Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-28T08:11:31.180Z Has data issue: false hasContentIssue false

The origin of the ultrahigh-pressure Tso Morari complex, NW Himalaya: implication for early Paleozoic rifting

Published online by Cambridge University Press:  25 January 2024

Takeshi Imayama*
Affiliation:
Research Institute of Frontier Science and Technology, Okayama University of Science, Okayama, Japan
Dripta Dutta
Affiliation:
Research Institute of Frontier Science and Technology, Okayama University of Science, Okayama, Japan Department of Earth Sciences, Indian Institute of Technology Kanpur, Kanpur, UP, India
Keewook Yi
Affiliation:
Geochronology Team, Korea Basic Science Institute, Ochang, Republic of Korea
*
Corresponding author: T. Imayama; Email: imayama@ous.ac.jp
Rights & Permissions [Opens in a new window]

Abstract

The origins and age distribution of the Himalayan high-pressure (HP) and ultrahigh-pressure (UHP) metamorphic rocks are critical for understanding the pre-Himalayan history. Although the protoliths to the UHP Tso Morari eclogites in Ladakh, NW Himalaya are believed to be the Permian Panjal volcanics, the geochronological evidence is absent. Here, we demonstrate that the protoliths of the UHP Tso Morari Complex formed in a continental rift setting at the Indian margin associated with the northern East Gondwana during the Early Paleozoic. Zircon U–Pb dates from eight gneisses and one garnet amphibolite indicate the Early Paleozoic bimodal magmatism of 493–476 Ma, which could be associated with the separation of South China from North India. Except for arc-related eclogites found in the Nidar ophiolite, the eclogites and amphibolites are rift-related, exhibiting enriched light rare earth elements and high concentrations of incompatible elements, along with evidence for crustal contamination. Our findings support the previously reported diversity in the sources and ages of the protoliths of the Himalayan HP–UHP metamorphic rocks along the orogen.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

High-pressure (HP) to ultrahigh-pressure (UHP) metamorphic rocks such as the eclogites are ubiquitous at the zones of oceanic plate subduction and continental collision in orogenic belts. They provide important geodynamic constraints on the regional tectono-metamorphic evolution. Numerous HP and UHP metamorphic rocks are present in the Tibet-Himalayan orogen and its vicinity (Fig. 1). The Himalayan eclogites can be broadly grouped as UHP coesite-bearing eclogites from the NW Himalaya (Tso Morari in Ladakh, NW India and Kaghan in NW Pakistan) (Guillot et al. Reference Guillot, de Sigoyer, Lardeaux and Mascle1997; Mukherjee & Sachan, Reference Mukherjee and Sachan2001, de Sigoyer et al. Reference de Sigoyer, Guillot and Dick2004, St-Onge et al. Reference St-Onge, Rayner, Palin, Searle and Waters2013; Rehman et al. Reference Rehman, Lee, Chung, Khan, O’Brien and Yamamoto2016) and HP eclogites from the central Himalaya Ama Drime Massif and Arun area in eastern Nepal (Lombardo & Rolfo, Reference Lombardo and Rolfo2000; Groppo et al. Reference Groppo, Lombardo, Rolfo and Pertusati2007; Cottle et al. Reference Cottle, Searle, Horstwood and Waters2009; Corrie et al. Reference Corrie, Kohn and Vervoort2010; Imayama et al. Reference Imayama, Uehara, Sakai, Yagi, Ikawa and Yi2020). Previous petrological and geochronological works demonstrate that the geothermal gradient of the UHP eclogites is significantly lower than that of their HP counterparts (Kohn, Reference Kohn2014; O’Brien, Reference O’Brien2019 and references therein). Although the differences in the origin of the two eclogite types have not received much attention, they are important to understand the pre-Himalayan geological history and the regional geological variations along the orogeny.

Figure 1. Topographic map (90 m SRTM SEM data) of the Himalaya-Tibet region illustrating the distributions of the HP-UHP metamorphic rocks (compiled from Zhang et al. Reference Zhang, Meng and Yang2005, Reference Zhang, Dong, Santosh and Zhao2014; Liou et al. Reference Liou, Ernst, Zhang, Tsujimori and Jahn2009; Yang et al. Reference Yang, Xu, Li, Xu, Li, Ren, Li, Chen and Robinson2009; Laskowski et al. Reference Laskowski, Kapp, Vervoort and Ding2016; Liang et al. Reference Liang, Wang, Yang, Ran, Zheng, Du and Li2017; Liu et al. Reference Liu, Zhang, Cao, Green, Yang, Xu, Liao and Kang2018, Reference Liu, Li, Xie, Santosh, Liu, Dong, Wang, Guo and Cao2022; Song et al. Reference Song, Bi, Qi, Yang, Allen, Niu, Su and Li2018; O’Brien, Reference O’Brien2019; Rehman, Reference Rehman2019). The major discontinuities (after Li et al. Reference Li, Wang, Dai, Xu, Hou and Li2015) and mountain peaks are also shown. Abbreviations (arranged alphabetically): AD – Ama Drime, ANP – Annapurna, CQT – Coqin Thrust, EV – Everest, GSCT – Gaize-Siling Co Thrust, KF – Karakoram fault, KGN – Kaghan, MBT – Main Boundary Thrust, MCT – Main Central Thrust, MFT – Main Frontal Thrust, NB – Namche Barwa, NP – Nanga Parbat, NQT – Northern Qaidam Thrust, SGAT – Shinquanhe-Gaize-Amdo Thrust, SQT – Southern Qiadam Thrust, SS – Sap-Shergole, STD – South Tibetan Detachment, TST – Tanggula Shan Thrust and UH – Ursi-Hinju.

Generally, the protoliths of the UHP eclogites in the NW Himalaya are thought to have been basalts associated with the Permian Panjal Traps formed in a continental rift setting (Spencer et al. Reference Spencer, Tonarini and Pognante1995; de Sigoyer et al. Reference de Sigoyer, Guillot and Dick2004). In the NW Himalaya, the Panjal volcanics regionally intruded into the High-Tethyan Himalayas (Shellnutt, Reference Shellnutt2016). The Kaghan eclogites (259±10 Ma) and associated gneisses in NW Pakistan are Early Permian in age (Rehman et al. Reference Rehman, Lee, Chung, Khan, O’Brien and Yamamoto2016). The Tso Morari eclogites and associated amphibolites also exhibit geochemical signatures of a rift environment (Rao & Rai, Reference Rao and Rai2006; Jonnalagadda et al. Reference Jonnalagadda, Karmalkar and Duraiswami2019). On the other hand, recent Nd-Sr isotope data (Ahmad et al. Reference Ahmad, Bhat, Tanaka, Bickle, Asahara, Chapman and Sachan2022) proposed their origin of ca. 289 Ma depleted Panjal volcanics and ca. 140 Ma Ladakh (Nidar) ophiolitic mafic rocks instead of the previously accepted origin from the enriched Panjal (ca. 289 Ma) and Phe volcanics (Zanskar). However, the only pre-Himalayan ages of ca. 479 Ma in the Tso Morari Crystallines (TMC) are reported for the undeformed Polokongka La granite and gneiss (Girard & Bussy, Reference Girard and Bussy1999), and the lack of geochronological study hinders our understanding of the origin of the UHP metamorphic rocks in the TMC. Here, we present the whole-rock geochemical analyses and zircon SHRIMP dating of the TMC gneisses and metabasites including eclogites. Our new data indicate that the UHP TMC formed in a continental rift setting during the Early Paleozoic. The results also demonstrate the diversity in the sources and ages of the HP-UHP Himalayan eclogites along the E-W direction of the orogen.

2. Geological setting

The TMC is located in the northwestern part of Indian Trans-Himalaya and is characterised by a sub-elliptical outline (Fig. 2a). It occurs as a northwesterly trending antiformal dome, separated from the Indus-Tsangpo suture including the Nidar ophiolite to the north and the Tetraogal Nappe to the south by detachment faults that dip away from the core of the dome. Further to the south, lies the Mata Nappe, which consists of two Early Paleozoic granite intrusions viz. Rupshu (ca. 482 Ma) and Nyimaling (ca. 460 Ma) (Epard & Steck, Reference Epard and Steck2008). The Tso Morari Gneiss forms the core of the TMC and is overlain by a metasedimentary sequence (Fig. 2a, (Buchs & Epard, Reference Buchs and Epard2019). The metabasites including eclogites occur as foliation-parallel boudins (Fig. 2b, c) within the ductilely deformed gneisses (Fig. 2d). Eclogites have been widely retrogressed to amphibolites (Fig. 2e). The tectono-thermal evolution of the TMC can be broadly classified into four stages: (i) prograde metamorphism during Neo-Tethys subduction, (ii) (U)HP metamorphism under eclogite-facies condition (>2 GPa, 500–760°C) at ca. 55 Ma, (iii) near-isothermal decompression under granulite-facies condition at the mid-crustal level at ca. 48–45 Ma, and, finally, (iv) retrograde amphibolite-facies conditions at ca. 31–29 Ma (Guillot et al. Reference Guillot, de Sigoyer, Lardeaux and Mascle1997; St-Onge et al. Reference St-Onge, Rayner, Palin, Searle and Waters2013; O’Brien, Reference O’Brien2019 and references therein). The Early Eocene Indo-Eurasia collision facilitated TMC extrusion along a northeasterly dipping channel (Dutta & Mukherjee, Reference Dutta and Mukherjee2021). At least three deformation phases affected the TMC during extrusion and produced the gently dipping gneissic foliations (de Sigoyer et al. Reference de Sigoyer, Guillot and Dick2004; Epard & Steck, Reference Epard and Steck2008).

Figure 2. Study area, outcrop photographs and photomicrographs. (a) Geological map of the Tso Morari region (reproduced from Epard and Steck, Reference Epard and Steck2008). Boudins of (b) eclogite and (c) K-rich metabasite in the Tso Morari gneiss parallel to the gneissic foliation. This photo was clicked at the same location as that of Fig. 3 by St-Onge et al. (Reference St-Onge, Rayner, Palin, Searle and Waters2013). The large eclogite boudin is about 4 m thick at its thickest portion. (d) Top-to-the SE ductile shear sense exhibited by a feldspar augen in the quartzofeldspathic gneiss. (e) Retrograded Grt amphibolite. (f) Augen-shaped aggregates of quartz and feldspar within the fine-grained quartzofeldspathic matrix. The mica grains of the gneiss define the foliations in (g) and (h). (i) Prismatic zoisite grain within the coarser phengite grains of the gneiss. (j) Quartz porphyroclasts are common in some gneisses. Photomicrographs of the metabasites (k-m) and Grt amphibolite (n).

Gneisses and metabasites samples were collected for petrographic, geochemical and geochronogical analyses (see Supplementary Materials Table S1 for sample details). Recently, Pan et al. (Reference Pan, Macris and Menold2023) investigated the fluid evolution of the TMC since the Eocene, and they sampled a traverse from the centre of an eclogite boudin out into the host orthogneiss. In this study, we focus on the protoliths of the TMC, and thus samples were not collected from the reaction zone to avoid the effects due to the metasomatism. Quartz and calcite veins are found in the outcrops but fist-sized samples without the veins were selected for whole-rock analyses to avoid the effect of veins. Gneisses consist mainly of Ph + Bt + Pl + Qz ± Mc ± Grt ± Zo (Fig. 2f–j, mineral abbreviations are as per Whitney & Evans, Reference Whitney and Evans2010). They are subdivided into Ph-Bt gneiss, Grt-Ph gneiss, Ph-rich gneiss and quartzofeldspathic gneiss depending on the modal amount of minerals and presence of garnet. Most of the gneisses are orthogneisses whereas the quartzofeldspathic gneisses may be paragneisses in part. The metabasites include (retrograded) eclogites, Grt amphibolites and K-rich metabasites. Eclogite-facies metamorphism is characterised by the mineral assemblage of Grt + Omp + Rt + Coe/Qz ± Ph ± Zo (Fig. 2k–m) with later phases of Amp + Pl ±Di. The eclogites have zoning patterns in garnets associated with prograde and peak metamorphism but compositions and textures related to metasomatism stages are rare (e.g., St-Onge et al. Reference St-Onge, Rayner, Palin, Searle and Waters2013). The Grt amphibolites and K-rich metabasites mainly consist of Grt + Amp + Pl + Qz + Ilm + Ttn (Fig. 2n) and Grt + Bt + Ph + Amp + Pl + Qz + Rt + Zo ± Ttn, respectively.

3. Major and trace element geochemistry

The ten metabasite samples have low SiO2 (44.09–51.20 wt%) and MgO (5.23–9.40 wt%), high TiO2 (1.09–3.12 wt%) relative to FeO/MgO ratio and variable alkali contents (1.61–7.21 wt%) (Supplementary Materials Table S2 and Figure S1). Unlike the retrograded eclogites and K-rich metabasites, the Grt amphibolites and eclogite boulder have low alkali contents. The analysed samples plot on the field of subalkaline basalts in the Zr/TiO2 vs. Nb/Y diagram, and the eclogite boulder has a low Nb/Y ratio (Supplementary Materials Figure S1).

The large ion lithophile elements (e.g., Rb and Ba) concentrations are variable (Fig. 3a, b) due to their mobility during surface weathering and alteration processes (Xia & Li, Reference Xia and Li2019). Thus it is important to check element mobility before applying tectonic discrimination diagrams (Polat & Hofmann, Reference Polat and Hofmann2003; Imayama et al. Reference Imayama, Oh, Jeon and Yi2021). The eclogites and amphibolites lack large Ce anomalies as shown by the Ce* values between 0.96 and 1.04 (Ce* = CeN/Sqrt(LaN×PrN), which is considered as immobile when the range is 0.9<Ce*<1.1. There is also no significant carbonisation or silicification due to the absence of carbonate and sulfide minerals in samples. These occurrences imply that the major chemical compositions of the sampled eclogites and amphibolites have not been strongly affected by hydrothermal alteration (Polat & Hofmann, Reference Polat and Hofmann2003; Imayama et al. Reference Imayama, Oh, Jeon and Yi2021).

Figure 3. Spider diagrams of (a) retrograde eclogite and K-rich metabasite and (b) Grt amphibolite and eclogite boulder. Chondrite-normalised rare earth element patterns for (c) retrograde eclogite and K-rich metabasite and (d) Grt amphibolite and eclogite boulder. The compositions of metabasites from the TMC plotted in (e) Nb*2–Zr/4–Y (Meschede, Reference Meschede1986), (f) Zr/Y vs. Zr (Pearce & Norry, Reference Pearce and Norry1979), (g) Nb/La vs. Nb and (h) Y/Nb vs. Zr/Nb diagrams (Wilson, Reference Wilson1989, updated by Xia & Li, Reference Xia and Li2019). Data sources of the N-MORB, E-MORB, mantle plume and depleted asthenosphere compositions are from Sun & McDonough (Reference Sun and McDonough1989) and Salters and Stracke (Reference Salters and Stracke2004) with the PetDB database (http://www.earthchem.org/petdb). A(22): Ahmad et al. (Reference Ahmad, Bhat, Tanaka, Bickle, Asahara, Chapman and Sachan2022), J(19): Jonnalagadda et al. (Reference Jonnalagadda, Karmalkar and Duraiswami2019), R&R(06): Rao and Rai (Reference Rao and Rai2006). WPA: within-plate alkali basalts, WPT: within-plate tholeiites, VAB: volcanic-arc basalts.

All the metabasite samples are enriched in high field strength elements (HFSE) with Nb-Ta negative anomalies, and their chondrite-normalised rare earth element (REE) patterns plot in the fields between oceanic island basalt (OIB) and enriched mid-ocean ridge basalt (E-MORB, Fig. 3a–d). The HFSE and REE contents are lowest in the eclogite boulder along with prominent Nb and Zr–Hf negative anomalies. In the Nb–Zr–Y diagram, the metabasites, except for the eclogite boulder, plot in the fields of within-plate tholeiite or volcanic-arc basalts (Fig. 3e). In the Zr/Y–Zr diagram (Fig. 3f), these metabasites have high Zr/Y representing the within-plate basalts and are thus distinguished from the island-arc origin. In contrast, the eclogite boulder plots in the fields where island-arc basalt and MORB overlap on both diagrams (Fig. 3e, f).

4. Zircon U-Pb geochronology

We separated the zircons from eight gneisses and one Grt amphibolite (Fig. 2a, Supplementary Materials Table S1) to obtain cathodoluminescence (CL) images and U–Pb ages (see Supplementary Materials Text S1). The CL images of the zircon grains from the gneisses show well-developed prismatic faces and internal oscillatory zoning, which is typical igneous-type zircons (Fig. 4). Some grains have modified grey-CL rims showing faint oscillatory zoning or patchy pattern of dark and bright colours in CL-image (Supplementary Materials Figure S2). Thin metamorphic rims showing dark-CL domains are observed in 19-2 and 19-4, but they are too thin to analyse. The U–Pb analyses of two Ph-Bt gneisses (15-3B and 9g) concentrated at a single population of concordant to near-concordant grains with a weighted mean 206Pb/238U date of 481.2±4.4 Ma (n = 14, mean squared weighted deviation or MSWD = 2.5, Fig. 4a) and 484.3±5.7 Ma (n = 14, MSWD = 2.1, Supplementary Materials Figure S3a), respectively. Some grains were possibly affected by Pb-loss resulting in young 206Pb/238U ages. The inherited grains yielded scattered dates ranging from 1402 Ma to 606 Ma. The zircons from a Grt-Ph gneiss (18-3B) yielded 206Pb/238U dates ranging from 500 to 470 Ma, but the weighted mean date was not calculated due to the large MSWD value (Supplementary Materials Figure S3b). The U–Pb analyses of concordant grains from the other two Grt-Ph gneisses (16-2A and 18-4) yielded the weighted mean 206Pb/238U date of 485.9±3.3 Ma (n = 12, MSWD = 1.7, Fig. 4b) and 493.4±5.2 Ma (n = 13, MSWD = 1.6, Supplementary Materials Figure S3c), respectively. Some inherited grains observed in three Grt-Ph gneisses yield 206Pb/238U dates of 1841 to 796 Ma. The igneous zircons of 15 concordant grains from Ph-rich gneiss (18-5B) yielded the weighted mean 206Pb/238U date of 493.3±5.4 Ma (n = 15, MSWD = 2.1, Supplementary Materials Figure S3d). The inherited domains gave 206Pb/238U dates of 1154 to 766 Ma. Some grains from two quartzofeldspathic gneisses (19-2 and 19-4) show discordant behaviour due to the disturbance of common Pb and Pb loss. The U–Pb analyses of the crystals that yielded concordant dates gave weighted mean 206Pb/238U dates of 480.4±4.4 Ma (n = 18, MSWD = 1.8, Fig. 4c) and 485.6±4.5 Ma (n = 14, MSWD = 3.0, Supplementary Materials Figure S3e), respectively.

Figure 4. Concordia diagrams from the SHRIMP zircon U–Pb analyses of (a) Ph-Bt gneiss (15-3B), (b) Grt-Ph gneiss (16-2A), (c) Quartzofeldspathic gneiss (19-2) and (d) Grt amphibolite (20-2) from the TMC. All error ellipses and weighted mean 206Pb/238U ages are quoted at the 1σ and 2σ levels, respectively. The white bars below the zircons are 20 μm long.

The zircon grains from Grt amphibolite (20-2) show well-developed prismatic faces and show bright-CL oscillatory zoning surrounding the inherited core (Fig. 4d). The igneous zircons of 13 concordant grains yielded the weighted mean 206Pb/238U date of 475.8±6.9 Ma (MSWD = 2.6, Fig. 4d). The inherited domains gave 206Pb/238U dates of 1039 to 813 Ma.

5. Discussion

Previous workers suggested that the protoliths of the Tso Morari eclogites are either the Permian Panjal basalts associated with a mantle plume and crustal contamination (Spencer et al. Reference Spencer, Tonarini and Pognante1995; de Sigoyer et al. Reference de Sigoyer, Guillot and Dick2004; Jonnalagadda et al. Reference Jonnalagadda, Karmalkar and Duraiswami2019) or the Late Jurassic to Early Cretaceous Ladakh ophiolitic mafic rocks in the supra-subduction zone (Ahmad et al. Reference Ahmad, Bhat, Tanaka, Bickle, Asahara, Chapman and Sachan2022). The metabasites we analysed have enriched light REE contents (Fig. 3c,d), and the mixing line for the lithospheric contamination in the Y/Nb–Zr/Nb diagram (Fig. 3h) starts from the mantle plume source, rather than the depleted MORB on the mantle array, precluding the possibility of N-MORB components in the mantle source. Although the enriched light REE patterns are comparable to both E-MORB and OIB, the concentrations of incompatible trace elements such as the HFSE are much higher than those of the E-MORB and resemble those of OIB (Fig. 3a, b). The signatures of within-plate basalts are marked by the high Zr/Y content in most metabasites (Fig. 3f). Although the Nb-Ta negative anomalies (Fig. 3a, b) and Nb/La<1 (Fig. 3g) in most of the TMC metabasites imply either continental intraplate basalts that experienced crustal contamination or island-arc basalts as their protoliths, the former is supported by high concentrations of incompatible trace elements (Xia and Li, Reference Xia and Li2019). Crustal contamination leads to low Nb content, increasing Y/Nb and Zr/Nb ratios (Fig. 3h). The trend has a different slope from that of the mantle array line from the plume, through MORB, to the depleted asthenosphere. Based on field occurrences, the contaminant lithology could partially consist of the Early Paleozoic orthogneisses surrounding metabasites. The Zr/Nb ratios from a few orthogneisses reach up to 30 (Ahmad et al. Reference Ahmad, Bhat, Tanaka, Bickle, Asahara, Chapman and Sachan2022), which exceed the average Zr/Nb ratios of around 10 for continental crust compositions (e.g., Wedepohl Reference Wedepohl1995; Rudnick & Gao, Reference Rudnick, Gao, Holland and Turekian2003). The crustal contamination is also supported by the presence of the inherited zircon cores in the garnet amphibolite (sample 20-2). Therefore, the TMC metabasites most likely formed in a continental rift setting and experienced lithospheric contamination. In contrast, the eclogite boulder with prominent Nb and Zr–Hf negative anomalies suggests its island-arc origin, which could correspond to the Nidar ophiolite mafic rocks formed by the subduction initiation in the forearc (Ahmad et al. Reference Ahmad, Bhat, Tanaka, Bickle, Asahara, Chapman and Sachan2022).

The U–Pb SHRIMP data from the eight gneisses suggest an Early Paleozoic (ca. 493–480 Ma) protolith, which is consistent with the previously reported U–Pb zircon date of 479±2 Ma from the Tso Morari Gneiss (Girard & Bussy, Reference Girard and Bussy1999). The youngest zircon date of 475.8±6.9 Ma was obtained from the Grt amphibolite, which represents a retrograde product of the eclogites. Oscillatory zoning is a common feature in zircons from felsic igneous rocks, but some zircon textures in amphibolites from the orogenic belts also include oscillatory zoning (Oh et al. Reference Oh, Imayama, Jeon and Yi2017; Kang et al. Reference Kang, Li, Kang, Dong, Jiang, Liang and Dong2020), representing the timing of mafic magmatism. It is unlikely that contamination of zircons from orthogneisses occurred during intrusion because the garnet amphibolite occurs within the paragneisses, not orthogneisses. The zircons may have crystalised from basaltic magmas if the rocks displayed high Zr contents (Shao et al. Reference Shao, Xia, Ding, Cai and Song2019). Therefore, the ca. 476 Ma zircon date from garnet amphibolite is interpreted as the timing of mafic magmatism and is roughly consistent with those of the gneisses within error, indicating the bimodal magmatism during the Early Paleozoic at least for some time. In terms of trace element geochemistry, continental basalt is similar to the OIB but the presence of coherent granites implies the tectonic setting of continental rift. These Early Paleozoic ages indicate that the TMC eclogites were not derived from the Permian Panjal Traps associated with the opening of the Neo-Tethys Ocean. The crystallisation ages (ca. 493 Ma) from the central part of the TMC near Puga are slightly older than those (ca. 486–476 Ma) in surrounding parts, implying multi-stage magmatism. Felsic rocks volumetrically dominate the TMC, whereas the volume of mafic rocks is quite small. This is also in contrast to what is observed in the NW Himalaya where the Panjal Traps are characterised by abundant mafic magmatism (Shellnutt, Reference Shellnutt2016). We alternatively propose that the protoliths of the UHP TMC formed in the extensional setting at the rifted Indian margin during the Early Paleozoic. During the Pan-African orogeny, the Indian shield was located in the northern East Gondwana (Gray et al. Reference Gray, Foster, Meert, Goscombe, Armstrong, Trouw and Passchier2008, Fig. 5). Epard and Steck (Reference Epard and Steck2008) pointed out that the sequences deposited in the north Indian margin in the Tso Morari region are affected by extensional structures, rather than the compressional structures of the Pan-African orogeny. The Early Paleozoic tectonic events are widely known in the Himalayas (Le Fort & Cronin, Reference Le Fort and Cronin1988; Gehrels et al. Reference Gehrels, DeCelles, Ojha and Upreti2006), but the investigation of the detailed tectonic evolution is beyond the scope of this study as here we only discuss the Early Paleozoic igneous events in the NW Himalaya. The report of A-type alkaline Kaghan metagranite of ca. 470 Ma (Trivedi et al. Reference Trivedi, Sharma and Gopalan1986) also supports the Early Paleozoic rifting at the passive margin of northern India. The Early Paleozoic felsic magmatism is also exposed in Pakistan (483–476 Ma, Ogasawara et al. Reference Ogasawara, Fukuyama, Siddiqui and Zhao2019; ca. 459 Ma, Mughal et al. Reference Mughal, Zhang, Hussain, Rehman, Du and Hameed2022) and in the Garhwal, NW India (495–490 Ma, Imayama et al. Reference Imayama, Bose, Yi, Jeong, Horie, Takehara and Kawabata2023; ca. 512 Ma, Sen et al. Reference Sen, Sen, Chatterjee, Choudhary and Dey2021), which are considered as S-type and A-type, respectively. Thus, the A-type Early Paleozoic granites in the Garhwal are more closely related to the Early Paleozoic rifting events in this study. Coupled with geochemical results in this study, we posit that the Early Paleozoic rifting after the Neoproterozoic Pan-African orogeny produced the protoliths of the TMC. Recently, a close relationship between North India and South China during the Neoproterozoic to Early Paleozoic has been discussed (Qi et al. Reference Qi, Cawood, Xu, Du, Zhang and Zhang2020; Imayama et al. Reference Imayama, Bose, Yi, Jeong, Horie, Takehara and Kawabata2023), and the Early Paleozoic rifting in NW India led to the separation of South China from North India (Fig. 5, Imayama et al. Reference Imayama, Bose, Yi, Jeong, Horie, Takehara and Kawabata2023). However, more study is required to confirm this assumption.

Figure 5. Map of Gondwana showing the positions of the Indian Shield and the rifting event inferred from this study. Modified from Gray et al. (Reference Gray, Foster, Meert, Goscombe, Armstrong, Trouw and Passchier2008), Meert and Lieberman (Reference Meert and Lieberman2008) and Imayama et al. (Reference Imayama, Bose, Yi, Jeong, Horie, Takehara and Kawabata2023). SF, São Francisco Craton; RP, Río de la Plata Craton.

Combined with the UHP Nanga Parbat eclogites (Rehman et al. Reference Rehman, Lee, Chung, Khan, O’Brien and Yamamoto2016), our results suggest multiple sources and ages for the protolith of the UHP metamorphic rocks in the NW Himalaya. Similarly, the protolith of the HP metamorphic rocks in the central Himalaya also originate from multiple sources and ages (Zhang et al. Reference Zhang, Wang, Webb, Zhang, Liu, Fu, Wu and Wang2022). The HP eclogites of the Ama Drime Massif originated from the Ordovician (∼480–430 Ma) rift-related or E-MORB-like magmatism (Wang et al. Reference Wang, Zhang, Li and Somerville2017; Dong et al. Reference Dong, Zhang, Tian, Niu and Zhang2022) and the Paleoproterozoic (∼1850 Ma) continental flood basalts (Zhang et al. Reference Zhang, Wang, Webb, Zhang, Liu, Fu, Wu and Wang2022). In Bhutan, the HP metabasite was derived from young Paleoproteorozoic rifting (∼1742 Ma, Chakungal et al. Reference Chakungal, Dostal, Grujic, Duchêne and Ghalley2010). These occurrences indicate the sources and ages of the protoliths of the Himalayan HP–UHP metamorphic rocks are diverse in the NW-SE direction along the orogen.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756824000025

Acknowledgements

The research was supported by the Japan Society for the Promotion of Science (22H01324) and the Korea Basic Science Institute under the R&D programme (C330430). We are grateful to Shinae Lee for the analytical work. We thank Dr. Johannes Pohlner and the anonymous reviewer for constructive and critical reviews that significantly helped to improve the manuscript. We also thank Dr. Simon Schorn for their careful editorial handling.

References

Ahmad, T, Bhat, IM, Tanaka, T, Bickle, M, Asahara, Y, Chapman, H and Sachan, HK (2022) Tso Morari Eclogites, Eastern Ladakh: isotopic and elemental constraints on their Protolith, genesis, and tectonic setting. The Journal of Geology 130, 231–52.CrossRefGoogle Scholar
Buchs, N and Epard, J-L (2019) Geology of the eastern part of the Tso Morari nappe, the Nidar Ophiolite and the surrounding tectonic units (NW Himalaya, India). Journal of Maps 15, 3848.Google Scholar
Chakungal, J, Dostal, J, Grujic, D, Duchêne, S and Ghalley, KS (2010) Provenance of the Greater Himalayan sequence: evidence from mafic granulites and amphibolites in NW Bhutan. Tectonophysics 480, 198212.CrossRefGoogle Scholar
Corrie, SL, Kohn, MJ and Vervoort, JD (2010) Young eclogite from the-Greater Himalayan Sequence, Arun Valley, eastern Nepal: P–T–t path and tectonic implications. Earth and Planetary Science Letters 289, 406–16.Google Scholar
Cottle, JM, Searle, MP, Horstwood, MSA and Waters, DJ (2009) Timing of Midcrustal Metamorphism, melting, and deformation in the Mount Everest region of Southern Tibet revealed by U(-Th)-Pb geochronology. The Journal of Geology 117, 643–64.Google Scholar
Dong, X, Zhang, Z, Tian, Z, Niu, Y and Zhang, L (2022) Protoliths and metamorphism of the central Himalayan eclogites: Zircon/titanite U–Pb geochronology, Hf isotope and geochemistry. Gondwana Research 104, 3953.Google Scholar
Dutta, D and Mukherjee, S (2021) Extrusion kinematics of UHP terrane in a collisional orogen: EBSD and microstructure-based approach from the Tso Morari Crystallines (Ladakh Himalaya). Tectonophysics 800, 228641.Google Scholar
Epard, J-L and Steck, A (2008) Structural development of the Tso Morari ultra-high pressure nappe of the Ladakh Himalaya. Tectonophysics 451, 242–64.CrossRefGoogle Scholar
Gehrels, GE, DeCelles, PG, Ojha, TP and Upreti, BN (2006) Geologic and U-Th-Pb geochronologic evidence for early Paleozoic tectonism in the Kathmandu thrust sheet, central Nepal Himalaya. Geological Society of America Bulletin 118, 185–98.Google Scholar
Girard, M and Bussy, F (1999) Late Pan-African magmatism in the Himalaya: new geochronological and geochemical data from the Ordovician Tso Morari metagranites (Ladakh, NW India). Schweizerische mineralogische und petrographische Mitteilungen 79, 399418.Google Scholar
Gray, DR, Foster, DA, Meert, JG, Goscombe, BD, Armstrong, R, Trouw, RAJ and Passchier, CW (2008) A Damara orogen perspective on the assembly of southwestern Gondwana. Geological Society, London, Special Publications 294, 257–78.Google Scholar
Groppo, C, Lombardo, B, Rolfo, F and Pertusati, P (2007) Clockwise exhumation path of granulitized eclogites from the Ama Drime range (Eastern Himalayas). Journal of Metamorphic Geology 25, 5175.CrossRefGoogle Scholar
Guillot, S, de Sigoyer, J, Lardeaux, JM and Mascle, G (1997) Eclogitic metasediments from the Tso Morari area (Ladakh, Himalaya): evidence for continental subduction during India-Asia convergence. Contributions to Mineralogy and Petrology 128, 197212.Google Scholar
Imayama, T, Uehara, S, Sakai, H, Yagi, K, Ikawa, C and Yi, K (2020) The absence of high-pressure metamorphism in the inverted Barrovian metamorphic sequences of the Arun area, eastern Nepal and its tectonic implication. International Journal of Earth Sciences 109, 465–88.Google Scholar
Imayama, T, Oh, CW, Jeon, J and Yi, K (2021) Neoproterozoic and middle Paleozoic geological events in the eastern Wolhyeonri complex of the southwestern Gyeonggi Massif, South Korea, and their tectonic correlations in northeastern Asia. Lithos 382–383, 105923.Google Scholar
Imayama, T, Bose, N, Yi, K, Jeong, Y-J, Horie, K, Takehara, M and Kawabata, R (2023) Zircon U–Pb, Hf, and O isotopic constraints on the tectonic affinity of the basement of the Himalayan orogenic belt: insights from metasedimentary rocks, orthogneisses, and leucogranites in Garhwal, NW India. Precambrian Research 397, 107183.Google Scholar
Jonnalagadda, MK, Karmalkar, NR and Duraiswami, RA (2019) Geochemistry of eclogites of the Tso Morari complex, Ladakh, NW Himalayas: insights into trace element behavior during subduction and exhumation. Geoscience Frontiers 10, 811–26.Google Scholar
Kang, W, Li, W, Kang, L, Dong, Y, Jiang, D, Liang, J and Dong, H (2020) Metamorphism and geochronology of garnet amphibolite from the Beishan Orogen, southern Central Asian Orogenic Belt: Constraints from P-T path and zircon U-Pb dating. Geoscience Frontiers 11, 1189–201.Google Scholar
Kohn, MJ (2014) Himalayan metamorphism and its tectonic implications. Annual Review of Earth and Planetary Sciences 42, 381419.CrossRefGoogle Scholar
Laskowski, AK, Kapp, P, Vervoort, JD and Ding, L (2016) High-pressure Tethyan Himalaya rocks along the India-Asia suture zone in southern Tibet. Lithosphere 8, 574–82.Google Scholar
Le Fort, P and Cronin, VS (1988) Granites in the tectonic evolution of the Himalaya, Karakoram and southern Tibet [and discussion]. Philosophical Transactions of the Royal Society of London Series A 326, 281–99.Google Scholar
Li, Y, Wang, C, Dai, J, Xu, G, Hou, Y and Li, X (2015) Propagation of the deformation and growth of the Tibetan–Himalayan orogen: a review. Earth-Science Reviews 143, 3661.CrossRefGoogle Scholar
Liang, X, Wang, G, Yang, B, Ran, H, Zheng, Y, Du, J and Li, L (2017) Stepwise exhumation of the Triassic Lanling high-pressure metamorphic belt in Central Qiangtang, Tibet: insights from a coupled study of metamorphism, deformation, and geochronology. Tectonics 36, 652–70.CrossRefGoogle Scholar
Liou, JG, Ernst, WG, Zhang, RY, Tsujimori, T and Jahn, BM (2009) Ultrahigh-pressure minerals and metamorphic terranes – the view from China. Journal of Asian Earth Sciences 35, 199231.CrossRefGoogle Scholar
Liu, L, Zhang, J-F, Cao, Y-T, Green, HW, Yang, W-Q, Xu, H-J, Liao, X-Y and Kang, L (2018) Evidence of former stishovite in UHP eclogite from the South Altyn Tagh, western China. Earth and Planetary Science Letters 484, 353–62.Google Scholar
Liu, Y, Li, S, Xie, C, Santosh, M, Liu, Y, Dong, Y, Wang, B, Guo, R and Cao, X (2022) Subduction–collision and exhumation of eclogites in the Lhasa terrane, Tibet Plateau. Gondwana Research 102, 394404.Google Scholar
Lombardo, B and Rolfo, F (2000) Two contrasting eclogite types in the Himalayas: implications for the Himalayan orogeny. Journal of Geodynamics 30, 3760.CrossRefGoogle Scholar
Meert, JG and Lieberman, BS (2008) The Neoproterozoic assembly of Gondwana and its relationship to the Ediacaran–Cambrian radiation. Gondwana Research 14, 521.CrossRefGoogle Scholar
Meschede, M (1986) A method of discriminating between different types of mid-ocean ridge basalts and continental tholeiites with the Nb-Zr-Y diagram. Chemical Geology 56, 207–18.CrossRefGoogle Scholar
Mughal, MS, Zhang, C, Hussain, A, Rehman, HU, Du, D and Hameed, F (2022) Pre-Himalayan crustal growth in the Indian plate: implications from U-Pb dating, geochemical and Sr-Nd isotopic systematics. Precambrian Research 383, 106920.Google Scholar
Mukherjee, BK and Sachan, HK (2001) Discovery of coesite from Indian Himalaya: a record of ultra-high pressure metamorphism in Indian Continental Crust. Current Science 81, 1358–61.Google Scholar
O’Brien, PJ (2019) Eclogites and other high-pressure rocks in the Himalaya: a review. Geological Society, London, Special Publications 483, 183213.Google Scholar
Ogasawara, M, Fukuyama, M, Siddiqui, RH and Zhao, Y (2019) Origin of the Ordovician Mansehra granite in the NW Himalaya, Pakistan: constraints from Sr–Nd isotopic data, zircon U–Pb age and Hf isotopes. Geological Society, London, Special Publications 481, 277–98.Google Scholar
Oh, CW, Imayama, T, Jeon, J and Yi, K (2017) Regional Middle Paleozoic metamorphism in the southwestern Gyeonggi Massif, South Korea: its implications for tectonics in Northeast Asia. Journal of Asian Earth Sciences 145, 542–64.Google Scholar
Pan, R, Macris, CA and Menold, CA (2023) Fluid evolution during burial and exhumation of the Tso Morari UHP complex, NW India: constraints from mineralogy, geochemistry, and thermodynamic modeling. Contributions to Mineralogy and Petrology 178, 3.CrossRefGoogle Scholar
Pearce, JA and Norry, MJ (1979) Petrogenetic implications of Ti, Zr, Y, and Nb variations in volcanic rocks. Contributions to Mineralogy and Petrology 69, 3347.CrossRefGoogle Scholar
Polat, A and Hofmann, AW (2003) Alteration and geochemical patterns in the 3.7–3.8 Ga Isua greenstone belt, West Greenland. Precambrian Research 126, 197218.Google Scholar
Qi, L, Cawood, PA, Xu, Y, Du, Y, Zhang, H and Zhang, Z (2020) Linking South China to North India from the late Tonian to Ediacaran: constraints from the Cathaysia Block. Precambrian Research 350, 105898.Google Scholar
Rao, DR and Rai, H (2006) Signatures of rift environment in the production of garnet-amphibolites and eclogites from Tso-Morari region, Ladakh, India: a geochemical study. Gondwana Research 9, 512–23.Google Scholar
Rehman, HU (2019) Geochronological enigma of the HP–UHP rocks in the Himalayan orogen. Geological Society, London, Special Publications 474, 183207.CrossRefGoogle Scholar
Rehman, HU, Lee, H-Y, Chung, S-L, Khan, T, O’Brien, PJ and Yamamoto, H (2016) Source and mode of the Permian Panjal Trap magmatism: evidence from zircon U–Pb and Hf isotopes and trace element data from the Himalayan ultrahigh-pressure rocks. Lithos 260, 286–99.CrossRefGoogle Scholar
Rudnick, RL and Gao, S (2003) 3.01 – composition of the continental crust. In Treatise on Geochemistry (eds Holland, HD & Turekian, KK), pp. 1–64. Oxford: Elsevier-Pergamon.Google Scholar
Salters, VJM and Stracke, A (2004) Composition of the depleted mantle. Geochemistry, Geophysics, Geosystems 5, Q05B07.CrossRefGoogle Scholar
Sen, A, Sen, K, Chatterjee, A, Choudhary, S and Dey, A (2021) Understanding pre- and syn-orogenic tectonic evolution in western Himalaya through age and petrogenesis of Palaeozoic and Cenozoic granites from upper structural levels of Bhagirathi Valley, NW India. Geological Magazine 159, 97123.Google Scholar
Shao, T, Xia, Y, Ding, X, Cai, Y and Song, M (2019) Zircon saturation in terrestrial basaltic melts and its geological implications. Solid Earth Sciences 4, 2742.CrossRefGoogle Scholar
Shellnutt, JG (2016) Igneous rock associations 21. The early Permian Panjal traps of the Western Himalaya. Geoscience Canada 43, 251.Google Scholar
de Sigoyer, J, Guillot, S and Dick, P (2004) Exhumation of the ultrahigh-pressure Tso Morari unit in eastern Ladakh (NW Himalaya): a case study. Tectonics 23, 118.Google Scholar
Song, S, Bi, H, Qi, S, Yang, L, Allen, MB, Niu, Y, Su, L and Li, W (2018) HP–UHP metamorphic belt in the East Kunlun Orogen: final closure of the Proto-Tethys ocean and formation of the Pan-North-China continent. Journal of Petrology 59, 2043–60.Google Scholar
Spencer, DA, Tonarini, S and Pognante, U (1995) Geochemical and Sr-Nd isotopic characterisation of Higher Himalayan eclogites (and associated metabasites). European Journal of Mineralogy, 7, 89102.Google Scholar
St-Onge, MR, Rayner, N, Palin, RM, Searle, MP and Waters, DJ (2013) Integrated pressure-temperature-time constraints for the Tso Morari dome (Northwest India): implications for the burial and exhumation path of UHP units in the western Himalaya. Journal of Metamorphic Geology 31, 469504.Google Scholar
Sun, S-S and McDonough, WF (1989) Chemical and isotopic systematics of ocean basalt: implications for mantle composition and processes. Geological Society Special Publication 42, 323–45.Google Scholar
Trivedi, JR, Sharma, KK and Gopalan, K (1986) Widespread Caledonian magmatism in Himalaya and its tectonic significance. Terra Cognita 6, 144.Google Scholar
Wang, Y-H, Zhang, L-F, Li, S-Z and Somerville, I (2017) Zircon U–Pb dating and phase equilibria modelling of gneisses from Dinggye area, Ama Drime Massif, central Himalaya. Geological Journal 52(S1), 476–94.Google Scholar
Wedepohl, KH (1995) The composition of the continental crust. Geochimica et Cosmochimica Acta 59, 1217–32.Google Scholar
Whitney, DL and Evans, BW (2010) Abbreviations for names of rock-forming minerals. American Mineralogist 95, 185–7.Google Scholar
Wilson, M (1989). Igneous Petrogenesis. London: Unwin Hyman, 464 pp.Google Scholar
Xia, L and Li, X (2019) Basalt geochemistry as a diagnostic indicator of tectonic setting. Gondwana Research 65, 4367.Google Scholar
Yang, J, Xu, Z, Li, Z, Xu, X, Li, T, Ren, Y, Li, H, Chen, S and Robinson, PT (2009) Discovery of an eclogite belt in the Lhasa block, Tibet: a new border for Paleo-Tethys? Journal of Asian Earth Sciences 34, 7689.Google Scholar
Zhang, G, Wang, J, Webb, AAG, Zhang, L, Liu, S, Fu, B, Wu, C and Wang, S (2022) The protoliths of central Himalayan eclogites. GSA Bulletin 134, 1949–66.Google Scholar
Zhang, J, Meng, F and Yang, J (2005) A new HP/LT metamorphic Terrane in the Northern Altyn Tagh, Western China. International Geology Review 47, 371–86.Google Scholar
Zhang, ZM, Dong, X, Santosh, M and Zhao, GC (2014) Metamorphism and tectonic evolution of the Lhasa terrane, Central Tibet. Gondwana Research 25, 170–89.Google Scholar
Figure 0

Figure 1. Topographic map (90 m SRTM SEM data) of the Himalaya-Tibet region illustrating the distributions of the HP-UHP metamorphic rocks (compiled from Zhang et al.2005, 2014; Liou et al.2009; Yang et al.2009; Laskowski et al.2016; Liang et al.2017; Liu et al.2018, 2022; Song et al.2018; O’Brien, 2019; Rehman, 2019). The major discontinuities (after Li et al.2015) and mountain peaks are also shown. Abbreviations (arranged alphabetically): AD – Ama Drime, ANP – Annapurna, CQT – Coqin Thrust, EV – Everest, GSCT – Gaize-Siling Co Thrust, KF – Karakoram fault, KGN – Kaghan, MBT – Main Boundary Thrust, MCT – Main Central Thrust, MFT – Main Frontal Thrust, NB – Namche Barwa, NP – Nanga Parbat, NQT – Northern Qaidam Thrust, SGAT – Shinquanhe-Gaize-Amdo Thrust, SQT – Southern Qiadam Thrust, SS – Sap-Shergole, STD – South Tibetan Detachment, TST – Tanggula Shan Thrust and UH – Ursi-Hinju.

Figure 1

Figure 2. Study area, outcrop photographs and photomicrographs. (a) Geological map of the Tso Morari region (reproduced from Epard and Steck, 2008). Boudins of (b) eclogite and (c) K-rich metabasite in the Tso Morari gneiss parallel to the gneissic foliation. This photo was clicked at the same location as that of Fig. 3 by St-Onge et al. (2013). The large eclogite boudin is about 4 m thick at its thickest portion. (d) Top-to-the SE ductile shear sense exhibited by a feldspar augen in the quartzofeldspathic gneiss. (e) Retrograded Grt amphibolite. (f) Augen-shaped aggregates of quartz and feldspar within the fine-grained quartzofeldspathic matrix. The mica grains of the gneiss define the foliations in (g) and (h). (i) Prismatic zoisite grain within the coarser phengite grains of the gneiss. (j) Quartz porphyroclasts are common in some gneisses. Photomicrographs of the metabasites (k-m) and Grt amphibolite (n).

Figure 2

Figure 3. Spider diagrams of (a) retrograde eclogite and K-rich metabasite and (b) Grt amphibolite and eclogite boulder. Chondrite-normalised rare earth element patterns for (c) retrograde eclogite and K-rich metabasite and (d) Grt amphibolite and eclogite boulder. The compositions of metabasites from the TMC plotted in (e) Nb*2–Zr/4–Y (Meschede, 1986), (f) Zr/Y vs. Zr (Pearce & Norry, 1979), (g) Nb/La vs. Nb and (h) Y/Nb vs. Zr/Nb diagrams (Wilson, 1989, updated by Xia & Li, 2019). Data sources of the N-MORB, E-MORB, mantle plume and depleted asthenosphere compositions are from Sun & McDonough (1989) and Salters and Stracke (2004) with the PetDB database (http://www.earthchem.org/petdb). A(22): Ahmad et al. (2022), J(19): Jonnalagadda et al. (2019), R&R(06): Rao and Rai (2006). WPA: within-plate alkali basalts, WPT: within-plate tholeiites, VAB: volcanic-arc basalts.

Figure 3

Figure 4. Concordia diagrams from the SHRIMP zircon U–Pb analyses of (a) Ph-Bt gneiss (15-3B), (b) Grt-Ph gneiss (16-2A), (c) Quartzofeldspathic gneiss (19-2) and (d) Grt amphibolite (20-2) from the TMC. All error ellipses and weighted mean 206Pb/238U ages are quoted at the 1σ and 2σ levels, respectively. The white bars below the zircons are 20 μm long.

Figure 4

Figure 5. Map of Gondwana showing the positions of the Indian Shield and the rifting event inferred from this study. Modified from Gray et al. (2008), Meert and Lieberman (2008) and Imayama et al. (2023). SF, São Francisco Craton; RP, Río de la Plata Craton.

Supplementary material: File

Imayama et al. supplementary material

Imayama et al. supplementary material 1

Download Imayama et al. supplementary material(File)
File 2.9 MB
Supplementary material: File

Imayama et al. supplementary material

Imayama et al. supplementary material 2

Download Imayama et al. supplementary material(File)
File 241.4 KB
Supplementary material: File

Imayama et al. supplementary material

Imayama et al. supplementary material 3

Download Imayama et al. supplementary material(File)
File 15.7 KB