eds.), The Archaean and
Saptarshi at IIT, Kharagpur provided access to the IIT Kharagpur National Proterozoic Terrains in Southern India within East Gondwana. Gondwana Research
EPMA Facility. Dr. Robert Holder is thanked for his helpful review. NS Group Memoir 3. Osaka. Field Science Publishers, pp. 157–165.
acknowledges funding from the Department of Science and Technology, Chetty, T.R.K., Bhaskar Rao, Y.J., Narayana, B.L., 2003. A structural cross section along
Krishnagiri-Palani Corridor, southern granulite terrain of India. In: Ramakrishnan, M.
Government of India (India) under the DST INSPIRE Faculty Scheme (Ed.), Tectonics of Southern Granulite Terrain, vol. 50. Geological Society of India,
(Grant: DST/INSPIRE/04/2014/000221). Memoir, pp. 255–278.
Clark, C., Collins, A.S., Santosh, M., Taylor, R., Wade, B., 2009. The P-T-t architecture of a
Gondwana Suture: REE, U-Pb and Ti-in-zircon thermometric constraints from the
Appendix A. Supplementary data Palaghat Cauvery shear system, South India. Precambrian Res. 174, 129–144.
Clark, C., Healy, D., Johnson, T., Collins, A.S., Taylor, R.J., Santosh, M., Timms, N.E.,
Supplementary data to this article can be found online at https://doi. 2015. Hot orogens and supercontinent amalgamation: a Gondwanan example from
southern India. Gondwana Res. 28, 1310–1328.
org/10.1016/j.gsf.2019.05.013. Clemens, J.D., 1990. The granulite granite connexion. In: Vielzeuf, D., Vidal, P. (Eds.),
Granulites and Crustal Differentiation. Kluwer Academic, Dordrecht, pp. 25–36.
References Cocherie, A., Albarede, F., 2001. An improved U–Th–Pb age calculation for electron
microprobe dating of monazite. Geochem. Cosmochim. Acta 65, 4509–4522.
Collins, A.S., Kroner, A., Fitzsimons, I.C.W., Razakamanana, T., 2003. Detrital footprint of
Bartoli, O., 2017. Phase equilibria modelling of residual migmatites and granulites: an
the Mozambique ocean: U-Pb SHRIMP and Pb evaporation zircon geochronology of
evaluation of the melt-reintegration approach. J. Metamorph. Geol. 35, 919–942.
meta-sedimentary gneisses in eastern Madagascar. Tectonophysics 375, 77–99.
Bhaskar Rao, Y.J., Janardhan, A.S., Vijaya Kumar, T., Narayana, B.L., Dayal, A.M.,
Taylor, P.N., Chetty, T.R.K., 2003. Sm-Nd model ages and Rb-Sr isotopic systematics
of charnockite gneisses across the Cauvery shear zone, south India: implication for

609
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Collins, A.S., Santosh, M., Braun, I., Clark, C., 2007. Age and sedimentary provinance of Jacobs, J., Elburg, M., L€aufer, A., Kleinhanns, I.C., Henjes-Kunst, F., Estrada, S.,
the Southern Granulites, South India: U-Th-Pb SHRIMP secondary ion mass Ruppel, A.S., Damaske, D., Montero, P., Bea, F., 2015. Two distinct late
spectrometry. Precambrian Res. 155, 125–138. Mesoproterozoic/early Neoproterozoic basement provinces in central/eastern
Collins, A.S., Kinny, P.D., Razakamanana, T., 2012. Depositional age, provenance and Dronning Maud Land, East Antarctica: the missing link, 15–21 E. Precambrian Res.
metamorphic age of metasedimentary rocks from southern Madagascar. Gondwana 265, 249–272.
Res. 21, 353–361. Johnson, T.E., Fischer, S., White, R.W., Brown, M., Rollinson, H.R., 2012. Archaean
Collins, A.S., Clark, C., Plavsa, D., 2014. Peninsular India in Gondwana: the intracrustal differentiation from partial melting of metagabbro—field and
tectonothermal evolution of the southern granulite terrain and its Gondwana geochemical evidence from the central region of the Lewisian complex, NW Scotland.
counterparts. Gondwana Res. 25, 190–203. J. Petrol. 53 (10), 2115–2138.
Connolly, J.A.D., 1990. Multivariable phase diagrams: an algorithm based on generalized Johnson, T.E., Clark, C., Taylor, R.J., Santosh, M., Collins, A.S., 2015. Prograde and
thermodynamics. Am. J. Sci. 290, 666–718. retrograde growth of monazite in migmatites: an example from the Nagercoil Block,
Corfu, F., Hanchar, J.M., Hoskin, P.W.O., Kinny, P., 2003. Atlas of zircon textures. In: southern India. Geosci. Front. 6, 373–387.
Hanchar, J.M., Hoskin, P.W.O. (Eds.), Zircons. Mineralogical Society of America and Jons, N., Schenk, V., 2008. Relics of the Mozambique ocean in the central East African
Geochemical Society, Reviews in Mineralogy and Geochemistry, vol. 53, orogen: evidence from the vohibory block of southern Madagascar. J. Metamorph.
pp. 469–495. Geol. 26, 17–28.
Crowley, J.L., Chatterjee, N., Bowring, S.A., Sylvester, P.J., Myers, J.S., Searle, M.P., Kelsey, D.E., Hand, M., 2015. On ultrahigh temperature crustal metamorphism: phase
2005. U–(Th)–Pb dating of monazite and xenotime by EMPA, LA-ICPMS, and IDTIMS: equilibria, trace element thermometry, bulk composition, heat sources, timescales
examples from the Yilgarn Craton and Himalayas. Geochem. Cosmochim. Acta 69, and tectonic settings. Geosci. Front. 6, 311–356.
19. A19. Kooijman, E., Upadhyay, D., Mezger, K., Raith, M.M., Berndt, J., Srikantappa, C., 2011.
Das, K., Dasgupta, S., Miura, H., 2001. Stability of osumilite coexisting with spinel solid Response of the U–Pb chronometer and trace elements in zircon to ultrahigh-
solution in metapelitic granulites at high oxygen fugacity. Am. Mineral. 86, temperature metamorphism: the Kadavur anorthosite complex, southern India.
1423–1434. Chem. Geol. 290, 177–188.
Das, K., Dasgupta, S., Miura, H., 2003. An experimentally constrained petrogenetic grid in Korhonen, F.J., Powell, R., Stout, J.H., 2012. Stability of sapphirineþ quartz in the
the silica-saturated portion of the system KFMASH at high temperatures and pressure. oxidized rocks of the Wilson Lake terrane, Labrador: calculated equilibria in
J. Petrol. 44, 105–1075. NCKFMASHTO. J. Metamorph. Geol. 30 (1), 21–36.
Das, K., Bose, S., Karmakar, S., Dunkley, D.J., Dasgupta, S., 2011. Multiple Koziol, A.M., Newton, R.C., 1989. Grossular activity-composition relationships in ternary
tectonometamorphic imprints in the lower crust: first evidence of c. 950 Ma garnets determined by reversed-displaced equilibrium experiments. Contrib. Mineral.
compressional reworking of older UHT metamorphosed aluminous granulites from Petrol. 103, 423–433.
the Eastern Ghats Belt, India. Geol. J. 46, 217–239. Kretz, R., 1983. Symbols for rock-forming minerals. Am. Mineral. 68, 277–279.
Dickinson, W.R., Gehrels, G.E., 2003. U–Pb ages of detrital zircons from Permian and Kriegsman, L.M., 1995. The Pan-African event in East Antarctica: a view from Sri Lanka
Jurassic eolian sandstones of the Colorado Plateau, USA: paleogeographic and the Mozambique belt. Precambrian Res. 75, 263–277.
implications. Sediment. Geol. 163, 29–66. Kr€
oner, A., Santosh, M., Wong, J., 2012. Zircon ages and Hf isotopic systematics reveal
Ferry, J.M., Spear, F.S., 1978. Experimental calibration of the partitioning of Fe and Mg vestiges of Mesoproterozoic to Archaean crust within the late
between biotite and garnet. Contrib. Mineral. Petrol. 66, 113–117. Neoproterozoic–Cambrian high-grade terrain of southernmost India. Gondwana Res.
Fyfe, W.S., 1973. The granulite facies, partial melting and the Archean crust. Philos. 21, 876–886.
Trans. R. Soc. London, Ser. A 273, 457–461. Kr€
oner, A., Santosh, M., Hegner, E., Shaji, E., Geng, H., Wong, J., Xie, H., Wan, Y.,
Ganguly, J., Cheng, W., Tirone, M., 1996. Thermodynamics of aluminosilicate garnet Shang, C.K., Liu, D., Sun, M., 2015. Palaeoproterozoic ancestry of Pan-African high-
solid solution: new experimental data, an optimized model, and thermometric grade granitoids in southernmost India: implications for Gondwana reconstructions.
applications. Contrib. Mineral. Petrol. 123, 137–151. Gondwana Res. 27, 1–37.
Ganguly, P., Bose, S., Das, K., Torimoto, J., Ghosh, G., 2017. Origin of Spinelþ quartz Kumar, T.V., Rao, Y.B., Plavsa, D., Collins, A.S., Tomson, J.K., Gopal, B.V.,
assemblage in a Si-undersaturated ultrahigh-temperature aluminous granulite and its Babu, E.V.S.S.K., 2017. Zircon U-Pb ages and Hf isotopic systematics of charnockite
implication for the P–T–fluid history of the Phulbani domain, Eastern Ghats Belt, gneisses from the Ediacaran–Cambrian high-grade metamorphic terranes, southern
India. J. Petrol. 58, 1941–1974. India: constraints on crust formation, recycling, and Gondwana correlations. GSA
Ghosh, J.G., de Wit, M.J., Zartman, R.E., 2004. Age and tectonic evolution of Bulletin 129 (5–6), 625–648.
Neoproterozoic ductile shear zones in the Southern Granulite Terrane of India, with Le Breton, N., Thompson, A.B., 1988. Fluid-absent (dehydration) melting of biotite in
implications for Gondwana studies. Tectonics 23, TC3006. metapelites in the early stages of crustal anatexis. Contrib. Mineral. Petrol. 99,
Grantham, G.H., Macey, P.H., Horie, K., Kawakami, T., Ishikawa, M., Satish-Kumar, M., 226–237.
Tsuchiya, N., Graser, P., Azevedo, S., 2013. Comparison of the metamorphic history Lee, H.Y., Ganguly, J., 1988. Equilibrium compositions of coexisting garnet and
of the Monapo Complex, northern Mozambique and Balchenfjella and Aussthameren orthopyroxene: experimental determinations in the system FeO-MgO-Al2O3-SiO2, and
areas, Sør Rondane, Antartica: implications for the Kuunga Orogeny and the applications. J. Petrol. 29 (1), 93–113.
amalgamation of N and S. Gondwana. Precambrian Res. 234, 85–135. Ludwig, K.R., 2008. User’s Manual for Isoplot 3.7. Special Publication No. 4. Berkeley
Harley, S.L., Nandakumar, V., 2014. Accessory mineral behaviour in granulite Geochronology Center, Berkley, p. 76.
migmatites: a case study from the Kerala Khondalite Belt, India. J. Petrol. 55, Moller, A., Mezger, K., Schenk, V., 2000. U-Pb dating of metamorphic minerals: Pan-
1965–2002. African metamorphism and prolonged slow cooling of high pressure granulites in
Harley, S.L., Nandakumar, V., 2016. New evidence for Palaeoproterozoic high grade Tanzania, East Africa. Precambrian Res. 104, 123–146.
metamorphism in the Trivandrum block, southern India. Precambrian Res. 280, Morimoto, T., Santosh, M., Tsunogae, T., Yoshimura, Y., 2004. Spinelþquartz association
120–138. from the Kerala khondalites, southern India: evidence for ultrahigh-temperature
Harley, S.L., Kelly, N.M., M€ oller, A., 2007. Zircon behaviour and the thermal histories of metamorphism. J. Mineral. Petrol. Sci. 99, 257–278.
mountain chains. Elements 3 (1), 25–30. Muhongo, S., Kroner, A., Nemchin, A.A., 2001. Single zircon evaporation and SHRIMP
Harris, N.B.W., Santosh, M., Taylor, P.N., 1994. Crustal evolution in South India: ages for granulite-facies rocks in the Mozambique belt of Tanzania. J. Geol. 109,
constraints from Nd isotopes. J. Geol. 102, 139–150. 171–189.
Hauzenberger, C.A., Sommer, H., Fritz, H., Bauernhofer, A., Kroner, A., Hoinkes, G., Nandakumar, V., Harley, S.L., 2000. A reappraisal of the pressure-temperature path
Wallbrecher, E., Thoni, M., 2007. SHRIMP U-Pb zircon and Sm-Nd garnet ages from of granulites from the Kerala Khondalite Belt, Southern India. J. Geol. 108,
the granulite-facies basement of SE Kenya: evidence for Neoproterozoic polycyclic 687–703.
assembly for the Mazambique belt. J. Geol. Soc. 164, 189–201. London. Nandakumar, V., Harley, S.L., 2019. Geochemical signatures of mid-crustal melting
Hensen, B.J., Harley, S.L., 1990. Graphical analysis of P - T - X relations in granulite facies processes and heat production in a hot orogen: the Kerala Khondalite Belt, Southern
metapelites. In: Ashworth, J.R., Brown, M. (Eds.), High Temperature Metamorphism India. Lithos 324–325, 479–500.
and Crustal Anatexis. Unwin Hyman, London, pp. 105–123. Navarro, M.S., Tonetto, E.M., Oliveira, E.P., 2015. LA-SF-ICP-MS U-Pb Zircon Dating at
Hokada, T., Horie, K., Adachi, T., Osanai, Y., Nakano, N., Baba, S., Toyoshima, T., 2013. University of Campinas, Brazil. Geonalysis 2015, Vienna, Austria. P-09.
Unraveling the metamorphic history at the crossing of Neoproterozoic orogens, Sør Newton, R.C., Haselton, H.T., 1981. Thermodynamics of the garnet-plagioclase- Al2SiO5-
Rondane Mountains, East Antarctica: constraints from U–Th–Pb geochronology, quartz geobarometer. In: Newton, R.C., Navrotsky, A., Wood, B.J. (Eds.),
petrography, and REE geochemistry. Precambrian Res. 234, 183–209. Thermodynamics of Minerals and Melts. Springer-Verlag, New York, pp. 131–147.
Holder, R.M., Hacker, B.R., Horton, F., Rakotondrazafy, A.M., 2018. Ultrahigh- Nichols, G.T., Berry, R.F., Green, D.H., 1992. Internally consistent gahnitic spinel-
temperature osumilite gneisses in southern Madagascar record combined heat cordierite-garnet equilibria in the FMASHZn system: geothermobarometry and
advection and high rates of radiogenic heat production in a long-lived high-T orogen. applications. Contrib. Mineral. Petrol. 111, 362–377.
J. Metamorph. Geol. 36, 855–880. Osanai, Y., Nogi, Y., Baba, S., Nakano, N., Adachi, T., Hokada, T., Toyoshima, T.,
Holland, T.J.B., Powell, R., 1998. An internally consistent thermodynamic dataset for Owada, M., Satish-Kumar, M., Kamei, A., Kitano, I., 2013. Geologic evolution of the
phases of petrological interest. J. Metamorph. Geol. 16, 309–343. Sør Rondane Mountains, East Antarctica: collision tectonics proposed based on
Holland, T.I.M., Powell, R., 2001. Calculation of phase relations involving haplogranitic metamorphic processes and magnetic anomalies. Precambrian Res. 234, 8–29.
melts using an internally consistent thermodynamic dataset. J. Petrol. 42, 673–683. Owona, S., Schulz, B., Ratschbacher, L., Ondoa, J.M., Ekodeck, G.E., Tchoua, F.M.,
Holtz, F., Johannes, W., Tamic, N., Behrens, H., 2001. Maximum and minimum water Affaton, P., 2011. Pan-African metamorphic evolution in southern Yaounde Group
contents of granitic melts generated in the crust: a reevaluation and implications. (Oubanguide Complex, Cameroon) as revealed by EMP-monzazite dating and
Lithos 6, 1–14. thermobarometry of garnet metapelites. J. Afr. Earth Sci. 59, 125–139.
Ishii, S., Tsunogae, T., Santosh, M., 2006. Ultrahigh-temperature metamorphism in the Palin, R.M., White, R.W., Green, E.C., 2016a. Partial melting of metabasic rocks and the
Achankovil Zone: implications for the correlation of crustal blocks in southern India. generation of tonalitic–trondhjemitic–granodioritic (TTG) crust in the Archaean:
Gondwana Res. 10, 99–114. constraints from phase equilibrium modelling. Precambrian Res. 287, 73–90.

610
N. Sorcar et al. Geoscience Frontiers 11 (2020) 597–611

Palin, R.M., White, R.W., Green, E.C., Diener, J.F., Powell, R., Holland, T.J., 2016b. High- Stuck, T.J., Diener, J.F., 2018. Mineral equilibria constraints on open-system melting in
grade metamorphism and partial melting of basic and intermediate rocks. metamafic compositions. J. Metamorph. Geol. 36 (2), 255–281.
J. Metamorph. Geol. 34 (9), 871–892. Tadokoro, H., Tsunogae, T., Santosh, M., 2008. Metamorphic PT path of the eastern
Pant, N.C., Kundu, A., Joshi, S., Dey, A., Bhandari, A., Joshi, A., 2009. Chemical dating of Trivandrum Granulite Block, southern India: implications for regional correlation of
monazite–testing of an analytical protocol against independently dated standards. lower crustal fragments. J. Mineral. Petrol. Sci. 103, 279–284.
Indian Journal of Geoscience 63, 311–318. Tajcmanova, L., Connolly, J.A.D., Cesare, B., 2009. A thermodynamic model for titanium
Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., Maas, R., 2010. and ferric iron solution in biotite. J. Metamorph. Geol. 27 (2), 153–165.
Improved laser ablation U-Pb zircon geochronology through robust downhole Tang, L., Rajesh, S., Santosh, M., Tsunogae, T., Pradeepkumar, A.P., Tsutsumi, Y.,
fractionation correction. Geochem. Geophys. Geosyst. 11 (3), Q0AA06. Takamura, Y., 2018. Metamorphic phase equilibria modelling and zircon U–Pb
Pattison, D.R., Chacko, T., Farquhar, J., Mcfarlane, C.R., 2003. Temperatures of granulite- geochronology of ultrahigh-temperature cordierite granulites from the Madurai
facies metamorphism: constraints from experimental phase equilibria and Block, India: implications for hot Gondwana crust. Int. Geol. Rev. 60 (1), 21–42.
thermobarometry corrected for retrograde exchange. J. Petrol. 44, 867–900. Taylor, R.J., Clark, C., Fitzsimons, I.C., Santosh, M., Hand, M., Evans, N., McDonald, B.,
Peterson, J.W., Newton, R.C., 1989. Reversed experiments on biotite-quartz-feldspar 2014. Post-peak, fluid-mediated modification of granulite facies zircon and monazite
melting in the system KMASH: implications for crustal anatexis. J. Geol. 97 (4), in the Trivandrum Block, southern India. Contrib. Mineral. Petrol. 168, 1044.
465–485. Tchato, D.T., Schulz, B., Nzenti, J.P., 2009. Electron microprobe dating and
Peucat, J.-J., Jayananda, M., Chardon, D., Capdevila, R., Fanning, C.M., Paquette, J.-L., thermobarometry of Neoproterozoic metamorphic events in the Kekem area, central
2013. The lower crust of the Dharwar craton, southern India: patchwork of Archean African fold belt of Cameroon. Neues Jahrbuch fur Mineralogie - Abhandlungen 186
granulitic domains. Precambrian Res. 227, 4–28. (1), 95–109.
Plavsa, D., Collins, A.S., Foden, J.F., Kropinski, L., Santosh, M., Chetty, T.R.K., Clark, C., Teale, W., Collins, A.S., Foden, J., Payne, J.L., Plavsa, D., Chetty, T.R.K., Santosh, M.,
2012. Delineating crustal domains in Peninsular India: age and chemistry of Fanning, M., 2011. Cryogenian (~830 Ma) mafic magmatism and metamorphism in
orthopyroxene bearing felsic gneiss in Madurai Block. Precambrian Research 198– the northern Madurai Block, southern India: a magmatic link between Sri Lanka and
199, 77–93. Madagascar? J. Asian Earth Sci. 42, 223–233.
Plavsa, D., Collins, A.S., Foden, J.D., Clark, C., 2015. The evolution of a Gondwanan Tenczer, V., Hauzenberger, C., Fritz, H., Hoinkes, G., Muhongo, S., Klotzli, U., 2013.
collisional orogen: a structural and geochronological appraisal from the Southern Crustal age domains and metamorphic reqorking of the deep crust in Northern-
Granulite terrane, south India. Tectonics 34, 820–857. Central Tanzania: a U/Pb zircon and monazite age study. Mineral. Petrol. 107,
Ravindra Kumar, G.R., Sreejith, C., 2016. Petrology and geochemistry of charnockites 679–707.
(felsic ortho- granulites) from the Kerala Khondalite Belt, Southern India: evidence Tomson, J.K., Bhaskar Rao, Y.J., Vijaya Kumar, T., Choudhary, A.K., 2013. Geochemistry
for intra-crustal melting, magmatic differentiation and episodic crustal growth. Lithos and neodymium model ages of Precambrian charnockites, Southern Granulite
262, 334–354. https://doi.org/10.1016/j.lithos.2016. 07.009. Terrain, India: constraints on terrain assembly. Precambrian Res. 227, 295–315.
Rosenberg, C.L., Handy, M.L., 2005. Experimental deformation of partial melted granite Tucker, R., Roig, J.-Y., Delor, C., Amelin, Y., Goncalves, P., Rabarimanana, M.,
revisited: implications for the continental crust. J. Metamorph. Geol. 23, 19–28. Ralison, A., Belcher, R., 2011. Neoproterozoic extension in the Greater Dharwar
Rubatto, D., 2002. Zircon trace element geochemistry: Partitioning with garnet and the Craton: a reevaluation of the “Betsimisaraka suture” in Madagascar. Can. J. Earth Sci.
link between U-Pb ages and metamorphism. Chemical Geology 184 (1), 123–138. 48, 389–417.
Santosh, M., 1986. Carbonic metamorphism of charnockites in the southwestern Indian Tucker, R.D., Roig, J.Y., Macey, P.H., Delor, C., Amelin, Y., Armstrong, R.A.,
Shield: a fluid inclusion study. Lithos 19, 1–10. Rabarimanana, M.H., Ralison, A.V., 2011a. A new geological framework for south-
Santosh, M., 1987. Cordierite gneisses of southern Kerala, India: petrology, fluid central Madagascar, and its relevance to the “out-of-Africa” hypothesis. Precambrian
inclusions and implications for crustal uplift history. Contrib. Mineral. Petrol. 96, Res. 185, 109–130.
343–356. Tucker, R.D., Roig, J.Y., Moine, B., Delor, C., Peters, S.G., 2014. A geological synthesis of
Santosh, M., Yokoyama, K., Biju-Sekhar, S., Rogers, J.J.W., 2003. Multiple tectonothermal the Precambrian shield in Madagascar. J. Afr. Earth Sci. 94, 9–30.
events in the granulite blocks of southern India revealed from EPMA dating: Van der Molen, I., Paterson, M.S., 1979. Experimental deformation of partially-melted
implications on the history of supercontinents. Gondwana Res. 6, 29–63. granite. Contrib. Mineral. Petrol. 70, 299–318.
Santosh, M., Collins, A.S., Morimoto, T., Yokoyama, K., 2005. Depositional constraints Verma, S.K., Verma, S.P., Oliveira, E.P., Singh, V.K., Moreno, J.A., 2016. LA-SF-ICP-MS
and age of metamorphism in southern India: U–Pb chemical (EPMA) and isotopic zircon U–Pb geochronology of granitic rocks from the central Bundelkhand
(SIMS) ages from the Trivandrum Block. Geol. Mag. 142, 255–268. greenstone complex, Bundelkhand craton, India. J. Asian Earth Sci. 118, 125–137.
Santosh, M., Morimoto, T., Tsutsumi, Y., 2006a. Geochronology of the Khondalite belt of Vielzeuf, D., Clemens, J.D., 1992. The fluid-absent melting of phlogopite þ quartz:
Trivandrum Block, southern India: electron probe ages and implications for experiments and models. Am. Mineral. 77, 1206–1222.
Gondwana tectonics. Gondwana Res. 9, 261–278. Vielzeuf, D., Holloway, J.R., 1988. Experimental determination of the fluid-absent
Santosh, M., Collins, A.S., Tamashiro, I., Koshimoto, S., Tsutsumi, Y., Yokoyama, K., melting relations in the pelitic system. Contrib. Mineral. Petrol. 98, 257–276.
2006b. The timing of ultrahigh-temperature metamorphism in southern India: UThPb Vielzeuf, D., Montel, J.-M., 1994. Partial melting of metagreywackes; Part 1, Fluid-absent
electron microprobe ages from zircon and monazite in sapphirine-bearing granulites. experiments and phase relationships. Contrib. Mineral. Petrol. 117, 375–393.
Gondwana Res. 10, 128–155. Vielzeuf, D., Clemens, J.D., Pin, C., Moinet, E., 1990. Granites, granulites, and crustal
Santosh, M., Maruyama, S., Sato, K., 2009. Anatomy of a Cambrian suture in Gondwana: differentiation. In: Vielzeuf, D., Vidal, Ph (Eds.), Granulites and Crustal Evolution.
implications for the timing of Gondwana assembly. Gondwana Res. 16, 321–341. Kluwer Academic, Dordrecht Boston London, pp. 59–85.
Santosh, M., Ziao, W.J., Tsunogae, T., Chetty, T.R.K., Yellappa, T., 2012. The Wendt, I., Carl, C., 1991. The statistical distribution of the mean squared weighted
Neoproterozoic subduction complex in southern India: SIMS zircon U–Pb ages and deviation. Chem. Geol. Isot. Geosci. 86, 275–285.
implications for Gondwana assembly. Precambrian Res. 192–195, 190–208. White, R.W., Powell, R., 2002. Melt loss and the preservation of granulite facies mineral
Satish-Kumar, M., Harley, S.L., 1998. Reaction textures in scapolite-wollastonite-grossular assemblages. J. Metamorph. Geol. 20, 621–632.
calc-silicate rock from the Kerala Khondalite Belt, southern India: evidence for high White, R.W., Powell, R., Holland, T.J.B., Worley, B.A., 2000. The effect of TiO2 and Fe2O3
temperature metamorphism and initial cooling. Lithos 44, 83–99. on metapelitic assemblages at greenschist and amphibolite facies conditions: mineral
Sato, K., Santosh, M., Tsunogae, T., Chetty, T.R.K., Hirata, T., 2011. equilibria calculations in the system K2O–FeO–MgO–Al2O3–SiO2–H2O–TiO2–Fe2O3.
Subduction–accretion–collision history along the Gondwana suture in southern India: J. Metamorph. Geol. 18, 497–511.
a laser ablation ICP-MS study of zircon chronology. J. Asian Earth Sci. 40, 162–171. White, R.W., Powell, R., Holland, T.J.B., 2001. Calculation of partial melting equilibria in
Shabeer, K.P., Satish-Kumar, M., Armstrong, R., Buick, I.S., 2005. Constraints on the the system Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O (NCKFMASH). J. Metamorph.
timing of Pan-African granulite-facies metamorphism in the Kerala khondalite belt of Geol. 19, 139–153.
southern India: SHRIMP mineral ages and Nd isotopic systematics. J. Geol. 113, White, R.W., Powell, R., Clarke, G.L., 2002. The interpretation of reaction textures in Fe-rich
95–106. metapelitic granulites of the Musgrave Block, central Australia: constraints from mineral
Shimizu, H., Tsunogae, T., Santosh, M., 2009. Spinel þ quartz assemblage in granulites equilibria calculations in the system K2O–FeO–MgO–Al2O3–SiO2–H2O–TiO2–Fe2O3.
from the Achankovil Shear Zone, southern India: implications for ultrahigh- J. Metamorph. Geol. 20, 41–55.
temperature metamorphism. J. Asian Earth Sci. 36, 209–222. White, R.W., Powell, R., Halpin, J.A., 2004. Spatially-focussed melt formation in
Shiraishi, K., Dunkley, D.J., Hokada, T., Fanning, C.M., Kagami, H., Hamamoto, T., 2008. aluminous metapelites from Broken Hill, Australia. J. Metamorph. Geol. 22, 825–845.
Geochronological constraints on the Late Proterozoic to Cambrian crustal evolution White, R.W., Stevens, G., Johnson, T.E., 2011. Is the crucible reproducible? Reconciling
of eastern Dronning Maud Land, East Antarctica: a synthesis of SHRIMP U-Pb age and melting experiments with thermodynamic calculations. Elements 7, 241–246.
Nd model age data. Geological Society, London, Special Publications 308 (1), 21–67. Whitehouse, M.J., Kumar, G.R., Rimsa, A., 2014. Behaviour of radiogenic Pb in zircon
Skjerlie, K.P., Pati~
no Douce, A.E., Johnston, A.D., 1993. Fluid absent melting of a layered during ultrahigh-temperature metamorphism: an ion imaging and ion tomography
crustal protolith: implications for the generation of anatectic granites. Contrib. case study from the Kerala Khondalite Belt, southern India. Contrib. Mineral. Petrol.
Mineral. Petrol. 114, 365–378. 168, 1042.
Sorcar, N., Hoppe, U., Dasgupta, S., Chakraborty, S., 2014. High-temperature cooling Wiedenbeck, M., Alle, P., Corfu, F., et al., 1995. Three natural zircon standards for U–Th-
histories of migmatites from the High Himalayan Crystallines in Sikkim, India: rapid Pb, Lu–Hf, trace element and REE analyses. Geostandard Newzletter 19, 1–23.
cooling unrelated to exhumation? Contrib. Mineral. Petrol. 167, 1–34. Williams, M.L., Jercinovic, M.J., Hetherington, C.J., 2007. Microprobe monazite
Spear, F.S., Kohn, M.J., Cheney, J.T., 1999. P–T paths from anatectic pelites. Contrib. geochronology: understanding geologic processes by integrating composition and
Mineral. Petrol. 134, 17–32. chronology. Annu. Rev. Earth Planet Sci. 35, 137–175.
Stevens, G., Clemens, J.D., Droop, G.T.R., 1997. Melt production during granulite–facies Yakymchuk, C., Kirkland, C.L., Clark, C., 2018. Th/U ratios in metamorphic zircon.
anatexis: experimental data from “primitive” metasedimentary protoliths. Contrib. J. Metamorph. Geol. 36, 715–737.
Mineral. Petrol. 128, 352–370.

611

You might also like