Hostname: page-component-76fb5796d-wq484 Total loading time: 0 Render date: 2024-04-28T19:30:15.914Z Has data issue: false hasContentIssue false

Geochronology, geochemistry and tectonic implications of early Carboniferous plutons in the southwestern Alxa Block

Published online by Cambridge University Press:  12 November 2021

Zeng-Zhen Wang
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Xuan-Hua Chen
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Zhao-Gang Shao*
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Bing Li
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Hong-Xu Chen
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Wei-Cui Ding
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Yao-Yao Zhang
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
Yong-Chao Wang
Affiliation:
Chinese Academy of Geological Sciences, Beijing100037, China SinoProbe Center, Chinese Academy of Geological Sciences and China Geological Survey, Beijing100037, China
*
Author for correspondence: Zhao-Gang Shao, Email: shaozhaogang@sina.com
Rights & Permissions [Opens in a new window]

Abstract

The southeastern Central Asian Orogenic Belt (CAOB) records the assembly process between several micro-continental blocks and the North China Craton (NCC), with the consumption of the Paleo-Asian Ocean (PAO), but whether the S-wards subduction of the PAO beneath the northern NCC was ongoing during Carboniferous–Permian time is still being debated. A key issue to resolve this controversy is whether the Carboniferous magmatism in the northern NCC was continental arc magmatism. The Alxa Block is the western segment of the northern NCC and contiguous to the southeastern CAOB, and their Carboniferous–Permian magmatism could have occurred in similar tectonic settings. In this contribution, new zircon U–Pb ages, elemental geochemistry and Sr–Nd isotopic analyses are presented for three early Carboniferous granitic plutons in the southwestern Alxa Block. Two newly identified aluminous A-type granites, an alkali-feldspar granite (331.6 ± 1.6 Ma) and a monzogranite (331.8 ± 1.7 Ma), exhibit juvenile and radiogenic Sr–Nd isotopic features, respectively. Although a granodiorite (326.2 ± 6.6 Ma) is characterized by high Sr/Y ratios (97.4–139.9), which is generally treated as an adikitic feature, this sample has highly radiogenic Sr–Nd isotopes and displays significantly higher K2O/Na2O ratios than typical adakites. These three granites were probably derived from the partial melting of Precambrian continental crustal sources heated by upwelling asthenosphere in lithospheric extensional setting. Regionally, both the Alxa Block and the southeastern CAOB are characterized by the formation of early Carboniferous extension-related magmatic rocks but lack coeval sedimentary deposits, suggesting a uniform lithospheric extensional setting rather than a simple continental arc.

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), 2021. Published by Cambridge University Press

1. Introduction

The Phanerozoic Central Asian Orogenic Belt (CAOB), one of the largest long-lived accretionary orogens worldwide, is situated to the north of the Tarim–North China cratons (Fig. 1a) and formed by complex subduction, accretion and collision processes related to the consumption of the Paleo-Asian Ocean (PAO), with significant crustal growth (Han et al. Reference Han, Wang, Jahn, Hong, Kagami and Sun1997, Reference Han, He, Wang and Guo2011; Jahn et al. Reference Jahn, Wu and Hong2000; Wu et al. Reference Wu, Jahn, Wilde, Lo, Yui, Lin, Ge and Sun2003; Windley et al. Reference Windley, Alexeiev, Xiao, Kröner and Badarch2007; Xiao et al. Reference Xiao, Windley, Han, Liu, Wan, Zhang, Ao, Zhang and Song2018). The southeastern CAOB records the Palaeozoic amalgamation between the North China Craton (NCC) in the south and Mongolia, Hunshandake and Songliao blocks within the CAOB in the north (Xu et al. Reference Xu, Charvet, Chen, Zhao and Shi2013; Zhao et al. Reference Zhao, Wang, Huang, Dong, Li, Zhang and Yu2018; Zhou et al. Reference Zhou, Wilde, Zhao and Han2018). The Permian–Early Triassic Solonker suture (Solonker–Xar Moron–Changchun suture) contains the youngest ophiolites within the southeastern CAOB and is usually regarded as the terminal closure site of the PAO (Eizenhöfer & Zhao, Reference Eizenhöfer and Zhao2018; Wilde & Zhou, Reference Wilde and Zhou2015; Xiao et al. Reference Xiao, Windley, Jie and Zhai2003). However, when and how the PAO finally closed in the southeastern CAOB is still controversial, and different opinions can be grouped into three models.

Fig. 1. (a) Tectonic location of the Alxa Block. (b) Schematic geological map showing the distribution of Palaeozoic intrusions and ophiolitic mélanges in the Alxa Block (modified after Dan et al. Reference Dan, Li, Wang, Tang and Liu2014). (c) Simplified geological map of the southwestern Alxa Block (modified after Wang et al. Reference Wang, Chen, Shao, Li, Ding, Zhang, Wang, Zhang, Xu and Qin2020).

In the first set of models, the subduction of the PAO was continuous from the early Palaeozoic Era to Late Permian–Early Triassic time and led to the successive accretion of micro-continental blocks and magmatic arcs to the northern NCC, with the northern margin of the NCC as a continental arc during Carboniferous–Permian time and the Solonker suture as the final closure site of the PAO (e.g. Xiao et al. Reference Xiao, Windley, Jie and Zhai2003, Reference Xiao, Windley, Huang, Han, Yuan, Chen, Sun, Sun and Li2009 b, Reference Xiao, Windley, Han, Liu, Wan, Zhang, Ao, Zhang and Song2018; Zhang et al. Reference Zhang, Zhao, Ye, Liu and Hu2014, Reference Zhang, Zhao, Liu and Hu2016d ). The second set of models propose the Late Devonian–early Carboniferous closure of the PAO, with the southeastern CAOB in a post-collisional setting since then (e.g. Xu et al. Reference Xu, Charvet, Chen, Zhao and Shi2013; Tong et al. Reference Tong, Jahn, Wang, Hong, Smith, Sun, Gao, Yang and Huang2015; Zhang et al. Reference Zhang, Yuan, Xue, Yan and Mao2015 b). The third set of models infer that the large-scale PAO closed before the Late Devonian Epoch, but a new orogenic cycle began with intra-continental rifting within the southeastern CAOB during early Carboniferous time and resulted in the formation of a Red-Sea-like limited ocean basin, with the Solonker suture marking its closure during the Early Triassic Epoch (e.g. Zhang et al. Reference Zhang, Wei and Chu2015 a; Luo et al. Reference Luo, Xu, Shi, Zhao, Faure and Chen2016; Pang et al. Reference Pang, Wang, Xu, Zhao, Feng, Wang, Luo and Liao2016; Zhao et al. Reference Zhao, Xu and Zhang2017; Xu et al. Reference Xu, Wang, Zhang, Wang, Yang and He2018). In the third model, the lithospheric extension may be triggered by slab break-off (Kozlovsky et al. Reference Kozlovsky, Yarmolyuk, Salnikova, Travin, Kotov, Plotkina, Kudryashova and Savatenkov2015; Zhang et al. Reference Zhang, Gao, Wang, Liu and Ma2012 a) and enhanced by slab avalanche-driven wet mantle upwelling rising from the hydrous mantle transition zone (Wang et al. Reference Wang, Wilde, Li and Yang2015 a, Reference Wang, Wilde, Xu and Pang2016 a).

To test the likelihood of one of these geodynamic models, a key question is whether the Carboniferous–Permian tectono-magmatic activity of the southeastern CAOB was dominated by the continued S-wards subduction of the PAO or by lithospheric extension. Accordingly, the tectonic setting of the Carboniferous magmatism in the northern margin of the NCC, either continental arc or lithospheric extension, can provide insights into the terminal evolutionary history of the southeastern CAOB.

The Alxa Block, also known as the Alxa Tectonic Belt (Song et al. Reference Song, Xiao, Collins, Glorie, Han and Li2018), connects the NCC to the east and the Tarim Craton to the west and lies between the CAOB to the north and the North Qilian Orogen to the south (Fig. 1a). Although this block is largely covered by deserts, numerous Phanerozoic plutons intruding into Precambrian metamorphic basement rocks crop out in its southwestern and northeastern parts (Fig. 1b). These plutons are mostly Palaeozoic in age, spanning Middle Ordovician–Early Devonian time (c. 458–394 Ma) and end Late Devonian–end Permian time (c. 359–252 Ma; Fig. 2a). Notably, this age pattern is quite similar to that of the southeastern CAOB (including the northern NCC), which includes two magmatic stages of middle Cambrian–Middle Devonian time (c. 508–386 Ma) and end Late Devonian–end Permian time (c. 362–252 Ma; Fig. 2b), indicating an operation of comparable tectonic processes. Further, the early magmatic stage in the southwestern Alxa Block could also be related to the North Qilian Orogen (Duan et al. Reference Duan, Li, Qian and Jiao2015; Zhang et al. Reference Zhang, Zhang, Zhang, Xiong, Luo, Yang, Pan, Zhou, Xu and Guo2017 a; Wang et al. Reference Wang, Chen, Shao, Li, Ding, Zhang, Wang, Zhang, Xu and Qin2020), but the Qilian orogenesis ended before the Late Devonian Epoch (Xiao et al. Reference Xiao, Windley, Yong, Yan, Yuan, Liu and Li2009 a; Song et al. Reference Song, Niu, Su and Xia2013). The Carboniferous magmatism within the Alxa Block was therefore most likely related to the tectono-magmatic activity of the southeastern CAOB.

Fig. 2. Statistical histograms of zircon U–Pb ages of Palaeozoic magmatic rocks in the (a) Alxa Block (data from this study and Qin, Reference Qin2012; Tang, Reference Tang2015; Gong et al. Reference Gong, Zhang, Wang, Yu and Wang2018 a; Zhang et al. Reference Zhang, Wang, Wang, Liu, Liu and Wu2018d ; Liu et al. Reference Liu, Pan, Liu, Wang, Wang and Xue2019; Pan, Reference Pan2019; Song et al. Reference Song, Xiao, Collins, Glorie and Han2019; Chen et al. Reference Chen, Zhao, Wang, Rong and Li2020; Wang et al. Reference Wang, Chen, Shao, Li, Ding, Zhang, Wang, Zhang, Xu and Qin2020; Zhao et al. Reference Zhao, Liu, Wang, Zhang and Guan2020) and (b) the southeastern Central Asian Orogenic Belt (data from Wang et al. Reference Wang, Han, Feng and Liu2015 b).

In this study, new geochronological, elemental and isotopic geochemical analyses of three early Carboniferous plutons in the southwestern Alxa Block are presented. These results, combined with regional correlations, suggest a lithospheric extensional setting rather than a simple continental arc for the development of early Carboniferous magmatism in both the Alxa Block and the southeastern CAOB.

2. Geological background

The Alxa Block is separated from the CAOB by the Enger Us Fault to the north, and from the North Qilian Orogen to the SW by the Longshoushan Fault (Fig. 1b). It is traditionally considered as the western part of the northern NCC (Fig. 1a), either the western part of the Yinshan Block (e.g. Zhao et al. Reference Zhao, Sun, Wilde and Li2005, Reference Zhao, Cawood, Li, Wilde, Sun, Zhang, He and Yin2012; Wan et al. Reference Wan, Song, Liu, Wilde, Wu, Shi, Yin and Zhou2006; Wang et al. Reference Wang, Han, Feng, Liu, Zheng and Kong2016 b, Reference Wang, Li, Ning, Kusky and Deng2019 a) or the western extension of the Khondalite Belt (e.g. Geng et al. Reference Geng, Wang, Wu and Zhou2010; Zhang et al. Reference Zhang, Gong, Yu, Li and Hou2013 a; Zhang & Gong, Reference Zhang and Gong2018). However, a close affinity of the Alxa Block to the Tarim or South China cratons had also been proposed (e.g. Tung et al. Reference Tung, Yang, Liu, Zhang, Tseng and Wan2007; Yuan & Yang, Reference Yuan and Yang2015; Song et al. Reference Song, Xiao, Collins, Glorie, Han and Li2017), and the amalgamation of this block with the NCC might have taken place during early–middle Palaeozoic time (Dan et al. Reference Dan, Li, Wang, Wang, Wyman and Liu2016; Zhang et al. Reference Zhang, Zhang and Zhao2016c ), although no ophiolitic mélanges have been recognized between them until now. Nevertheless, in any of the proposed models, the Alxa Block has been considered as part of the northern NCC, having been amalgamated at least since the Carboniferous Period.

Three ophiolitic mélanges have been reported in Alxa area (Fig. 1b). Two of them crop out in the NE, including the c. 302 Ma Enger Us and the c. 275 Ma Quagan Qulu ophiolitic mélanges, with their basaltic rocks exhibiting normal mid-ocean-ridge basalt (N-MORB) and boninite-like geochemical features (Zheng et al. Reference Zheng, Wu, Zhang, Xu, Meng and Zhang2014), respectively. The Tepai ophiolitic mélange in the SW is also characterized by boninite-like basaltic rocks, but its formation age is either c. 278 Ma (Zheng et al. Reference Zheng, Li, Xiao and Wang2018) or c. 437–448 Ma (Pan, Reference Pan2019).

The southwestern Alxa Block between the Longshoushan Fault and the Badain Jaran Desert involves the NW–SE-trending Beidashan and Longshoushan–Helishan mountains (Fig. 1c). The widespread Precambrian basement rocks in this area include the Neoarchean Beidashan complex (Gong et al. Reference Gong, Zhang, Yu, Li and Hou2012; Zhang et al. Reference Zhang, Gong, Yu, Li and Hou2013 a) and Palaeoproterozoic Longshoushan Group (Tung et al. Reference Tung, Yang, Liu, Zhang, Tseng and Wan2007; Gong et al. Reference Gong, Zhang and Yu2011). They consist of amphibolite- to greenschist-facies metamorphosed igneous and sedimentary rocks and are overlain unconformably by Neoproterozoic greenschist-facies meta-sedimentary rocks (Zhang & Gong, Reference Zhang and Gong2018). Recently, syenite of age c. 1.87 Ga and granitic gneiss of age c. 1.2 Ga were recognized in the Helishan area (Song et al. Reference Song, Xiao, Collins, Glorie, Han and Li2017; Wang et al. Reference Wang, Chen, Li, Zhang and Xu2019 b).

Lower Palaeozoic sedimentary rocks in the southwestern Alxa area crop out only to the south of the Longshoushan Fault (Fig. 1c). They are known as the Dahuangshan Formation and are composed of unmetamorphosed or greenschist-facies marine clastic and carbonate rocks (Zhang et al. Reference Zhang, Zhang, Zhang, Zhao, Wang and Nie2016 a). In contrast, the upper Carboniferous–middle Permian sedimentary rocks are widely distributed (Fig. 1c). The upper Carboniferous succession consists of interbedded volcanic and clastic rocks in the lower part and shallow-marine bioclastic limestones and sandstones in the upper part, and is conformably overlain by lower–middle Permian strata, which include, from bottom to top, conglomerates, pebbly coarse sandstone, sandstone and siltstone, with volcanic interlayers. Mesozoic terrigenous clastic rocks are extensively distributed in this area (Fig. 1c).

Phanerozoic plutons are voluminous and widely exposed in the southwestern Alxa Block (Fig. 1c), with two magmatic periods of Middle Ordovician–Early Devonian and early Carboniferous–late Permian. Plutons of the earlier period are generally felsic granitoids (Qin, Reference Qin2012; Wei et al. Reference Wei, Hao, Lu, Zhao, Zhao and Shi2013; Tang, Reference Tang2015; Liu et al. Reference Liu, Zhao, Sun, Han, Eizenhöfer, Hou, Zhang, Zhu, Wang, Liu and Xu2016 b; Zhou et al. Reference Zhou, Zhang, Luo, Pan, Zhang and Guo2016; Zhang et al. Reference Zhang, Wang, Wang, Liu, Liu and Wu2018d ; Wang et al. Reference Wang, Chen, Shao, Li, Ding, Zhang, Wang, Zhang, Xu and Qin2020), with only a few dolerite dykes (c. 424 Ma) in eastern Longshoushan (Duan et al. Reference Duan, Li, Qian and Jiao2015). In contrast, plutons of the later period are widely distributed and include peridotite, gabbro, diorite, tonalite, granodiorite, monzogranite and granite (Chen et al. Reference Chen, Zhou, Chen, Liu, Wang, Fang, He and Zhang2013; Jiao et al. Reference Jiao, Jin, Rui, Zhang, Ning and Shao2017; Liu et al. Reference Liu, Zhao, Han, Eizenhöfer, Zhu, Hou, Zhang and Wang2017; Xue et al. Reference Xue, Ling, Liu, Zhang and Sun2017; Gong et al. Reference Gong, Zhang, Wang, Yu and Wang2018 a, b; Huo, Reference Huo2019; Song et al. Reference Song, Xiao, Collins, Glorie and Han2019). In addition, several Triassic plutons crop out in the western Beidashan (Fig. 1c; Gu, Reference Gu2012).

3. Samples and petrography

In this study, three granitic plutons were investigated and sampled in the southwestern Alxa Block; all are massive and salmon-pink to off-white in colour (Fig. 3). A medium- to coarse-grained alkali-feldspar granite in western Beidashan (17WAL-07; Fig. 1c) is composed of quartz (c. 30%), plagioclase (c. 20%), alkali-feldspar (c. 40%), biotite (c. 10%) and minor hornblende (Fig. 3b). The other two plutons are located in Longshoushan to the north of Shandan County (Fig. 1c). One is medium-grained granodiorite (17WAL-35) and composed of quartz (c. 20%), plagioclase (c. 40%), alkali-feldspar (c. 20%) and biotite (c. 20%; Fig. 3d). The other sample is coarse-grained monzogranite (17WAL-39), with similar mineral assemblage of quartz (c. 25%), plagioclase (c. 25%), alkali-feldspar (c. 30%) and biotite (20%; Fig. 3f). Accessory minerals of zircon, apatite and titanite are present in all three plutons.

Fig. 3. Field photographs and mineral assemblages under microscope (cross-polarized light) of the studied late early Carboniferous plutons in the southwestern Alxa Block. (a, b) 17WAL-17, alkali-feldspar granite; (c, d) 17WAL-35, granodiorite; (e, f) 17WAL-39, monzogranite. Afs – alkali-feldspar; Bt – biotite; Hbl – hornblende; Pl – plagioclase; Qtz – quartz.

4. Analytical methods

4.a. Whole-rock major- and trace-element analyses

Fresh granitoid samples were first crushed and then ground to 200 mesh in a tungsten carbide cup and ball mill, and then analysed geochemically at the National Research Center of Geoanalysis, China Geological Survey. Whole-rock major-element oxides were measured using a Malvern Panalytical Axios PW4400 x-ray fluorescence spectrometer (XRF), and the analytical uncertainties are generally between 1% and 5%. The concentrations of trace and rare earth elements were determined by a PerkinElmer NexION 300Q inductively coupled plasma mass spectrometer (ICP-MS), with analytical precision generally better than 5%.

4.b. Zircon U–Pb dating

Zircon grains were firstly separated by conventional heavy liquid and magnetic techniques, and then hand-picked under a binocular microscope. The selected zircon crystals were mounted in epoxy resin and polished to half thickness. Potential analytical spots were determined based on morphological features and internal structures of zircons on optical and cathodoluminescence (CL) images. Zircon U–Pb analyses on mineral separates from the three samples were conducted in Tianjin Institute of Geology and Mineral Resources, China Geological Survey, China. A Thermo Fisher Scientific multi-collector inductively coupled plasma mass spectrometer (MC-ICP-MS; Neptune) was coupled to a New Wave 193 nm ArF excimer laser ablation system. Detailed procedures are reported by Cui et al. (Reference Cui, Zhou, Geng, Li and Li2012). Zircon standard GJ-1 was employed as an external standard (Jackson et al. Reference Jackson, Pearson, Griffin and Belousova2004), and measurements of zircon standard Plešovice, which was used as an unknown, yielded a weighted mean 206Pb/238U age of 335.5 ± 2.6 Ma (n = 12; 2σ). This result is in good agreement with the recommended value within error (337.13 ± 0.37 Ma; Sláma et al. Reference Sláma, Košler, Condon, Crowley, Gerdes, Hanchar, Horstwood, Morris, Nasdala, Tubrett and Whitehouse2008). The corrections of common lead were carried out using the method of Andersen (Reference Andersen2002). Concordia diagrams and ages were obtained using ISOPLOT 4.15 (Ludwig, Reference Ludwig2012). Uncertainties of individual measurements were at the 1σ level, but the weighted mean ages and concordia diagrams were given at the 2σ level (95% confidence level).

4.c. Sr–Nd isotopic analyses

The whole-rock Sr and Nd isotopic compositions were determined using a Finnigan MAT-262 mass spectrometer and a Nu Plasma high-resolution MC-ICP-MS, respectively, at the Institute of Geology, Chinese Academy of Geological Sciences, China. The measured 87Sr/86Sr ratio of the SrCO3 standard SRM 987 was 0.710243 ± 0.000012 (2σ), in good agreement with the recommended value within error (0.710251 ± 0.000018; Coombs et al. Reference Coombs, Clague, Moore and Cousens2004). Two standards of JMC Nd2O3 (reference value = 0.511137 ± 0.000008; Jahn et al. Reference Jahn, Bernard-Griffiths, Charlot, Cornichet and Vidal1980) and GSB 04-3258-2015 (certified value = 0.512438; Tang et al. Reference Tang, Li, Liang, Zhang, Li, Pu, Li, Yang, Chu, Zhang, Hou and Wang2017) were employed during Nd isotopic analyses, with measured 143Nd/144Nd ratios of 0.511123 ± 0.000010 and 0.512441 ± 0.000012 at the 2σ level, respectively. Detailed analytical procedures for both Sr and Nd isotopic compositions are described by Tang et al. (Reference Tang, Li, Pan, Liu and Yan2021). All measured ratios were corrected for mass fractionation by normalizing to 88Sr/86Sr = 8.37521 and 146Nd/144Nd = 0.7219, respectively.

5. Results

Whole-rock major- and trace-element concentrations, LA-ICP-MS zircon U–Pb data and Sr–Nd isotopic compositions are given in online Supplementary Tables S1S3 (available at http://journals.cambridge.org/geo), respectively.

5.a. Whole-rock major and trace elements

All three plutons have high SiO2 (68.49–77.01 wt%) and K2O + Na2O (8.07–8.25 wt%; Fig. 4a) and low MgO (0.17–0.82 wt%) and MnO (0.03–0.06 wt%), show peraluminous features (A/CNK = 1.04–1.13), and belong to the high-K calc-alkaline series (Fig. 4b). Alkali-feldspar granite 17WAL-07 and monzogranite 17WAL-39 display lower CaO (0.59–0.61 wt%), higher K2O (K2O/Na2O = 1.35–1.55), higher total rare earth element (REE) concentrations (257.58–275.96 ppm) and distinct negative Eu anomalies (δEu = 0.17–0.37; Fig. 5a), with enrichments in large-ion-lithophile elements (LILEs; e.g. Cs, Rb, Th and Pb) and depletions in Nb, Ta, Ba and Sr (Fig. 5b). In comparison, granodiorite 17WAL-35 displays relatively higher CaO (1.52–2.51 wt%) and lower total REE concentrations (114.30–206.16 ppm), with significantly enriched light rare earth elements (LREEs; (La/Yb)N = 35.75–56.24) and positive Eu anomalies (δEu = 1.12–1.14; Fig. 5c). Moreover, it is characterized by enriched LILEs (Cs, Rb, Ba, Th, Pb and Sr) and depleted high-field-strength elements (HFSEs; Y, Yb and Lu), with negative Nb–Ta and positive Zr–Hf anomalies, respectively (Fig. 5d).

Fig. 4. (a) Na2O + K2O versus SiO2 and (b) K2O versus SiO2 diagrams for the early Carboniferous plutons in the Alxa Block. Data sources include Wang et al. (Reference Wang, Han, Feng and Liu2015 b), Dan et al. (Reference Dan, Li, Wang, Wang, Wyman and Liu2016), Liu et al. (Reference Liu, Zhang, Xiong, Zhao, Di, Wang and Zhou2016 a) and Xue et al. (Reference Xue, Ling, Liu, Zhang and Sun2017).

Fig. 5. (a, c) Chondrite-normalized REE patterns and (b, d) primitive mantle-normalized trace-element diagrams for the late early Carboniferous plutons in the southwestern Alxa Block. Compositions of C1 chondrite and primitive mantle after Sun & McDonough (Reference Sun and McDonough1989).

5.b. Zircon U–Pb ages

Zircon grains from the studied samples are transparent, euhedral and short columnar or prismatic in shape. They exhibit well preserved concentric magmatic oscillatory zoning, with a few inherited zircon cores appearing occasionally in samples 17WAL-35 and 39 (Fig. 6). For alkali-feldspar granite 17WAL-07, all 24 spots are concordant and cluster together (Fig. 7a). Their Th/U ratios are 0.33–0.51 and they yield a concordia age of 331.6 ± 1.6 Ma (mean square weighted deviation (MSWD) = 4.2; 2σ, decay-constant errors included), which is consistent with the weighted mean 206Pb/238U age (331.7 ± 1.5 Ma; MSWD = 1.01; 2σ). With the exception of four discordant spots (16, 17, 18 and 22), concordant analyses of the other 21 granodiorite 17WAL-35 spots have Th/U ratios of 0.36–0.83 but form two age clusters (Fig. 7b). The older population includes 17 spots with a weighted mean 206Pb/238U age of 344.1 ± 2.2 Ma (MSWD = 1.40; 2σ; Fig. 7b1) and the younger population includes 4 spots with a weighted mean 206Pb/238U age of 326.2 ± 6.6 Ma (MSWD = 1.50; 2σ; Fig. 7b2). Furthermore, monzogranite 17WAL-39 has six discordant spots (1, 8, 11, 16, 17 and 21) and one concordant age cluster (Fig. 7c), which yields a consistent concordia age of 331.8 ± 1.7 Ma (MSWD = 4.8; 2σ, decay-constant errors included) and weighted mean 206Pb/238U age of 331.9 ± 1.7 Ma (MSWD = 0.88; n = 17; 2σ), with Th/U ratios of 0.43–1.03.

Fig. 6. Cathodoluminescence (CL) images of representative zircon grains from the studied late early Carboniferous plutons in the southwestern Alxa Block.

Fig. 7. (a–c) Concordia diagrams showing LA-ICP-MS zircon U–Pb data of the studied late early Carboniferous plutons in the southwestern Alxa Block (all the diagrams and calculations are at the 2σ level).

5.c. Whole-rock Sr–Nd isotopes

The 87Rb/86Sr and 147Sm/144Nd ratios of three granitic samples were calculated using the measured whole-rock Rb, Sr, Sm and Nd concentrations. The alkali-feldspar granite (17WAL-07; t = 332 Ma) has the lowest initial 87Sr/86Sr (0.700128) and highest initial 143Nd/144Nd (0.512219) ratios among the three plutons, with positive ϵNd(t) value (0.16) and Mesoproterozoic Nd model age (T DM = 1207 Ma; Fig. 8). The initial 87Sr/86Sr ratio of the granodiorite (17WAL-35; t = 326 Ma) is low (0.705102), and its initial 143Nd/144Nd ratio and ϵNd(t) value are 0.511358 and −16.80, respectively (Fig. 8). As its ƒSm/Nd (−0.59) significantly deviates from that of the average crust (−0.40; DePaolo et al. Reference Depaolo, Linn and Schubert1991), both T DM (1847 Ma) and T DM2 (2446 Ma) were calculated. For the monzogranite (17WAL-39; t = 332 Ma), its initial 87Sr/86Sr and 143Nd/144Nd ratios are 0.717670 and 0.511706, respectively, with negative ϵNd(t) value (−9.85) and Palaeoproterozoic T DM2 (1889 Ma; Fig. 8).

Fig. 8. Sr–Nd isotopic features of early Carboniferous plutons in the Alxa Block. Symbols and data sources as for Figure 4.

6. Discussion

The well preserved concentric magmatic oscillatory zoning (Fig. 6) and high Th/U ratios (0.33–1.03) of dated zircon grains indicate their magmatic origin (Corfu et al. Reference Corfu, Hanchar, Hoskin and Kinny2003); the concordia and weighted mean 206Pb/238U ages are therefore interpreted as crystallization ages (Fig. 7). Because several spots from the older age cluster of granodiorite (17WAL-35) are located within the inherited zircon cores (e.g. spot 24 in Fig. 6b), the younger age cluster is employed. The three granitic plutons in the southwestern Alxa Block were therefore formed during late early Carboniferous time (c. 332–326 Ma).

6.a. Petrogenesis of the studied late early Carboniferous granitic plutons

The alkali-feldspar granite (17WAL-07) and monzogranite (17WAL-39) have similar geochemical features, such as high K2O + Na2O (8.10–8.25 wt%), FeOT (1.49–1.51 wt%) and FeOT/MgO (4.38–8.87), low CaO (0.59–0.61 wt%), MgO (0.17–0.43 wt%) and P2O5 (< 0.06 wt%), high total REE concentrations (257.58–275.96 ppm) with V-type REE patterns (Fig. 5a), and strongly depleted Ba and Sr (Fig. 5b). These characteristics indicate A-type granite nature, which can be clearly identified on the discrimination diagrams (e.g. Fig. 9b, c; Whalen et al. Reference Whalen, Currie and Chappell1987; King et al. Reference King, White, Chappell and Allen1997). A-type granites may originate from the fractionation of mantle-derived basaltic magmas (Eby, Reference Eby1990, Reference Eby1992; Bonin, Reference Bonin2007), the mixing of mantle- and crust-derived magmas (Yang et al. Reference Yang, Wu, Chung, Wilde and Chu2006), or the partial melting of crust at high temperatures (Whalen et al. Reference Whalen, Currie and Chappell1987; King et al. Reference King, White, Chappell and Allen1997; Wu et al. Reference Wu, Sun, Li, Jahn and Wilde2002). If rhyolitic magmas were derived from fractional crystallization of coeval basaltic magmas, the two components would commonly be spatially and temporally associated (Whitaker et al. Reference Whitaker, Nekvasil, Lindsley and McCurry2008). If the plutons had their origin by magma mixing, then they would have intermediate compositions with the presence of profuse mafic microgranular enclaves (MMEs; Yang et al. Reference Yang, Wu, Chung, Wilde and Chu2006, Reference Yang, Wu, Wilde, Xie, Yang and Liu2007; Zhang et al. Reference Zhang, Wang, Castro, Zhang, Shi, Tong, Zhang, Guo, Yang and Iaccheri2016 b), although the MMEs may be also cogenetic with their host granitoids (Zhang & Zhao, Reference Zhang and Zhao2017). The two A-type granites in the southwestern Alxa Block are rhyolitic in composition (Fig. 4a), but no MMEs were observed (Fig. 3a, e) and their coeval mafic intrusions crop out far away in the northeastern Alxa Block (Wang et al. Reference Wang, Han, Feng and Liu2015 b; Liu et al. Reference Liu, Zhang, Xiong, Zhao, Di, Wang and Zhou2016 a). They are also characterized by high SiO2 (73.89–77.01 wt%) and K2O/Na2O (1.35–1.55) and are peraluminous (A/CNK = 1.04–1.13), similar to aluminous A-type granites with continental crustal sources (King et al. Reference King, White, Chappell and Allen1997). Moreover, the alkali-feldspar granite has low positive ϵNd(t) value (0.16) and Mesoproterozoic Nd model age (1207 Ma; Fig. 8b), which is close to the protolith crystallization age of a granitic gneiss in the Helishan (c. 1200 Ma; Song et al. Reference Song, Xiao, Collins, Glorie, Han and Li2017). Its unusually low initial 87Sr/86Sr value (0.700128; Fig. 8a) may be caused by the strong depletion of Sr (Fig. 5b), as the initial 87Sr/86Sr value was calculated based on the measured whole-rock Sr concentration. The monzogranite has radiogenic Sr–Nd isotopes (Fig. 8a) and a Palaeoproterozoic Nd model age (1889 Ma; Fig. 8b). The Palaeoproterozoic basement rocks are commonly observed in Longshoushan (Tung et al. Reference Tung, Yang, Liu, Zhang, Tseng and Wan2007; Gong et al. Reference Gong, Zhang and Yu2011), in addition to a c. 1872 Ma syenite in Helishan (Wang et al. Reference Wang, Chen, Li, Zhang and Xu2019 b). The two aluminous A-type granites were therefore most probably the high-temperature partial melts of Palaeo- and Mesoproterozoic crustal materials.

Fig. 9. (a) Tectonic discrimination diagrams of Rb versus (Y + Nb) for the early Carboniferous felsic plutons in the Alxa Block (Pearce, Reference Pearce1996). (b) Plot of (K2O + Na2O)/CaO versus Zr + Nb + Ce + Y and (c) plot of Ce versus 10 000×Ga/Al for A-type granites (Whalen et al. Reference Whalen, Currie and Chappell1987). (d) Nb–Y–Ce diagram for distinguishing between A1 and A2 granites (Eby, Reference Eby1992). Symbols and data sources as for Figure 4.

The granodiorite (17WAL-35) is also high-K calc-alkaline (Fig. 4b) and weakly peraluminous (A/CNK = 1.07–1.08) and has depleted HREEs and HFSEs (Fig. 5c, d). It is chemically characterized by high Sr (522.0–918.0 ppm) and low Y (5.36–6.56 ppm) and Yb (0.62–0.75 ppm) concentrations, with high Sr/Y ratios (97.4–139.9). Although high Sr/Y ratio (> 40) usually occurs in adakitic rocks, the high K2O contents (3.59–4.12 wt%) and K2O/Na2O ratios (0.80–1.03) of this granodiorite are more ‘continental’ than typical adakites (Defant & Drummond, Reference Defant and Drummond1990; Martin et al. Reference Martin, Smithies, Rapp, Moyen and Champion2005; Moyen, Reference Moyen2009). The coexistence of negative Nb–Ta and positive Zr–Hf anomalies (Fig. 5d) and highly radiogenic Sr–Nd isotopes (Fig. 8a) also suggest a continental crustal source (Rudnick & Gao, Reference Rudnick, Gao, Holland and Turekian2003). The enrichments of Eu, Ba and Sr are attributed to the large proportion of plagioclase (c. 40%), whereas the low Y concentration may suggest the presence of garnet in the residue, so that the high Sr/Y ratios indicate a deeper crustal level of magma source (Ducea et al. Reference Ducea, Saleeby and Bergantz2015). In addition, c. 2.5 Ga basement rocks and magmatic activity are commonly observed in the southwestern Alxa Block (Zhang et al. Reference Zhang, Gong, Yu, Li and Hou2013 a; Zhang & Gong, Reference Zhang and Gong2018; Wang et al. Reference Wang, Chen, Li, Zhang and Xu2019 b), which is coeval with the two-stage Nd model age of this granodiorite (c. 2446 Ma; Fig. 8b). This granodiorite of high Sr/Y ratio may therefore have its origin in the partial melting of upper Neoarchean lower crust.

6.b. Tectonic setting of the early Carboniferous magmatism in the Alxa Block

Two different tectonic processes accounting for the early Carboniferous magmatism within the Alxa Block were proposed previously: continental arc magmatism induced by the S-wards subduction of the PAO (Liu et al. Reference Liu, Zhang, Xiong, Zhao, Di, Wang and Zhou2016 a; Xue et al. Reference Xue, Ling, Liu, Zhang and Sun2017; Gong et al. Reference Gong, Zhang, Wang, Yu and Wang2018 a), or the collision and amalgamation between the Alxa Block and the NCC (Zhang et al. Reference Zhang, Li, Xiao, Wang and Qi2013 b; Dan et al. Reference Dan, Li, Wang, Wang, Wyman and Liu2016). Noticeably, whether a Palaeozoic suture between the Alxa Block and the NCC existed or not is still in debate, especially with no associated ophiolitic mélanges observed (e.g. Dan et al. Reference Dan, Li, Wang, Wang, Wyman and Liu2016; Zhang & Gong, Reference Zhang and Gong2018; Wang et al. Reference Wang, Chen, Li, Zhang and Xu2019 b), and the early Carboniferous magmatic rocks are widely distributed, rather than along a linear trend in the eastern margin of the Alxa Block (Fig. 1b), so they are less likely attributed to such an amalgamation process. Furthermore, the argument of continental arc magmatism is mainly based on their arc-like geochemical signatures, such as calc–alkaline characteristics (Fig. 4b), negative Nb–Ta anomalies and high Sr/Y ratios (e.g. Liu et al. Reference Liu, Zhang, Xiong, Zhao, Di, Wang and Zhou2016 a; Xue et al. Reference Xue, Ling, Liu, Zhang and Sun2017). However, these signatures can also be inherited from magma sources (Wang et al. Reference Wang, Wilde, Xu and Pang2016 a), and most granites of high Sr/Y ratio in this area exhibit high K2O/Na2O ratios (0.92–3.70), positive Zr–Hf anomalies and radiogenic Nd–Hf isotopes, indicating derivation by the partial melting of lower continental crust (Fig. 8a; Dan et al. Reference Dan, Li, Wang, Wang, Wyman and Liu2016; Xue et al. Reference Xue, Ling, Liu, Zhang and Sun2017); this can occur not only in continental arc belts but also in lithospheric extensional environments.

It is noteworthy that the early Carboniferous plutons within the Alxa Block are mostly basic or acidic in silica content (Fig. 4), resembling bimodal associations. The felsic plutons plot not only in volcanic arc but also in within-plate and post-collision granite fields (Fig. 9a), with most of them exhibiting radiogenic Sr–Nd isotopes (Fig. 8a). They are characterized by the coexistence of A-type granites, peraluminous granites and calc-alkaline I-type granitoids (Dan et al. Reference Dan, Li, Wang, Wang, Wyman and Liu2016; Liu et al. Reference Liu, Zhang, Xiong, Zhao, Di, Wang and Zhou2016 a; Xue et al. Reference Xue, Ling, Liu, Zhang and Sun2017; Zheng et al. Reference Zheng, Li, Zhang, Xiao and Li2019), which mostly occur in extensional settings (Maniar & Piccoli, Reference Maniar and Piccoli1989). A-type granites usually indicate high-temperature anatectic conditions related to asthenospheric upwelling in a lithospheric extensional setting (Whalen et al. Reference Whalen, Currie and Chappell1987; Eby, Reference Eby1992). The mafic plutons plot mostly in the MORB and within-plate basalt fields, similar to the rift-related Basin-and-Range basalts (Fig. 10), and display juvenile or weakly radiogenic Sr–Nd isotopes (Fig. 8a). It is noteworthy that several of the mafic plutons in the northeastern Alxa Block have hornblende as the dominant mafic mineral and resemble appinitic intrusions in geochemistry (Wang et al. Reference Wang, Han, Feng and Liu2015 b). Generally, mafic appinitic melts were most likely produced by the partial melting of subduction-modified sub-continental lithospheric mantle (Fig. 10c) and the melting may be triggered by asthenospheric upwelling following slab break-off or delamination after a subduction event (Murphy, Reference Murphy2013). The generation of both the mafic and felsic early Carboniferous plutons within the Alxa Block therefore most likely resulted from the asthenospheric upwelling at that time. Although an upwelling asthenosphere may also occur in a continental arc setting, continental arc magmatism is typically characterized by linear tracks within a specific tectonic unit and dominated by andesitic rocks, with continued major elemental compositions from basalts to rhyolites but without compositional gaps (Ducea et al. Reference Ducea, Saleeby and Bergantz2015). Evidently, this is not the case for the early Carboniferous plutons within the Alxa Block (Figs 1b, 4a), meaning that their formation in a continental arc is less likely, but rather more likely in a lithospheric extensional setting.

Fig. 10. Petrogenetic discrimination diagrams of (a) V–(Ti/1000) (Shervais, Reference Shervais1982), (b) (Zr/Y)–Zr (Pearce & Norry, Reference Pearce and Norry1979), (c) (La/Ba)–(La/Nb) (Saunders et al. Reference Saunders, Storey, Kent and Norry1992), and (d) (Zr/Sm)–(Sr/Nd)–(Ti/V) (Wang et al. Reference Wang, Wilde, Xu and Pang2016 a) for the early Carboniferous mafic rocks in the Alxa Block. The Basin-and-Range rift-related basalt field refers to Wang et al. (Reference Wang, Wilde, Xu and Pang2016 a). Symbols and data sources as for Figure 4.

Furthermore, A-type granites are a good indicator of lithospheric extension, but the specific extensional setting could be varied (Sain et al. Reference Sain, Saha, Joy, Jelsma and Armstrong2017), including not only rift-related (intraplate) extension (Whalen et al. Reference Whalen, Currie and Chappell1987; Eby, Reference Eby1992) but also back-arc extension (Karsli et al. Reference Karsli, Caran, Dokuz, Çoban, Chen and Kandemir2012; Bickford et al. Reference Bickford, Schmus, Karlstrom, Mueller and Kamenov2015). The two early Carboniferous aluminous A-type granites in the southwestern Alxa Block are A2 type (Fig. 9d) and therefore represent magmas derived from continental crust that has been through an orogenic cycle of arc magmatism and collision (Eby, Reference Eby1992). The geochemical similarities between early Carboniferous mafic plutons in the Alxa Block and Basin-and-Range basalts (Fig. 10), which were generated in back-arc extensional setting to the Sierra Nevada arc (Cousens et al. Reference Cousens, Henry, Stevens, Varve, John and Wetmore2019), also suggest a subduction-related tectonic setting. In back-arc extensional setting, the asthenospheric upwelling could be induced by the foundering of arc root during the roll-back process of subducting slab (DeCelles et al. Reference DeCelles, Ducea, Kapp and Zandt2009; DeCelles & Graham, Reference DeCelles and Graham2015). Another possibility is the intra-continental extensional setting, because the sub-continental lithospheric mantle and lower continental crust of the Alxa Block had been modified by subduction during Middle Ordovician–Early Devonian time (Liu et al. Reference Liu, Zhao, Sun, Han, Eizenhöfer, Hou, Zhang, Zhu, Wang, Liu and Xu2016 b; Zhou et al. Reference Zhou, Zhang, Luo, Pan, Zhang and Guo2016), and the subduction-related geochemical signatures of later magmas may be inherited from the subduction-modified magma sources (Wang et al. Reference Wang, Wilde, Xu and Pang2016 a). Moreover, the extension-related rock associations of calc-alkaline I-type granites, aluminous A2-type granites and peralkaline granites were present in the southwestern Alxa Block from late Silurian–Early Devonian time, following earlier arc magmatism and implying post-collisional setting (Wang et al. Reference Wang, Chen, Shao, Li, Ding, Zhang, Wang, Zhang, Xu and Qin2020). In addition, the cyclical magmatic flare-ups and lulls within each Palaeozoic magmatic stage of the Alxa Block (Fig. 2a) are quite similar to those of Cordilleran arcs in terms of time span and frequency (DeCelles et al. Reference DeCelles, Ducea, Kapp and Zandt2009), but the magmatic hiatus between the two magmatic stages is relatively too long for one single subduction event. The two magmatic stages of the Alxa Block may therefore represent two orogenic cycles and the early Carboniferous extension, as the initiation of the second orogenic cycle, may suggest intra-continental extensional setting. Although more geological evidence is urgently needed to discriminate between the two kinds of extensional settings, a simple continental arc model is less likely for the early Carboniferous magmatism within the Alxa Block.

Additionally, continental arc magmatism is usually accompanied by syn-arc sedimentation in fore-arc or back-arc basins (Ducea et al. Reference Ducea, Saleeby and Bergantz2015), but lower Carboniferous strata are absent from the Alxa Block based on available geological reports. Although a few outcrops in the northern Alxa Block were previously identified as lower Carboniferous deposits, they were recently reassigned as lower–middle Permian strata (Zhang et al. Reference Zhang, Niu, Wei, Shi and Song2018c ). By contrast, the upper Carboniferous–middle Permian strata are widely distributed. The sedimentary facies show a distinct change from terrestrial alluvial fan and delta in the lower stratigraphic sections to platform, littoral and shallow-marine in the upper stratigraphic sections, with abundant fossils (e.g. plants, fusulinids, brachiopods, corals) and volcanic interlayers (Bu et al. Reference Bu, Niu, Wu and Duan2012; Han et al. Reference Han, Liu, Li and Shi2012; Yin et al. Reference Yin, Zhou, Zhang, Zheng and Wang2016; Song et al. Reference Song, Xiao, Collins, Glorie, Han and Li2018). Such a transgression sequence is consistent with the further development of the lithospheric extension.

6.c. Tectonic implications for the development of southeastern CAOB

Even if the Alxa Block was separated from the NCC during the Precambrian Eon, sedimentologic, magmatic and structural evidences (Li et al. Reference Li, Zhang and Qu2012 a; Dan et al. Reference Dan, Li, Wang, Wang, Wyman and Liu2016; Zhang et al. Reference Zhang, Li, Xiao, Wang and Qi2013 b, Reference Zhang, Zhang and Zhao2016 c) all suggest that their amalgamation occurred before early Carboniferous time. Palaeomagnetic studies also suggest that the Precambrian micro-continental blocks within the southeastern CAOB (e.g. Mongolia, Songliao and Hunshandake blocks) may have already accreted to the northern NCC by early Carboniferous time (Pruner, Reference Pruner1992; Li et al. Reference Li, Zhang, Gao, Li, Zhao, Li and Guan2012 b; Zhao et al. Reference Zhao, Chen, Xu, Faure, Shi and Choulet2013; Zhang et al. Reference Zhang, Huang, Zhao and Zhang2018 a). Furthermore, the Palaeozoic magmatic episodes of the Alxa Block and the southeastern CAOB (including the northern margin of the NCC) are very similar (Fig. 2), indicating comparable tectonic processes. Consequently, the whole region had been experiencing a uniform tectonic regime since early Carboniferous time and, if there was on-going S-wards subduction of the large-scale PAO at that time, the arc-trench system was most likely located to the north of these micro-continental blocks.

Regionally, the early Carboniferous is the initial period of the second magmatic stage (Fig. 2), and magmatic rocks during this period are characterized by the mafic–ultramafic complexes in northern Inner Mongolia (Jian et al. Reference Jian, Kröner, Windley, Shi, Zhang, Zhang and Yang2012; Zhang et al. Reference Zhang, Li, Li, Tang, Chen and Luo2015c ; Li et al. Reference Li, Wang, Santosh, Wang, Dong and Li2018), the appinitic intrusions in the northern NCC (Zhou et al. Reference Zhou, Zhang, Liu, Liu and Liu2009; Zhang et al. Reference Zhang, Gao, Wang, Liu and Ma2012 a; Wang et al. Reference Wang, Han, Feng and Liu2015 b), the calc-alkaline I-type and peraluminous granites with crustal origins throughout the southeastern CAOB (Bao et al. Reference Bao, Zhang, Wu, Wang, Li, Sang and Liu2007; Zhang et al. Reference Zhang, Zhao, Song, Yang, Hu and Wu2007, Reference Zhang, Liu, Chai, Xu, Zhao and Xu2011; Liu et al. Reference Liu, Chi, Zhang, Ma, Zhao, Wang, Hu and Zhao2009, Reference Liu, Zhang, Xiong, Zhao, Di, Wang and Zhou2016 a; Blight et al. Reference Blight, Crowley, Petterson and Cunningham2010; Dan et al. Reference Dan, Li, Guo, Liu and Wang2012; Xue et al. Reference Xue, Ling, Liu, Zhang and Sun2017), and the A-type granites newly identified in the southwestern Alxa Block (this study). Such rock associations are commonly associated with asthenospheric upwelling in lithospheric extensional setting. Although some of the basaltic rocks from the mafic–ultramafic complexes exhibit subduction-related geochemical features (Jian et al. Reference Jian, Kröner, Windley, Shi, Zhang, Zhang and Yang2012; Zhang et al. Reference Zhang, Li, Li, Tang, Chen and Luo2015c ; Li et al. Reference Li, Wang, Santosh, Wang, Dong and Li2018), these features can also be imprinted by crustal contamination (Xia, Reference Xia2014) or inherited from magma sources that have been modified by earlier subduction fluids or melts (Wang et al. Reference Wang, Wilde, Xu and Pang2016 a). Further, the coeval intrusions are widely distributed (Xu et al. Reference Xu, Zhao, Bao, Zhou, Wang and Luo2014) rather than along one or two specific ribbons as would be expected for a magmatic arc, supporting their formation in an extensional tectonic setting. Moreover, if this lithospheric extension occurred in back-arc, then the remnants of the large-scale PAO may be represented by the early Carboniferous Erenhot–Hegenshan ophiolitic mélanges to the north of the micro-continental blocks (Zhang et al. Reference Zhang, Li, Li, Tang, Chen and Luo2015c ; Li et al. Reference Li, Wang, Santosh, Wang, Dong and Li2018). Otherwise, the early Carboniferous extension of the southeastern CAOB was probably developed in an intra-continent environment and may represent the initiation of the second orogenic cycle (Xu et al. Reference Xu, Wang, Zhang, Wang, Yang and He2018).

In addition to the intrusions, the early Carboniferous sedimentary rocks are mostly absent from the southeastern CAOB, indicating regional uplift related to asthenospheric upwelling during the initial stage of the lithospheric extension. The Carboniferous metamorphic rocks are high-temperature–low-pressure and show a clockwise PT path, involving pre-peak heating with slight decompression, peak and post-peak cooling stages, also suggesting an extension process (Zhang et al. Reference Zhang, Wei and Chu2018 b).

Subsequently, the late Carboniferous–Permian magmatism in the southeastern CAOB became intense (Fig. 2) with the formation of the widespread bimodal volcanic rocks, continental basaltic intrusions, calc-alkaline I-type granites, peraluminous S-type granites, A-type granites and several peralkaline magmatic belts (e.g. Jahn et al. Reference Jahn, Litvinovsky, Zanvilevich and Reichow2009; Zhang et al. Reference Zhang, Xue, Yuan, Ma and Wilde2012 b, Reference Zhang, Yuan, Xue, Yan and Mao2015 b, Reference Zhang, Zhao, Liu and Hu2016 d, Reference Zhang, Chen, Li, Li, Yang and Qian2017 b; Pang et al. Reference Pang, Wang, Xu, Zhao, Feng, Wang, Luo and Liao2016, Reference Pang, Wang, Xu, Luo and Liu2017; Zhao et al. Reference Zhao, Jahn, Xu, Liao and Wang2016 a; Ji et al. Reference Ji, Ge, Yang, Tian, Chen and Zhang2018; Wang et al. Reference Wang, Han, Feng, Liu, Zheng, Kong and Qi2021 b), implying further development of the early Carboniferous extension. This is also consistent with the occurrence of many late Carboniferous–Permian mafic dykes (Fig. 3a) with MORB or within-plate basalt geochemical signatures in this region (Lin et al. Reference Lin, Xiao, Wan, Windley, Ao, Han, Feng, Zhang and Zhang2014). Accordingly, the late Carboniferous–Permian Solonker, Enger Us and Quagan Qulu ophiolitic mélanges (Jian et al. Reference Jian, Liu, Kröner, Windley, Shi, Zhang, Zhang, Miao, Zhang and Tomurhuu2010; Zheng et al. Reference Zheng, Wu, Zhang, Xu, Meng and Zhang2014), which contain MORB-type intrusions, continental basalts and terrigenous sediments (Luo et al. Reference Luo, Xu, Shi, Zhao, Faure and Chen2016; Shi et al. Reference Shi, Song, Wang, Huang, Zhang and Tang2016), may represent the newly opened limited ocean basins and mark the strongest extension (Xu et al. Reference Xu, Zhao, Bao, Zhou, Wang and Luo2014, Reference Xu, Wang, Zhang, Wang, Yang and He2018). The late Carboniferous–Permian sedimentary sequences are also widely exposed throughout the southeastern CAOB. They vary from plant fossil-bearing terrigenous clastic rocks to shallow-marine clastic and carbonate depositions, with basal conglomerates, and are transgression sequences related to regional extension (Zhao et al. Reference Zhao, Xu, Tong, Chen and Faure2016 b; Ji et al. Reference Ji, Zhang, Yang, Chen and Tang2020; Wang et al. Reference Wang, Xu, Song, Zhao, Zhang and Yan2021 a).

To summarize, we propose a lithospheric extensional process rather than a simple continental arc for the tectono-magmatic development of the southeastern CAOB during early Carboniferous time (Fig. 11). The early Carboniferous extension-related magmatism and the absence of coeval sedimentary successions may reflect the onset of asthenospheric upwelling and regional uplift, and therefore mark the initiation of the lithospheric extension. Nevertheless, the asthenospheric upwelling could be induced by either slab roll-back or slab break-off of the subducted PAO; more geological, geochemical, geophysical and palaeontological evidence is therefore needed to further constrain the specific tectonic setting of this extension, either back-arc or intra-continental.

Fig. 11. Extensional tectonics of the Alxa Block and the southeastern CAOB during early Carboniferous time. (a) Micro-continental blocks within the southeastern CAOB had already been accreted to the northern NCC (Alxa Block) before early Carboniferous time. (b) During early Carboniferous time, the asthenospheric upwelling induced by either the roll-back or the break-off of the subducted PAO slab heated both the subduction-modified lithospheric mantle and the overlying crust, leading to the generation of the mafic and felsic plutons, respectively.

7. Conclusions

The early Carboniferous (c. 332–326 Ma) granodiorite with high Sr/Y ratio, A-type monzogranite and A-type alkali-feldspar granite in the southwestern Alxa Block were most likely formed by partial melting of Neoarchean, Palaeoproterozoic and Mesoproterozoic crustal sources heated by upwelling asthenosphere in an lithospheric extensional setting. According to regional geological correlations, a uniform lithospheric extensional setting, either back-arc or intra-continental, but not a simple continental arc, is suggested for both the Alxa Block and the southeastern CAOB during early Carboniferous time, with the development of extension-related magmatism and the absence of coeval sedimentary rocks.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756821000984

Acknowledgements

We are grateful to Yurong Cui (Tianjin Institute of Geology and Mineral Resources) and Suohan Tang (Institute of Geology, Chinese Academy of Geological Sciences) for their help with zircon U–Pb dating and isotopic analyses, respectively. Special thanks are extended to Professor Peter Clift and anonymous reviewers for their constructive comments. This work was financially supported by China Geological Survey (grant number DD20190011), Chinese Academy of Geological Sciences (grant number JKY202011) and National Key Research and Development Program of China (grant number 2018YFC0603701).

Declaration of interest

The authors declare that they have no known conflicts of interests or personal relationships that could have appeared to influence the work reported in this paper.

References

Andersen, T (2002) Correction of common lead in U–Pb analyses that do not report 204Pb. Chemical Geology 192, 5979.CrossRefGoogle Scholar
Bao, Q, Zhang, C, Wu, Z, Wang, H, Li, W, Sang, J and Liu, Y (2007) SHRIMP U-Pb zircon geochronology of a Carboniferous quartzdiorite in Baiyingaole area, Inner Mongolia and its implications. Journal of Jilin University (Earth Science Edition) 37, 1523 (in Chinese with English abstract).Google Scholar
Bickford, ME, Schmus, WR, Karlstrom, KE, Mueller, PA and Kamenov, GD (2015) Mesoproterozoic-trans-Laurentian magmatism: a synthesis of continent-wide age distributions, new SIMS U–Pb ages, zircon saturation temperatures, and Hf and Nd isotopic compositions. Precambrian Research 265, 286312.CrossRefGoogle Scholar
Blight, JHS, Crowley, QG, Petterson, MG and Cunningham, D (2010) Granites of the Southern Mongolia Carboniferous Arc: new geochronological and geochemical constraints. Lithos 116, 3552.CrossRefGoogle Scholar
Bonin, B (2007) A-type granites and related rocks: evolution of a concept, problems and prospects. Lithos 97, 129.CrossRefGoogle Scholar
Bu, J, Niu, Z, Wu, J and Duan, X (2012) Sedimentary characteristics and age of Amushan formation in Ejin Banner and its adjacent areas, western Inner Mongolia. Geological Bulletin of China 31, 1669–83 (in Chinese with English abstract).Google Scholar
Chen, W, Zhou, W, Chen, K, Liu, M, Wang, T, Fang, M, He, H and Zhang, B (2013) Subduction-related Early Permian granodiorite in Jinchangshan of Alashan, Inner Mongolia: Evidences from zircon U-Pb geochronology and geochemistry. Journal of Mineralogy and Petrology 33, 5360 (in Chinese with English abstract).Google Scholar
Chen, Y, Zhao, R, Wang, G, Rong, X and Li, T (2020) Geochronology, geochemical characteristics and significances of quartz monzonite in Niujiaogou, Longshoushan, Gansu. Journal of East China University of Technology (Natural Science) 43, 2129 (in Chinese with English abstract).Google Scholar
Coombs, ML, Clague, DA, Moore, GF and Cousens, BL (2004) Growth and collapse of Waianae Volcano, Hawaii, as revealed by exploration of its submarine flanks. Geochemistry, Geophysics, Geosystems 5, Q05006, doi: 10.1029/2004GC000717.CrossRefGoogle Scholar
Corfu, F, Hanchar, JM, Hoskin, PWO and Kinny, P (2003) Atlas of zircon textures. Reviews in Mineralogy and Geochemistry 53, 469500.CrossRefGoogle Scholar
Cousens, BL, Henry, CD, Stevens, C, Varve, S, John, DV and Wetmore, S (2019) Igneous rocks in the Fish Creek Mountains and environs, Battle Mountain area, north-central Nevada: a microcosm of Cenozoic igneous activity in the northern Great Basin, Basin and Range Province, USA. Earth-Science Reviews 192, 403–44.CrossRefGoogle Scholar
Cui, Y, Zhou, H, Geng, J, Li, H and Li, H (2012) In-site LA-MC-ICP-MS U–Pb isotopic dating of monazite. Acta Geologica Sinica 33, 865–76 (in Chinese with English abstract).Google Scholar
Dan, W, Li, XH, Guo, J, Liu, Y and Wang, XC (2012) Paleoproterozoic evolution of the eastern Alxa Block, westernmost North China: evidence from in situ zircon U–Pb dating and Hf–O isotopes. Gondwana Research 21, 838–64.CrossRefGoogle Scholar
Dan, W, Li, XH, Wang, Q, Tang, GJ and Liu, Y (2014) An Early Permian (ca. 280 Ma) silicic igneous province in the Alxa Block, NW China: a magmatic flare-up triggered by a mantle-plume? Lithos 204, 144–58.CrossRefGoogle Scholar
Dan, W, Li, XH, Wang, Q, Wang, XC, Wyman, DA and Liu, Y (2016) Phanerozoic amalgamation of the Alxa Block and North China Craton: evidence from Paleozoic granitoids, U–Pb geochronology and Sr–Nd–Pb–Hf–O isotope geochemistry. Gondwana Research 32, 105–21.CrossRefGoogle Scholar
DeCelles, PG, Ducea, MN, Kapp, P and Zandt, G (2009) Cyclicity in Cordilleran Orogenic systems. Nature Geoscience 2, 251–7.CrossRefGoogle Scholar
DeCelles, PG and Graham, SA (2015) Cyclical processes in the North American Cordilleran Orogenic system. Geology 43, 499502.CrossRefGoogle Scholar
Defant, MJ and Drummond, MS (1990) Derivation of some modern arc magmas by melting of young subducted lithosphere. Nature 347, 662–5.CrossRefGoogle Scholar
Depaolo, DJ, Linn, AM and Schubert, G (1991) The continental crustal age distribution: methods of determining mantle separation ages from Sm-Nd isotopic data and application to the Southwestern United States. Journal of Geophysical Research 96, 2071–88.CrossRefGoogle Scholar
Duan, J, Li, C, Qian, Z and Jiao, J (2015) Geochronological and geochemical constraints on the petrogenesis and tectonic significance of Paleozoic dolerite dykes in the southern margin of Alxa Block, North China Craton. Journal of Asian Earth Sciences 111, 244–53.CrossRefGoogle Scholar
Ducea, MN, Saleeby, JB and Bergantz, G (2015) The architecture, chemistry, and evolution of continental magmatic arcs. Annual Review of Earth and Planetary Sciences 43, 10.110.33.CrossRefGoogle Scholar
Eby, GN (1990) The A-type granitoids: a review of their occurrence and chemical characteristics and speculations on their petrogenesis. Lithos 26, 115–34.10.1016/0024-4937(90)90043-ZCrossRefGoogle Scholar
Eby, GN (1992) Chemical subdivision of A-type granitoids: petrogenesis and tectonic implications. Geology 20, 641–4.10.1130/0091-7613(1992)020<0641:CSOTAT>2.3.CO;22.3.CO;2>CrossRefGoogle Scholar
Eizenhöfer, PR and Zhao, G (2018) Solonker Suture in east Asia and its bearing of the final closure of the eastern segment of the Palaeo-Asian Ocean. Earth-Science Reveiws 186, 153–72.10.1016/j.earscirev.2017.09.010CrossRefGoogle Scholar
Geng, Y, Wang, X, Wu, C and Zhou, X (2010) Late-Paleoproterozoic tectonothermal events of the metamorphic basement in Alxa area: evidence from geochronology. Acta Petrologica Sinica 26, 1159–70 (in Chinese with English abstract).Google Scholar
Gong, J, Zhang, J, Wang, Z, Yu, S and Wang, D (2018a) Late Ordovician–Carboniferous tectonic evolutionary history of the Alxa Block: constrained by the multistage magmatic-metamorphic-deformation events in Beidashan area. Acta Petrologica et Mineralogica 37, 771–98 (in Chinese with English abstract).Google Scholar
Gong, J, Zhang, J, Wang, Z, Yu, S, Wang, D and Zhang, H (2018b) Zircon U-Pb dating, Hf isotopic and geochemical characteristics of two suites of gabbros in the Beidashan region, western Alxa Block: its implications for evolution of the Central Asian Orogenic Belt. Acta Geologica Sinica 92, 1369–88 (in Chinese with English abstract).Google Scholar
Gong, J, Zhang, J and Yu, S (2011) The origin of Longshoushan Group and associated rocks in the southern part of the Alxa Block: constraint from LA-ICP-MS U-Pb zircon dating. Acta Petrologica et Mineralogica 30, 795818 (in Chinese with English abstract).Google Scholar
Gong, J, Zhang, J, Yu, S, Li, H and Hou, K (2012) Ca. 2.5 Ga TTG rocks in the western Alxa Block and their implications. Chinese Science Bulletin 57, 4064–76 (in Chinese with English abstract).10.1007/s11434-012-5315-8CrossRefGoogle Scholar
Gu, G (2012) The preliminary studies on the origin and tectonic setting of Early Mesozoic granites in the southwest margin of Alxa. M.Sc. thesis, Lanshou University, China. Published thesis (in Chinese with English abstract).Google Scholar
Han, BF, He, GQ, Wang, XC and Guo, ZJ (2011) Late Carboniferous collision between the Tarim and Kazakhstan–Yili terranes in the western segment of the South Tian Shan Orogen, Central Asia, and implications for the Northern Xinjiang, western China. Earth-Science Reviews 109, 7493.CrossRefGoogle Scholar
Han, BF, Wang, SG, Jahn, BM, Hong, DW, Kagami, H and Sun, YL (1997) Depleted-mantle source for the Ulungur River A-type granites from North Xinjiang, China: geochemistry and Nd–Sr isotopic evidence, and implications for Phanerozoic crustal growth. Chemical Geology 138, 135–59.10.1016/S0009-2541(97)00003-XCrossRefGoogle Scholar
Han, W, Liu, X, Li, J and Shi, J (2012) Sedimentary environment of Carboniferous–Permian Amushan Formation in Wulanaobao area of Urad Rear Banner, Inner Mongolia. Geological Bulletin of China 31, 1684–91 (in Chinese with English abstract).Google Scholar
Huo, Y (2019) Geochemical characteristics and geological significance of Late Paleozoic intrusive rocks in Ubud, Beidashan, Alxa. M.Sc. thesis, China University of Geosciences (Beijing), China (in Chinese with English abstract). Published thesis.Google Scholar
Jackson, SE, Pearson, NJ, Griffin, WL and Belousova, EA (2004) The application of laser ablation-inductively coupled plasma-mass spectrometry to in situ U–Pb zircon geochronology. Chemical Geology 211, 4767.CrossRefGoogle Scholar
Jahn, BM, Bernard-Griffiths, J, Charlot, R, Cornichet, J and Vidal, F (1980) Nd and Sr isotopic compositions and REE abundances of Cretaceous MORB (Holes 417D and 418A, LEGS 51, 52 and 53). Earth and Planetary Science Letters 48, 171184.10.1016/0012-821X(80)90180-6CrossRefGoogle Scholar
Jahn, BM, Litvinovsky, BA, Zanvilevich, AN and Reichow, M (2009) Peralkaline granitoid magmatism in the Mongolian–Transbaikalian Belt: evolution, petrogenesis and tectonic significance. Lithos 113, 521–39.10.1016/j.lithos.2009.06.015CrossRefGoogle Scholar
Jahn, BM, Wu, F and Hong, D (2000) Important crustal growth in the Phanerozoic: isotopic evidence of granitoids from east-central Asia. Journal of Earth System Science 109, 520.CrossRefGoogle Scholar
Ji, Z, Ge, WC, Yang, H, Tian, DX, Chen, HJ and Zhang, YL (2018) Late Carboniferous–Early Permian high- and low-Sr/Y granitoids of the Xing’an Block, northeastern China: implications for the late Paleozoic tectonic evolution of the eastern Central Asian Orogenic Belt. Lithos 322, 179–96.CrossRefGoogle Scholar
Ji, Z, Zhang, Z, Yang, J, Chen, Y and Tang, J (2020) Carboniferous–Early Permian sedimentary rocks from the north-eastern Erenhot, North China: implications on the tectono-sedimentary evolution of the south-eastern Central Asian Orogenic Belt. Geological Journal 55(3), 2383–401, doi: 10.1002/gj.3763.CrossRefGoogle Scholar
Jian, P, Kröner, A, Windley, BF, Shi, Y, Zhang, W, Zhang, L and Yang, W (2012) Carboniferous and Cretaceous mafic–ultramafic massifs in Inner Mongolia (China): A SHRIMP zircon and geochemical study of the previously presumed integral ‘Hegenshan ophiolite’. Lithos 142–143, 4866.10.1016/j.lithos.2012.03.007CrossRefGoogle Scholar
Jian, P, Liu, D, Kröner, A, Windley, BF, Shi, Y, Zhang, W, Zhang, F, Miao, L, Zhang, L and Tomurhuu, D (2010) Evolution of a Permian intraoceanic arc–trench system in the Solonker suture zone, Central Asian Orogenic Belt, China and Mongolia. Lithos 118, 169–90.CrossRefGoogle Scholar
Jiao, J, Jin, S, Rui, H, Zhang, G, Ning, Q and Shao, L (2017) Petrology, geochemistry and chronology study of the Xiaokouzi mafic-ultramafic intrusion in the eastern section of Longshou Mountains, Gansu. Acta Geologica Sinica 91, 736–47 (in Chinese with English abstract).Google Scholar
Karsli, O, Caran, Ş, Dokuz, A, Çoban, H, Chen, B and Kandemir, R (2012) A-type granitoids from the Eastern Pontides, NE Turkey: records for generation of hybrid A-type rocks in a subduction-related environment. Tectonophysics 530–531, 208–24.CrossRefGoogle Scholar
King, PL, White, AJR, Chappell, BW and Allen, CM (1997) Characterization and origin of aluminous A-type granites from the Lachlan Fold Belt, southeastern Australia. Journal of Petrology 38, 371–91.CrossRefGoogle Scholar
Kozlovsky, AM, Yarmolyuk, VV, Salnikova, EB, Travin, AV, Kotov, AB, Plotkina, JV, Kudryashova, EA and Savatenkov, VM (2015) Late Paleozoic anorogenic magmatism of the Gobi Altai (SW Mongolia): tectonic position, geochronology and correlation with igneous activity of the Central Asian Orogenic Belt. Journal of Asian Earth Sciences 113, 524–41.CrossRefGoogle Scholar
Li, J, Zhang, J and Qu, J (2012a) Amalgamation of the North China Craton with Alxa Block in the late of Early Paleozoic: evidence from sedimentary sequences in the Niushou Mountain, Ningxia Hui Autonomous Region, NW China. Geological Review 58, 208–14 (in Chinese with English abstract).Google Scholar
Li, P, Zhang, S, Gao, R, Li, H, Zhao, Q, Li, Q and Guan, Y (2012b) New Upper Carboniferous–Lower Permian paleomagnetic results from the central Inner Mongolia and their geological implications. Journal of Jilin University (Earth Science Edition) 42, 423–40 (in Chinese with English abstract).Google Scholar
Li, Y, Wang, G, Santosh, M, Wang, J, Dong, P and Li, H (2018) Supra-subduction zone ophiolites from Inner Mongolia, North China: Implications for the tectonic history of the southeastern Central Asian Orogenic Belt. Gondwana Research 59, 126–43.CrossRefGoogle Scholar
Lin, L, Xiao, WJ, Wan, B, Windley, BF, Ao, S, Han, C, Feng, J, Zhang, J and Zhang, Z (2014) Geochronologic and geochemical evidences for persistence of south-dipping subduction to Late Permian time, Langshan area, Inner Mongolia (China): significance for termination of accretionary orogenesis in the southern Altaids. American Journal of Science 314, 679703.CrossRefGoogle Scholar
Liu, J, Chi, X, Zhang, X, Ma, Z, Zhao, Z, Wang, T, Hu, Z and Zhao, X (2009) Geochemical characteristic of Carboniferous quartz-diorite in the southern Xiwuqi area, Inner Mongolia and its tectonic significance. Acta Geologica Sinica 83, 365–76 (in Chinese with English abstract).Google Scholar
Liu, M, Zhang, D, Xiong, G, Zhao, H, Di, Y, Wang, Z and Zhou, Z (2016a) Zircon U–Pb age, Hf isotope and geochemistry of Carboniferous intrusions from the Langshan area, Inner Mongolia: petrogenesis and tectonic implications. Journal of Asian Earth Sciences 120, 139–58.CrossRefGoogle Scholar
Liu, Q, Zhao, G, Han, Y, Eizenhöfer, PR, Zhu, Y, Hou, W, Zhang, X and Wang, B (2017) Geochronology and geochemistry of Permian to Early Triassic granitoids in the Alxa Terrane: constraints on the final closure of the Paleo-Asian Ocean. Lithosphere 9, L646.641.Google Scholar
Liu, Q, Zhao, G, Sun, M, Han, Y, Eizenhöfer, PR, Hou, W, Zhang, X, Zhu, Y, Wang, B, Liu, D and Xu, B (2016b) Early Paleozoic subduction processes of the Paleo-Asian Ocean: insights from geochronology and geochemistry of Paleozoic plutons in the Alxa Terrane. Lithos 262, 546–60.CrossRefGoogle Scholar
Liu, W, Pan, J, Liu, X, Wang, K, Wang, G and Xue, P (2019) Petrogenesis and tectonic implication of Qingshanbao pluton in Longshou Moutains, Gansu: constraints from elemental geochemistry, zircon U–Pb age and Sr-Nd isotopes. Journal of Mineralogy and Petrology 39, 2640 (in Chinese with English abstract).Google Scholar
Ludwig, KR (2012) User’s Manual for Isoplot 3.75: A Geochronological Toolkit for Microsoft Excel. Berkeley: Berkeley Geochronology Center, Special Publication no. 5, 75 pp.Google Scholar
Luo, ZW, Xu, B, Shi, GZ, Zhao, P, Faure, M and Chen, Y (2016) Solonker ophiolite in Inner Mongolia, China: a late Permian continental margin-type ophiolite. Lithos 261, 7291.CrossRefGoogle Scholar
Maniar, PD and Piccoli, PM (1989) Tectonic discrimination of granitoids. Geological Society of America Bulletin 101, 635–43.2.3.CO;2>CrossRefGoogle Scholar
Martin, H, Smithies, RH, Rapp, R, Moyen, JF and Champion, D (2005) An overview of adakite, tonalite–trondhjemite–granodiorite (TTG), and sanukitoid: relationships and some implications for crustal evolution. Lithos 79, 124.CrossRefGoogle Scholar
Moyen, JF (2009) High Sr/Y and La/Yb ratios: the meaning of the ‘adakitic signature’. Lithos 112, 556–74.CrossRefGoogle Scholar
Murphy, JB (2013) Appinite suites: a record of the role of water in the genesis, transport, emplacement and crystallization of magma. Earth-Science Reviews 119, 3559.CrossRefGoogle Scholar
Pan, X (2019) Petrogenesis of Tebai basic-ultrabasic pluton in the Beidashan area in the southern margin of the Alxa block and its tectonic significance. M.Sc. thesis, Chang’an University, China. Published thesis (in Chinese with English abstract).Google Scholar
Pang, CJ, Wang, XC, Xu, B, Luo, ZW and Liu, YZ (2017) Hydrous parental magmas of Early to Middle Permian gabbroic intrusions in western Inner Mongolia, North China: new constraints on deep-Earth fluid cycling in the Central Asian Orogenic Belt. Journal of Asian Earth Sciences 144, 184204.CrossRefGoogle Scholar
Pang, CJ, Wang, XC, Xu, B, Zhao, JX, Feng, YX, Wang, YY, Luo, ZW and Liao, W (2016) Late Carboniferous N-MORB-type basalts in central Inner Mongolia, China: products of hydrous melting in an intraplate setting? Lithos 261, 5571.CrossRefGoogle Scholar
Pearce, JA (1996) Source and settings of granitic rocks. Episodes 19, 120–5.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
Pruner, P (1992) Palaeomagnetism and palaeogeography of Mongolia from the Carboniferous to the Cretaceous—final report. Physics of the Earth and Planetary Interiors 70, 169–77.CrossRefGoogle Scholar
Qin, H (2012) Petrology of Early Paleozoic granites and their relation to tectonic evolution of orogen in the North Qilian Orogenic Belt. Ph.D. thesis, Chinese Academy of Geological Sciences, China. Published thesis (in Chinese with English abstract).Google Scholar
Rudnick, RL and Gao, S (2003) Composition of the continental crust. In Treatise on Geochemistry, volume 3 (eds Holland, HD and Turekian, KK), pp. 164. Amsterdam: Elsevier Science Ltd.Google Scholar
Sain, A, Saha, D, Joy, S, Jelsma, H and Armstrong, R (2017) New SHRIMP age and microstructures from a deformed A-type granite, Kanigiri, Southern India: constraining the Hiatus between orogenic closure and postorogenic rifting. The Journal of Geology 125, 241–59.CrossRefGoogle Scholar
Saunders, AD, Storey, M, Kent, RW and Norry, MJ (1992) Consequences of plume-lithosphere interactions. In Magmatism and the Causes of Continental Break-up (eds BC Storey, T Alabaster and RJ Pankhurst), pp. 4160. Geological Society of London, Special Publication no. 68.CrossRefGoogle Scholar
Shervais, JW (1982) Ti-V plots and the petrogenesis of modern and ophiolitic lavas. Earth and Planetary Science Letters 59, 101–18.CrossRefGoogle Scholar
Shi, G, Song, G, Wang, H, Huang, C, Zhang, L and Tang, J (2016) Late Paleozoic tectonics of the Solonker Zone in the Wuliji area, Inner Mongolia, China: insights from stratigraphic sequence, chronology, and sandstone geochemistry. Journal of Asian Earth Sciences 127, 100–18.CrossRefGoogle Scholar
Sláma, J, Košler, J, Condon, DJ, Crowley, JL, Gerdes, A, Hanchar, JM, Horstwood, MSA, Morris, GA, Nasdala, L, Tubrett, MN and Whitehouse, MJ (2008) Plešovice zircon–A new natural reference material for U–Pb and Hf isotopic microanalysis. Chemical Geology 249, 135.CrossRefGoogle Scholar
Song, D, Xiao, W, Collins, A, Glorie, S and Han, C (2019) Late Carboniferous–Early Permian arc magmatism in the southwestern Alxa Tectonic Belt (NW China): constraints on the Late Palaeozoic subduction history of the Paleo-Asian Ocean. Geological Journal 54, 1046–63.CrossRefGoogle Scholar
Song, D, Xiao, W, Collins, AS, Glorie, S, Han, C and Li, Y (2017) New chronological constrains on the tectonic affinity of the Alxa Block, NW China. Precambrian Research 299, 230–43.CrossRefGoogle Scholar
Song, D, Xiao, W, Collins, AS, Glorie, S, Han, C and Li, Y (2018) Final subduction processes of the Paleo-Asian Ocean in the Alxa Tectonic Belt (NW China): constraints from field and chronological data of Permian arc-related volcano-sedimentary rocks. Tectonics 37, 1658–87.CrossRefGoogle Scholar
Song, S, Niu, Y, Su, L and Xia, X (2013) Tectonics of the North Qilian orogen, NW China. Gondwana Research 23, 1378–401.CrossRefGoogle Scholar
Sun, SS and McDonough, W (1989) Chemical and isotopic systematics of oceanic basalts: implications for mantle composition and processes. In Magmatism in the Ocean Basins (eds AD Saunders and MJ Norry), pp. 313–45. Geological Society of London, Special Publication no. 42.CrossRefGoogle Scholar
Tang, L (2015) Granites’ characteristics and zircon LA-ICP-MS U-Pb dating of Jiling area in Longshoushan, Gansu province. M.Sc. thesis, East China Institute of Technology, China. Published thesis (in Chinese with English abstract).Google Scholar
Tang, S, Li, J, Liang, X, Zhang, L, Li, G, Pu, W, Li, C, Yang, Y, Chu, Z, Zhang, J, Hou, K and Wang, X (2017) Reference material preparation of 143Nd/144Nd isotope ratio. Rock and Mineral Analysis 36, 163–70 (in Chinese with English abstract)Google Scholar
Tang, S, Li, J, Pan, CX, Liu, H and Yan, B (2021) Production and certification of the reference material GBW04139, GBW04140 and GBW04141 as Rb-Sr and Sm-Nd isotope analysis references. Rock and Mineral Analysis 40, 284–94 (in Chinese with English abstract).Google Scholar
Tong, Y, Jahn, BM, Wang, T, Hong, DW, Smith, EI, Sun, M, Gao, JF, Yang, QD and Huang, W (2015) Permian alkaline granites in the Erenhot–Hegenshan belt, northern Inner Mongolia, China: model of generation, time of emplacement and regional tectonic significance. Journal of Asian Earth Sciences 97, Part B, 320–36.CrossRefGoogle Scholar
Tung, K, Yang, H, Liu, D, Zhang, J, Tseng, C and Wan, Y (2007) SHRIMP U-Pb geochronology of the detrital zircons from the Longshoushan Group and its tectonic significance. Chinese Science Bulletin 52, 1414–25.CrossRefGoogle Scholar
Wan, Y, Song, B, Liu, D, Wilde, SA, Wu, J, Shi, Y, Yin, X and Zhou, H (2006) SHRIMP U–Pb zircon geochronology of Palaeoproterozoic metasedimentary rocks in the North China Craton: evidence for a major Late Palaeoproterozoic tectonothermal event. Precambrian Research 149, 249–71.CrossRefGoogle Scholar
Wang, J, Li, X, Ning, W, Kusky, T and Deng, H (2019a) Geology of a Neoarchean suture: evidence from the Zunhua ophiolitic mélange of the Eastern Hebei Province, North China Craton. Geological Society of America Bulletin 131, 11–2.CrossRefGoogle Scholar
Wang, XC, Wilde, SA, Li, QL and Yang, YN (2015a) Continental flood basalts derived from the hydrous mantle transition zone. Nature Communications 6, 7700, doi: 10.1038/ncomms8700.CrossRefGoogle ScholarPubMed
Wang, XC, Wilde, SA, Xu, B and Pang, CJ (2016a) Origin of arc-like continental basalts: implications for deep-Earth fluid cycling and tectonic discrimination. Lithos 261, 545.CrossRefGoogle Scholar
Wang, Y, Xu, B, Song, S, Zhao, P, Zhang, J and Yan, L (2021a) A late Paleozoic extension basin constrained by sedimentology and geochronology in eastern Central Asia Orogenic Belt. Gondwana Research 89, 265–86.CrossRefGoogle Scholar
Wang, ZZ, Chen, X, Li, B, Zhang, Y and Xu, S (2019b) The discovery of the Paleoproterozoic syenite in Helishan, Gansu Province, and its implications for the tectonic attribution of the Alxa Block. Geology in China 46, 1094–104 (in Chinese with English abstract).Google Scholar
Wang, ZZ, Chen, X, Shao, Z, Li, B, Ding, W, Zhang, Y, Wang, Y, Zhang, Y, Xu, S and Qin, X (2020) Petrogenesis of the Late Silurian–Early Devonian granites in the Longshoushan–Helishan area, Gansu Province, and its tectonic implications for the Early Paleozoic evolution of the southwestern Alxa Block. Acta Geologica Sinica 94, 2243–61 (in Chinese with English abstract).Google Scholar
Wang, ZZ, Han, BF, Feng, LX and Liu, B (2015b) Geochronology, geochemistry and origins of the Paleozoic–Triassic plutons in the Langshan area, western Inner Mongolia, China. Journal of Asian Earth Sciences 97, Part B, 337–51.CrossRefGoogle Scholar
Wang, ZZ, Han, BF, Feng, LX, Liu, B, Zheng, B and Kong, LJ (2016b) Tectonic attribution of the Langshan area in western Inner Mongolia and implications for the Neoarchean–Paleoproterozoic evolution of the western North China Craton: evidence from LA-ICP-MS zircon U–Pb dating of the Langshan basement. Lithos 261, 278–95.CrossRefGoogle Scholar
Wang, ZZ, Han, BF, Feng, LX, Liu, B, Zheng, B, Kong, LJ and Qi, CY (2021b) Early–Middle Permian plutons in the Langshan area, western Inner Mongolia, China, and their tectonic implications. Lithos 382–383, 105934.CrossRefGoogle Scholar
Wei, Q, Hao, L, Lu, J, Zhao, Y, Zhao, X and Shi, H (2013) LA-MC-ICP-MS zircon U-Pb dating of Hexipu granite and its geological implications. Bulletin of Mineralogy, Petrology and Geochemistry 32, 729–35 (in Chinese with English abstract).Google Scholar
Whalen, J, Currie, K and Chappell, B (1987) A-type granites: geochemical characteristics, discrimination and petrogenesis. Contributions to Mineralogy and Petrology 95, 407–19.CrossRefGoogle Scholar
Whitaker, ML, Nekvasil, H, Lindsley, DH and McCurry, M (2008) Can crystallization of olivine tholeiite give rise to potassic rhyolites?—an experimental investigation. Bulletin of Volcanology 70, 417–34.CrossRefGoogle Scholar
Wilde, SA & Zhou, JB (2015) The late Paleozoic to Mesozoic evolution of the eastern margin of the Central Asian Orogenic Belt in China. Journal of Asian Earth Sciences 113, 909–21.CrossRefGoogle Scholar
Windley, BF, Alexeiev, D, Xiao, W, Kröner, A and Badarch, G (2007) Tectonic models for accretion of the Central Asian Orogenic Belt. Journal of the Geological Society 164, 3147.CrossRefGoogle Scholar
Wu, FY, Jahn, BM, Wilde, SA, Lo, CH, Yui, TF, Lin, Q, Ge, WC and Sun, DY (2003) Highly fractionated I-type granites in NE China (II): isotopic geochemistry and implications for crustal growth in the Phanerozoic. Lithos 67, 191204.CrossRefGoogle Scholar
Wu, FY, Sun, DY, Li, H, Jahn, BM and Wilde, S (2002) A-type granites in northeastern China: age and geochemical constraints on their petrogenesis. Chemical Geology 187, 143–73.CrossRefGoogle Scholar
Xia, LQ (2014) The geochemical criteria to distinguish continental basalts from arc related ones. Earth-Science Reviews 139, 195212.CrossRefGoogle Scholar
Xiao, W, Windley, BF, Han, C, Liu, W, Wan, B, Zhang, Je, Ao, S, Zhang, Z and Song, D (2018) Late Paleozoic to early Triassic multiple roll-back and oroclinal bending of the Mongolia collage in Central Asia. Earth-Science Reviews 186, 94128.CrossRefGoogle Scholar
Xiao, W, Windley, BF, Jie, H and Zhai, M (2003) Accretion leading to collision and the Permian Solonker suture, Inner Mongolia, China: termination of the central Asian orogenic belt. Tectonics 22, 120.CrossRefGoogle Scholar
Xiao, W, Windley, BF, Yong, Y, Yan, Z, Yuan, C, Liu, C and Li, J (2009a) Early Paleozoic to Devonian multiple-accretionary model for the Qilian Shan, NW China. Journal of Asian Earth Sciences 35, 323–33.CrossRefGoogle Scholar
Xiao, WJ, Windley, BF, Huang, BC, Han, CM, Yuan, C, Chen, HL, Sun, M, Sun, S and Li, JL (2009b) End-Permian to mid-Triassic termination of the accretionary processes of the southern Altaids: implications for the geodynamic evolution, Phanerozoic continental growth, and metallogeny of Central Asia. International Journal of Earth Sciences 98, 1189–217.CrossRefGoogle Scholar
Xu, B, Charvet, J, Chen, Y, Zhao, P and Shi, G (2013) Middle Paleozoic convergent orogenic belts in western Inner Mongolia (China): framework, kinematics, geochronology and implications for tectonic evolution of the Central Asian Orogenic Belt. Gondwana Research 23, 1342–64.CrossRefGoogle Scholar
Xu, B, Wang, Z, Zhang, L, Wang, Z, Yang, Z and He, Y (2018) The Xing-Meng intracontinent orogenic belt. Acta Petrologica Sinica 34, 2819–44 (in Chinese with English abstract).Google Scholar
Xu, B, Zhao, P, Bao, Q, Zhou, Y, Wang, Y and Luo, Z (2014) Preliminary study on the pre-Mesozoic tectonic unit division of the Xing-Meng Orogenic Belt (XMOB). Acta Petrologica Sinica 30, 1841–57 (in Chinese with English abstract).Google Scholar
Xue, S, Ling, MX, Liu, YL, Zhang, H and Sun, W (2017) The genesis of early Carboniferous adakitic rocks at the southern margin of the Alxa Block, North China. Lithos 278–281, 181–94.CrossRefGoogle Scholar
Yang, JH, Wu, FY, Chung, SL, Wilde, SA and Chu, MF (2006) A hybrid origin for the Qianshan A-type granite, northeast China: geochemical and Sr–Nd–Hf isotopic evidence. Lithos 89, 89106.CrossRefGoogle Scholar
Yang, JH, Wu, FY, Wilde, SA, Xie, LW, Yang, YH and Liu, XM (2007) Tracing magma mixing in granite genesis: in situ U-Pb dating and Hf-isotope analysis of zircons. Contributions to Mineralogy & Petrology 153, 177–90.CrossRefGoogle Scholar
Yin, H, Zhou, H, Zhang, W, Zheng, X and Wang, S (2016) Late Carboniferous to early Permian sedimentary–tectonic evolution of the north of Alxa, Inner Mongolia, China: evidence from the Amushan Formation. Geoscience Frontiers 7, 733–41.CrossRefGoogle Scholar
Yuan, W and Yang, Z (2015) The Alashan Terrane was not part of North China by the Late Devonian: evidence from detrital zircon U–Pb geochronology and Hf isotopes. Gondwana Research 27, 1270–82.CrossRefGoogle Scholar
Zhang, B, Zhang, J, Zhang, Y, Zhao, H, Wang, Y and Nie, F (2016a) Tectonic affinity of the Alxa Block, Northwest China: constrained by detrital zircon U–Pb ages from the early Paleozoic strata on its southern and eastern margins. Sedimentary Geology 339, 289303.CrossRefGoogle Scholar
Zhang, D, Huang, B, Zhao, Q and Zhang, Y (2018a) Paleomagnetic results from Lower Devonian sandstones of the Niqiuhe Formation in the Duobaoshan area and its constraints on paleoposition of the Xing’an block. Chinese Science Bulletin 63, 1502–14 (in Chinese with English abstract).CrossRefGoogle Scholar
Zhang, J and Gong, J (2018) Revisiting the nature and affinity of the Alxa Block. Acta Petrologica Sinica 34, 940–62 (in Chinese with English abstract).Google Scholar
Zhang, J, Gong, J, Yu, S, Li, H and Hou, K (2013a) Neoarchean–Paleoproterozoic multiple tectonothermal events in the western Alxa block, North China Craton and their geological implication: evidence from zircon U–Pb ages and Hf isotopic composition. Precambrian Research 235, 3657.CrossRefGoogle Scholar
Zhang, J, Li, J, Xiao, W, Wang, Y and Qi, W (2013b) Kinematics and geochronology of multistage ductile deformation along the eastern Alxa block, NW China: new constraints on the relationship between the North China Plate and the Alxa block. Journal of Structural Geology 57, 3857.CrossRefGoogle Scholar
Zhang, J, Wang, T, Castro, A, Zhang, L, Shi, X, Tong, Y, Zhang, Z, Guo, L, Yang, Q and Iaccheri, LM (2016b) Multiple mixing and hybridization from magma source to final emplacement in the Permian Yamatu Pluton, the northern Alxa Block, China. Journal of Petrology 57, 933–80.CrossRefGoogle Scholar
Zhang, J, Wei, C and Chu, H (2015a) Blueschist metamorphism and its tectonic implication of Late Paleozoic–Early Mesozoic metabasites in the mélange zones, central Inner Mongolia, China. Journal of Asian Earth Sciences 97, Part B, 352–64.CrossRefGoogle Scholar
Zhang, J, Wei, C and Chu, H (2018b) New model for the tectonic evolution of Xing’an-Inner Mongolia Orogenic Belt: evidence from four different phases of metamorphism in Central Inner Mongolia. Acta Petrologica Sinica 34, 2857–572 (in Chinese with English abstract).Google Scholar
Zhang, J, Zhang, B and Zhao, H (2016c) Timing of amalgamation of the Alxa Block and the North China Block: constraints based on detrital zircon U–Pb ages and sedimentologic and structural evidence. Tectonophysics 668–669, 6581.CrossRefGoogle Scholar
Zhang, L, Zhang, H, Zhang, S, Xiong, Z, Luo, B, Yang, H, Pan, F, Zhou, X, Xu, W and Guo, L (2017a) Lithospheric delamination in post-collisional setting: evidence from intrusive magmatism from the North Qilian orogen to southern margin of the Alxa block, NW China. Lithos 288–289, 2034.CrossRefGoogle Scholar
Zhang, Q, Liu, Z, Chai, S, Xu, Z, Zhao, Q and Xu, X (2011) Geochronology and geochemistry of granodiorites from Wulan area of Urad Zhongqi, Inner Mongolia. Journal of Mineralogy and Petrology 31, 714 (in Chinese with English abstract).Google Scholar
Zhang, S and Zhao, Y (2017) Cogenetic origin of mafic microgranular enclaves in calc-alkaline granitoids: the Permian plutons in the northern North China Block. Geosphere 13, 482517.CrossRefGoogle Scholar
Zhang, SH, Zhao, Y, Liu, JM and Hu, ZC (2016d) Different sources involved in generation of continental arc volcanism: the Carboniferous–Permian volcanic rocks in the northern margin of the North China block. Lithos 240–243, 382401.CrossRefGoogle Scholar
Zhang, SH, Zhao, Y, Song, B, Yang, ZY, Hu, JM and Wu, H (2007) Carboniferous granitic plutons from the northern margin of the North China block: implications for a late Palaeozoic active continental margin. Journal of the Geological Society 164, 451–63.CrossRefGoogle Scholar
Zhang, SH, Zhao, Y, Ye, H, Liu, JM and Hu, ZC (2014) Origin and evolution of the Bainaimiao arc belt: implications for crustal growth in the southern Central Asian orogenic belt. GSA Bulletin 126, 1275–300.CrossRefGoogle Scholar
Zhang, X, Gao, Y, Wang, Z, Liu, H and Ma, Y (2012a) Carboniferous appinitic intrusions from the northern North China Craton: geochemistry, petrogenesis and tectonic implications. Journal of the Geological Society 169, 337–51.CrossRefGoogle Scholar
Zhang, X, Xue, F, Yuan, L, Ma, Y and Wilde, SA (2012b) Late Permian appinite–granite complex from northwestern Liaoning, North China Craton: petrogenesis and tectonic implications. Lithos 155, 201–17.CrossRefGoogle Scholar
Zhang, X, Yuan, L, Xue, F, Yan, X and Mao, Q (2015b) Early Permian A-type granites from central Inner Mongolia, North China: magmatic tracer of post-collisional tectonics and oceanic crustal recycling. Gondwana Research 28, 311–27.CrossRefGoogle Scholar
Zhang, Y, Niu, Y, Wei, J, Shi, J and Song, B (2018c) Chronology of the Haobiru Formation in the Haobiru area of northern Alxa, Inner Mongolia and its geological implications. Geological Bulletin of China 37, 5162 (in Chinese with English abstract).Google Scholar
Zhang, Z, Chen, Y, Li, K, Li, J, Yang, J and Qian, X (2017b) Geochronology and geochemistry of Permian bimodal volcanic rocks from central Inner Mongolia, China: implications for the late Palaeozoic tectonic evolution of the south-eastern Central Asian Orogenic Belt. Journal of Asian Earth Sciences 135, 370–89.CrossRefGoogle Scholar
Zhang, Z, Li, K, Li, J, Tang, W, Chen, Y and Luo, Z (2015c) Geochronology and geochemistry of the Eastern Erenhot ophiolitic complex: implications for the tectonic evolution of the Inner Mongolia–Daxinganling Orogenic Belt. Journal of Asian Earth Sciences 97, 279–93.CrossRefGoogle Scholar
Zhang, Z, Wang, K, Wang, G, Liu, X, Liu, W and Wu, B (2018d) Petrogenesis and tectonic significances of the Paleozoic Jiling syenite in the mountain Longshou area, Gansu province. Geological Review 64, 1017–29 (in Chinese with English abstract).Google Scholar
Zhao, G, Cawood, PA, Li, S, Wilde, SA, Sun, M, Zhang, J, He, Y and Yin, C (2012) Amalgamation of the North China Craton: key issues and discussion. Precambrian Research 222–223, 5576.CrossRefGoogle Scholar
Zhao, G, Sun, M, Wilde, SA and Li, SZ (2005) Late Archean to Paleoproterozoic evolution of the North China Craton: key issues revisited. Precambrian Research 136, 177202.CrossRefGoogle Scholar
Zhao, G, Wang, Y, Huang, B, Dong, Y, Li, S, Zhang, G and Yu, S (2018) Geological reconstructions of the East Asian blocks: from the breakup of Rodinia to the assembly of Pangea. Earth-Science Reviews 186, 262–86.CrossRefGoogle Scholar
Zhao, P, Chen, Y, Xu, B, Faure, M, Shi, G and Choulet, F (2013) Did the Paleo-Asian Ocean between North China Block and Mongolia Block exist during the late Paleozoic? First paleomagnetic evidence from central-eastern Inner Mongolia, China. Journal of Geophysical Research: Solid Earth 118, 1873–94.CrossRefGoogle Scholar
Zhao, P, Jahn, BM, Xu, B, Liao, W and Wang, Y (2016a) Geochemistry, geochronology and zircon Hf isotopic study of peralkaline-alkaline intrusions along the northern margin of the North China Craton and its tectonic implication for the southeastern Central Asian Orogenic Belt. Lithos 261, 92108.CrossRefGoogle Scholar
Zhao, P, Xu, B, Tong, Q, Chen, Y and Faure, M (2016b) Sedimentological and geochronological constraints on the Carboniferous evolution of central Inner Mongolia, southeastern Central Asian Orogenic Belt: Inland sea deposition in a post-orogenic setting. Gondwana Research 31, 253–70.CrossRefGoogle Scholar
Zhao, P, Xu, B and Zhang, C (2017) A rift system in southeastern Central Asian Orogenic Belt: constraint from sedimentological, geochronological and geochemical investigations of the Late Carboniferous-Early Permian strata in northern Inner Mongolia (China). Gondwana Research 47, 342–57.CrossRefGoogle Scholar
Zhao, X, Liu, C, Wang, J, Zhang, S and Guan, Y (2020) Geochemistry, geochronology and Hf isotope of granitoids in the northern Alxa region: implications for the Late Paleozoic tectonic evolution of the Central Asian Orogenic Belt. Geoscience Frontiers 11, 1711–25.CrossRefGoogle Scholar
Zheng, R, Li, J, Xiao, W and Wang, L (2018) A new ophiolitic mélange containing boninitic blocks in Alxa region: implications for Permian subduction events in southern CAOB. Geoscience Frontiers 9, 1355–67.CrossRefGoogle Scholar
Zheng, R, Li, J, Zhang, J, Xiao, W and Li, Y (2019) Early Carboniferous high Ba-Sr granitoid in southern Langshan of northeastern Alxa: implications for accretionary tectonics along the southern Central Asian Orogenic Belt. Acta Geologica Sinica (English Edition) 93, 820–44.CrossRefGoogle Scholar
Zheng, R, Wu, T, Zhang, W, Xu, C, Meng, Q and Zhang, Z (2014) Late Paleozoic subduction system in the northern margin of the Alxa block, Altaids: geochronological and geochemical evidences from ophiolites. Gondwana Research 25, 842–58.CrossRefGoogle Scholar
Zhou, JB, Wilde, SA, Zhao, GC and Han, J (2018) Nature and assembly of microcontinental blocks within the Paleo-Asian Ocean. Earth-Science Reviews 186, 7693.CrossRefGoogle Scholar
Zhou, XC, Zhang, HF, Luo, BJ, Pan, FB, Zhang, SS and Guo, L (2016) Origin of high Sr/Y-type granitic magmatism in the southwestern of the Alxa Block, Northwest China. Lithos 256–257, 211–27.CrossRefGoogle Scholar
Zhou, Z, Zhang, H, Liu, H, Liu, C and Liu, W (2009) Zircon U-Pb dating of basic intrusions in Siziwangqi area of middle Inner Mongolia, China. Acta Petrologica Sinica 25, 1519–28 (in Chinese with English abstract).Google Scholar
Figure 0

Fig. 1. (a) Tectonic location of the Alxa Block. (b) Schematic geological map showing the distribution of Palaeozoic intrusions and ophiolitic mélanges in the Alxa Block (modified after Dan et al.2014). (c) Simplified geological map of the southwestern Alxa Block (modified after Wang et al.2020).

Figure 1

Fig. 2. Statistical histograms of zircon U–Pb ages of Palaeozoic magmatic rocks in the (a) Alxa Block (data from this study and Qin, 2012; Tang, 2015; Gong et al.2018a; Zhang et al.2018d; Liu et al.2019; Pan, 2019; Song et al.2019; Chen et al.2020; Wang et al.2020; Zhao et al.2020) and (b) the southeastern Central Asian Orogenic Belt (data from Wang et al.2015b).

Figure 2

Fig. 3. Field photographs and mineral assemblages under microscope (cross-polarized light) of the studied late early Carboniferous plutons in the southwestern Alxa Block. (a, b) 17WAL-17, alkali-feldspar granite; (c, d) 17WAL-35, granodiorite; (e, f) 17WAL-39, monzogranite. Afs – alkali-feldspar; Bt – biotite; Hbl – hornblende; Pl – plagioclase; Qtz – quartz.

Figure 3

Fig. 4. (a) Na2O + K2O versus SiO2 and (b) K2O versus SiO2 diagrams for the early Carboniferous plutons in the Alxa Block. Data sources include Wang et al. (2015b), Dan et al. (2016), Liu et al. (2016a) and Xue et al. (2017).

Figure 4

Fig. 5. (a, c) Chondrite-normalized REE patterns and (b, d) primitive mantle-normalized trace-element diagrams for the late early Carboniferous plutons in the southwestern Alxa Block. Compositions of C1 chondrite and primitive mantle after Sun & McDonough (1989).

Figure 5

Fig. 6. Cathodoluminescence (CL) images of representative zircon grains from the studied late early Carboniferous plutons in the southwestern Alxa Block.

Figure 6

Fig. 7. (a–c) Concordia diagrams showing LA-ICP-MS zircon U–Pb data of the studied late early Carboniferous plutons in the southwestern Alxa Block (all the diagrams and calculations are at the 2σ level).

Figure 7

Fig. 8. Sr–Nd isotopic features of early Carboniferous plutons in the Alxa Block. Symbols and data sources as for Figure 4.

Figure 8

Fig. 9. (a) Tectonic discrimination diagrams of Rb versus (Y + Nb) for the early Carboniferous felsic plutons in the Alxa Block (Pearce, 1996). (b) Plot of (K2O + Na2O)/CaO versus Zr + Nb + Ce + Y and (c) plot of Ce versus 10 000×Ga/Al for A-type granites (Whalen et al.1987). (d) Nb–Y–Ce diagram for distinguishing between A1 and A2 granites (Eby, 1992). Symbols and data sources as for Figure 4.

Figure 9

Fig. 10. Petrogenetic discrimination diagrams of (a) V–(Ti/1000) (Shervais, 1982), (b) (Zr/Y)–Zr (Pearce & Norry, 1979), (c) (La/Ba)–(La/Nb) (Saunders et al.1992), and (d) (Zr/Sm)–(Sr/Nd)–(Ti/V) (Wang et al.2016a) for the early Carboniferous mafic rocks in the Alxa Block. The Basin-and-Range rift-related basalt field refers to Wang et al. (2016a). Symbols and data sources as for Figure 4.

Figure 10

Fig. 11. Extensional tectonics of the Alxa Block and the southeastern CAOB during early Carboniferous time. (a) Micro-continental blocks within the southeastern CAOB had already been accreted to the northern NCC (Alxa Block) before early Carboniferous time. (b) During early Carboniferous time, the asthenospheric upwelling induced by either the roll-back or the break-off of the subducted PAO slab heated both the subduction-modified lithospheric mantle and the overlying crust, leading to the generation of the mafic and felsic plutons, respectively.

Supplementary material: File

Wang et al. supplementary material

Wang et al. supplementary material

Download Wang et al. supplementary material(File)
File 251.4 KB