Hostname: page-component-7bb8b95d7b-2h6rp Total loading time: 0 Render date: 2024-09-25T09:39:24.025Z Has data issue: false hasContentIssue false

RETRACTED-Late Permian to early Triassic gabbro in North Lhasa, Tibet: evidence for plume – subduction-zone interaction of the Palaeo-Tethys ocean

Published online by Cambridge University Press:  19 January 2023

Meng-Long Duan
Affiliation:
College of Earth Sciences, Jilin University, Changchun 130061, China Research Center for Tibetan Plateau, Jilin University, Changchun, Jilin 130061, China
Chao-Ming Xie*
Affiliation:
College of Earth Sciences, Jilin University, Changchun 130061, China Key Laboratory of Mineral Resources Evaluation in Northeast Asia, Ministry of Natural Resources, Jilin University, Changchun 130061, China Research Center for Tibetan Plateau, Jilin University, Changchun, Jilin 130061, China
Bin Wang
Affiliation:
College of Earth Sciences, Jilin University, Changchun 130061, China Research Center for Tibetan Plateau, Jilin University, Changchun, Jilin 130061, China
Yu-Hang Song
Affiliation:
College of Earth Sciences, Jilin University, Changchun 130061, China Research Center for Tibetan Plateau, Jilin University, Changchun, Jilin 130061, China
Wen-qing Li
Affiliation:
Key Laboratory of Mineral Resources Evaluation in Northeast Asia, Ministry of Natural Resources, Jilin University, Changchun 130061, China
Yu-jie Hao
Affiliation:
Key Laboratory of Mineral Resources Evaluation in Northeast Asia, Ministry of Natural Resources, Jilin University, Changchun 130061, China
*
Author for correspondence: Chao-Ming Xie, Email: xcmxcm1983@126.com
Rights & Permissions [Opens in a new window]

Abstract

The Palaeo-Mesozoic geodynamic evolution of the Tangjia–Sumdo accretionary complex belt, which separates the North and South Lhasa Terrane, remains controversial. Moreover, the lack of geological records restricts the understanding of the evolution of the Sumdo Palaeo-Tethys Ocean from the middle Permian until the middle Triassic. Here we present zircon U–Pb geochronology, whole-rock geochemistry and Sr–Nd–Hf isotopic compositions of the Yeqing gabbro. Zircon U–Pb geochronology yields ages from 254 ± 1 to 249 ± 1 Ma. In situ Hf isotopic analyses yield ϵ Hf(t) values of −0.2 to +6.3. These samples have high TiO2 (3.69 wt %) and P2O5 (0.78 wt %) contents, with typical patterns like ocean island basalt (OIB). Besides, they are classified as high-Nb basalts (HNBs) based on the high content of Nb (45.3–113.5 ppm). Whole-rock Sr–Nd isotopic compositions are similar to OIB, with initial 87Sr/86Sr of 0.7047–0.7054, 143Nd/144Nd of 0.512526–0.512647 and ϵ Nd(t) of 0.3–2.7. These signatures suggest that the Yeqing gabbro is mainly derived from low-degree melting of the garnet lherzolite mantle. Based on field observations of HNBs intruding into the continental margin and their geochemical characteristics, we infer that the Yeqing gabbro was generated in a subduction environment. Combined with the regional geology of the subduction environment and the evolution of oceanic islands in the Sumdo Palaeo-Tethys Ocean, we propose that the Yeqing gabbro may represent a product of the asthenosphere upwelling through a slab window produced by subduction of seismic ridge in the Sumdo Palaeo-Tethys Ocean, called plume – subduction-zone interaction, during the late Permian to early Triassic.

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

1. Introduction

The Palaeo-Tethys Ocean was a large ancient ocean located between the supercontinents of Gondwana and Laurasia (Şengor, Reference Şengor1987; Metcalfe, Reference Metcalfe2013). A series of banded terranes (Qiangtang, Lhasa, Himalaya, Cimmerides, Sibumasu, etc.) derived from the north edge of Gondwana drifted northward and were accreted into Laurasia along with the closure of the Palaeo-Tethys Ocean (Yin & Harrison, Reference Yin and Harrison2000; Dilek & Furnes, Reference Dilek and Furnes2011; Metcalfe, Reference Metcalfe2013), forming a huge orogenic belt, known as the ‘Eastern Tethys System’, in the northern and southeast edge of the Tibetan Plateau and the central orogenic belt between North China and the Yangtze (Xu et al. Reference Xu, Dilek, Cao, Yang, Robinson, Ma, Li, Jolivet, Roger and Chen2015 and references therein). The Palaeo-Tethys Ocean in the Eastern Tethys System, which probably opened in the Middle Cambrian and continued to grow throughout the Palaeozoic and closed in the later Triassic, is mainly represented by the Jinshajiang and Longmu Co – Shuanghu suture zone in the northern Tibetan Plateau, and the Changning–Menglian and Ailaoshan suture zone of the Sanjiang Tethys realm in the eastern margin of the Tibetan Plateau (Fig. 1a ; Yin & Harrison, Reference Yin and Harrison2000; Li et al. Reference Li, Zhai, Dong and Huang2006, Reference Li, Dong, Zhai, Wang, Yan, Wu and He2008; Zhai et al. Reference Zhai, Wang, Li and Su2010, Reference Zhai, Jahn, Wang, Su, Mo, Wang, Tang and Lee2013, Reference Zhai, Jahn, Wang, Hu, Chung, Lee, Tang and Tang2016; Metcalfe, Reference Metcalfe2013; Fan et al. Reference Fan, Li, Xu and Wang2014, Reference Fan, Li, Xie, Wang and Chen2015, Reference Fan, Li, Wang, Liu and Xie2017; M Wang et al. Reference Wang, Li, Wu and Xie2014, Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Zhang et al. Reference Zhang, Dong, Wang, Dan, Zhang, Xu and Huang2017; Xie et al. Reference Xie, Li, Ren, Wang and Su2017 a, b).

The discovery of the late Palaeozoic Sumdo high/ultra-high-pressure (HP/UHP) metamorphic belt in Lhasa terrane reveals that there are records of an oceanic subduction zone, which may represent the southernmost branch of the Palaeo-Tethys Ocean (Fig. 1a; Yang et al. Reference Yang, Xu, Geng, Li, Xu, Li, Ren, Li, Cai, Liang and Chen2006, Reference Yang, Xu, Li, Xu, Li, Ren and Robinson2009; Liu et al. Reference Liu, Liu, Theye and Massonne2009; Xu et al. Reference Xu, Dilek, Cao, Yang, Robinson, Ma, Li, Jolivet, Roger and Chen2015). The fact that Sumdo HP/UHP metamorphic belt is located between Indus – Yarlung Zangbo (Neo-Tethys Ocean) and Bangong Co – Nujiang (Meso-Tethys Ocean) Suture Zone is incongruent with the common view that the Tethys Ocean Suture Zone becomes gradually younger from north to south (Xu et al. Reference Xu, Dilek, Cao, Yang, Robinson, Ma, Li, Jolivet, Roger and Chen2015). In recent years, further evidence for the evolution of the Sumdo Palaeo-Tethys Ocean (SPTO) has been established in Sumdo and adjacent regions, with examples such as ophiolites (Fig. 1b; Chen et al. Reference Chen, Yang, Li and Xu2010; Duan et al. Reference Duan, Xie, Fan, Wang and Hao2019; Wang et al. Reference Wang, Xie, Dong, Fan, Yu and Duan2021), oceanic islands (Fig. 1b; B Wang et al. Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Zhong et al. Reference Zhong, Zhu, Yang, Xie, Mai, Zhang, Li and Zhou2021; Duan et al. Reference Duan, Xie, Wang, Song and Hao2022), eclogite and blueschist (Fig. 1b ; Yang et al. Reference Yang, Xu, Geng, Li, Xu, Li, Ren, Li, Cai, Liang and Chen2006, Reference Yang, Xu, Li, Xu, Li, Ren and Robinson2009; Liu et al. Reference Liu, Liu, Theye and Massonne2009), arc magmatism (Fig. 1b ; Geng et al. Reference Geng, Sun, Pan, Zhu and Wang2009; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; C Wang, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022) and flysch-like sedimentary strata (Fig. 1b ; Xie et al. Reference Xie, Song, Wang, Fan and Hao2019, Reference Xie, Duan, Song and Wang2021). Thus, this belt is also called the Tangjia–Sumdo accretionary complex belt (TSACB) (Fig. 1a; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Xie et al. Reference Xie, Duan, Song and Wang2021). Previous research suggests that the SPTO may have opened before the early Carboniferous and subducted initially before the early Permian, then soon after the late Triassic at latest (Cheng et al. Reference Cheng, Liu, Vervoort and Lu2015; Duan et al. Reference Duan, Xie, Fan, Wang and Hao2019, Reference Duan, Xie, Wang, Song and Hao2022; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Dong, Fan, Yu and Duan2021, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Liu et al. Reference Liu, Li, Xie, Santosh, Liu, Dong, Wang, Guo and Cao2022). Based on the study of eclogite and arc magma, we know that the SPTO subducted during the early–middle Permian and middle–late Triassic periods (Cheng et al. Reference Cheng, Zhang, Vervoort, Lu, Wang and Cao2012, Reference Cheng, Liu, Vervoort and Lu2015; Zhang et al. Reference Zhang, Bader, Zhang, Shen, Li and Li2018 a; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Song et al. Reference Song, Xie, Gao, Yu, Wang, Duan and Hao2022; C Wang et al. Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022). However, many aspects of the subduction evolution of the SPTO remain unclear, especially during the late Permian to early Triassic owing to gaps in the geological record (Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; Cheng et al. Reference Cheng, Zhang, Vervoort, Lu, Wang and Cao2012, Reference Cheng, Liu, Vervoort and Lu2015; Zhang et al. Reference Zhang, Bader, Van, Yang, Shen, Qiu, Li and Zheng2018 b; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Li et al. Reference Li, Yang, Zhu, Xie, Zhong, Mai, Zhou and Zhang2022; C Wang et al. Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022).

The Yeqing gabbro is discovered near the north edge of the TSACB, whose zircon U–Pb ages vary from 254 ± 1 to 249 ± 1 Ma in this study, coincident with the active period of the SPTO. Here, we present zircon U–Pb geochronology, whole-rock geochemistry, as well as zircon Hf and whole-rock Sr–Nd isotopic data, which are significant in resolving the diagenetic age and petrogenesis of the Yeqing gabbro. We also discuss the tectonic setting of the intrusion and draw implications regarding the subduction evolution of the SPTO during the late Permian to early Triassic.

2. Geological background

The Tibetan Plateau is located in the eastern part of the Tethys tectonic domain. The closure of the Tethys oceans created four major suture zones associated with the Tibetan Plateau (Fig. 1a;Yin & Harrison, Reference Yin and Harrison2000; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; Torsvik & Cocks, Reference Torsvik and Cocks2013; Zhai et al. Reference Zhai, Jahn, Wang, Hu, Chung, Lee, Tang and Tang2016). These suture zones divide the Tibetan Plateau, from north to south, into the Songpan–Ganzi, Northern Qiangtang, Southern Qiangtang, Lhasa and Himalayan terranes (Li et al. Reference Li, Zhai, Dong and Huang2006, Reference Li, Dong, Zhai, Wang, Yan, Wu and He2008; Zhai et al. Reference Zhai, Wang, Li and Su2010, Reference Zhai, Jahn, Wang, Su, Mo, Wang, Tang and Lee2013; Fan et al. Reference Fan, Li, Xu and Wang2014, Reference Fan, Li, Xie, Wang and Chen2015, Reference Fan, Li, Wang, Liu and Xie2017; M Wang et al. Reference Wang, Li, Wu and Xie2014, Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Xie et al. Reference Xie, Li, Ren, Wang and Su2017 a, b; Zhang et al. Reference Zhang, Dong, Wang, Dan, Zhang, Xu and Huang2017; Hu et al. Reference Hu, Zhai, Wang, Tang, Wang and Hou2018, Reference Hu, Zhai, Zhao, Wang, Tang, Zhu and Wu2019). The Lhasa terrane is further divided into North and South Lhasa Terrane by the SPTO, marked by the TSACB (Yang et al. Reference Yang, Xu, Li, Xu, Li, Ren and Robinson2009).

The TSACB, correlated spatially with the Luobadui–Milashan Fault (Fig. 1a; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010), is dominated by scattered fragments of the SPTO remnants (Fig. 1b). In the study area, the Nuoco Formation, formed in a continental margin environment, mainly consists of sandstone or metasedimentary rock on the north of the TSACB (Zhu et al. Reference Zhu, Zhao, Niu, Hou and Mo2013). The strata exposed in the TSACB is mainly Sumdo Formation, which is a set of low-grade-metamorphosed terrigenous clastic sandstones and mudstones formed in an initial fore-arc basin environment (Xie et al. Reference Xie, Song, Wang, Fan and Hao2019, Reference Xie, Duan, Song and Wang2021). The Luobadui Formation exposes a little in the study area and is mainly composed of limestone, terrigenous sediments and arc-type volcanic rocks (Geng et al. Reference Geng, Sun, Pan, Zhu and Wang2009; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; C Wang et al. Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022).

The SPTO remnants are abundant in the Sumdo Formation as slices, including Permian–Triassic eclogites (Li et al. Reference Li, Yang, Xu, Li, Xu, Ren and Robinson2009; Yang et al. Reference Yang, Xu, Li, Xu, Li, Ren and Robinson2009; Zeng et al. Reference Zeng, Liu, Gao, Chen and Xie2009), late Carboniferous – middle Triassic ophiolites (Chen et al. Reference Chen, Yang, Li and Xu2010; Duan et al. Reference Duan, Xie, Fan, Wang and Hao2019; Wang et al. Reference Wang, Xie, Dong, Fan, Yu and Duan2021) and early Carboniferous – middle Permian oceanic islands (B Wang et al. Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Zhong et al. Reference Zhong, Zhu, Yang, Xie, Mai, Zhang, Li and Zhou2021; Duan et al. Reference Duan, Xie, Wang, Song and Hao2022). Previous studies have suggested that there are at least two types of eclogites, with ages of the metamorphic peak in the middle Permian (274–261 Ma) and middle–late Triassic (238–227 Ma) (Li et al. Reference Li, Yang, Xu, Li, Xu, Ren and Robinson2009; Yang et al. Reference Yang, Xu, Li, Xu, Li, Ren and Robinson2009; Zeng et al. Reference Zeng, Liu, Gao, Chen and Xie2009; Chen et al., Reference Chen, Yang, Li and Xu2010; Cheng et al. Reference Cheng, Zhang, Vervoort, Lu, Wang and Cao2012, Reference Cheng, Liu, Vervoort and Lu2015; Weller et al. Reference Weller, St-Onge, Rayner, Waters, Searle and Palin2016; Cao et al. Reference Cao, Cheng, Zhang and Wang2017; Zhang et al. Reference Zhang, Bader, Zhang, Shen, Li and Li2018 a, b). There is also plenty of Permian arc magmatism found in the south edge of the North Lhasa Terrane (Fig. 1a; Geng et al. Reference Geng, Sun, Pan, Zhu and Wang2009; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010) and a part in the TSACB (Fig. 1b, 278–262 Ma; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Mai et al. Reference Mai, Zhu, Yang, Xie, Tong, Hao and Zhong2021; Li et al. Reference Li, Yang, Zhu, Xie, Zhong, Mai, Zhou and Zhang2022; C Wang et al. Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022). In addition, the middle–late Triassic (230–200 Ma) granite with typical arc magmatism characteristics may be related to the northward subduction of the SPTO (Li et al. Reference Li, Zhang, Wu, Xie, Zhu and Han2020; Song et al. Reference Song, Xie, Gao, Yu, Wang, Duan and Hao2022).

3. Field observations and petrology

The study area is located between Tangjia and Sumdo (Fig. 1b). The Yeqing gabbro intrusions expose as near east–west-trending dike in the Nuoco Formation (Fig. 1b). The larger intrusion is c. 8 km long and 50 m thick, while the smaller intrusion is c. 3 km long and 20 m thick. The host rocks of the Yeqing gabbro contains quartzite and sandstone of the Nuoco Formation (Fig. 2a). We can observe the obvious chilled margin between gabbro and quartzite (Fig. 2b).

Fig. 2. Photographs of Yeqing gabbro. (a) Macro outcrop photo of Yeqing gabbro intruding into Sumdo Formation. (b) Close-up photo of the boundary between Yeqing gabbro and Sumdo Formation. (c, d) Micrograph of Yeqing gabbro, Plagioclase is replaced by sericite, and pyroxene is replaced by hornblende, in part of our samples. Pl – plagioclase; Px – pyroxene; hor – hornblende; Sre – sericite; Mag – magnetite.

The gabbro samples are fine- to medium-grained and consist mainly of pyroxene and plagioclase (Fig. 2b). Petrographic observations under the microscope reveal that the gabbro shows crystalline texture, consisting of dominant mineralogy of pyroxene (35 %), plagioclase (60 %) and magnetite (<5 %), with slight metamorphism (Fig. 2c, d). Some pyroxene has been replaced by chlorite or hornblende, and the plagioclase is weakly altered to sericite, which shows the sericite microcrystalline aggregates under a microscope with orthogonal polarized light (Fig. 2d).

4. Analytical methods

4.a. Whole-rock geochemistry

Whole-rock geochemical analysis was performed at the experimental centre of the Academy of Sciences, China University of Geosciences, Beijing, China. Major-element analysis was performed in the inductively coupled plasma – optical emission spectroscopy (ICP-OES) laboratory using an X-ray fluorescence spectrometer. The precision is better than 1 % for all elements. Trace-element analysis was performed using an Agilent 7500a ICP – mass spectrometry (ICP-MS) instrument. During the analysis, standard samples AGV-2, W2 and BHOV from the United States Geological Survey and rock samples R-1 and R-3 from the China Geological Testing Center were used to monitor the analytical precision, which is better than 5 % for most elements. Further details of the experimental method are presented by Zhai et al. (Reference Zhai, Jahn, Wang, Su, Mo, Wang, Tang and Lee2013).

4.b. Zircon U–Pb geochronology

Zircon grains were separated at the premises of the Yuheng Mineral Technology Service, Langfang, China, by conventional heavy liquid and magnetic techniques. Cathodoluminescence (CL) images were taken at the Institute of Geology, Chinese Academy of Geological Sciences, Beijing, China. U–Pb isotopic and trace-element analyses of zircon were carried out by laser-ablation – ICP-MS (LA-ICP-MS) at the Key Laboratory of Mineral Resources Evaluation in Northeast Asia, Ministry of Natural Resources, Jilin University, Changchun, China. The laser beam spot diameter was 32 µm, and helium was used as the carrier gas. Details of the procedure are reported by Liu et al. (Reference Liu, Gao, Hu, Gao, Zong and Wang2010). Fractionation correction of isotopic ratios was performed using zircon 91500 as the external standard. NIST SRM 610 was used as the external standard for correction of elemental abundances, and 29Si was used as the internal standard. Data were processed using Glitter software (version 4.4; Griffin et al. Reference Griffin, Powell, Pearson and O’Reilly2008). Details of specific experimental data reduction techniques are given by Chang et al. (Reference Chang, Vervoort, McClelland and Knaack2006). The software package Isoplot 4.15 was used to calculate weighted-mean U–Pb ages of the samples and to generate Concordia diagrams (Ludwig, Reference Ludwig2003).

4.c. Zircon Hf isotopes

Zircon Hf isotopic analyses were conducted at the facility of Beijing Createch Testing Technology, Beijing, China. Hf isotopic analyses were performed on the same spots or in the same age domains (as identified by CL images) as used for U–Pb dating. An NWR-213 (nm) laser ablation microprobe coupled to a Neptune Plus multi-collector (MC)-ICP-MS instrument was used for isotopic analysis. The laser beam spot was 40 µm in diameter, with an energy density of 10–11 J cm−2 and a frequency of 10 Hz. The ablated material was carried to the mass spectrometer by high-purity helium gas. Zircon GJ-1 was used as the reference standard during analyses, whose weighted mean 176Hf/177Hf ratio (0.282000 ± 32; 2σ; n = 17) is similar to the commonly accepted weighted mean 176Hf/177Hf ratio of 0.282013 ± 19 (2σ) reported for in situ analysis by Elhlou et al. (Reference Elhlou, Belousova, Griffin, Pearson and O’Reilly2006). Technical procedures and instrument operational parameters are described by Hu et al. (Reference Hu, Liu, Gao, Liu, Zhang, Tong, Lin, Zong, Li, Chen, Zhou and Yang2012). Data reduction methods followed those presented by Bouvier et al. (Reference Bouvier, Vervoort and Patchett2008).

4.d. Whole-rock Sr–Nd isotopic analysis

Whole-rock Sr–Nd isotopic analyses were carried out using a Thermo Fisher Scientific Neptune Plus MC-ICP-MS instrument at the facility of Beijing Createch Testing Technology. 87Sr/86Sr ratios were corrected for instrumental mass fractionation using an exponential fractionation law and assuming 88Sr/86Sr = 8.375209. 143Nd/144Nd ratios were corrected for instrumental mass fractionation using an exponential fractionation law and assuming 146Nd/144Nd = 0.7219. The Sr isotope international standard NBS 987 was repeatedly tested to monitor accuracy, yielding a mean 87Sr/86Sr value of 0.710248 ± 9 (2SD, n = 11). Stability assessment for 143Nd/144Nd was conducted with the in-house standard GSB-Nd, yielding a value of 0.512195 ± 6 (2SD, n = 12). Detailed analytical procedures are given by Hu et al. (Reference Hu, Zhai, Wang, Tang, Wang and Hou2018).

5. Results

5.a. U–Pb zircon geochronology

LA-ICP-MS zircon U–Pb data and zircon trace element of the Yeqing gabbro are presented in Supplementary Tables S1 and S2, respectively. Zircon grains separated from gabbro samples (ST30, ST31, ST38 and ST39) are semi-transparent and columnar to granular. The lengths of the zircon crystals are 100–250 μm, with aspect ratios of 1:1–3:1. Some zircon grains exhibit oscillatory zoning in CL imaging (Fig. 3). The range of Th and U contents is relatively wide (Th = 36–1053 ppm, U = 37–450 ppm) and the Th/U ratios are quite high (0.56–2.57), which is consistent with the magmatic origin (Hoskin & Schaltegger, Reference Hoskin and Schaltegger2003). Positive Ce and negative Eu anomalies are observed in chondrite-normalized rare earth element (REE) patterns of zircon grains from these samples (Fig. 3).

Fig. 3. Cathodoluminescence (CL) images and chondrite-normalized REE patterns diagram of representative zircon grains from Yeqing gabbro. The yellow circle is the location of U–Pb isotope analysis, and the red is Lu–Hf isotope analysis. Values of chondrite are after Sun and McDonough (Reference Sun, McDonough, Saunders and Norry1989).

U–Pb data from all four samples are concordant (Fig. 4). The data yield weighted-mean ages of 249 ± 1 Ma for ST30 (n = 25, MSWD = 0.34, 1σ), 251 ± 1 Ma for ST31 (n = 25, MSWD = 0.75, 1σ), 252 ± 1 Ma for ST39 (n = 28, MSWD = 0.51, 1σ) and 254 ± 1 Ma for ST38 (n = 23, MSWD = 0.92, 1σ).

Fig. 4. U–Pb zircon Concordia of representative zircon grains from Yeqing gabbro. MSWD = mean squared weighted deviation.

5.b. Whole-rock geochemistry

Whole-rock major- and trace-element geochemical data are listed in Supplementary Table S3. These samples contain variational content of SiO2 (41.91–50.12 wt %) and MgO (4.15–8.56 wt %). The contents of Fe2O3 T (Fe2O3 total) and Al2O3 are relatively concentrated, with means of 13.34 wt % and 15.39 wt % respectively. In particular, the contents of TiO2, P2O5 and (Na2O + K2O) are comparatively high, with means of 3.69 wt %, 0.78 wt % and 4.36 wt % respectively. These samples plot mostly in the alkaline basalt field in the SiO2 vs Nb/Y diagram (Fig. 5a).

Fig. 5. (a) SiO2 vs Nb/Y (Winchester & Floyd, Reference Winchester and Floyd1977) plot. (b) Nb/La vs MgO (Kepezhinskas et al. Reference Kepezhinskas, Defant and Drummond1996) plot. (c) Nb*2 vs Zr/4 vs Y figure (Meschede, Reference Meschede1986). Within-plate alkali basalts – AI, AII within-plate tholeiites – AII, C; plume-type MORB- B; N-type MORB- D; volcanic arc basalts – C, D. (d) TiO2 vs MnO2*10 vs P2O5*10 figure (Mullen, Reference Mullen1983). MORB – mid-ocean ridge basalt; IAT – island arc tholeiite; CAB – calc-alkaline basalt; OIT – ocean island tholeiite; OIA – ocean island alkaline. Data of the NEBs, HNBs from Baja, Nicaragua, Renso, Dubuzha, Tuotuohe and Gerze are after Storey et al. (Reference Storey, Rogers, Saunders and Terrell1989), Luhr et al. (Reference Luhr, Aranda-Gomez and Housh1995), Benoit et al. (Reference Benoit, Aguilloân-Robles, Calmus, Maury, Bellon, Cotton, Bourgois and Michaud2002), Wang et al. (Reference Wang, Wyman, Xu, Wan, Li, Zi, Jiang, Qiu, Chu, Zhao and Dong2007), Gazel et al. (Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011), Li et al. (Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016) and Hao et al. (Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018).

The REE contents of these samples are relatively higher (168.54–559.62 ppm) than mid-ocean ridge basalt (MORB) (39.11 ppm; Sun & McDonough, Reference Sun, McDonough, Saunders and Norry1989), with light REE (LREE) enrichment (LaN/YbN = 11.51–27.10) in chondrite-normalized REE patterns (Fig. 6a). Positive Nb and Ta anomalies and negative Th, Zr and Hf anomalies are observed in primitive-mantle-normalized trace-element spider diagrams, which is similar to the typical OIB and the NEBs and HNBs from Baja (California), Nicaragua, Renso and Duobuzha (Tibet), and different from those from Tuotuohe and Gerze (Tibet) (Fig. 6b, d). The higher Nb contents (45.3–113.5 ppm) of the Yeqing gabbro are similar to those of Nb-enrichment basalts (NEBs; 20 > Nb > 5 ppm) and high-Nb basalts (HNBs; Nb > 20 ppm) which are enriched in LILEs, LREEs and HFSEs and have weakly negative or positive primitive-mantle-normalized Nb and Ta anomalies (Castillo et al. Reference Castillo, Rigby and Solidum2007; Hastie et al. Reference Hastie, Mitchell, Kerr, Minifie and Millar2011); these samples plot in the NEBs and HNBs field in the Nb/La vs MgO plot (Fig. 5b). In addition, the Yeqing gabbro shows characteristics of intraplate alkaline basalts in the tectonic discrimination diagrams (Fig. 5c, d).

Fig. 6. Chondrite-normalized REE patterns and primitive-mantle-normalized spider diagrams. Values of chondrite, primitive mantle, OIB and E-MORB are after Sun and McDonough (Reference Sun, McDonough, Saunders and Norry1989).

5.c. Zircon Hf and whole-rock Sr–Nd isotopic analysis

Zircon Hf and whole-rock Sr–Nd isotope data are listed in Supplementary Tables S4 and S5, respectively. The Hf isotope analyses of the zircon grains from the gabbro samples show low 176Lu/177Hf ratios of 0.0003–0.0021 and 176Hf/177Hf ratios of 0.282608–0.282794. Calculated small negative or positive ϵ Hf(t) values range from −0.2 to +6.3. Initial Sr isotope ratios and ϵ Nd(t) values were calculated using ages of c. 254 Ma and c. 251 Ma reported in this study. These samples have a narrow range of initial (87Sr/86Sr) ratios of 0.7047–0.7054, 87Sr/86Sr ratios of 0.7047–0.7054 and 143Nd/144Nd ratios of 0.512526–0.512647. Calculated small positive ϵ Nd(t) values range from 0.3 to 2.7.

6. Discussion

6.a. Petrogenesis

Low loss-on-ignition (LOI) values of 1.14–1.77 wt % and metamorphic minerals under the microscope indicate that the samples have undergone low-level metamorphism or alteration. This process may have modified the contents of mobile elements (e.g. Na, K, Rb, Ba and Sr), whereas the REE and high-field-strength elements (HFSE; e.g. Th, Zr, Hf, Nb, Ta, Ti and Y) should preserve primary magma compositions (Barnes et al. Reference Barnes, Naldrett and Gorton1985; Jochum et al. Reference Jochum, Arndt and Hofmann1991). In fact, most mobile elements and HFSE display good correlations with MgO (Fig. 8, further below), indicating there is almost no significant disturbance by metamorphism or alteration on most elements.

The Yeqing gabbro exhibits clear positive Nb and Ta and negative Th, U, Zr and Hf anomalies, with significant differences compared with continental crust (Rudnick & Gao, Reference Rudnick and Gao2003; Niu, Reference Niu2009). The Th/Ta ratios of these samples are 0.1–1.29, similar to those of volcanic rocks derived from a primitive mantle source (Th/Ta = 2.3), and much lower than that of the upper crust (Th/Ta > 10) (Thompson et al. Reference Thompson, Morrison, Hendry, Parry, Simpson, Hutchison and O’Hara1984; Condie, Reference Condie1993). The (Th/Ta)PM (∼0.24) and (La/Nb)PM (∼0.85) ratios of these samples are both less than 1, indicating that the crustal assimilation is negligible (Peng et al. Reference Peng, Mahoney, Hooper, Harris and Beane1994). Furthermore, these samples plot in the oceanic basalts field without the trend of crustal assimilation in the (Th/Nb)PM vs (La/Nb)PM diagram (Fig. 7).

Fig. 7. (La/Nb)PM vs (Th/Nb)PM plot (Neal et al. Reference Neal, Mahoney and Chazey2002). Middle crust and lower crust data are after Rudnick and Gao (Reference Rudnick and Gao2003). The ‘most oceanic basalts’ data are after Neal et al. (Reference Neal, Mahoney and Chazey2002 and references therein).

The Mg# values (=100 × Mg2+/(Mg2+ + Fe2+)) (43.8–61.1) and the Cr (2.9–253.9 ppm) and Ni (2.2–81.9 ppm) contents of the Yeqing gabbro are lower than those of primitive mantle (Mg# = 68–76, Cr = 300–500 ppm, Ni = 300–400 ppm), indicating that they may have undergone fractional crystallization of olivine, pyroxene and chromite (Wilson, Reference Wilson1989; Jung & Masberg, Reference Jung and Masberg1998). In the process of fractional crystallization, Ni preferentially integrates into the olivine phase, and Cr preferentially integrates into the pyroxene phase (Wilson, Reference Wilson1989; Rollinson, Reference Rollinson1993). The strongly positive relationships between FeOT, Ni and MgO suggest that the magma underwent obvious fractional crystallization of olivine (Fig. 8a, h). In addition, the higher content and weak positive relationship with MgO of Cr element illustrate that there is a little or no fractional crystallization of pyroxene (Fig. 8g). The negative relationships between Sr, Na2O, Al2O3 and MgO indicate that these samples underwent almost no fractional crystallization of plagioclase (Fig. 8b, c, i; Fodor & Vetter, Reference Fodor and Vetter1984; Baker et al. Reference Baker, Menzies, Thirlwall and Macpherson1997), consistent with the absence of Eu anomaly in the chondrite-normalized REE patterns (Fig. 6a).

Fig. 8 Harker variation diagrams for the Yeqing gabbro. The red arrows represent variation trend of part of major (wt %) and trace (ppm) elements towards the increase of MgO (wt %).

As presented above, the Yeqing gabbro is classified as HNBs with clear LREE enrichment and high Nb content (45.3–113.5 ppm). Two possible mantle sources have been generally proposed to generate HNBs, namely (1) OIB-type mantle source with mixing depleted normal-MORB (N-MORB) type components (Reagan & Gill, Reference Reagan and Gill1989; Storey et al. Reference Storey, Rogers, Saunders and Terrell1989; Luhr et al. Reference Luhr, Aranda-Gomez and Housh1995; Castillo et al. Reference Castillo, Rigby and Solidum2007; Castillo, Reference Castillo2008; Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016) and (2) a mantle wedge metasomatized by slab melt (Defant et al. Reference Defant, Jackson, Drummond, De Bore, Bellon, Feigenson, Maury and Stewart1992; Kepezhinskas et al. Reference Kepezhinskas, Defant and Drummond1996; Sajona et al. Reference Sajona, Maury, Bellon, Cotton and Defant1996; Wang et al. Reference Wang, Wyman, Xu, Wan, Li, Zi, Jiang, Qiu, Chu, Zhao and Dong2007; Hastie et al. Reference Hastie, Mitchell, Kerr, Minifie and Millar2011; Xu et al. Reference Xu, Li, Wang, Fan, Wu and Li2017; Hao et al. Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018). In REE patterns and trace-element spider diagrams, the NEBs/HNBs derived from OIB-type mantle source with mixing depleted N-MORB type components show obvious LREE enrichment and positive Nb and Ta anomalies reported in Baja (California), Nicaragua, Renso and Duobuzha (Tibet), called a-type NEBs/HNBs in the following discussion (Fig. 6a, b). In contrast, the NEBs/HNBs derived from mantle wedge with mixing slab melt show an almost flat curve and negative Nb and Ta anomalies reported in Tuotuohe and Gerze (Tibet), called b-type NEBs/HNBs (Fig. 6c, d).

The HNBs that originated from a mantle wedge source metasomatized by slab melt still display some arc geochemical signatures, such as negative Nb–Ta anomalies and enrichment of Th (Fig. 6d; Kepezhinskas et al. Reference Kepezhinskas, Defant and Drummond1996; Sajona et al. Reference Sajona, Maury, Bellon, Cotton and Defant1996; Wang et al. Reference Wang, Wyman, Xu, Wan, Li, Zi, Jiang, Qiu, Chu, Zhao and Dong2007; Hao et al. Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018). The Yeqing gabbro samples actually exhibit clearly positive Nb–Ta anomalies and depletion of Th (Fig. 6b, d). In the REE patterns and trace-element spider diagrams, the Yeqing gabbro shows a large slope like the OIB and a-type NEBs/HNBs, as distinct from the b-type NEBs/HNBs (Fig. 6b, d). In the Th/Yb vs Nb/Yb program (Fig. 9a), the gabbro samples plot in the OIB array like the a-type NEBs/HNBs, while the b-type NEBs/HNBs plot in the OIB to enriched MORB (E-MORB) array with a tendency to convert to continental arc, indicating that the magma of Yeqing gabbro generated without any addition of subduction fluids and melts. In the TiO2/Yb vs Nb/Yb diagram (Fig. 9b), these samples are plotted as OIBs (alkaline) like the a-type NEBs/HNBs, while the b-type NEBs/HNBs show characteristics of shallow melting, indicating that the Yeqing gabbro may be the product of deep melting where garnet is stable. In addition, the Sr and Nd isotopic compositions of the Yeqing gabbro and Sumdo eclogite differ greatly (Fig. 9d), while the metasomatism of the mantle wedge by slab-derived melt would require the formation of Nb-rich magma with Sr and Nd isotopic characteristics similar to oceanic slab (Castillo et al. Reference Castillo, Rigby and Solidum2007). Above all, the Yeqing gabbro is notably distinct from the NEBs and HNBs from mantle wedge metasomatized by slab melt with negligible characteristics of arc, precluding the possibility of a mantle wedge source metasomatized by slab melt.

Fig. 9 (a) Th/Yb vs Nb/Yb plot and (b) TiO2/Yb vs Nb/Yb plot (Pearce, Reference Pearce2014). (c) La/Sm vs Sm/Yb plot. Mantle array (heavy line) defined by depleted MORB mantle (DMM; McKenzie & O’Nions, Reference McKenzie and O’Nions1991) and primitive mantle (PM; Sun & McDonough, Reference Sun, McDonough, Saunders and Norry1989); Melting curves for spinel lherzolite and garnet peridotite with both DMM and PM compositions are after Aldanmaz et al. (Reference Aldanmaz, Pearce and Thirlwall2000). Numbers along lines represent the degree of partial melting. (d) ϵ Nd(t) vs (87Sr/86Sr) i plot. Mantle arrays are after Zindler and Hart (Reference Zindler and Hart1986). Bulk Earth is after Depaolo (Reference DePaolo1988). OIB is after Wilson (Reference Wilson1989). The date of Sumdo eclogite is from Li et al. (Reference Li, Yang, Xu, Li, Xu, Ren and Robinson2009). Dates of Halberstadt, Gerze NEBs and HNBs are after Hastie et al. (Reference Hastie, Mitchell, Kerr, Minifie and Millar2011) and Hao et al. (Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018).

Compared to normal arc basalts, the higher TiO2 (2.38–5.37 wt %), P2O5 (0.36–1.42 wt %), Nb (45.3–113.5 ppm) and Nb/Yb (14.6–38.1) suggest a deep mantle origin for the Yeqing gabbro. These samples exhibit LREE-enriched chondrite-normalized REE patterns and high (La/Yb)PM (11.51–27.12) and (Ce/Yb)PM (9.91–21.87) ratios, similar to those of basalts derived from garnet lherzolite (Hart & Dunn, Reference Hart and Dunn1993; Hauri et al. Reference Hauri, Wagner and Grove1994). These samples also plot in the garnet lherzolite field with a low degree of partial melting in the La/Sm vs Sm/Yb plot near the a-type NEBs/HNBs and far from b-type NEBs/HNBs (Fig. 9c; Aldanmaz et al. Reference Aldanmaz, Pearce and Thirlwall2000), indicating that the magma may be derived from a garnet-stable region (>85 km; Robinson & Wood, Reference Robinson and Wood1998) with lesser addition of N-MORB components, where it is generally considered to generate OIB (Niu, Reference Niu2009). However, according to mantle heterogeneity, the regional evolution and compotation of the study area must be taken into account when considering a ‘true’ OIB or the depleted and enriched mantle model (Hastie et al. Reference Hastie, Mitchell, Kerr, Minifie and Millar2011). There are several pieces of evidence about oceanic islands of the SPTO found in the TSACB, indicating that a plume had indeed existed in the SPTO (B Wang et al. Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Zhong et al. Reference Zhong, Zhu, Yang, Xie, Mai, Zhang, Li and Zhou2021; Duan et al. Reference Duan, Xie, Wang, Song and Hao2022). Consequently, there exists a true enriched OIB mantle beneath the SPTO crust as the material source of the Yeqing gabbro.

In addition, the Yeqing gabbro has elevated FeOT (11.54–14.72 wt %) and TiO2 (2.38–5.37 wt %) like the Fe–Ti basalts which are defined by >12 wt % FeOT and >2 wt % TiO2 (Sinton et al. Reference Sinton, Wilson, Christie, Hey and Delaney1983). Most of the Fe–Ti basalts show MORB affinity, such as lower concentrations of MgO, CaO and Al2O3. rather than N-MORB (Hollis et al. Reference Hollis, Roberts, Cooper, Earls, Herrington, Condon, Cooper, Archibald and Piercey2012). The gabbro in this study has a slightly lower MgO (average of 6.06 wt %) and CaO (average of 10.89 wt %) than N-MORB (MgO with average of 7.60 wt %; CaO with average of 11.48 wt %), which may be related to the fractional crystallization of olivine and a little pyroxene. The Al2O3 (average of 15.39 wt %) is slightly higher than N-MORB (average of 14.85 wt %). Besides, the Yeqing gabbro shows apparent LREE enrichment and HREE depletion in the REE diagram, and large slope curve in the trace-element diagram, in contrast to the Fe–Ti basalts with flat–modest slope curve in the REE and trace-element diagrams (Sinton et al. Reference Sinton, Wilson, Christie, Hey and Delaney1983; Hollis et al. Reference Hollis, Roberts, Cooper, Earls, Herrington, Condon, Cooper, Archibald and Piercey2012).

Relative to MORB, the zircon Hf isotopic composition of OIB has a narrow range and a lower ϵ Hf(t) value, and the mantle typically has a positive ϵ Hf(t) value (Nowell et al. Reference Nowell, Kempton, Noble, Fitton, Saunders, Mahoney and Taylor1998; Dobosi et al. Reference Dobosi, Kempton, Downes, Embey-Isztin, Thirlwall and Greenwood2003; Wu et al. Reference Wu, Li, Zheng and Gao2007). Zircon ϵ Hf(t) values of these samples are −0.2–6.3 which is lower than those of HNBs from Duobuzha and Rena Tso (2.44–11.64 and 1.9–7.6 respectively; Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016) and similar to those of OIBs. The whole-rock Sr–Nd isotopic data of these samples exhibit slightly high (87Sr/86Sr) i (0.70468–0.70542), low 143Nd/144Nd (0.51253–0.51265) and low ϵ Nd(t) (0.3–2.7) relative to some other HNBs from mantle wedge (87Sr/86Sr of 0.70341–0.70484; 143Nd/144Nd of 0.51292–0.51307; ϵ Nd(t) of 2.57–6.80) (Wang et al. Reference Wang, Wyman, Xu, Wan, Li, Zi, Jiang, Qiu, Chu, Zhao and Dong2007; Hastie et al. Reference Hastie, Mitchell, Kerr, Minifie and Millar2011; Xu et al. Reference Xu, Li, Wang, Fan, Wu and Li2017; Hao et al. Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018). As we can see, the Yeqing gabbro has a more enriched Sr–Nd isotopic composition than Sumdo eclogite and NEBs/HNBs from Halberstadt in the ϵNd(t) vs (87Sr/86Sr) i plot (Fig. 9d). The Halberstadt NEBs and HNBs exhibit relatively depleted features, which were generated from the mantle wedge metasomatized by slab melt (Fig. 9d; Hastie et al. Reference Hastie, Mitchell, Kerr, Minifie and Millar2011). An analogous Sr–Nd isotopic composition could be found between Yeqing gabbro and Gerze HNBs, which was produced by partial melting of an OIB-type source component involving upwelling asthenosphere mantle (Fig. 9d; Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016). In conclusion, the geochemistry and isotopic signatures of the Yeqing gabbro samples can be interpreted as the production of a low-degree partial melting of garnet lherzolite mantle with the negligible contribution of subducted oceanic crust.

6.b. Tectonic setting

HNBs are not intraplate lavas like OIB and other alkaline lavas (Adam & Green, Reference Adam and Green2010); they are only found in subduction zone environments (Defant et al. Reference Defant, Jackson, Drummond, De Bore, Bellon, Feigenson, Maury and Stewart1992), such as Zamboanga Peninsula (Philippines; Sajona et al. Reference Sajona, Maury, Bellon, Cotton and Defant1996), Sulu (southern Philippines; Castillo et al. Reference Castillo, Rigby and Solidum2007), Baja (California; Luhr et al. Reference Luhr, Aranda-Gomez and Housh1995; Aguillón-Robles et al. Reference Aguillón-Robles, Calmus, Benoit, Bellon, Maury, Cotten, Bourgois and Michaud2001; Castillo, Reference Castillo2008), Halberstadt (Germany; Hastie et al. Reference Hastie, Mitchell, Kerr, Minifie and Millar2011), Kamchatka (Russia; Kepezhinskas et al. Reference Kepezhinskas, Defant and Drummond1996), Nicaragua (Gazel et al. Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011), Tuotuohe (Tibet; Wang et al. Reference Wang, Wyman, Xu, Wan, Li, Zi, Jiang, Qiu, Chu, Zhao and Dong2007), Rena Tso (Tibet; Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016), Duobuzha (Tibet; Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016; Xu et al. Reference Xu, Li, Wang, Fan, Wu and Li2017) and Gerze (Tibet; Hao et al. Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018). Actually, there are multiple hypotheses about the formation of HNBs and NEBs, which are linked with some specific subduction processes, like flat subduction, slab rollback, slab break-off, ridge subduction or plume – subduction-zone interaction (Thorkelson, Reference Thorkelson1996; Wang et al. Reference Wang, Wyman, Xu, Wan, Li, Zi, Jiang, Qiu, Chu, Zhao and Dong2007; Gazel et al. Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011; Thorkelson et al. Reference Thorkelson, Madsen and Sluggett2011; Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016; Xu et al. Reference Xu, Li, Wang, Fan, Wu and Li2017; Hao et al. Reference Hao, Wang, Zhang, Ou, Yang, Dan and Jiang2018; Wu et al. Reference Wu, Sun, Liu, Chu and Ding2018). In the case of the HNBs-type gabbro investigated in this study, the hypothesis of plume – subduction-zone interaction is invoked to interpret its generation based on the following evidence.

  1. (1) The Yeqing gabbro is a part of the arc-volcanism system of the SPTO, which is supported by its spatial distribution and intrusion time.

In fact, lots of HNBs and NEBs have been reported on the east coast of the Central Pacific as abnormal arc magmatism within the modern oceanic island arc system (Castillo, Reference Castillo2008; Hoernle et al. Reference Hoernle, Abt, Fischer, Nichols, Hauff, Abers, van den Bogaard, Heydolph, Alvarado, Protti and Strauch2008; Gazel et al. Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011; Fletcher & Wyman, Reference Fletcher and Wyman2015). The subduction polarity of the SPTO is from south to north (ZL Li et al. Reference Li, Yang, Xu, Li, Xu, Ren and Robinson2009; Yang et al. Reference Yang, Xu, Li, Xu, Li, Ren and Robinson2009; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; Mai et al. Reference Mai, Zhu, Yang, Xie, Tong, Hao and Zhong2021; YM Li et al. Reference Li, Yang, Zhu, Xie, Zhong, Mai, Zhou and Zhang2022), proved by the spatial distribution characteristics of the oceanic crust (ophiolites, oceanic islands; Chen et al. Reference Chen, Yang, Li and Xu2010; Duan et al. Reference Duan, Xie, Fan, Wang and Hao2019, Reference Duan, Xie, Wang, Song and Hao2022; B Wang et al. Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019, Reference Wang, Xie, Dong, Fan, Yu and Duan2021; Zhong et al. Reference Zhong, Zhu, Yang, Xie, Mai, Zhang, Li and Zhou2021), trench and initial fore-arc basin (Xie et al. Reference Xie, Song, Wang, Fan and Hao2019, Reference Xie, Duan, Song and Wang2021) and arc magmatism (Geng et al. Reference Geng, Sun, Pan, Zhu and Wang2009; Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; GM Li et al. Reference Li, Zhang, Wu, Xie, Zhu and Han2020; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020; Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Mai et al. Reference Mai, Zhu, Yang, Xie, Tong, Hao and Zhong2021; N Li et al. Reference Li, Yang, Zhu, Xie, Zhong, Mai, Zhou and Zhang2022; Song et al. Reference Song, Xie, Gao, Yu, Wang, Duan and Hao2022; C Wang et al. Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022) from south to north. The intrusion time of the Yeqing gabbro is coincident with the period of the SPTO subduction in the early–middle Permian and middle–late Triassic by studying the eclogites (Li et al. Reference Li, Yang, Xu, Li, Xu, Ren and Robinson2009; Cheng et al. Reference Cheng, Zhang, Vervoort, Lu, Wang and Cao2012, Reference Cheng, Liu, Vervoort and Lu2015; Zhang et al. Reference Zhang, Bader, Zhang, Shen, Li and Li2018 a, b), which is also proved by the contemporaneous arc magmatism mentioned above (Fig. 10).

  1. (2) The flat subduction and slab rollback could scarcely happen in the SPTO.

The young ocean slab with slighter density subducted with a low angle like flat subduction in the early stage due to greater buoyancy, while the density of the subduction slab increased after the subduction slab dehydrated and then slab rollback occurred (Klein & Langmuir, Reference Klein and Langmuir1987; Hawkins et al. Reference Hawkins, Lonsdale, MacDougall and Volpe1990). However, the SPTO had subducted to a deep mantle in the middle Permian (Li et al. Reference Li, Yang, Xu, Li, Xu, Ren and Robinson2009; Cheng et al. Reference Cheng, Zhang, Vervoort, Lu, Wang and Cao2012, Reference Cheng, Liu, Vervoort and Lu2015; Zhang et al. Reference Zhang, Bader, Van, Yang, Shen, Qiu, Li and Zheng2018 b), indicating that slab rollback and flat subduction is unlikely to have taken place during the late Permian to early Triassic. A magma belt parallel to the subduction belt is commonly required to respond to the asthenosphere upwelling after slab rollback, whereas no analogous magma was discovered in the Sumdo area to constitute a magma belt with the Yeqing gabbro.

  1. (3) The slab break-off and ridge subduction are unlikely to lead to formation of the Yeqing gabbro.

Slab break-off generally occurs c. 10 Ma after a continental collision (Benoit et al. Reference Benoit, Aguilloân-Robles, Calmus, Maury, Bellon, Cotton, Bourgois and Michaud2002; Wu et al. Reference Wu, Sun, Liu, Chu and Ding2018). The closure of the SPTO occurred later at least than the late Triassic, supported by the discovery of middle Triassic ophiolite (232–231 Ma; Duan et al. Reference Duan, Xie, Fan, Wang and Hao2019) in Sumdo and a late Triassic – early Jurassic medium-pressure metamorphic belt (225–192 Ma; Dong et al. Reference Dong, Zhang, Liu, Wang, Yu and Shen2011, Reference Dong, Xu, Li, Liu and Li2015; Lin et al. Reference Lin, Zhang, Dong, Xiang and Yan2013; Zhang et al. Reference Zhang, Dong, Santosh and Zhao2014, Reference Zhang, Bader, Van, Yang, Shen, Qiu, Li and Zheng2018 b) between the South and North Lhasa terranes accompanied by the coeval magmatism with a geochemical affinity to syn- or post-collisional plutons (227–180 Ma; HF Zhang et al. Reference Zhang, Xu, Guo, Zong, Cai and Yuan2007; Zhu et al. Reference Zhu, Zhao, Niu, Mo, Chung, Hou, Wang and Wu2011; Li et al. Reference Li, Xu, Yang and Tang2012; C Zhang et al. Reference Zhang, Bader, Van, Yang, Shen, Qiu, Li and Zheng2018 b). Thus, slab break-off did not occur in the SPTO. The Yeqing gabbro is also unlikely to have formed in ridge subduction, which requires a huge volume of magmatism response, such as A-type granite, adakite, high-Mg andesite and high-temperature metamorphic rocks (Hole et al. Reference Hole, Rogers, Saunders and Storey1991; McCrory et al. Reference McCrory, Wilson and Stanley2009; Xu et al. Reference Xu, Li, Wang, Fan, Wu and Li2017), that is missing in the Sumdo area.

  1. (4) The back-arc basin may not be a better position to form the Yeqing gabbro.

Geochemistry characteristics of the Yeqing gabbro show Fe–Ti–P enrichment, and lack of arc- and contamination signature. Fe–Ti-enriched basalts are confined to extensional settings with upwelling of the asthenosphere and have been reported from back-arc basins in the subduction belt (Hollis et al. Reference Hollis, Roberts, Cooper, Earls, Herrington, Condon, Cooper, Archibald and Piercey2012). However, there have been no reports about the sedimentary rock of a back-arc basin in the Sumdo area. On the other hand, the back-arc basin basalts are characterized by mainly strong arc affinity in the early stage, and clear MORB affinity in the late stage (Klein & Langmuir, Reference Klein and Langmuir1987; Hawkins et al. Reference Hawkins, Lonsdale, MacDougall and Volpe1990). In fact, the arc magmatism in the Sumdo area shows typical island-arc characteristics (Zhu et al. Reference Zhu, Mo, Zhao, Niu, Wang, Chu, Pan, Xu and Zhou2010; B Wang et al. Reference Wang, Xie, Dong, Fan and Yu2020, Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022; Mai et al. Reference Mai, Zhu, Yang, Xie, Tong, Hao and Zhong2021; Li et al. Reference Li, Yang, Zhu, Xie, Zhong, Mai, Zhou and Zhang2022; C Wang et al. Reference Wang, Xie, Spier, Dong, Yu, Song, Duan and Yakymchuk2022), which also supports the absence of a back-arc basin.

  1. (5) The plume – subduction-zone interaction may be a better model to explain the formation of Yeqing gabbro.

As discussed above, the Yeqing gabbro derived from a deep garnet lherzolite mantle with the negligible contribution of subducted oceanic crust. The intrusion from deep garnet lherzolite mantle in the subduction belt is usually related to asthenosphere upwelling in an extension environment, which is caused by a slab window or slab rollback. The slab rollback model has been excluded already. Besides, the Yeqing gabbro samples plot in within-plate alkali basalts and ocean island alkaline field like the a-type NEBs/HNBs that are caused by ridge subduction or plume – subduction-zone interaction, different from the b-type NEBs/HNBs caused by flat subduction or slab rollback (Fig. 5c, d). In the P2O5 vs TiO2 and Nb/Yb vs Nb figure (Fig. 11a, b), most of these samples plot in the slab window field. The slab window is common in slab break-off or ridge subduction, which are not suitable for this study. Moreover, the seismic ridge, a series of seamount island chains formed by oceanic crust moving over fixed hotspot/mantle plume, could bring about a slab window while subducting (Gazel et al. Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011; Fletcher & Wyman, Reference Fletcher and Wyman2015). Actually, researchers have reported multiple ocean islands formed from the latest early Carboniferous to middle Permian (B Wang et al. Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Zhong et al. Reference Zhong, Zhu, Yang, Xie, Mai, Zhang, Li and Zhou2021; Duan et al. Reference Duan, Xie, Wang, Song and Hao2022), implying the presence of seismic ridge in the SPTO. At the same time, the hotspot/mantle plume could provide a perfect mantle source for the Yeqing gabbro and favourable conditions for the possibility of plume – subduction-zone interaction.

Fig. 11 (a) P2O5 vs TiO2 figure (Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016). (b) Nb/Yb vs Nb figure (Li et al. Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016). Slab window basalts are from Li et al. (Reference Li, Zhu, Wang, Zhao, Zhang, Liu, Chang, Lu, Dai and Zheng2016 and references therein).

6.c. Geological significance for the subduction evolution of the SPTO

According to the discussion of petrogenesis and tectonic setting above, we propose that the hotspot/mantle plume began to interact with the subduction zone in the SPTO after the middle Permian, and the hotter plume material upwelled through a slab window intruding into the continental edge during the late Permian to early Triassic. Under this framework, the formation of the Yeqing gabbro and subduction evolution of the SPTO can be briefly described as follows combined with the regional geology.

Research into ophiolites and oceanic islands has revealed that the SPTO had already developed an initial ocean basin in the early Carboniferous (Wang et al. Reference Wang, Xie, Dong, Fan, Yu and Duan2021; Duan et al. Reference Duan, Xie, Wang, Song and Hao2022). The SPTO began to subduct not later than the early Permian and subducted to a greater depth beneath the north Lhasa terrane during the middle Permian (Fig. 12a; Yang et al. Reference Yang, Xu, Geng, Li, Xu, Li, Ren, Li, Cai, Liang and Chen2006; Cheng et al. Reference Cheng, Zhang, Vervoort, Lu, Wang and Cao2012; Weller et al. Reference Weller, St-Onge, Rayner, Waters, Searle and Palin2016). Meanwhile, the middle Permian Wenmulang and Ewulang ocean islands imply that the plume under the ocean plate was still active during the middle Permian (Fig. 12a; B Wang et al. Reference Wang, Li, Zeng, Li, Fan, Xie and Hao2019; Zhong et al. Reference Zhong, Zhu, Yang, Xie, Mai, Zhang, Li and Zhou2021). Until the late Permian, the seismic ridge in the SPTO, a structurally weak position of oceanic slab, subducted into the trench, and the slab was torn in the frail place forming a slab window (Gazel et al. Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011) (Fig. 12b, c). The detached slab was replaced by a hot and buoyant asthenosphere mantle, which generated the Yeqing gabbro in this study (Fig. 12b, c). This is the hypothesis put forward in this study to explain the generation of the Yeqing gabbro. Nevertheless, it is mainly based on petrological, geochronological and geochemical observations. Further studies will be required to verify this hypothesis.

Fig. 12 Reconstructed palaeogeography and subduction model of the Sumdo Palaeo–Tethys Ocean during the middle Permian (a) and late Permian to early Triassic (b) (modified after Torsvik & Cocks, Reference Torsvik and Cocks2013; Xie et al. Reference Xie, Duan, Song and Wang2021). (c) The schematic model of the geological processes of plume – subduction-zone interactions required to explain the formation of Yeqing gabbro (modified after Gazel et al. Reference Gazel, Hoernle, Carr, Herzberg, Saginor, Bogaard, Hauff, Feigenson and Swisher2011). GI, Greater India; S, Sibumasu; NL, North Lhasa; SL, South Lhasa; T, Tengchong.

Conclusions

  1. (1) LA-ICP-MS U–Pb ages of zircon from Yeqing gabbro are 254–249 Ma, late Permian to early Triassic, which represents the magmatic crystallization age of the Yeqing gabbro.

  2. (2) The Yeqing gabbro exhibits positive Nb–Ta anomalies, Fe–Ti–P enrichment, lack of arc- and contamination signature, similar to those of OIB and HNBs, indicating that the Yeqing gabbro may be the product of a low degree of partial melting of garnet lherzolite mantle generated from an extensional environment in the subduction belt.

  3. (3) Considering the regional geology of the SPTO, a slab window produced by the plume – subduction-zone interaction is a better explanation for the formation of Yeqing gabbro, proving the SPTO continued to subduct during late Permian to early Triassic.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756822001182

Acknowledgements

This study was financially supported by the National Natural Science Foundation of China (Grant No. 42172226), the Second Tibetan Plateau Scientific Expedition and Research (STEP) Program (Grant No. 2019QZKK0703), and the Independent research fund of Key Laboratory of Mineral Resources Evaluation in Northeast Asia, Department of Natural Resources (DBY-ZZ-18-06).

Conflict of interest

None.

References

Adam, J and Green, TH (2010) Trace element partitioning between mica- and amphibole-bearing garnet lherzolite and hydrous basanitic melt: 2. Tasmanian Cainozoic basalts and the origins of intraplate basaltic magmas. Contributions to Mineralogy and Petrology 161, 883–99.CrossRefGoogle Scholar
Aguillón-Robles, A, Calmus, T, Benoit, M, Bellon, H, Maury, RC, Cotten, J, Bourgois, J and Michaud, F (2001) Late Miocene adakites and Nb-enriched basalts from Vizcaino Peninsula, Mexico: indicators of East Pacific Rise subduction below southern Baja California? Geology 29, 531–4.2.0.CO;2>CrossRefGoogle Scholar
Aldanmaz, E, Pearce, JA and Thirlwall, MF (2000) Petrogenetic evolution of late Cenozoic, post-collision volcanism in western Anatolia, Turkey. Journal of Volcanology and Geothermal Research 102, 6795.CrossRefGoogle Scholar
Baker, JA, Menzies, MA, Thirlwall, MF and Macpherson, CG (1997) Petrogenesis of Quaternary intraplate volcanism, Sana’a, Yemen: implications for plume–lithosphere interaction and polybaric melt hybridization. Journal of Petrology 38, 1359–90.CrossRefGoogle Scholar
Barnes, SJ, Naldrett, AJ and Gorton, MP (1985) The origin of the fractionation of platinum-group elements in terrestrial magmas. Chemical Geology 53, 303–23.CrossRefGoogle Scholar
Benoit, M, Aguilloân-Robles, A, Calmus, T, Maury, RC, Bellon, H, Cotton, J, Bourgois, J and Michaud, F (2002) Geochemical diversity of Late Miocene volcanism in Southern Baja California, Mexico: implication of mantle and crustal sources during the opening of an asthenospheric window. The Journal of Geology 110, 627–48.CrossRefGoogle Scholar
Bouvier, A, Vervoort, JD and Patchett, PJ (2008) The Lu-Hf and Sm-Nd isotopic composition of CHUR: constraints from unequilibrated chondrites and implications for the bulk composition of terrestrial planets. Earth and Planetary Science Letters 273, 4857.CrossRefGoogle Scholar
Castillo, PR (2008) Origin of the adakite–high-Nb basalt association and its implications for postsubduction magmatism in Baja California, Mexico. Geological Society of America Bulletin 120, 451–62.CrossRefGoogle Scholar
Castillo, PR, Rigby, SJ and Solidum, RU (2007) Origin of high field strength element enrichment in volcanic arcs: geochemical evidence from the Sulu Arc, southern Philippines. Lithos 97, 271–88.CrossRefGoogle Scholar
Cao, DD, Cheng, H, Zhang, LM and Wang, K (2017) Post-peak metamorphic evolution of the Sumdo eclogite from the Lhasa terrane of southeast Tibet. Journal of Asian Earth Sciences 143, 156–70.CrossRefGoogle Scholar
Chang, ZS, Vervoort, JD, McClelland, WC and Knaack, C (2006) U-Pb dating of zircon by LA–ICP–MS. Geochemistry, Geophysics, Geosystems 7, 114.CrossRefGoogle Scholar
Chen, SY, Yang, JS, Li, Y and Xu, XZ (2010) Ultramafic terranes in Sumdo region, Lhasa terrane, Eastern Tibet Plateau: an ophiolite unit. Journal of Earth Science 20, 332–47.CrossRefGoogle Scholar
Cheng, H, Liu, YM, Vervoort, JD and Lu, HH (2015) Combined U-Pb, Lu-Hf, Sm-Nd and Ar-Ar multichronometric dating on the Bailang eclogite constrains the closure timing of the paleo-Tethys ocean in the Lhasa terrane, Tibet. Gondwana Research 28, 1482–99.CrossRefGoogle Scholar
Cheng, H, Zhang, C, Vervoort, JD, Lu, HH, Wang, C and Cao, DD (2012) Zircon U–Pb and garnet Lu-Hf geochronology of eclogites from the Lhasa terrane, Tibet. Lithos 155, 341–59.CrossRefGoogle Scholar
Condie, KC (1993) Chemical composition and evolution of the upper continental crust: contrasting results from surface samples and shales. Chemical Geology 104, 137.CrossRefGoogle Scholar
Defant, MJ, Jackson, TE, Drummond, MS, De Bore, JZ, Bellon, H, Feigenson, MD, Maury, RC and Stewart, RH (1992) The geochemistry of young volcanism throughout western Panama and southeastern Costa Rica: an overview. Journal of the Geological Society 149, 569–79.CrossRefGoogle Scholar
DePaolo, DJ (1988) Neodymium isotope geochemistry. Minerals and Rocks 20, 159–74.Google Scholar
Dilek, Y and Furnes, H (2011) Ophiolite genesis and global tectonics: geochemical and tectonic fingerprinting of ancient oceanic lithosphere. Geological Society of America Bulletin 123, 387411.CrossRefGoogle Scholar
Dobosi, G, Kempton, PD, Downes, H, Embey-Isztin, A, Thirlwall, M and Greenwood, P (2003) Lower crustal granulite xenoliths from the Pannonian Basin, Hungary, Part 2: Sr-Nd-Pb-Hf and O isotope evidence for formation of continental lower crust by tectonic emplacement of oceanic crust. Contributions to Mineralogy and Petrology 144, 671–83.CrossRefGoogle Scholar
Dong, HW, Xu, ZQ, Li, Y, Liu, Z and Li, HQ (2015) The Mesozoic metamorphic–magmatic events in the Medog area, the Eastern Himalayan Syntaxis: constraints from zircon U–Pb geochronology, trace elements and Hf isotope compositions in granitoids. International Journal of Earth Sciences 104, 6174.CrossRefGoogle Scholar
Dong, X, Zhang, ZM, Liu, F, Wang, W, Yu, F and Shen, K (2011) Zircon U–Pb geochronology of the Nyainqêntanglha Group from the Lhasa terrane: new constraints on the Triassic orogeny of the south Tibet. Journal of Asian Earth Sciences 42, 732–9.Google Scholar
Duan, ML, Xie, CM, Fan, JJ, Wang, B and Hao, YJ (2019) Identification of the Middle Triassic oceanic crust of the Sumdo in the Tibet Plateau and its constraints on the evolution of the Sumdo Paleo-Tethys Ocean. Earth Science 44, 2249–64 (in Chinese with English abstract).Google Scholar
Duan, ML, Xie, CM, Wang, B, Song, YH and Hao, YJ (2022) Ocean island rock assembly and its tectonic significance in Tangga–Sumdo area, Tibet. Earth Science 47, 2968–84 (in Chinese with English abstract).Google Scholar
Elhlou, S, Belousova, E, Griffin, WL, Pearson, NJ and O’Reilly, SY (2006) Trace element and isotopic composition of GJ-red zircon standard by laser ablation. Geochimica et Cosmochimica Acta 70, A158.CrossRefGoogle Scholar
Fan, JJ, Li, C, Wang, M, Liu, LM and Xie, CM (2017) Remnants of a Late Triassic ocean island in the Gufeng area, northern Tibet: implications for the opening and early evolution of the Bangong–Nujiang Tethys Ocean. Journal of Asian Earth Sciences 2016, 135.CrossRefGoogle Scholar
Fan, JJ, Li, C, Xie, CM, Wang, M and Chen, JW (2015) The evolution of the Bangong–Nujiang Neo-Tethys ocean: evidence from zircon U-Pb and Lu-Hf isotopic analyses of Early Cretaceous oceanic islands and ophiolites. Tectonophysics 655, 2740.CrossRefGoogle Scholar
Fan, JJ, Li, C, Xu, JX and Wang, M (2014) Petrology, geochemistry, and geological significance of the Nadong ocean island, Banggongco-Nujiang suture, Tibetan Plateau. International Geology Review 56, 915–28.CrossRefGoogle Scholar
Fletcher, M and Wyman, DA (2015) Mantle plume–subduction zone interactions over the past 60 Ma. Lithos 233, 162–73.CrossRefGoogle Scholar
Fodor, RV and Vetter, SK (1984) Rift-zone magmatism: petrology of basaltic rocks transitional from CFB to MORB, southeastern Brazil margin. Contributions to Mineralogy and Petrology 88, 307–21.CrossRefGoogle Scholar
Gazel, E, Hoernle, K, Carr, MJ, Herzberg, C, Saginor, I, Bogaard, PVD, Hauff, F, Feigenson, M and Swisher, C III (2011) Plume–subduction interaction in southern Central America: mantle upwelling and slab melting. Lithos 121, 117–34.CrossRefGoogle Scholar
Geng, QR, Sun, ZM, Pan, GT, Zhu, DC and Wang, LQ (2009) Origin of the Gangdise(Transhimalaya) Permian arc in southern Tibet: stratigraphic and volcanic geochemical constraints. Island Arc 18, 467–87.CrossRefGoogle Scholar
Griffin, WL, Powell, WJ, Pearson, NJ and O’Reilly, SY (2008) GLITTER: data reduction software for laser ablation ICP-MS (appendix). In Laser Ablation-ICP-MS in the Earth Sciences (ed. P Sylvester), 204–7. Mineralogical Association of Canada Short Course Series 40.Google Scholar
Hao, LL, Wang, Q, Zhang, CF, Ou, Q, Yang, JH, Dan, W and Jiang, ZQ (2018) Oceanic plateau subduction during closure of the Bangong–Nujiang Tethyan Ocean: insights from central Tibetan volcanic rocks. Geological Society of America Bulletin 131, 864–80.CrossRefGoogle Scholar
Hart, SR and Dunn, T (1993) Experimental cpx/melt partitioning of 24 trace elements. Contributions to Mineralogy and Petrology 113, 18.CrossRefGoogle Scholar
Hastie, AR, Mitchell, SF, Kerr, AC, Minifie, MJ and Millar, IL (2011) Geochemistry of rare high–Nb basalt lavas: are they derived from a mantle wedge metasomatised by slab melts? Geochimica et Cosmochimica Acta 75, 5049–72.CrossRefGoogle Scholar
Hauri, EH, Wagner, TP and Grove, TL (1994) Experimental and natural partitioning of Th, U, Pb and other trace elements between garnet, clinopyroxene and basaltic melts. Chemical Geology 117, 149–66.CrossRefGoogle Scholar
Hawkins, JW, Lonsdale, PF, MacDougall, JD and Volpe, AM (1990) Petrology of the axial ridge of the Mariana Trough backarc spreading center. Earth and Planetary Science Letters 100, 226–50.CrossRefGoogle Scholar
Hoernle, K, Abt, DL, Fischer, KM, Nichols, H, Hauff, F, Abers, GA, van den Bogaard, P, Heydolph, K, Alvarado, G, Protti, M and Strauch, W (2008) Arc-parallel flow in the mantle wedge beneath Costa Rica and Nicaragua. Nature 451, 1094–7.CrossRefGoogle ScholarPubMed
Hole, MJ, Rogers, G, Saunders, AD and Storey, M (1991) Relation between alkalic volcanism and slab–window formation. Geology 19, 657–60.2.3.CO;2>CrossRefGoogle Scholar
Hollis, SP, Roberts, S, Cooper, MR, Earls, G, Herrington, R, Condon, DJ, Cooper, MJ, Archibald, SM and Piercey, SJ (2012) Episodic arc-ophiolite emplacement and the growth of continental margins: late accretion in the Northern Irish sector of the Grampian-Taconic orogeny. Geological Society of America Bulletin 124, 1702–23.CrossRefGoogle Scholar
Hoskin, PWO and Schaltegger, U (2003) The composition of zircon and igneous and metamorphic petrogenesis. Reviews of Mineralogy and Geochemistry 53, 2762.CrossRefGoogle Scholar
Hu, PY, Zhai, QG, Wang, J, Tang, Y, Wang, HT and Hou, KJ (2018) Precambrian origin of the North Lhasa terrane, Tibetan Plateau: constraint from early Cryogenian backarc magmatism. Precambrian Research 313, 5167.CrossRefGoogle Scholar
Hu, PY, Zhai, QG, Zhao, GC, Wang, J, Tang, Y, Zhu, ZC and Wu, H (2019) The North Lhasa terrane in Tibet was attached with the Gondwana before it was drafted away in Jurassic: evidence from detrital zircon studies. Journal of Asian Earth Sciences 185, 104055.CrossRefGoogle Scholar
Hu, ZC, Liu, YS, Gao, S, Liu, WG, Zhang, W, Tong, XR, Lin, L, Zong, KQ, Li, M, Chen, HH, Zhou, L and Yang, L (2012) Improved in situ Hf isotope ratio analysis of zircon using newly designed X skimmer cone and Jet sample cone in combination with the addition of nitrogen by laser ablation multiple collector ICP–MS. Journal of Analytical Atomic Spectrometry 27, 1391–9.CrossRefGoogle Scholar
Jochum, KP, Arndt, NT and Hofmann, AW (1991) Nb–Th–La in komatiites and basalts: constraints on komatiite petrogenesis and mantle evolution. Earth and Planetary Science Letters 107, 272–89.CrossRefGoogle Scholar
Jung, S and Masberg, P (1998) Major- and trace-element systematics and isotope geochemistry of Cenozoic mafic volcanic rocks from the Vogelsberg (central Germany). Journal of Volcanology and Geothermal Research 86, 151–77.CrossRefGoogle Scholar
Kepezhinskas, P, Defant, MJ and Drummond, MS (1996) Progressive enrichment of island arc mantle by melt–peridotite interaction inferred from Kamchatka xenoliths. Geochimica et Cosmochimica Acta 60, 1217–29.CrossRefGoogle Scholar
Klein, EM and Langmuir, CH (1987) Global correlations of ocean ridge basalt chemistry with axial depth and crustal thickness. Journal of Geophysical Research 92, 8089–115.CrossRefGoogle Scholar
Li, C, Dong, YS, Zhai, QG, Wang, LQ, Yan, QR, Wu, YW and He, TT (2008) Discovery of Eopaleozoic ophiolite in the Qiangtang of Tibet Plateau: evidence from SHRIMP U–Pb dating and its tectonic implications. Yanshi Xuebao 24, 336 (in Chinese with English abstract).Google Scholar
Li, C, Zhai, QG, Dong, YS and Huang, X (2006) Discovery of eclogite and its geological significance in Qiangtang area, central Tibet. Chinese Science Bulletin 51, 1095–100 (in Chinese with English abstract).CrossRefGoogle Scholar
Li, GM, Zhang, LK, Wu, JY, Xie, CM, Zhu, LD and Han, FL (2020) Geological reconstruction and scientific significance of the oceanic plate in the southern Qinghai–Tibet plateau. Sedimentary Geology and Tethyan Geology 40, 114 (in Chinese with English abstract).Google Scholar
Li, HQ, Xu, ZQ, Yang, JS and Tang, ZM (2012) Indosinian orogenesis in the Lhasa Terrane, Tibet: new Muscovite 40Ar–39Ar geochronology and evolutionary process. Acta Geologica Sinica – English Edition 86, 1116–27.Google Scholar
Li, N, Yang, WG, Zhu, LD, Xie, L, Zhong, Y, Mai, YJ, Zhou, Y and Zhang, HL (2022) Permian arc magmatism in southern Tibet: implications for the subduction and accretion of the Zhikong–Sumdo Paleo-Tethys Ocean. Gondwana Research 111, 265–79.CrossRefGoogle Scholar
Li, SM, Zhu, DC, Wang, Q, Zhao, ZD, Zhang, LL, Liu, SA, Chang, QS, Lu, YH, Dai, JG and Zheng, YC (2016) Slab-derived adakites and sub-slab asthenosphere-derived OIB-type rocks at 156±2 Ma from the north of Gerze, central Tibet: records of the Bangong–Nujiang oceanic ridge subduction during the Late Jurassic. Lithos 262, 456–69.CrossRefGoogle Scholar
Li, ZL, Yang, JS, Xu, ZQ, Li, TF, Xu, XZ, Ren, YF and Robinson, PT (2009) Geochemistry and Sm-Nd and Rb-Sr isotopic composition of eclogite in the Lhasa terrane, Tibet, and its geological significance. Lithos 109, 240–7.CrossRefGoogle Scholar
Lin, YH, Zhang, ZM, Dong, X, Xiang, H and Yan, R (2013) Early Mesozoic metamorphism and tectonic significance of the eastern segment of the Lhasa terrane, south Tibet. Journal of Asian Earth Sciences 78, 160–83.CrossRefGoogle Scholar
Liu, Y, Liu, HF, Theye, T and Massonne, HJ (2009) Evidence for oceanic subduction at the NE Gondwana margin during Permo-Triassic times. Terra Nova 21, 195202.CrossRefGoogle Scholar
Liu, YM, Li, SZ, Xie, CM, Santosh, M, Liu, YJ, Dong, YC, Wang, B, Guo, RH and Cao, XZ (2022) Subduction–collision and exhumation of eclogites in the Lhasa terrane, Tibet Plateau. Gondwana Research 102, 394404.CrossRefGoogle Scholar
Liu, YS, Gao, S, Hu, ZC, Gao, CG, Zong, KQ and Wang, DB (2010) Continental and oceanic crust recycling-induced melt-peridotite interactions in the Trans-North China Orogen: U–Pb dating, Hf isotopes and trace elements in zircons from mantle xenoliths. Journal of Petrology 51, 537–71.CrossRefGoogle Scholar
Ludwig, KR (2003) User’s manual for isoplot/Ex, version 3.0//A Geochronological Toolkit for Microsoft Excel. Berkeley Geochronology Center Special Publication 4, 170.Google Scholar
Luhr, JF, Aranda-Gomez, JJ and Housh, TB (1995) San Quintin volcanic field, Baja California Norte, Mexico: geology, petrology and geochemistry. Journal of Geophysical Research 100, 353–80.CrossRefGoogle Scholar
Mai, YJ, Zhu, LD, Yang, WG, Xie, L, Tong, X, Hao, JY and Zhong, Y (2021) Zircon U⁃Pb and Hf isotopic composition of Permian felsic tuffs in southeastern margin of Lhasa, Tibet. Earth Science 46, 3880–91 (in Chinese with English abstract).Google Scholar
McCrory, PA, Wilson, DS and Stanley, RG (2009) Continuing evolution of the Pacific–Juan de Fuca–North America slab window system: a trench-ridge-transform example from the Pacific Rim. Tectonophysics 464, 3042.CrossRefGoogle Scholar
McKenzie, D and O’Nions, RK (1991) Partial melt distributions from inversion of rare earth element concentrations. Journal of Petrology 32, 1021–91.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
Metcalfe, I (2013) Gondwana dispersion and Asian accretion: tectonic and palaeogeographic evolution of eastern Tethys. Journal of Asian Earth Sciences 66, 133.CrossRefGoogle Scholar
Mullen, ED (1983) MnO/TiO2/P2O5: a minor element discriminant for basaltic rocks of oceanic environments and its implications for petrogenesis. Earth and Planetary Science Letters 62, 5362.CrossRefGoogle Scholar
Neal, CR, Mahoney, JJ and Chazey, WJ (2002) Mantle sources and the highly variable role of continental lithosphere in basalt petrogenesis of the Kerguelen Plateau and Broken Ridge LIP: results from ODP Leg 183. Journal of Petrology 43, 1177–205.CrossRefGoogle Scholar
Niu, YL (2009) Some basic concepts and problems on the petrogenesis of intra-plate ocean island basalts. Chinese Science Bulletin 54, 418460.CrossRefGoogle Scholar
Nowell, GM, Kempton, PD, Noble, SR, Fitton, JG, Saunders, AD, Mahoney, JJ and Taylor, RN (1998) High precision Hf isotope measurements of MORB and OIB by thermal ionisation mass spectrometry: insights into the depleted mantle. Chemical Geology 149, 211–33.CrossRefGoogle Scholar
Pearce, JA (2014) Immobile element fingerprinting of ophiolites. Elements 10, 101–8.CrossRefGoogle Scholar
Peng, ZX, Mahoney, J, Hooper, P, Harris, C and Beane, J (1994) A role for lower continental crust in flood basalt genesis? Isotopic and incompatible element study of the lower six formations of the western Deccan Traps. Gochimica et Cosmochimica Acta 58, 267–88.CrossRefGoogle Scholar
Reagan, MK and Gill, JB (1989) Coexisting calcalkaline and high-niobium basalts from Turrialba Volcano, Costa Rica: implications for residual titanates in arc magma sources. Journal of Geophysical Research: Solid Earth 94, 4619–33.CrossRefGoogle Scholar
Robinson, JAC and Wood, BJ (1998) The depth of the spinel to garnet transition at the peridotite solidus. Earth and Planetary Science Letters 164, 277–84.CrossRefGoogle Scholar
Rollinson, HR (1993) Using Geochemical Data Evaluation, Presentation, Interpretation. London: Longman Geochemistry Society.Google Scholar
Rudnick, RL and Gao, S (2003) Composition of the continental crust. Treatise on Geochemistry 3, 164.Google Scholar
Sajona, FG, Maury, RC, Bellon, H, Cotton, J and Defant, MJ (1996) High field strength element enrichment of Pliocene–Pleistocene island arc basalts, Zamboanga Peninsula, Western Mindanao (Philippines). Journal of Petrology 37, 693726.CrossRefGoogle Scholar
Şengor, AMC (1987) Tectonics of the Tethysides-Orogenic collage development in a collisional setting. Annual Review of Earth and Planetary Science 15, 213–44.CrossRefGoogle Scholar
Sinton, JM, Wilson, DS, Christie, DM, Hey, RN and Delaney, JR (1983) Petrologic consequences of rift propagation on oceanic spreading ridges. Earth and Planetary Science Letters 62, 193207.CrossRefGoogle Scholar
Song, YH, Xie, CM, Gao, ZW, Yu, YP, Wang, B, Duan, ML and Hao, YJ (2022) Tectonic transition from Paleo- to Neo-Tethyan Ocean in Tangjia–Sumdo area, Southern Tibet: constraints from early Jurassic magmatism. Gondwana Research 105, 1224.CrossRefGoogle Scholar
Storey, M, Rogers, G, Saunders, AD and Terrell, DJ (1989) San Quintin volcanic field, Baja California, Mexico: “Within-plate” magmatism following ridge subduction. Terra Nova 1, 195202.CrossRefGoogle Scholar
Sun, S-S and McDonough, WF (1989) Chemical and isotopic systematics of oceanic basalt: implications for mantle composition and processes. In Magmatism in the Ocean Basins (eds Saunders, AD and Norry, MJ), pp. 313–45. Geological Society of London, Special Publication no. 42.Google Scholar
Thompson, RN, Morrison, MA, Hendry, GL, Parry, SJ, Simpson, PR, Hutchison, R and O’Hara, MJ (1984) An assessment of the relative roles of crust and mantle in magma genesis: an elemental approach [and discussion]. Philosophical Transactions of the Royal Society of London 310, 549–90.Google Scholar
Thorkelson, DJ (1996) Subduction of diverging plates and the principles of slab window formation. Tectonophysics 255, 4763.CrossRefGoogle Scholar
Thorkelson, DJ, Madsen, JK and Sluggett, CL (2011) Mantle flow through the Northern Cordilleran slab window revealed by volcanic geochemistry. Geology 39, 267–70.CrossRefGoogle Scholar
Torsvik, TH and Cocks, LRM (2013) Gondwana from top to base in space and time. Gondwana Research 24, 9991030.CrossRefGoogle Scholar
Wang, B, Xie, CM, Dong, YS, Fan, JJ and Yu, YP (2020) Middle Permian adakitic granite dikes in the Sumdo region, Central Lhasa Terrane, Central Tibet: implications for the subduction of the Sumdo Paleo-Tethys Ocean. Journal of Asian Earth Sciences 205, 104610.CrossRefGoogle Scholar
Wang, B, Xie, CM, Dong, YS, Fan, JJ, Yu, YP and Duan, ML (2021) Middle–late Permian mantle plume/hotspot–ridge interaction in the Sumdo Paleo-Tethys Ocean region, Tibet: evidence from mafic rocks. Lithos 390–391, 106128.CrossRefGoogle Scholar
Wang, B, Xie, CM, Fan, JJ, Wang, M, Dong, YC and Hao, YJ (2019) Genesis and tectonic setting of Middle Permian OIB-type mafic rocks in the Sumdo area, southern Lhasa terrane. Lithos, 324–325, 429–38.Google Scholar
Wang, B, Xie, CM, Spier, CA, Dong, YS, Yu, YP, Song, YH, Duan, ML and Yakymchuk, C (2022) Middle Permian magmatism in the Tangjia-Sumdo region, Tibet: evidence for intra-oceanic subduction. International Geology Review. doi: 10.1080/00206814.2022.2056718.Google Scholar
Wang, C, Ding, L, Cai, FL, Wang, HQ, Zhang, LY and Yue, YH (2022) Evolution of the Sumdo Paleo-Tethyan Ocean: constraints from Permian Luobadui Formation in Lhasa terrane, South Tibet. Palaeogeography, Palaeoclimatology, Palaeoecology 595, 110974.CrossRefGoogle Scholar
Wang, M, Li, C, Wu, YW and Xie, CM (2014) Geochronology, geochemistry, Hf isotopic compositions and formation mechanism of radial mafic dikes in northern Tibet. International Geology Review 56, 187205.CrossRefGoogle Scholar
Wang, M, Li, C, Zeng, XW, Li, H, Fan, JJ, Xie, CM and Hao, YJ (2019) Petrogenesis of the southern Qiangtang mafic dykes, Tibet: link to a late Paleozoic mantle plume on the northern margin of Gondwana? Geological Society of America Bulletin 131, 1907–19.CrossRefGoogle Scholar
Wang, Q, Wyman, DA, Xu, J, Wan, YS, Li, CF, Zi, F, Jiang, ZQ, Qiu, HN, Chu, ZY, Zhao, ZH and Dong, YH (2007) Triassic Nb-enriched basalts, magnesian andesites, and adakites of the Qiangtang terrane (Central Tibet): evidence for metasomatism by slab-derived melts in the mantle wedge. Contributions to Mineralogy & Petrology 155, 473–90.CrossRefGoogle Scholar
Weller, OM, St-Onge, MR, Rayner, N, Waters, DJ, Searle, MP and Palin, RM (2016) U–Pb zircon geochronology and phase equilibria modelling of a mafic eclogite from the Sumdo complex of south-east Tibet: insights into prograde zircon growth and the assembly of the Tibetan Plateau. Lithos 262, 729–41.CrossRefGoogle Scholar
Wilson, M (1989) Igneous Petrogenesis: A Global Approach. London: Unwin Hyman Dostal, 466 pp.CrossRefGoogle Scholar
Winchester, JA and Floyd, PA (1977) Geochemical discrimination of different magma series and their differentiation products using immobile elements. Chemical Geology 20, 325–43.CrossRefGoogle Scholar
Wu, FY, Li, XH, Zheng, YF and Gao, S (2007) Lu-Hf isotope systematics and their application in petrology. Acta Petrologica Sinica 23, 185220 (in Chinese with English abstract).Google Scholar
Wu, H, Sun, SL, Liu, HY, Chu, H and Ding, W (2018) An early Cretaceous slab window beneath central Tibet, SW China: evidence from OIB-like alkaline gabbros in the Duolong area. Terra Nova 31, 6775.Google Scholar
Xie, CM, Duan, ML, Song, YH and Wang, B (2021) Provenance and tectonic setting of the Sumdo Formation in the Lhasa Terrane, Tibet: implications for early subduction evolution of the Sumdo Paleo-Tethys Ocean. Palaeogeography, Palaeoclimatology, Palaeoecology 584, 110712.CrossRefGoogle Scholar
Xie, CM, Li, C, Fan, JJ and Su, L (2017b) Ordovician sedimentation and bimodal volcanism in the Southern Qiangtang terrane of northern Tibet: implications for the evolution of the northern Gondwana margin. International Geology Review 59, 2078–105.CrossRefGoogle Scholar
Xie, CM, Li, C, Ren, YS, Wang, M and Su, L (2017a) Detrital provenance, depositional environment, and palaeogeographic implications of Lower Triassic marine sediments in central Tibet. International Geology Review 60, 113.Google Scholar
Xie, CM, Song, YH, Wang, M, Fan, JJ and Hao, YJ (2019) Age and provenance of Sumdo formation in central Gangdise, Tibetan plateau: detrital Zircon U–Pb geochronological evidence. Earth Science 44, 2224–33 (in Chinese with English abstract).Google Scholar
Xu, W, Li, C, Wang, M, Fan, JJ, Wu, H and Li, X (2017) Subduction of a spreading ridge within the Bangong Co–Nujiang Tethys Ocean: evidence from early Cretaceous mafic dykes in the Duolong porphyry Cu–Au deposit, western Tibet. Gondwana Research 41, 128–41.CrossRefGoogle Scholar
Xu, ZQ, Dilek, Y, Cao, H, Yang, JS, Robinson, P, Ma, CQ, Li, HQ, Jolivet, M, Roger, F and Chen, XJ (2015) Paleo-Tethyan evolution of Tibet as recorded in the East Cimmerides and West Cathaysides. Journal of Asian Earth Sciences 105, 320–37.CrossRefGoogle Scholar
Yang, JS, Xu, ZQ, Geng, QR, Li, ZL, Xu, XZ, Li, TF, Ren, YF, Li, HQ, Cai, ZH, Liang, FH and Chen, SY (2006) A possible new HP/UHP metamorphic belt in China: discovery of eclogite in the Lhasa terrane. Tibet. Acta Geologica Sinica 80, 1787–92 (in Chinese with English abstract).Google Scholar
Yang, JS, Xu, ZQ, Li, ZL, Xu, XZ, Li, TF, Ren, YF and Robinson, PT (2009) Discovery of an eclogite belt in the Lhasa terrane, Tibet: a new border for Paleo-Tethys? Journal of Asian Earth Sciences 34, 7689.CrossRefGoogle Scholar
Yin, A and Harrison, TM (2000) Geologic evolution of the Himalayan-Tibetan orogen. Annual Review of Earth and Planetary Sciences 28, 211–80.CrossRefGoogle Scholar
Zeng, LS, Liu, J, Gao, LE, Chen, FY and Xie, KJ (2009) Early Mesozoic high–pressure metamorphism within the Lhasa terrane, Tibet and implications for regional tectonics. Earth Science Frontiers 16, 140–51.CrossRefGoogle Scholar
Zhai, QG, Jahn, BM, Wang, J, Hu, PY, Chung, SL, Lee, HY, Tang, SH and Tang, Y (2016) Oldest Paleo-Tethyan ophiolitic mélange in the Tibetan Plateau. Geological Society of America 128, 119.CrossRefGoogle Scholar
Zhai, QG, Jahn, BM, Wang, J, Su, L, Mo, XX, Wang, KL, Tang, SH and Lee, HY (2013) The Carboniferous ophiolite in the middle of the Qiangtang terrane, northern Tibet: SHRIMP U–Pb dating, geochemical and Sr-Nd-Hf isotopic characteristics. Lithos 168, 186–99.CrossRefGoogle Scholar
Zhai, QG, Wang, J, Li, C and Su, L (2010) SHRIMP U–Pb dating and Hf isotopic analyses of Middle Ordovician meta-cumulate gabbro in central Qiangtang, northern Tibetan Plateau. Science in China, Series D, Earth Science 53, 657–64.CrossRefGoogle Scholar
Zhang, C, Bader, T, Van, HR, Yang, JS, Shen, TT, Qiu, T and Li, P (2018b) The metamorphic evolution and tectonic significance of the Sumdo HP-UHP metamorphic terrane, central–south Lhasa Block, Tibet. In Metamorphism and Tectonic Evolution of Orogenic Belts (ed. Zheng, LF), pp. 209–29. Geological Society of London Special Publication no. 474,.Google Scholar
Zhang, C, Bader, T, Zhang, LG, Shen, TT, Li, P and Li, XP (2018a) Metamorphic evolution and age constraints of the garnet-bearing mica schist from the Xindaduo area of the Sumdo (U)HP metamorphic belt, Tibet. Geological Magazine 156, 1175–89.CrossRefGoogle Scholar
Zhang, HF, Xu, WC, Guo, JQ, Zong, KQ, Cai, HM and Yuan, HL (2007) Indosinian orogenesis of the Gangdise Terrane: evidences from Zircon U–Pb dating and petrogenesis of granitoids. Earth Science 32, 155–66 (in Chinese with English abstract).Google Scholar
Zhang, XZ, Dong, YS, Wang, Q, Dan, W, Zhang, C, Xu, W and Huang, ML (2017) Metamorphic records for subduction erosion and subsequent underplating processes revealed by garnet–staurolite–muscovite schists in central Qiangtang, Tibet. Geochemistry, Geophysics, Geosystems 18, 266–79.CrossRefGoogle 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.CrossRefGoogle Scholar
Zhong, Y, Zhu, LD, Yang, WG, Xie, L, Mai, YJ, Zhang, HL, Li, N and Zhou, Y (2021) Geological significance of the newly discovered Middle Permian ocean island basalt-type gabbros in Ewulang, Nianqing–Sumdo, Tibet. Geological Journal 56, 4523–37.CrossRefGoogle Scholar
Zhu, DC, Mo, XX, Niu, YL, Zhao, ZD, Yang, YH and Wang, LQ (2009) Zircon U–Pb dating and in-situ Hf isotopic analysis of Permian peraluminous granite in the Lhasa terrane, southern Tibet: implications for Permian collisional orogeny and paleogeography. Tectonophysics 469, 4860.CrossRefGoogle Scholar
Zhu, DC, Mo, XX, Zhao, ZD, Niu, YL, Wang, LQ, Chu, QH, Pan, GT, Xu, JF and Zhou, CY (2010) Presence of Permian extension- and arc-type magmatism in southern Tibet: paleogeographic implications. Geological Society of America Bulletin 122, 979–93.CrossRefGoogle Scholar
Zhu, DC, Zhao, ZD, Niu, YL, Hou, ZQ and Mo, XX (2013) The origin and pre-Cenozoic evolution of the Tibetan Plateau. Gondwana Research 23, 1429–54.CrossRefGoogle Scholar
Zhu, DC, Zhao, ZD, Niu, YL, Mo, XX, Chung, SL, Hou, ZQ, Wang, LQ and Wu, FY (2011) The Lhasa Terrane: record of a microcontinent and its histories of drift and growth. Earth and Planetary Science Letters 301, 241–55.CrossRefGoogle Scholar
Zindler, A and Hart, S (1986) Chemical geodynamics. Annual Review of Earth and Planetary Sciences 14, 493571.CrossRefGoogle Scholar
Figure 0

Fig. 1. (a) Tectonic framework of the Tibetan Plateau (modified after Li et al.2006 and Zhu et al.2010). (b) Geological sketch map of the Tanga–Sumdo area; published age data are after Chen et al. (2009), Yang et al. (2009), Cheng et al. (2012, 2014), Weller et al. (2016), Cao et al. (2017), Duan et al. (2019, 2022), B Wang et al. (2019) and Song et al. (2022).

Figure 1

Fig. 2. Photographs of Yeqing gabbro. (a) Macro outcrop photo of Yeqing gabbro intruding into Sumdo Formation. (b) Close-up photo of the boundary between Yeqing gabbro and Sumdo Formation. (c, d) Micrograph of Yeqing gabbro, Plagioclase is replaced by sericite, and pyroxene is replaced by hornblende, in part of our samples. Pl – plagioclase; Px – pyroxene; hor – hornblende; Sre – sericite; Mag – magnetite.

Figure 2

Fig. 3. Cathodoluminescence (CL) images and chondrite-normalized REE patterns diagram of representative zircon grains from Yeqing gabbro. The yellow circle is the location of U–Pb isotope analysis, and the red is Lu–Hf isotope analysis. Values of chondrite are after Sun and McDonough (1989).

Figure 3

Fig. 4. U–Pb zircon Concordia of representative zircon grains from Yeqing gabbro. MSWD = mean squared weighted deviation.

Figure 4

Fig. 5. (a) SiO2 vs Nb/Y (Winchester & Floyd, 1977) plot. (b) Nb/La vs MgO (Kepezhinskas et al.1996) plot. (c) Nb*2 vs Zr/4 vs Y figure (Meschede, 1986). Within-plate alkali basalts – AI, AII within-plate tholeiites – AII, C; plume-type MORB- B; N-type MORB- D; volcanic arc basalts – C, D. (d) TiO2 vs MnO2*10 vs P2O5*10 figure (Mullen, 1983). MORB – mid-ocean ridge basalt; IAT – island arc tholeiite; CAB – calc-alkaline basalt; OIT – ocean island tholeiite; OIA – ocean island alkaline. Data of the NEBs, HNBs from Baja, Nicaragua, Renso, Dubuzha, Tuotuohe and Gerze are after Storey et al. (1989), Luhr et al. (1995), Benoit et al. (2002), Wang et al. (2007), Gazel et al. (2011), Li et al. (2016) and Hao et al. (2018).

Figure 5

Fig. 6. Chondrite-normalized REE patterns and primitive-mantle-normalized spider diagrams. Values of chondrite, primitive mantle, OIB and E-MORB are after Sun and McDonough (1989).

Figure 6

Fig. 7. (La/Nb)PM vs (Th/Nb)PM plot (Neal et al.2002). Middle crust and lower crust data are after Rudnick and Gao (2003). The ‘most oceanic basalts’ data are after Neal et al. (2002 and references therein).

Figure 7

Fig. 8 Harker variation diagrams for the Yeqing gabbro. The red arrows represent variation trend of part of major (wt %) and trace (ppm) elements towards the increase of MgO (wt %).

Figure 8

Fig. 9 (a) Th/Yb vs Nb/Yb plot and (b) TiO2/Yb vs Nb/Yb plot (Pearce, 2014). (c) La/Sm vs Sm/Yb plot. Mantle array (heavy line) defined by depleted MORB mantle (DMM; McKenzie & O’Nions, 1991) and primitive mantle (PM; Sun & McDonough, 1989); Melting curves for spinel lherzolite and garnet peridotite with both DMM and PM compositions are after Aldanmaz et al. (2000). Numbers along lines represent the degree of partial melting. (d) ϵNd(t) vs (87Sr/86Sr)i plot. Mantle arrays are after Zindler and Hart (1986). Bulk Earth is after Depaolo (1988). OIB is after Wilson (1989). The date of Sumdo eclogite is from Li et al. (2009). Dates of Halberstadt, Gerze NEBs and HNBs are after Hastie et al. (2011) and Hao et al. (2018).

Figure 9

Fig. 10 Distribution diagram of ages related to the Sumdo Palaeo-Tethys Ocean (data are quoted from Geng et al.2009; Yang et al.2009; Zhu et al.2009, 2010; Cheng et al.2012, 2014; Zhang et al.2018; Duan et al.2019, 2022; Xie et al.2019, 2021; B Wang et al.2019, 2020, 2021,2022; Li et al.2020, 2022; Mai et al.2021; Zhong et al.2021; Song et al.2022; C Wang et al.2022).

Figure 10

Fig. 11 (a) P2O5 vs TiO2 figure (Li et al.2016). (b) Nb/Yb vs Nb figure (Li et al.2016). Slab window basalts are from Li et al. (2016 and references therein).

Figure 11

Fig. 12 Reconstructed palaeogeography and subduction model of the Sumdo Palaeo–Tethys Ocean during the middle Permian (a) and late Permian to early Triassic (b) (modified after Torsvik & Cocks, 2013; Xie et al.2021). (c) The schematic model of the geological processes of plume – subduction-zone interactions required to explain the formation of Yeqing gabbro (modified after Gazel et al.2011). GI, Greater India; S, Sibumasu; NL, North Lhasa; SL, South Lhasa; T, Tengchong.

Supplementary material: File

Duan et al. supplementary material

Duan et al. supplementary material

Download Duan et al. supplementary material(File)
File 90.4 KB