Hostname: page-component-7bb8b95d7b-fmk2r Total loading time: 0 Render date: 2024-09-25T06:43:17.287Z Has data issue: false hasContentIssue false

Granitic magmatism associated with gold mineralization: evidence from the Baizhangzi gold deposit, in the northern North China Craton

Published online by Cambridge University Press:  21 June 2023

Zhi-xiong Zhao
Affiliation:
School of Earth Science and Resources, China University of Geosciences, Beijing, China Ulanqab Key Laboratory of Geospatial Big Data Application and Environmental Monitoring, Jining Normal University, Ulanqab, China
Wei-jun Xu
Affiliation:
School of Earth Science and Resources, China University of Geosciences, Beijing, China Rixing Mining Co. Limited of Lingyuan, Liaoning, China
Guo-chen Dong*
Affiliation:
School of Earth Science and Resources, China University of Geosciences, Beijing, China
M Santosh
Affiliation:
School of Earth Science and Resources, China University of Geosciences, Beijing, China Department of Earth Science, University of Adelaide, Adelaide, Australia
Hong-bin Li
Affiliation:
School of Earth Science and Resources, China University of Geosciences, Beijing, China
Ze-guang Chang
Affiliation:
School of Earth Science and Resources, China University of Geosciences, Beijing, China
*
Corresponding author: Guo-chen Dong; Email: donggc@cugb.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

The relationship between magmatism and gold mineralization has been a topic of interest in understanding the formation of ore deposits. The Baizhangzi gold deposit, located in the northern margin of the North China Craton, is hosted by the Baizhangzi granite (BZG) and provides a case to evaluate the relation between granite and gold mineralization in Late Triassic. In this study, we present petrography, bulk geochemistry, zircon U-Pb isotope and trace elements data, as well as major elements of biotite and plagioclase for the BZG to evaluate the petrogenesis and link with gold mineralization. The BZG comprises biotite monzogranite, biotite-bearing monzogranite and monzogranite (BZGs). Zircon U-Pb geochronology shows that all the granitoids of BZGs were coeval with a formation age of 232 Ma. The granitoids, with high SiO2, Al2O3 and Sr, while low Y and Yb, show adakitic affinity. They are enriched in LILFs (e.g., Rb, Ba, Th, U and Sr) and LREEs, while depletion in HFSEs (e.g., Nb, Ta, P and Ti). The geochemical and mineral chemical data suggest that the granitoids have experienced the fractional crystallization of biotite + plagioclase + K- feldspar + apatite. Crystallization temperature is estimated as ca. 700°C, and pressure is between 0.71 kbar and 1.60 kbar. The monzogranite shows higher values of logfO2, △FMQ and △NNO than the biotite-bearing monzogranite, ranging from −19.76 to −11.71, −4.93 to +3.67 and −5.48 to +3.11, respectively. The fractional crystallization, together with high fO2, K-metasomatism and low evolution degree, provided favourable conditions for gold mineralization.

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

1. Introduction

The intrusion-related gold deposits (IRGDs) exhibit that the gold mineralization is spatially and temporally associated with the granitoids (Goldfarb et al. Reference Goldfarb, Groves and Gardoll2001; Lang & Baker, Reference Lang and Baker2001). They mainly contain intrusion-hosted, proximal and distal deposit styles (Dressel et al. Reference Dressel, Chauvet, Trzaskos, Biondi, Bruguier, Monié, Villanova and Newton2018). The characteristics of the latter two types are the relationship between the structural setting and their ore-forming process (Cepedal et al. Reference Cepedal, Fuertes-Fuente, Martín-Izard, García-Nieto and Boiron2013; Tuduri et al. Reference Tuduri, Chauvet, Barbanson, Labriki, Dubois, Trapy, Lahfid, Poujol, Melleton, Badra, Ennaciri and Maacha2018), for instance, the Fuwen Au-dominated Au-Ag deposit (Xu et al. Reference Xu, Wang, Wu, Zhou, Shan, Hou, Fu and Zhang2017) and Num˜ao gold deposit (Leal et al. Reference Leal, Lima, Morris, Pedro and Noronha2022). However, the intrusion-hosted deposits are characterized by the direct effect of magmatism on gold mineralization (Lang & Baker, Reference Lang and Baker2001), such as the Linares deposit (Cepedal et al. Reference Cepedal, Fuertes-Fuente, Martín-Izard, García-Nieto and Boiron2013), Bilihe gold deposit (Yang et al. Reference Yang, Chang, Hou and Meffre2016) and the Julia deposit (Soloviev et al. Reference Soloviev, Kryazhev, Semenova, Kalinin, Dvurechenskaya and Sidorova2022). The IRGDs have been intensively studied, and the role of magmatism in gold mineralization is a subject of debate. However, there is a consensus on the basically consistent time of magmatism and mineralization (e.g. Yang et al. Reference Yang, Chang, Hou and Meffre2016; Wang et al. Reference Wang, Wang, Bi, Tao and Lan2020; Liu et al. Reference Liu, Zhao and Liu2021; Soloviev et al. Reference Soloviev, Kryazhev, Semenova, Kalinin, Dvurechenskaya and Sidorova2022). In summary, four different aspects have been stressed: (1) the origin of magma, such as reworking of an ancient crust is related to the formation of gold in the Argun massif (Liu et al. Reference Liu, He, Lai, Wang and Li2022); (2) fractional crystallization, for instance, highly fractionated quartz diorite porphyry is related to the Bilihe gold deposit (Yang et al. Reference Yang, Chang, Hou and Meffre2016); (3) composition (type, volatiles, K2O content and alkalinity), for example adakitic granitoids have genetic links to the Fuwen Au-Ag ore deposit (e.g. Zorin et al. Reference Zorin, Zorina, Spiridonov and Rutshtein2001; Xu et al. Reference Xu, Wang, Wu, Zhou, Shan, Hou, Fu and Zhang2017); ilmenite-series igneous rocks are related to the Bilihe gold deposit (Yang et al. Reference Yang, Chang, Hou and Meffre2016); higher volatiles, K and alkaline are more conducive to the gold mineralization in the northern margin of the North China Craton and Xiaoqinling (e.g. N’dri et al. Reference N’dri, Zhang, Zhang, Tamehe, Kouamelan, Wu, Assie, Koua, Kouamelan and Zhang2021; Liu et al. Reference Liu, He, Lai, Wang and Li2022); and (4) high oxygen fugacity. Furthermore, Blevin (Reference Blevin2004) has assessed the relationships of magmatic-hydrothermal mineralization associated with an igneous suite, suggesting that (1) silica distributions ‘skewed’ to high values are typically associated with Au, (2) K/Rb ratios greater than 200 are associated with Cu-Au systems, (3) the degree and type of fractionation determine the potential for mineralization and the type of mineralization and (4) the relative oxidation state control the compatible or incompatible nature of the ore. These investigations suggest that the composition, magmatic process and physiochemical conditions are effective parameters to assess the relationship between magmatism and gold mineralization.

The North China Craton (NCC) is one of the major gold-producing regions in China, and over 60% of the gold deposits are hosted by or are related to Phanerozoic intrusions (Nie et al. Reference Nie, Jiang and Liu2004; Deng & Wang, Reference Deng and Wang2016; Yang & Santosh, Reference Yang and Santosh2020). These gold deposits have been classified as either IRGDs (Sillitoe & Thompson, Reference Sillitoe and Thompson1998) or orogenic gold deposits (Goldfarb et al. Reference Goldfarb, Groves and Gardoll2001; Miao et al. Reference Miao, Qiu, Fan, Zhang and Zhai2005; Goldfarb & Groves, Reference Goldfarb and Groves2011; Groves & Santosh, Reference Groves and Santosh2016; Santosh & Groves, Reference Santosh and Groves2022). It is, therefore, of great significance to study the relationship between magmatism and gold mineralization.

There are many gold deposits defining gold ore concentration in the eastern Hebei – western Liaoning area in the northern margin of the NCC (Kong et al. Reference Kong, Xu, Yin, Chen, Li, Guo, Yang and Shao2015). These deposits are mostly medium to large scale in the ore reserves and are distributed as ‘satellites’ around the Dushan batholith (DSB) (Fig. 1b), such as Yu’erya, Jinchangyu, Huajian, Baizhangzi, etc. They are closely related to the Mesozoic alkaline to calc-alkaline intrusive rocks, directly hosted by, or located in the vicinity of igneous rocks with almost the same mineralization age as that of magmatism (Chen et al. Reference Chen, Ye, Wang, He, Zhang and Wang2019; Zhang et al. Reference Zhang, Zhang, Danyushevsky, Wu, Alexis, Liao and Zhang2020), indicating a typical IRGD characteristic according to Lang and Baker (Reference Lang and Baker2001) and Goldfarb et al. (Reference Goldfarb, Groves and Gardoll2001). These gold deposits have long been studied (e.g. Sillitoe & Thompson, Reference Sillitoe and Thompson1998; Miao et al. Reference Miao, Qiu, Fan and Zhang2008; Chen et al. Reference Chen, Ye, Wang, Zhang, Lu and Hu2014; Song et al. Reference Song, Jiang, Bagas, Li, Hu, Zhang, Zhou and Ding2016; Bai et al. Reference Bai, Zhu, Zhang, Huang and Li2019), with the understanding that the gold mineralization is related to the magmatism. For example, the Jinchangyu gold deposit was regarded as IRGD type by Sillitoe and Thompson (Reference Sillitoe and Thompson1998) and further considered to be related to the DSB (Luo et al. Reference Luo, Guan, Qiu and Miao2001), while the Yu’erya gold deposit together with its hosted intrusion is identified as the product of partial melting in the lower crust during early Yanshan Movement (Chen et al. Reference Chen, Ye, Wang, Zhang, Lu and Hu2014), and the Baizhangzi gold deposit is thought to be related to the Late Triassic magmatism (Miao et al. Reference Miao, Qiu, Fan and Zhang2008).

Figure 1. (a) Tectonic map of the North China Craton (Jiang et al. Reference Jiang, Guo and Chang2013), CAOB – Central Asian Orogenic Belt, SLS – Solonker suture, QDOB – Qinling-Dabie Orogenic Belt. (b) Sketch geological map of the LX district showing the distribution of gold deposits and igneous rocks with ages (modified from Xiong et al. Reference Xiong, Shi, Li, Tian, Chen, Zhou, Zhao and Li2017; Zhang et al. Reference Zhang, Zhang, Danyushevsky, Wu, Alexis, Liao and Zhang2020). (c) Sketch map of BZG at the 30m level with photos of contact relationship. Intrusive rocks: BJ – Baijiadian, DS – Dushan, DSZ – Dashizhuzi, GJ – Gaojiadian, GSZ – Gushanzi, HK – Hekanzi, JG – Jinbaogou, JJ – Jiajiashan, LW – Luowenyu, MJ – Maojiagou, NX – Niuxinshan, QS – Qingshankou, SJ – Sanjia, TZ – Tangzhangzi, XY – Xiaoyingzi, YE – Yu’erya, YZ-Yangzhangzi, ZJ-Zhaojiazhuang.

The Baizhangzi gold deposit, one of the typical IRGDs, is hosted by the Baizhangzi granite (BZG). Previous studies reported zircon SHRIMP (222 ± 3 Ma) (Luo et al. Reference Luo, Li, Guan, Qiu, Qiu, McNaughton and Groves2004) and LA-ICP-MS age (233 ± 3 Ma) (Xiong et al. Reference Xiong, Shi, Li, Tian, Chen, Zhou, Zhao and Li2017) of the BZG. The ore formation mechanism and metallogenic prognosis (Wang, Reference Wang1989; Wei et al. Reference Wei, Xuan, Ma, Diao, Liu, Zheng, Guo and Zhang2016; Wang et al. Reference Wang2018) were also reported. However, the factors which are conducive to the gold mineralization during the magmatic process have not been clarified. In this study, we present detailed petrography, bulk geochemistry, zircon U-Pb isotope and trace element data, as well as EPMA mineral compositions, with a view to evaluate the magmatism associated with gold mineralization to provide guidelines for the future mineral exploration in the area.

2. Geological setting and petrology

2.a. Geological setting

The NCC, as one of the oldest cratons in the world with a basement rocks as old as 3.8 Ga (e.g. Zhai & Santosh, Reference Zhai and Santosh2011), is bordered by the Central Asian Orogenic Belt (CAOB) in the north, the Qinling-Dabie Orogenic Belt (QDOB) in the south and the Sulu Orogen in the east (Fig. 1a). The basement of the NCC is composed of metamorphosed Archaean and Paleoproterozoic rocks, including tonalite-trondhjemite-granodiorite (TTG) gneisses, amphibolite and a few supracrustal rocks (Lu et al. Reference Lu, Zhao, Wang and Hao2008; Kong et al. Reference Kong, Xu, Yin, Chen, Li, Guo, Yang and Shao2015). These are overlain by post-Mesoproterozoic and Cambrian-Ordovician marine clastic and carbonate rocks, Middle and Late Carboniferous marine and continental facies, Permian to Triassic fluvial and delta facies sedimentary rocks and Jurassic-Cretaceous sedimentary and volcanic rocks. After the final cratonization of the NCC during the late Paleoproterozoic at about 1.85 Ga (Wilde et al. Reference Wilde, Zhao and Sun2002; Zhao & Zhai, Reference Zhao and Zhai2013), the craton remained mostly stable prior to Triassic. During Mesozoic, the NCC witnessed intense magmatic activity during six stages: Early Triassic, Middle-Late Triassic, Early Jurassic-earliest Middle Jurassic, Middle-Late Jurassic, Early Cretaceous and Late Cretaceous (Zhang et al. Reference Zhang, Zhao, Davis, Ye and Wu2014). Among them, Permian-Triassic magmatic rocks are widely distributed in the Yinshan and Yanshan belts along the northern margin of the NCC (Zhang et al. Reference Zhang, Zhao, Liu, Hu, Song, Liu and Wu2010). Middle–Late Triassic intrusive rocks are mainly distributed in the northern and eastern NCC, Liaodong and Korean Peninsula (Wu et al. Reference Wu, Yang and Liu2005a, Reference Wu, Han, Yang, Wilde and Zhai2007). Early Jurassic–earliest Middle Jurassic igneous rocks are mainly distributed in the southern Yanbian area (Wu et al. Reference Wu, Sun, Ge, Zhang, Grant, Wilde and Jahn2011) and Western Hills of Beijing in the northern NCC (Liu et al. Reference Liu, Li, Guo, Hou and He2012). Middle–Late Jurassic magmatic rocks are widely distributed in the Yanshan Belt in the northern NCC, the Liaodong area and Jiaodong Peninsula in the eastern NCC, the southern Yanbian-Liaobei area in the northeastern NCC and in the northern part of North Korea (Wu et al. Reference Wu, Sun, Ge, Zhang, Grant, Wilde and Jahn2011). Early Cretaceous igneous rocks are widely distributed throughout the eastern and central NCC and are considered to be a ‘giant igneous event’ in eastern China (Wu et al. Reference Wu, Lin, Wilde, Zhang and Yang2005b). Late Cretaceous magmatic rocks are relatively minor and distributed in the eastern part of the NCC (Zhang et al. Reference Zhang, Ma, Liao, Zhang and She2011).

The eastern Hebei – western Liaoning district, located in the northern margin of the NCC, is underlain by gneiss, amphibolite and granulites of the Mesoarchean Badaohe Group (Fig. 1b), with Proterozoic carbonate and Quaternary cover (Miao et al. Reference Miao, Qiu, Fan and Zhang2008; Zhang et al. Reference Zhang, Zhang, Danyushevsky, Wu, Alexis, Liao and Zhang2020). Dozens of intrusive rocks occur in this area, including the Dushan, Dashizhu, Hekanzi, Xiaoyingzi, Qingshan and BZGs, etc. (Fig. 1b), which belong to three periods of magmatism (Miao et al. Reference Miao, Qiu, Fan and Zhang2008): ∼220 Ma, 190 Ma–200 Ma and ∼170 Ma. Among them, the DSB is the largest with nearly 30 km in NE-SW and ∼17 km in E-W and an elliptic shape area of over 450 km2 (Miao et al. Reference Miao, Qiu, Fan and Zhang2008). Previous zircon U-Pb dating reported ages of 215.3 Ma (Rao et al. Reference Rao2002), 223 Ma (Luo et al. Reference Luo, Miao, Guan, Qiu, Qiu, McNaughton and Groves2003), 221–222 Ma (Ye et al. Reference Ye, Zhang, Zhao and Wu2014), 218 Ma (Jiang et al. Reference Jiang, Bagas, Liu and Zhang2018), 162 Ma and 170 Ma (Jiang et al. Reference Jiang, Bagas, Liu and Zhang2018), indicating magmatism during Late Triassic and Middle-Late Jurassic. There are a series of E-, NE- and NNE-trending faults and folds (Fig. 1b, e.g. EW-trending Xinglong-Xifengkou-Qinglong fault, Lingyuan fault extending NE-SW and Malanyu anticline) in the western Liaoning region (Zhang et al. Reference Zhang, Zhang, Danyushevsky, Wu, Alexis, Liao and Zhang2020). The E-trending faults are generally parallel to the northern margin of the NCC, which is believed to be related to the southward subduction of the Paleo-Asian Ocean (Davis et al. Reference Davis, Zheng, Wang, Darby, Zhang and Gehrels2001). The NE- to NNE-trending faults are in general parallel to the Tanlu fault, considered to be related to the subduction of the Pacific plate (Cox et al. Reference Cox, Dabiche, Engebretson and Ben-Avraham1989; Miao et al. Reference Miao, Qiu, Fan and Zhang2008).

2.b. The Baizhangzi granitoid

The BZG, located 3 km north of the Dashizhu (DSZ) granite (Fig. 1b), is the main ore-bearing rock of the Baizhangzi gold deposit. It occurs as a stock or dike along the NNE-trending fault, with an exposed area of 1 km2, intruding the Proterozoic clastic rocks and Triassic diorite (Luo et al. Reference Luo, Li, Guan, Qiu, Qiu, McNaughton and Groves2004; Xiong et al. Reference Xiong, Shi, Li, Tian, Chen, Zhou, Zhao and Li2017). Our field investigation and microscopic observation revealed that the Baizhangzi stock intrusion is composed of medium to fine biotite monzogranite, biotite-bearing monzogranite and monzogranite (BZGs) (Fig. 1c). These granites are separated by transitional boundaries (Fig. 1c).

Biotite monzogranite occurs as medium to fine-grained, dark red rocks (Fig. 2a), composed of plagioclase (35–40%), K- feldspar (30–35%), quartz (ca. 20%) and biotite (10–15%) (Fig. 2b). Zircon, apatite and magnetite are accessory minerals. Plagioclase is often euhedral with prismatic to lath-like form, and a few crystals are selectively sericitized (Fig. 2c). K- feldspar is present as sub-euhedral to anhedral. Quartz commonly is anhedral. Biotite occurs as euhedral to sub-euhedral with strong pleochroism from yellowish brown to dark reddish brown, and a few of them are altered into chlorite.

Figure 2. Photographs of BZGs. (a) and (b) Biotite Monzogranite. (c) Sericitization. (d) Biotite-bearing monzogranite. (e) and (f) Monzogranite. (g) Silicification (quartz metasomatic K- feldspar). (h) Potassium (K- feldspar metasomatic plagioclase). Ser – sericite; Bi – biotite; Kf – K- feldspar; Pl – plagioclase; Q – quartz.

Biotite-bearing monzogranite contains much less plagioclase (30–35%) and biotite (5 - 10%), while little more quartz (20–25%) (Fig. 2d). The accessory mineral types and alteration characteristics are the same as those of biotite monzogranite.

Monzogranite is constituted mainly by plagioclase (30–35%), K- feldspar (35–40%) and quartz (25%±), with <5% biotite (Fig. 2e, f). The accessory minerals are zircon, apatite and magnetite. These rocks occur as white or red colours, which is interpreted as silicification (Fig. 2g), and potassium alteration (Fig. 2h).

3. Analytical methods

3.a. LA-ICP-MS Zircon trace elemental and U-Pb dating

Zircon separations were performed at the Xinhang Institute of Surveying and Mapping in Hebei Province through panning, heavy liquid and magnetic separation. Around 180 zircons, with complete crystal shape, with no cracks and no inclusions, were selected under binocular and mounted with epoxy resin on a glass plate and polished. The transparent, reflected light and cathodoluminescence (CL) images were obtained at the Beijing Zhongke Mineral Research Testing Technology Co., Ltd. Zircon trace element and U-Pb analyses were conducted in the Beijing Zirconian Pilot Technology Co., Ltd., Beijing, China, using the UP 213 nm laser ablation system and Agilent 7500A ICP-MS produced by New Wave. Helium gas was used as carrier gas, and the laser beam diameter was 30 μm. One 91500 standard sample was used for calibration of test data at every 10 sample points. Andersen (Reference Andersen2002) was used for the calibration of ordinary lead. ICPMSDataCal program (Liu et al. Reference Liu, Hu, Zong, Gao and Chen2010) was used to process the data, and U-Pb age concordance plots were plotted using Isoplot 3.0 (Ludwig, Reference Ludwig2003).

3.b. Whole-rock major and trace elements

The least altered samples were selected and crushed to less than 200 mesh. The whole-rock major and trace element compositions were analysed at the Wuhan SampleSolution Analytical Technology Co., Ltd., Wuhan, China. Major element oxides were measured by X-ray fluorescence spectrometer (XRF). Trace elements were determined using an Agilent 7700a ICP–MS. The standards for major and trace elements are GBW07103, GBW07105, GBW07111, and BHVO-2, BCR-2, RGM-2, respectively, and the relative standard deviation is <5% for both of them.

3.c. Electron probe microanalyses

The mineral compositions of the studied rocks were analysed using an electron microprobe analyser (EPMA; JXA-8100, JEOL) with a 15 kV accelerating voltage, 20 nA probe current and 5 μm beam diameter, at the Institute of Geology, Chinese Academy of Geological Science. SiO2, Al2O3, MgO, MnO, CaO, Na2O, K2O, FeO, TiO2, Cr2O3 and NiO were analysed. Line and spectrometer crystal used for major element are as follows: Si (Kα, TAP), Na (Kα, TAP), K (Kα, PETJ), Al (Kα, TAP), Mg (Kα, TAP), Mn (Kα, LIFH), Fe (Kα, LIFH), Cr (Kα, LIFH), Ni (Kα, LIFH), Ti (Kα, PETJ) and Ca (Kα, PETJ). Count times were 10s for peak and 5s for background per element. Natural and synthetic minerals used for standardization are as follows: (1) for plagioclase: albite (Na and Al), K- feldspar (K), rutile (Ti), NiO (Ni), Cr2O3 (Cr), Fe2O3 (Fe), spessartite (Mn), diopside (Ca and Mg), plagioclase/K- feldspar (Si); (2) for biotite: albite (Na and Al), K- feldspar (K), rutile (Ti), NiO (Ni), Cr2O3 (Cr), Fe2O3 (Fe), spessartite (Mn), diopside (Ca), Biotite (Si and Mg). ZAF corrections were carried out. The estimated precisions for major elements are ±2%.

4. Results

4.a. Zircon trace elemental and U-Pb ages

One sample each from the biotite-bearing monzogranite (No. 0-8) and monzogranite (No. -30-7) were selected for LA-ICP-MS zircon U-Pb dating and trace elements. The analytical data are listed in Supplementary Tables S1, S2.

The zircon grains from both the biotite-bearing monzogranite and monzogranite are colourless and euhedral to subhedral. They show oscillatory zoning, without inherited cores in the CL images (Fig. 3a, c). The abundances of thorium and uranium of analysed spots range from 280 ppm to 1648 ppm, 245 ppm to 1390 ppm, respectively, corresponding to Th/U ratios between 0.92 and 1.35 (>0.4). Zircon shares light rare earth element (LREE)-inclined patterns, with varying contents of LREE, characteristics of the Ce-positive anomalies (Ce/Ce*) and Eu-negative anomalies (Eu/Eu*) (Fig. 3b, d), which are consistent with the characteristics of magmatic zircons (Hoskin & Ireland, Reference Hoskin and Ireland2000). All of such indicate a magmatic origin (Christopher et al. Reference Christopher, Keith and Robert2003; Jan & Sylvester, Reference Jan and Sylvester2003; Randall & Stephen, Reference Randall and Stephen2003).

Figure 3. U–Pb concordia diagrams with cathodoluminescence (CL) images of representative zircons analysed and zircon chondrite-normalized REE patterns from the BZGs. (a) and (b) Biotite-bearing monzogranite (sample No. 0-8). (c) and (d) Monzogranite (sample No. -30-7). Chondrite values are from Sun and McDonough (Reference Sun, Donough, Saunders and Norry1989).

Fifteen zircons from sample 0–8 yielded 206Pb/238U ages from 230 to 238 Ma, with a weighted mean age of 231.7 ± 3.6 Ma (MSWD = 0.69) (Fig. 3a). Fifteen zircons in sample -30-7 yielded 206Pb/238U ages from 229 to 241 Ma, with a weighted mean age of 231.7 ± 3.3 Ma (MSWD = 2.9) (Fig. 3c). The two ages are almost identical to each other.

4.b. Whole-rock major and trace element compositions

Two biotite monzogranites, four biotite-bearing monzogranites and seven monzogranites of the BZGs were analysed in this work. Whole-rock major and trace element compositions of these samples are listed in Supplementary Table S3. The following data of major elements are recalculated after removing the LOI.

The SiO2 content of biotite monzogranite samples ranges from 67.57 to 67.81 wt%, and Na2O +K2O ranges from 9.77 to 9.99 wt%. The data fall in the monzogranite field in the QAP diagram (Fig. 4a) and plot to the granite field in the An-Ab-Or diagram (Fig. 4b). The values of Na2O +K2O −CaO range from 8.03 to 8.18 wt%, and the two data plot along with the alkaline series in a (Na2O+K2O −CaO) versus SiO2 diagram (Fig. 4c). The content of Al2O3 is between 15.75 and 16.38 wt%, with A/CNK values ranging from 0.96 to 0.98, and A/NK values between 1.20 and 1.22, identifying with the metaluminous field in an A/CNK versus A/NK diagram (Fig. 4d). The content of MgO ranges from 0.87 to 1.18 wt%, with the Mg values (nMg2+/ (nMg2+ +nFe2+)*100) from 40.48 to 42.36. The biotite monzogranite shows similar chondrite-normalized rare earth element (REE) patterns (Fig. 5a), with being enriched in LREE (LaN = 356.03–357.66, YbN = 6.21–6.53 and LaN/YbN = 54.8–57.4) and insignificant Eu anomaly (Eu/Eu* = EuN/[(SmN) ×(GdN)]1/2 = 0.94–1.01). A primitive-mantle-normalized trace element diagram (Fig. 5b) shows negative anomalies for high field strength elements (HFSEs) (such as Nb, Ta, P and Ti) and variable enrichment in large ion lithophilic elements (LILEs) (for instance, Rb, Ba, Th, U and Sr), with high Sr (793–855 ppm).

Figure 4. Classification diagrams of samples from the BZGs. (a) QAP diagram (Le Maitre, Reference Le Maitre2002). (b) An-Ab-Or diagram (Barker & Millard, Reference Barker and Millard1979). (c) (Na2O +K2O-CaO) vs. SiO2 (Frost et al. Reference Frost, Barnes, Collins, Arculus, Ellis and Frost2001). (d) A/NK vs. A/CNK diagram (Manilar & Piccoli, Reference Manilar and Piccoli1989).

Figure 5. Chondrite-normalized rare earth element (REE) patterns (a, c, and e) and primitive-mantle-normalized trace element spider diagrams (b, d, and f) are shown. Chondrite values and primitive-mantle values are from Sun and McDonough (Reference Sun, Donough, Saunders and Norry1989).

The monzogranite exhibits higher SiO2 (from 68.70 to 70.51 wt%), and Na2O +K2O (9.87–10.25 wt%). In a QAP diagram (Fig. 4a), they all fall into the monzogranite field, which is consistent with plotted in the granite field in the An-Ab-Or diagram (Fig. 4b). The Na2O +K2O −CaO contents vary widely, from 8.15 to 9.02, and the data mainly plot along with the alkalic series in a (Na2O+K2O −CaO) versus SiO2 diagram (Fig. 4c). The Al2O3 content ranges from 15.16 to 15.56, with A/CNK and A/NK values 0.88- 0.96, 1.11- 1.16, respectively, corresponding to the metaluminous field in an A/CNK versus A/NK diagram (Fig. 4d). The MgO content ranges from 0.46 to 0.82 wt% (average of 0.66 wt%), with the Mg values (nMg2+/ (nMg2+ +nFe2+) *100) from 29.23 to 47.31. The monzogranite shows similar characteristics in the chondrite-normalized REE diagram (Fig. 5c) and primitive-mantle-normalized trace element diagram (Fig. 5d) with the biotite monzogranite, while the lower Sr (420–612 ppm) and Eu/Eu* (0.77–0.89) relatively, suggesting that the crystallization of plagioclase may have occurred.

The biotite-bearing monzogranite displays a coherent behaviour of major and trace elements with the other rock types (Figs. 4, 5) suggesting that the biotite monzogranite, biotite-bearing monzogranite and monzogranite are likely cogenetic, and magma evolution may have played an important role in their formation.

For the BZGs, we used the samples of Xiong et al. (Reference Xiong, Shi, Li, Tian, Chen, Zhou, Zhao and Li2017) and Luo et al. (Reference Luo, Li, Guan, Qiu, Qiu, McNaughton and Groves2004).

4.c. Mineral geochemistry

Representative grains of plagioclase and biotite were analysed by microprobe for mineral chemistry.

4.c.1. Plagioclase

The following samples were analysed: thirteen plagioclases from the three biotite monzogranite samples (20BZ125, 20BZ126, 20BZ07), fourteen plagioclases from the three biotite-bearing monzogranite samples (120-1, 19BZ38, 20BZ16) and thirty-five plagioclases from the nine monzogranite samples (150-1, 0-12, 20BZ31, 20BAZ36, 108-12, -30-5, 20BZ01, 19BZ41, 19BZ35). Plagioclase in the BZGs is generally fresh to slightly altered and euhedral to subhedral tabular crystals, exhibiting obvious polysynthetic twinning under cross-polarized light. In this study, unaltered regions of plagioclase were selected for chemical composition analysis; therefore, the analysis results are little or not affected by alteration. Some representative data are listed in Supplementary Table S4.

The major element contents of plagioclase from the biotite monzogranite are as follows: SiO2 (63.5–67.9 wt%, average 65.7%), Al2O3 (19.5–23.0 wt%, average 20.8%), CaO (0.42–3.07 wt%, average 1.67%), FeO (0.08–0.58 wt%, average 0.23%), Na2O (9.2–11.8 wt%, average 10.6%) and K2O (0.1–0.8 wt%, average 0.36%), while An values range from An2.01 to An14.06, with an average of An7.84, corresponding to four of them belong to oligoclase and the other nine belong to albite in Ab-An-Or diagram (Fig. 6).

Figure 6. Ternary plots of plagioclase compositions (after Smith, Reference Smith1974), and the right is a 5× magnification of the projection area of the left.

The plagioclase from the monzogranite has higher SiO2 (ranges from 64.9 to 68.9wt%, average 67.3%), Na2O (9.3–12.3 wt%, average 11.3%) and K2O (0.06–3.12 wt%, average 0.41%); however, lower Al2O3 (19.0–21.5 wt%, average 19.9%), CaO (0.006–1.7 wt%, average 0.4%), FeO (0.003–0.66 wt%, average 0.17%) and An values (ranged from An0.03 to An7.78, with an average of An2.02), corresponding to all of them belong to albite (Fig. 6).

The average major element contents of plagioclase from the biotite-bearing monzogranite are as follows: SiO2 (66.8%), Al2O3 (20.2%), CaO (0.9%), Na2O (10.9) and An4.22, and all belong to albite (Fig. 6).

The composition and chemical formula of plagioclase listed in Supplementary Table S5 show that the number of Na, Ca, Al, Si from the biotite monzogranite ranges from 0.7922 to 1.0030 (average 0.9095), from 0.0198 to 0.1466 (average 0.0791), from 1.0098 to 1.2033 (average 1.0857), from 2.8205 to 2.9817 (average 2.9087), respectively. There are more Na and Si (average 0.9667, 2.9650, respectively), while less Ca and Al (average 0.0207, 1.0323, respectively) of the monzogranite. The number of cations (Na, Ca, Al, Si) lie between biotite monzogranite and monzogranite (average 0.9370, 0.0420, 1.0534, 2.9467, respectively) from the biotite-bearing monzogranite. The mean chemical formula of plagioclase from three types of rock are (Na0.9095, Ca0.0791) [Al1.0857Si2.9087O8], (Na0.9370, Ca0.0420) [Al1.0534Si2.9467O8] and (Na0.9667, Ca0.0207) [Al1.0323Si2.9650O8], respectively.

On the An vs. SiO2 Al2O3, CaO, Na2O, Ab, Or diagram (Fig. 7), plagioclase from the biotite-bearing monzogranite mostly plot between biotite monzogranite and monzogranite, exhibiting the same characteristic as the major and trace elements of the BZGs, which also suggests that the BZGs are products from cogenetic magma.

Figure 7. Correlation diagram of plagioclase compositions.

4.c.2. Biotite

The EPMA data and chemical formula of biotite are presented in Supplementary Table S6. The number of cations and related parameters is calculated using the Li (Li et al. Reference Li, Li, Xu, Yang, Wang, Xu and Ning2020) method, based on the general formula of A1M3T4O10W2.

The biotite from the biotite monzogranite shows SiO2 from 34.61 wt% to 36.32wt% (average 35.67%), Al2O3 in the range of 13.54–14.59 wt% (average 13.86 wt%), FeO of 18.73–23.88 wt%, (average 20.97 wt%), MgO of 9.80–12.15 wt% (average 11.55 wt%), TiO2 of 3.60–4.23 wt% (average 3.92 wt%) and K2O of 7.98–8.96 wt% (average 8.58 wt%).

The biotite from the biotite-bearing monzogranite shows SiO2 (35.27wt%), Al2O3 (13.29 wt%,), FeO (21.23 wt%), MgO (12.92 wt%), TiO2 (4.56 wt%) and K2O (7.59 wt%).

The biotite from the monzogranite has the highest SiO2 (36.92wt%) and MgO (13.16 wt%), while the lowest FeO (18.24 wt%) and TiO2 (4.56 wt%).

Fe /(Fe+Mg) ratio ranges from 0.44 to 0.58, suggesting that they were not been altered by fluids (Stone, Reference Stone2000), consistent with falling in the primary biotite field in the 10×TiO2-FeO-MgO diagram (Fig. 8a). The Mg/(Fe+Mg) ratio is relatively high, corresponding to magnesium biotite in the Mg-(AlVI+Fe3++Ti) –(Fe2++Mn) diagram (Fig. 8b). The Mg/(Fe+Mg) ratio increases gradually from the biotite monzogranite (average 0.49) to the monzogranite (0.56), indicating that the biotite became more Mg-rich as the magma evolved.

Figure 8. Genesis (a, after Nachit et al. Reference Nachit, Ibhi, Abia and Ohoud2005) and classification (b, after Foster, Reference Foster1960) diagram of biotite.

5. Discussion

5.a. Magmatic evolution

The contents of TiO2, Al2O3, FeO, CaO, P2O5 and MgO (Fig. 9a–f) decrease with increasing SiO2 from the biotite monzogranite to the monzogranite, attributed to the separation of some minerals. As previously mentioned, the biotite monzogranite is composed of plagioclase, K- feldspar, quartz, biotite, minor zircon, apatite and magnetite. Among them, biotite in these rocks has high FeO, MgO, Al2O3 and TiO2 (Table S6). Significant fractionation of biotite can cause the variation as observed in Fig. 9. The relatively stable concentration of TiO2 with decreasing Nb/Ta ratios (Fig. 10a) precludes any major role of magnetite, as the separation of magnetite would lead to a negative correlation between Ti and Nb/Ta (Green & Pearson, Reference Green and Pearson1987; Wang et al. Reference Wang, Zeng, Chen, Cai, Yan, Hou and Wang2017). Apatite fractionation was weak during the magmatic evolution, because the P2O5 content shows only slight decrease from the biotite monzogranite (average 0.25%) to the monzogranite (average 0.14%). The P2O5 also shows decrease with respect to the REE (Fig. 10b). Plagioclase is a major mineral in the monzogranite and has high Al2O3, CaO and Na2O (Table S4). The obvious separation of plagioclase, supported by lower Sr, Eu/Eu* for the monzogranite, the weak decrease in Rb concentrations and decrease of Ba concentrations with decreasing Sr concentrations (Fig. 10c, d), the increasing Rb/Ba and Rb/Sr ratios with decreasing Eu and Ba concentrations (Fig. 10e, f), are also typical. It has been known that fractional crystallization of plagioclase could result in decreasing of Sr of the residual magma, while the declining Ba is due to the separation of K- feldspar, attributed to the high partition coefficient of Sr for plagioclase and Ba for K- feldspar. The lower Ba (average 1284 ppm) and Sr (average 510 ppm) concentrations for the monzogranite than the biotite monzogranite (average 1637 ppm and 824 ppm, respectively) also indicate fractional crystallization of plagioclase and K- feldspar. Collectively, the BZGs are the products of magmatic evolution, and fractional crystallization of biotite, plagioclase, K- feldspar and apatite is the major magmatic process.

Figure 9. Variation of ω(SiO2) % vs. ω(TiO2) %, ω(Al2O3) %, ω(FeOT) %, ω(CaO) %, ω(P2O5) %, ω(MgO) %.

Figure 10. Diagrams of (a) TiO2 (wt%) versus Nb/Ta ratios, (b) P2O5 (wt%) versus REE (ppm) (after Wang et al. Reference Wang, Zeng, Chen, Cai, Yan, Hou and Wang2017), (c) Sr (ppm) versus Rb (ppm), (d) Sr (ppm) versus Ba (ppm), (e) Eu (ppm) versus Rb/Ba ratios and (F) Ba (ppm) versus Rb/Sr ratios, showing the fractional crystallization of different minerals. See Supplementary Table S7 for the partition of those minerals. Symbols are: Mt = magnetite; Sp = spinel; Hb = hornblende; Pl = plagioclase; Kf = K- feldspar; BI = biotite; Ap = apatite.

5.b. Petrogenetic model

The BZGs have high SiO2(≥56 %), Al2O3 (≥14.5%) and Sr(>400 ppm), low Y(<18 ppm) and Yb (<1.9 ppm), corresponding with the high Sr/Y (>35) and La/Yb (>60) features of characteristics of adakites (Defant & Drummond, Reference Defant and Drummond1990; Sajona et al. Reference Sajona, Maury, Bellon, Cotten, Defant, Pubellier and Rangin1993). Also, they plot in the adakite area in the Sr/Y-Y and (La/Yb)N-YbN diagram (Fig. 11). Both of them indicate that our studied rocks have the properties of adakite. Various models have been proposed for the genesis of adakites including: (1) partial melting of a subducting slab (e.g. Kay, Reference Kay1978; Saunders et al. Reference Saunders, Rogers, Marriner, Terrell and Verma1987); (2) partial melting of the thickened lower crust (e.g. Yang et al. Reference Yang, Chang, Hou and Meffre2016; Xu et al. Reference Xu, Gu, Wang, Zhang, He, Zhou and Liu2018); (3) mixing of mantle-derived and crust-derived magma (e.g. Xu et al. Reference Xu, Ma and Zhang2012; Zhang et al. Reference Zhang, Ma, Holtz, Koepke, Wolff and Berndt2013); and (4) fractional crystallization and assimilation (AFC) of mantle-derived basaltic magma (e.g. Chiaradia et al. Reference Chiaradia, Fontboté and Beate2004; Macpherson et al. Reference Macpherson, Dreher and Thirwall2006; Foley et al. Reference Foley, Pearson, Rushmer, Turner and Adam2013; Dai et al. Reference Dai, Zheng, Zhou and Griffin2017; Azizi et al. Reference Azizi, Stern, Topuz, Asahara and Moghadam2019; Temizel et al. Reference Temizel, Arslan, Yücel, Yazar, Kaygusuz and Aslan2020; Wang et al. Reference Wang, Wang, Bi, Tao and Lan2020).

Figure 11. (a) Variation of Sr/Y vs. Y. Fields for adakites and island arcs are from Defant and Drummond (Reference Defant and Drummond1990). (b)Variation of (La/Yb)N vs. YbN. Data are normalized (N) to C1 (chondrite) values of Sun & McDonough (1989). Fields for adakites and island arc magma are from Martin (Reference Martin1999).

In general, adakites formed by partial melting of the subducted slab are characterized by high Na2O and low K2O contents (Na2O >3.5%, K2O/Na2O≈0.4) (Richards & Kerrich, Reference Richards and Kerrich2007), which is not consistent with our samples because K2O/Na2O ratios range from 1.12 to 1.13 (average 1.12), 0.99 to 1.53 (average 1.18) and 0.90 to 1.57 (average 1.12), respectively. As a consequence, partial melting of subducted slab is unlikely. Second, there is no evidence for the required regional crustal uplift in response to the delamination of a thickened lower crust (Li et al. Reference Li, Li, Li, Wang and Gao2013). Furthermore, the BZGs are massive and homogeneous, and no direct evidence has been found in the field for magma mixing, such as the presence of mafic enclaves. The limited Sr-Nd isotopic composition variation with (87Sr/86Sr(t) ranging from 0.70445 to 0.70524, and ϵNd(t) from −7.3 to −1.7) (Xiong et al. Reference Xiong, Shi, Li, Tian, Chen, Zhou, Zhao and Li2017), and the lack of any correlation between Sr-Nd and SiO2 also precludes the magma mixing model (Li et al. Reference Li, Li, Li, Wang and Gao2013).

In summary, we infer that the adakitic BZGs were products of fractional crystallization and assimilation (AFC) of mantle-derived basaltic magma.

5.c. Magmatic physiochemical conditions

5.c.1. Temperature

The distribution of some trace elements in minerals is a function of temperature and is used as a geological thermometer (Albuquerque, Reference Albuquerque1973). Henry et al. (Reference Henry, Guidotti and Thomson2005) proposed the thermometers for titanium content in biotite: T = {[ln (Ti)-a-c (X Mg)3]/b}0.333. The procedure follows T = temperature (°C); Ti = the number of Ti cations in biotite based on 22 oxygen atoms, calculated using the Luhr method (Luhr et al. Reference Luhr, Carmichael and Varekamp1984); X Mg = Mg/ (Mg+Fe); a = −2.3594, b = 4.6484×10−9, c = −1.7283. The calculation results are listed in Supplementary Table S8, indicating that the crystallization temperature of the biotite ranges from 700 °C to 744 °C (average 723 °C). Besides, zircon crystallization temperatures of biotite-bearing monzogranite and monzogranite were estimated using the equations of Waston et al. (Reference Watson, Wark and Thomas2006): [log10Ti in zircon (ppm)] = (6.01 ± 0.03)–(5080 ± 30)/T(K), listed in Table S2. The biotite-bearing monzogranite shows zircon temperature in the range of 679–732°C (n = 15, average 705°C). The zircon temperatures of monzogranite are relatively concentrated (except for the -30-7-6 of 783 °C), ranging from 693 to727°C (n = 14, average 709 °C), suggesting that they have similar high crystallization temperature. Both estimates are consistent with the features shown by the biotite, indicating that the BZGs were formed at a temperature of ca. 700°C.

5.c.2. Pressure

The Al content of biotite is positively correlated with the pressure of granitoids according to Uchida et al. (Reference Uchida, Endo and Makino2007). They proposed the empirical formula: P(kb) = 3.03×T Al−6.53 (±0.33) (T Al = the number of Al cations in biotite based on 22 oxygen atoms, calculated using the Luhr method) (Luhr et al. Reference Luhr, Carmichael and Varekamp1984). The results show that the BZG emplacement pressure is between 0.71 kbar and 1.60 kbar (average 0.97 kbar), corresponding to paleodepths of 2.7 to 6.0 km under lithostatic pressure.

5.c.3. Oxygen fugacity

The concentration of Fe3+, Fe2+, Mg2+ in biotite can be used to estimate fO2 (e.g. Dymek, Reference Dymek1983; Wones, Reference Wones1989). The Fe2+-Fe3+-Mg2+ ternary diagram (Fig. 12) shows that the biotite from BZGs plot in the compositional fields between the Ni-NiO and Fe2O3-Fe3O4 buffer, and closer to Fe2O3-Fe3O4 line, indicating a high fO2 in the environment of biotite crystallization from the BZGs parent magma. Moreover, zircon fO2 was calculated using the Ti in zircon of Geo- fO2 software (Li et al. Reference Li, Cheng and Yang2019), listed in Table S2. LogfO2, △FMQ, and △NNO for the biotite-bearing monzogranite vary from −22.39 to −13.44 (average of −17.02), −6.31 to +1.30 (average of −1.67), and −6.88 to +0.74 (average of −2.22), respectively. The monzogranite shows higher values of logfO2, △FMQ and △NNO, ranging from −19.76 to −11.71 (average of −15.14), −4.93 to +3.67 (average of −0.03) and −5.48 to +3.11 (average of −0.58), respectively. These data suggest that the magma may have evolved to a more oxidized state from the biotite-bearing monzogranite to monzogranite.

Figure 12. Fe2+-Fe3+-Mg2+ ternary diagram (after Wones, Reference Wones1989) of the biotite from the BZGs.

5.d. Potential link between magmatism and mineralization

5.d.1. Conductive factors

Fractional crystallization is considered to be a critical component for mineralization (e.g. Blevin, Reference Blevin2004; Chiaradia et al. Reference Chiaradia, Ulianov, Kouzmanov and Beate2012; Hronsky et al. Reference Hronsky, Groves, Loucks and Begg2012). As discussed in previous sections, fractional crystallization played a dominant role during magma differentiation of the BZGs.

Besides, other three parameters have been proposed to evaluate the magmatic-hydrothermal associated with gold-bearing granitoids by Blevin (Reference Blevin2004): (1) redox state; (2) compositions, for example, K2O and alkalinity; (3) degree of magma evolution.

On the redox state classification diagram of Fe2+-Fe3+-Mg2+ (Fig. 12), the studied rocks are closer to Fe2O3-Fe3O4, indicating that they belong to the magnetite series. The magnetite-bearing granitoids are generally related to Au, Cu and Mo deposits (Ishihara, Reference Ishihara1977). Furthermore, the BZGs have a high fO2, as evidenced by EPMA data of biotite and trace element characteristics of zircon, which are favourable for gold mineralization (Park et al. Reference Park, Campbell, Kim and Moon2015; Xu et al. Reference Xu, Cao, Du, Wang and Pang2021).

The BZGs are alkali-rich, with Na2O + K2O values ranging from 9.77 to 9.99% (average 9.88%), 9.93 to 10.16% (average 10.07%), 9.87 to 10.25% (average 10.07%), respectively. On the (Na2O+K2O −CaO) versus SiO2 diagram, they all plot as alkaline series (Fig. 5c). On the other hand, potassium alteration is also observed (Fig. 2h), which is a significant feature of gold deposits related to alkaline magmatism (Jensen & Barton, Reference Jensen and Barton2000), and is favourable for the transport of gold (N’dri et al. Reference N’dri, Zhang, Zhang, Tamehe, Kouamelan, Wu, Assie, Koua, Kouamelan and Zhang2021).

The Rb/Sr and K/Rb ratios are two useful parameters to evaluate the degree of compositional evolution. The Rb/Sr ratio of the studied rocks is low (0.16–0.18, 0.14–0.22 and 0.22–0.32, respectively), indicating that these rocks are less evolved (Blevin, Reference Blevin2004). Previously studied Au-Cu deposits are related to low differentiation granites (Thompson et al. Reference Thompson, Sillitoe, Baker, Lang and Mortensen1999; Blevin et al. Reference Blevin, Chappell and Jones2003). In addition, granites associated with Cu and Cu-Au deposits have a K/Rb ratio >200 (Blevin, Reference Blevin2004). The K/Rb ratio is all greater than 200 of the BZGs, ranging from 296 to 318 (average 307), 275 to 326 (average 301) and 306 to 350 (average 324), respectively. The above features suggest that the BZGs are less evolved and conducive to gold mineralization according to Blevin (Reference Blevin2004).

5.d.2. Timing of magmatism and gold mineralization

The northern margin of the NCC is an important gold mineralization belt, and there are multiple magmatic events associated with numerous gold mineralization (e.g. N’dri et al. Reference N’dri, Zhang, Zhang, Tamehe, Kouamelan, Wu, Assie, Koua, Kouamelan and Zhang2021). These gold ores are hosted by both Precambrian metamorphic rocks and Variscan, Indosinian and Yanshanian granites (Zhou et al. Reference Zhou, Goldfarb and Phillips2002). Many gold deposits, related to the Mesozoic intrusive rocks, are distributed in the eastern Hebei – western Liaoning area in the northern margin of the NCC (Kong et al. Reference Kong, Xu, Yin, Chen, Li, Guo, Yang and Shao2015). These deposits have almost the same mineralization age as that of magmatism (Chen et al. Reference Chen, Ye, Wang, He, Zhang and Wang2019; Zhang et al. Reference Zhang, Zhang, Danyushevsky, Wu, Alexis, Liao and Zhang2020), such as Yuerya and Jinchangyu (Wang et al. Reference Wang, Wang, Bi, Tao and Lan2020).

In this study, the zircon U-Pb dates of 231.7 ± 3.6 Ma (biotite-bearing monzogranite sample 0–8) and 231.7 ± 3.3 Ma (monzogranite sample -30-7) have been obtained from the BZGs. They are very similar to those reported previously, where LA-ICP-MS zircon U-Pb ages show 233 ± 3 Ma (Xiong et al. Reference Xiong, Shi, Li, Tian, Chen, Zhou, Zhao and Li2017) and SHRIMP zircon U-Pb ages yielded 222 ± 3 Ma (Luo et al. Reference Luo, Li, Guan, Qiu, Qiu, McNaughton and Groves2004). All these data indicate that the time of magmatism is Late Triassic. We therefore speculate that the Baizhangzi gold deposit, one of the typical IRGDs, was also formed in the Late Triassic. It was also demonstrated by a number of investigated gold deposits hosted by igneous rocks, including Bilihe (Yang et al. Reference Yang, Chang, Hou and Meffre2016), Daxiyingzi (Liu et al. Reference Liu, Zhao and Liu2021), Julia (Soloviev et al. Reference Soloviev, Kryazhev, Semenova, Kalinin, Dvurechenskaya and Sidorova2022) and Xiajinbao (Wang et al. Reference Wang, Wang, Bi, Tao and Lan2020).

In summary, the BZGs experienced fractional crystallization, together with high fO2, alkali-rich, K-metasomatism and low evolution degree, which are conductive for gold mineralization. Furthermore, the monzogranite probably has high metallogenic potential due to its more oxidized state. We expect that this investigation will provide an important guide for regional gold exploration in the area.

6. Conclusions

  1. (1) Zircon U−Pb dating shows that the biotite-bearing monzogranite and monzogranite were formed at 231.7 ± 3.6 Ma and 231.7 ± 3.3 Ma, respectively. The dating, supporting previous geochronology, represents the age of magmatism and gold mineralization in the area.

  2. (2) Bulk and mineral geochemistry data indicate that the BZGs are cogenetic and fractional crystallization is the dominant magmatic process.

  3. (3) Crystallization temperature is ca. 700°C, and pressure is between 0.71 kb and 1.60 kb of the BZGs. Zircon trace elements suggest that the magma may have evolved to a more oxidized state.

  4. (4) The fractional crystallization, together with high fO2, K-metasomatism and low evolution degree, are considered favourable conditions for gold mineralization.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S0016756823000341

Acknowledgements

This research was financially supported by the Natural Science Foundation of China (grant No. 92062217 and 42121002). We thank Xiaohong Mao of the Institute of Geology, Chinese Academy of Geological Science for the assistance in geochemical and isotope analysis.

Competing interests

The authors declare that all authors have contributed to this manuscript, and they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

Albuquerque, CAR (1973) Geochemistry of biotite from granitic rocks, northern Portugal. Geochemical et Cosmochimica Acta 37, 1779–802.CrossRefGoogle Scholar
Andersen, T (2002) Correction of common lead in U-Pb analyses that do not report 204Pb. Chemical Geology 192, 5979.CrossRefGoogle Scholar
Azizi, H, Stern, RJ, Topuz, G, Asahara, Y and Moghadam, HS (2019) Late Paleocene adakitic granitoid from NW Iran and comparison with adakites in the NE Turkey: Adakitic melt generation in normal continental crust. Lithos 346–347, e105151. doi: 10.1016/j.lithos.2019.105151.CrossRefGoogle Scholar
Bai, Y, Zhu, MT, Zhang, LC, Huang, K and Li, WJ (2019) Auriferous pyrite Re-Os geochronology and He-Ar isotopic compositions of the Jinchangyu Au deposit in the northern margin of the North China Craton. Ore Geology Reviews 111, e102948. doi: 10.1016/j.oregeorev.2019.102948.CrossRefGoogle Scholar
Barker, F and Millard, HT (1979) Geochemistry of the type Trondhjemite and three associated rocks, Norway. Developments in Petrology 6, 517–29.CrossRefGoogle Scholar
Blevin, PL (2004) Redox and compositional parameters for interpreting the granitoid metallogeny of Eastern Australia: implications for good-rich ore systems. Resource Geology 54, 241–52.CrossRefGoogle Scholar
Blevin, PL, Chappell, BW and Jones, M (2003) Magmas to mineralisation: the Ishihara Symposium. Abstract volume GEMOC, Macquarie University, July 22–24.Google Scholar
Cepedal, A, Fuertes-Fuente, M, Martín-Izard, A, García-Nieto, J and Boiron, MC (2013) An intrusion-related gold deposit (IRGD) in the NW of Spain, the Linares deposit: igneous rocks, veins and related alterations, ore features and fluids involved. Journal of Geochemical Exploration 124, 101–26.CrossRefGoogle Scholar
Chen, SC, Ye, HT, Wang, YT, He, W, Zhang, XK and Wang, NN (2019) 40Ar-39Ar age of altered sericite from Yuerya Au deposit in eastern Hebei Province and its geological significance. Mineral Deposits 38, 557–70 (in Chinese with English abstract).Google Scholar
Chen, SC, Ye, HT, Wang, YT, Zhang, XK, Lu, DY and Hu, HB (2014) Re-Os age of molybdenite from the Yuerya Au deposit in eastern Hebei Province and its geological significance. Geology in China 41, 1565–76 (in Chinese with English abstract).Google Scholar
Chiaradia, M, Fontboté, L and Beate, B (2004) Cenozoic continental arc magmatism and associated mineralization in Ecuador. Mineralium Deposita 39, 204–22.CrossRefGoogle Scholar
Chiaradia, M, Ulianov, A, Kouzmanov, K and Beate, B (2012) Why large porphyry Cu deposits like high Sr/Y magmas. Scientific Reports 2, 15.CrossRefGoogle ScholarPubMed
Christopher, MF, Keith, NS and Robert, HR (2003) Detrital zircon analysis of the sedimentary record. Mineralogy and Geochemistry 53, 277303.Google Scholar
Cox, A, Dabiche, MG and Engebretson, DC (1989) Terrane trajectories and plate interactions along continental margins in the north Pacific basin. In The Evolution of the Pacific Ocean Margins, (ed Ben-Avraham, Z.), pp. 2035. New York, NY, Oxford University Press.Google Scholar
Dai, HK, Zheng, JP, Zhou, X and Griffin, WL (2017) Generation of continental adakitic rocks: crystallization modeling with variable bulk partition coefficients. Lithos 272–273, 222–31.CrossRefGoogle Scholar
Davis, GA, Zheng, YD, Wang, C, Darby, BJ, Zhang, C and Gehrels, G (2001) Mesozoic tectonic evolution of the Yanshan fold and thrust belt, with emphasis on Hebei and Liaoning provinces, northern China. In: Hendrix, M.S., Davis, G.A. (Eds.), Paleozoic and Mesozoic Tectonic Evolution of Central Asia: from Continental Assembly to Intracontinental Deformation. Geological Society of America Memoir 194, 171–97.Google 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
Deng, J and Wang, Q (2016) Gold mineralization in China: metallogenic provinces, deposit types and tectonic framework. Gondwana Research 36, 219–74.CrossRefGoogle Scholar
Dressel, BC, Chauvet, A, Trzaskos, B, Biondi, JC, Bruguier, O, Monié, P, Villanova, SN and Newton, JB (2018) The Passa Três lode gold deposit (Paraná State, Brazil): an example of structurally-controlled mineralisation formed during magmatic-hydrothermal transition and hosted within granite. Ore Geology Reviews 102, 701–27.CrossRefGoogle Scholar
Dymek, RF (1983) Titanium, aluminum and interlayer cation substitutions in biotite from high-grade gneisses, West Greenland. American Mineralogist 68, 880–99.Google Scholar
Foley, FV, Pearson, NJ, Rushmer, T, Turner, S and Adam, J (2013) Magmatic evolution and magma mixing of quaternary Adakites at Solander and Little Solander Islands, New Zealand. Journal of Petrology 54, 703–44.CrossRefGoogle Scholar
Foster, MD (1960) Interpretation of the composition of trioctahedral micas. US Geological Survey Professor Paper 354B, 149.Google Scholar
Frost, BR, Barnes, CG, Collins, WJ, Arculus, RJ, Ellis, DJ and Frost, CD (2001) A geochemical classification for granitic rocks. Journal of Petrology 43, 2033–48.CrossRefGoogle Scholar
Goldfarb, RJ and Groves, DI (2011) Orogenic gold: common or evolving fluid and metal sources through time. Lithos 233, 226.CrossRefGoogle Scholar
Goldfarb, RJ, Groves, DI and Gardoll, S (2001) Orogenic gold and geologic time: a global synthesis. Ore Geology Reviews 18, 175.CrossRefGoogle Scholar
Green, TH and Pearson, NJ (1987) An experimental study of Nb and Ta partitioning between Ti-rich minerals and silicate liquids at high pressure and temperature. Geochimica et Cosmochimica Acta 51, 5562.CrossRefGoogle Scholar
Groves, DI and Santosh, M (2016) The giant Jiaodong gold province: the key to a unified model for orogenic gold deposits?. Geoscience Frontiers 7, 409–17.CrossRefGoogle Scholar
Henry, DJ, Guidotti, CV and Thomson, JA (2005)The Ti-saturation surface for low- to- medium pressure metapelitic biotite: implications for geothermometry and Ti-substitution mechanisms. Journal of American Mineralogist 90, 316–28.CrossRefGoogle Scholar
Hoskin, PWO and Ireland, TR (2000) Rare earth element chemistry of zircon and its use as a provenance indicator. Geology 28, 627–30.2.0.CO;2>CrossRefGoogle Scholar
Hronsky, JM, Groves, DI, Loucks, RR and Begg, GC (2012) A unified model for gold mineralization in accretionary orogens and implications for regional–scale exploration targeting methods. Mineralium Deposita 47, 339–58.CrossRefGoogle Scholar
Ishihara, SI (1977) The magnetite-series and Ilmenite-series Granitic Rocks. Mining Geology 27, 293305.Google Scholar
Jan, K and Sylvester, PJ (2003) Present trends and the future of zircon in geochronology: laser ablation ICPMS. Mineralogy and Geochemistry 53, 243–75.Google Scholar
Jensen, EP and Barton, M (2000) Gold deposits related to alkaline magmatism. Reviews in Economic Geology 13, 279314.Google Scholar
Jiang, N, Guo, JH and Chang, GH (2013) Nature and evolution of the lower crust in the eastern North China craton: a review. Earth Science Reviews 122, 19.CrossRefGoogle Scholar
Jiang, SH, Bagas, L, Liu, YF and Zhang, LL (2018) Geochronology and petrogenesis of the granites in Malanyu Anticline in eastern North China Block. Lithos 312, 2137.CrossRefGoogle Scholar
Kay, RW (1978) Aleutian magnesian andesites: melts from subducted Pacific Ocean crust. Journal of Volcanology and Geothermal Research 4, 117–32.CrossRefGoogle Scholar
Kong, DX, Xu, JF, Yin, JW, Chen, JL, Li, J, Guo, Y, Yang, HT and Shao, XK (2015) Electron microprobe analyses of ore minerals and H–O, S isotope geochemistry of the Yuerya gold deposit, eastern Hebei, China: implications for ore genesis and mineralization. Ore Geology Reviews 69, 199216.CrossRefGoogle Scholar
Lang, JR and Baker, T (2001) Intrusion-related gold systems: the present level of understanding. Mineralium Deposita 36, 477–89.CrossRefGoogle Scholar
Le Maitre, RW (2002) Igneous Rocks: A Classification and Glossary of Terms Second. Cambridge, UK: Cambridge University Press, pp. 1236.CrossRefGoogle Scholar
Leal, S, Lima, A, Morris, J, Pedro, M and Noronha, F (2022) Num˜ao gold deposit in the Iberian Variscan belt, Northern Portugal: ore features and mineralization controls. A gold deposit in a W-Sn metallogenic province. Ore Geology Reviews 144, e104815. doi: 10.1016/j.oregeorev.2022.104815.CrossRefGoogle Scholar
Li, H, Li, YJ, Xu, XY, Yang, GX, Wang, ZP, Xu, Q and Ning, WT (2020) Geochemistry, petrogenesis and geological significance of Early Carboniferous adakite in Sawuer region, West Junggar, Xinjiang. Acta Petrologica Sinica 36, 2017–34 (in Chinese with English abstract).Google Scholar
Li, WK, Cheng, YQ and Yang, ZM (2019) Geo–fO2: Integrated software for analysis of magmatic oxygen fugacity. Geochemistry, Geophysics, Geosystems 20, 2542–55.Google Scholar
Li, X, Zhang, C, Behrens, H and Holtz, F (2020) Calculating biotite formula from electron microprobe analysis data using a machine learning method based on principal components regression. Lithos 356–357. doi: 10.1016/j.lithos.2020.105371.Google Scholar
Li, XH, Li, ZX, Li, WX, Wang, XC and Gao, YY (2013) Revisiting the “C-type adakites” of the Lower Yangtze River Belt, central eastern China: in-situ zircon Hf–O isotope and geochemical constraints. Chemical Geology 345: 115.CrossRefGoogle Scholar
Liu, CF, Zhao, SH and Liu, WC (2021) The relationship between gold mineralization, high K calc-alkaline to alkaline volcanic rocks, and A-type granite: formation of the Daxiyingzi gold deposit in northern North China Craton. Ore Geology Reviews 138, 104383. doi: 10.1016/j.oregeorev.2021.104383.CrossRefGoogle Scholar
Liu, J, He, JC, Lai, CK, Wang, XT and Li, TG (2022) Time and Hf isotopic mapping of Mesozoic igneous rocks in the Argun massif, NE China: implication for crustal architecture and its control on polymetallic mineralization. Ore Geology Reviews 141, 104648. doi: 10.1016/j.oregeorev.2021.104648.CrossRefGoogle Scholar
Liu, SA, Li, S, Guo, S, Hou, Z and He, Y (2012) The Cretaceous adakitic– basaltic– granitic magma sequence on south- eastern margin of the North China Craton: implications for lithospheric thinning mechanism. Lithos 134–135, 163–78.CrossRefGoogle Scholar
Liu, YS, Hu, ZC Zong, KQ, Gao, CG and Chen, CC (2010) Reappraisement and refinement of zircon U-Pb isotope and trace element analyses by LA-ICP-MS. Chinese Science Bulletin 55, 1535–46.CrossRefGoogle Scholar
Lu, SN, Zhao, GC, Wang, HC and Hao, GJ (2008) Precambrian metamorphic basement and sedimentary cover of the North China Craton: a review. Precambrian Reviews 160, 7793.CrossRefGoogle Scholar
Ludwig, KR (2003) User’s Manual for Isoplot v. 3.00: A Geochronological Toolkit for Microsoft Excel. Berkeley, CA: Berkeley Geochronology Center, pp. 6061.Google Scholar
Luhr, JF, Carmichael, ISE and Varekamp, JC (1984) The 1982 eruptions of El Chichón volcano, Chiapas, Mexico: Mineralogy and petrology of the anhydrite phenocryst-bearing pumices. Journal of Volcanology & Geothermal Research, 23, 69108.CrossRefGoogle Scholar
Luo, ZK, Guan, K, Qiu, YS and Miao, LC (2001) Zircon SHRIMP U-Pb dating of albite dyke in Jinchangyu gold mine, Jidong area, Hebei, China. Geological Prospecting Theory Cluster 16, 226–31 (in Chinese with English abstract).Google Scholar
Luo, ZK, Li, JJ, Guan, K, Qiu, YS, Qiu, YM, McNaughton, NJ and Groves, DI (2004) SHRIMP zircon U-Pb age of the granite at Baizhangzi gold field in Lingyuan, Liaoning Province. Geological Survey and Research 27, 82–5 (in Chinese with English abstract).Google Scholar
Luo, ZK, Miao, LC, Guan, K, Qiu, YS, Qiu, YM, McNaughton, NJ and Groves, DI (2003) SHRIMP U-Pb zircon dating of the Dushan ganitic batholith and related granite-porphyry dyke, eastern Hebei Province, China, and their geological significance. Geochemica 32, 173–80 (in Chinese with English abstract).Google Scholar
Macpherson, CG, Dreher, ST and Thirwall, MF (2006) Adakites without slab melting: high pressure processing of basaltic island arc magma, indanao, the Philippines. Earth and Planetary Science Letters 243, 581–93.CrossRefGoogle Scholar
Manilar, 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 (1999) Adakitic magmas: modern analogues of Archaean granitoids. Lithos 46, 411–29.CrossRefGoogle Scholar
Miao, L, Qiu, Y, Fan, W and Zhang, F (2008) Mesozoic multi–phase magmatism and gold mineralization in the early Precambrian North China craton, eastern Hebei Province, China: SHRIMP zircon U-Pb evidence. International Geology Review 50, 826–47.CrossRefGoogle Scholar
Miao, LC, Qiu, YM, Fan, WM, Zhang, FQ and Zhai, MG (2005) Geology, geochronology, and tectonic setting of the Jiapigou gold deposits, southern Jilin Province, China. Ore Geology Reviews 26, 137–65.CrossRefGoogle Scholar
N’dri, KA, Zhang, DH, Zhang, T, Tamehe, LS, Kouamelan, KS, Wu, MQ, Assie, KR, Koua, KAD, Kouamelan, AN and Zhang, JL (2021) Gold mineralization in the northern margin of the North China Craton: influence of alkaline magmatism and regional tectonic during Middle Paleozoic-Mesozoic. Ore Geology Reviews 133, e103969. doi: 10.1016/j.oregeorev.2020.103969.Google Scholar
Nachit, H, Ibhi, A, Abia, EH and Ohoud, MB (2005) Discrimination between primary magmatic biotites, re-equilibrated biotites and neoformed biotites. Comptes Rendus Geoscience 337, 1415–20.CrossRefGoogle Scholar
Nie, FJ, Jiang, SH and Liu, Y (2004) Intrusion-related gold deposits of North China Craton, People’s Republic of China. Resource Geology 54, 299324.CrossRefGoogle Scholar
Park, JW, Campbell, LH, Kim, J and Moon, JW (2015) The role of late sulfide saturation in the formation of a Cu and Au-rich magma: insights from the platinum group element geochemistry of Niuatahi-Motutahi Lavas, Tonga Rear Arc. Journal of Petrology 56, 5981.CrossRefGoogle Scholar
Randall, RP and Stephen, RN (2003) Zircon U-Th-Pb Geochronology by isotope dilution—thermal ionization mass spectrometry. Mineralogy and Geochemistry 53, 183213.Google Scholar
Rao, YX (2002) Discussion on some problems related with granite in east section of Yanshan area. Mineral Resources and Geology 16, 327–31 (in Chinese with English abstract).Google Scholar
Richards, JR and Kerrich, R (2007) Special paper: adakite-like rocks: their diverse origins and questionable role in metallogenesis. Economic Geology 102, 537–76.CrossRefGoogle Scholar
Sajona, FG, Maury, RC, Bellon, H, Cotten, J, Defant, MJ, Pubellier, M and Rangin, C (1993) Initiation of subduction and the generation of slab melts in western and eastern Mindanao, Philippines. Geology 21, 1007–10.2.3.CO;2>CrossRefGoogle Scholar
Santosh, M and Groves, DI (2022) Global metallogeny in relation to secular evolution of the earth and supercontinent cycles. Gondwana Research 107, 395422.CrossRefGoogle Scholar
Saunders, AD, Rogers, G, Marriner, GF, Terrell, DJ and Verma, SP (1987) Geochemistry of Cenozoic volcanic rocks, Baja California, Mexico: implications for the petrogenesis of post-subduction magmas. Journal of Volcanology and Geothermal Resource 32, 223–45.CrossRefGoogle Scholar
Sillitoe, RH and Thompson, JFH (1998) Intrusionrelated vein gold deposits: types, tectono-magmatic settings, and difficulties of distinction from orogenic gold deposits. Resource Geology 48, 237–50.CrossRefGoogle Scholar
Smith, JV (1974) Feldspar Minerals. New York: Springer Verlag, pp. 112–45.Google Scholar
Soloviev, SG, Kryazhev, SG, Semenova, DV, Kalinin, YA, Dvurechenskaya, SS and Sidorova, NV (2022) Geology, mineralization, igneous geochemistry, and zircon U-Pb geochronology of the early Paleozoic shoshonite-related Julia skarn deposit, SW Siberia, Russia: toward a diversity of Cu-Au-Mo skarn to porphyry mineralization in the Altai-Sayan orogenic system. Ore Geology Reviews 142, e104706. doi: 10.1016/j.oregeorev.2022.104706.CrossRefGoogle Scholar
Song, Y, Jiang, SH, Bagas, L, Li, C, Hu, JZ, Zhang, Q, Zhou, W and Ding, HY (2016) The geology and geochemistry of Jinchangyu gold deposit, North China Craton: implications for metallogenesis and geodynamic setting. Ore Geology Reviews 73, 313–29.CrossRefGoogle Scholar
Stone, D (2000) Temperature and pressure variations in suites of Archean felsic plutonic rocks, Berensriver area, northwest Superior province, Ontario, Canada. The Canadian Mineralogist 38, 455–70.CrossRefGoogle Scholar
Sun, SS and Donough, WF (1989) Chemical and isotopic systematics of ocean basalts: implications for mantle composition and process. In Magmatism in Ocean Basins (eds Saunders, AD and Norry, MJ), pp. 313–45. London: Geological Society Special Publication, no. 42.Google Scholar
Temizel, I, Arslan, M, Yücel, C, Yazar, EA, Kaygusuz, A and Aslan, Z (2020) Eocene tonalite-granodiorite from the Havza (Samsun) area, northern Turkey: adakite-like melts of lithospheric mantle and crust generated in a post-collisional setting. International Geology Review 62, 1131–58.CrossRefGoogle Scholar
Thompson, JFH, Sillitoe, RH, Baker, T, Lang, JR and Mortensen, JK (1999) Intrusionrelated gold deposits associated with tungsten-tin provinces. Mineralium Deposita 34, 323–34.CrossRefGoogle Scholar
Tuduri, J, Chauvet, A, Barbanson, L, Labriki, M, Dubois, M, Trapy, PH, Lahfid, A, Poujol, M, Melleton, J, Badra, L, Ennaciri, A and Maacha, L (2018) Structural control, magmatic-hydrothermal evolution and formation of hornfels-hosted, intrusion-related gold deposits: insight from the Thaghassa deposit in Eastern Anti-Atlas, Morocco. Ore Geology Reviews 97, 171–98.CrossRefGoogle Scholar
Uchida, E, Endo, S and Makino, M (2007) Relationship between solidification depth of granitic rocks and formation of hydrothermal ore deposits. Resource Geology 57, 4756.CrossRefGoogle Scholar
Wang, C (2018) Study on geological characteristics and metallogenic prognosis of BaiZhangZi gold deposit. Mineral Resources 12, 136–7 (in Chinese with English abstract).Google Scholar
Wang, HQ, Gao, WS, Deng, XD and Li, JW (2020) Zircon U-Pb dating reveals Late Jurassic gold mineralization in the Jidong district of the northern North China Craton. Ore Geology Reviews 126, 103798. doi: 10.1016/j.oregeorev.2020.103798.CrossRefGoogle Scholar
Wang, QD (1989) Research on the mechanism of the transportation and precipitation of gold in Baizhangzi gold deposit. Mining and geology 3, 4953 (in Chinese with English abstract).Google Scholar
Wang, YJ, Wang, XS, Bi, XW, Tao, Y and Lan, TG (2020) Intraplate adakite-like rocks formed by differentiation of mantle-derived mafic magmas: a case study of Eocene intermediate-felsic porphyries in the Machangqing porphyry Cu-Au mining district, SE Tibetan plateau. Journal of Asian Earth Sciences 196, e104364. doi: 10.1016/j.jseaes.2020.104364.CrossRefGoogle Scholar
Wang, YY, Zeng, LS, Chen, FK, Cai, JH, Yan, GH, Hou, KJ and Wang, Q (2017) The evolution of high Ba-Sr granitoid magmatism from“crust-mantle”interaction: a record from the Laoshan and Huyanshan complexes in the central North China Craton. Acta Petrologica Sinica 33, 3873–96 (in Chinese with English abstract).Google Scholar
Watson, EB, Wark, DA and Thomas, JB (2006) Crystallization thermometers for zircon and rutile. Contributions to Mineralogy & Petrology 151, 413–33.CrossRefGoogle Scholar
Wei, Q, Xuan, L, Ma, L, Diao, WY, Liu, ZM, Zheng, YX, Guo, TC and Zhang, JF (2016) Prospecting new progress and significance in Maojiadian-Baizhangzi gold-concentrated area. Non-ferrous Mining and Metallurcy 32, 49 (in Chinese with English abstract).Google Scholar
Wilde, SA, Zhao, GC and Sun, M (2002) Development of the North China Craton during the Late Archaean and its final amalgamation at 1.8 Ga; some speculations on its position within a global Palaeoproterozoic supercontinent. Gondwana Research 5, 8594.CrossRefGoogle Scholar
Wones, DR (1989) Significance of the assemblage titanite +magnetite +quartz in granitic rocks. American Mineralogist 74, 744–9.Google Scholar
Wu, FY, Han, RH, Yang, JH, Wilde, SA and Zhai, MG (2007) Initial constraints on the timing of granitic magmatism in North Korea using U–Pb zircon geochronology. Chemical Geology 238, 232–48.CrossRefGoogle Scholar
Wu, FY, Lin, JQ, Wilde, SA, Zhang, XO and Yang, JH (2005b) Nature and significance of the early Cretaceous giant igneous event in Eastern China. Earth and Planetary Science Letters 233, 103–19.CrossRefGoogle Scholar
Wu, FY, Sun, DY, Ge, WC, Zhang, YB, Grant, ML, Wilde, SA and Jahn, BM (2011) Geochronology of the Phanerozoic granitoids in northeastern China. Journal of Asian Earth Sciences 41, 130.CrossRefGoogle Scholar
Wu, FY, Yang, JH and Liu, XM (2005a) Geochronological framework of the Mesozoic granitic magmatism in the Liaodong Peninsula, Northeast China. Geological Journal of China Universities 11, 305–17 (in Chinese with English abstract).Google Scholar
Xiong, L, Shi, WJ, Li, H, Tian, N, Chen, C, Zhou, HZ, Zhao, SQ and Li, PY (2017) Geochemistry, Sr-Nd-Hf isotopes and petrogenesis of Mid-Late Triassic Baizhangzi granitic intrusive rocks in eastern Hebei-western Liaoning province. Earth Science 42, 207–22 (in Chinese with English abstract).Google Scholar
Xu, DR, Wang, ZL, Wu, CJ, Zhou, YQ, Shan, Q, Hou, MZ, Fu, YR and Zhang, XW (2017) Mesozoic gold mineralization in Hainan Province of South China: genetic types, geological characteristics and geodynamic settings. Journal of Asian Earth Sciences 137, 80108.CrossRefGoogle Scholar
Xu, HJ, Ma, CQ and Zhang, JF (2012) Generation of early Cretaceous high-Mg adakitic host and enclaves by magma mixing, Dabie orogen, Eastern China. Lithos 142–143, 182200.CrossRefGoogle Scholar
Xu, JC, Gu, XX, Wang, JL, Zhang, YM, He, L, Zhou, C and Liu, RP (2018) Zircon U-Pb geochronology, geochemistry and geological significance of granites from the Sanfengshan Cu-Au deposit in Southern Beishan Orogenic Belt, Xinjiang. Bulletin of Mineralogy, Petrology and Geochemistry 37, 294306 (in Chinese with English abstract).Google Scholar
Xu, JW, Cao, Y, Du, YS, Wang, GW and Pang, ZS (2021) Generation and redox state of two episodes of Early Cretaceous granitoid in the Xiaoqinling area, North China Craton: implications for lithosphere thinning and metallogenic potential. Lithos 406–407, e106538. doi: 10.1016/j.lithos.2021.106538.CrossRefGoogle Scholar
Yang, CX and Santosh, M (2020) Ancient deep roots for Mesozoic world-class gold deposits in the north China craton: an integrated genetic perspective. Geoscience Frontiers 11, 203–14.CrossRefGoogle Scholar
Yang, DB, Xu, WL, Zhao, GC, Huo, TF, Shi, JP and Yang, HT (2016). Tectonic implications of Early Cretaceous low- Mg adakitic rocks generated by partial melting of thickened lower continental crust at the southern margin of the central North China Craton. Gondwana Research 38, 220–37.CrossRefGoogle Scholar
Yang, ZM, Chang, ZS, Hou, ZQ and Meffre, S (2016) Age, igneous petrogenesis, and tectonic setting of the Bilihe gold deposit, China, and implications for regional metallogeny. Gondwana Research 34, 296314.CrossRefGoogle Scholar
Ye, H, Zhang, SH, Zhao, Y and Wu, F (2014) Petrogenesis and emplacement deformation of the Late Triassic Dushan composite batholith in the Yanshan fold and thrust belt: implications for the tectonic settings of the northern margin of the North China Craton during the Early Mesozoic. Earth Science Frontiers 21, 275–92 (in Chinese with English abstract).Google Scholar
Zhai, MG and Santosh, M (2011) The early Precambrian odyssey of the North China Craton: a synoptic overview. Gondwana Research 20, 625.CrossRefGoogle Scholar
Zhang, C, Ma, CQ, Holtz, F, Koepke, J, Wolff, PE & Berndt, J (2013) Mineralogical and geochemical constraints on contribution of magma mixing and fractional crystallization to high-Mg adakite-like diorites in eastern Dabie orogen, East China. Lithos 172–173, 118–38.CrossRefGoogle Scholar
Zhang, C, Ma, CQ, Liao, QA, Zhang, JY and She, ZB (2011) Implications of subduction and subduction zone migration of the Paleo-Pacific Plate beneath eastern North China, based on distribution, geochronology, and geochemistry of Late Mesozoic volcanic rocks. International Journal of Earth Science 100, 1665–84.CrossRefGoogle Scholar
Zhang, SH, Zhao, Y, Davis, GA, Ye, H and Wu, F (2014) Temporal and spatial variations of Mesozoic magmatism and deformation in the North China Craton: implications for lithospheric thinning and decratonization. Earth Science Reviews 131, 4987.CrossRefGoogle Scholar
Zhang, SH, Zhao, Y, Liu, JM, Hu, JM, Song, B, Liu, J and Wu, H (2010) Geochronology, geochemistry and tectonic setting of the Late Paleozoic–early Mesozoic magmatism in the northern margin of the North China block: a preliminary review. Acta Petrologica et Mineralogica 29, 824–42 (in Chinese with English abstract).Google Scholar
Zhang, T, Zhang, DH, Danyushevsky, LV, Wu, MQ, Alexis, NK, Liao, YZ and Zhang, JL (2020) Timing of multiple magma events and duration of the hydrothermal system at the Yu’erya gold deposit, eastern Hebei Province, China: constraints from U–Pb and Ar–Ar dating. Ore Geology Reviews 127, e103804. doi: 10.1016/j.oregeorev.2020.103804.CrossRefGoogle Scholar
Zhao, G and Zhai, M (2013) Lithotectonic elements of Precambrian basement in the North China Craton: review and tectonic implications. Gondwana Research 23, 1207–40.CrossRefGoogle Scholar
Zhou, T, Goldfarb, RJ and Phillips, NG (2002) Tectonics and distribution of gold deposits in China – an overview. Mineralium Deposita 37, 249–82.CrossRefGoogle Scholar
Zorin, YA, Zorina, LD, Spiridonov, AM and Rutshtein, IG (2001) Geodynamic setting of gold deposits in Eastern and Central Trans-Baikal (Chita Region, Russia). Ore Geology Reviews 17, 215–32. doi: 10.1016/S0169-1368(00)00015-9.CrossRefGoogle Scholar
Figure 0

Figure 1. (a) Tectonic map of the North China Craton (Jiang et al.2013), CAOB – Central Asian Orogenic Belt, SLS – Solonker suture, QDOB – Qinling-Dabie Orogenic Belt. (b) Sketch geological map of the LX district showing the distribution of gold deposits and igneous rocks with ages (modified from Xiong et al.2017; Zhang et al.2020). (c) Sketch map of BZG at the 30m level with photos of contact relationship. Intrusive rocks: BJ – Baijiadian, DS – Dushan, DSZ – Dashizhuzi, GJ – Gaojiadian, GSZ – Gushanzi, HK – Hekanzi, JG – Jinbaogou, JJ – Jiajiashan, LW – Luowenyu, MJ – Maojiagou, NX – Niuxinshan, QS – Qingshankou, SJ – Sanjia, TZ – Tangzhangzi, XY – Xiaoyingzi, YE – Yu’erya, YZ-Yangzhangzi, ZJ-Zhaojiazhuang.

Figure 1

Figure 2. Photographs of BZGs. (a) and (b) Biotite Monzogranite. (c) Sericitization. (d) Biotite-bearing monzogranite. (e) and (f) Monzogranite. (g) Silicification (quartz metasomatic K- feldspar). (h) Potassium (K- feldspar metasomatic plagioclase). Ser – sericite; Bi – biotite; Kf – K- feldspar; Pl – plagioclase; Q – quartz.

Figure 2

Figure 3. U–Pb concordia diagrams with cathodoluminescence (CL) images of representative zircons analysed and zircon chondrite-normalized REE patterns from the BZGs. (a) and (b) Biotite-bearing monzogranite (sample No. 0-8). (c) and (d) Monzogranite (sample No. -30-7). Chondrite values are from Sun and McDonough (1989).

Figure 3

Figure 4. Classification diagrams of samples from the BZGs. (a) QAP diagram (Le Maitre, 2002). (b) An-Ab-Or diagram (Barker & Millard, 1979). (c) (Na2O +K2O-CaO) vs. SiO2 (Frost et al.2001). (d) A/NK vs. A/CNK diagram (Manilar & Piccoli, 1989).

Figure 4

Figure 5. Chondrite-normalized rare earth element (REE) patterns (a, c, and e) and primitive-mantle-normalized trace element spider diagrams (b, d, and f) are shown. Chondrite values and primitive-mantle values are from Sun and McDonough (1989).

Figure 5

Figure 6. Ternary plots of plagioclase compositions (after Smith, 1974), and the right is a 5× magnification of the projection area of the left.

Figure 6

Figure 7. Correlation diagram of plagioclase compositions.

Figure 7

Figure 8. Genesis (a, after Nachit et al.2005) and classification (b, after Foster, 1960) diagram of biotite.

Figure 8

Figure 9. Variation of ω(SiO2) % vs. ω(TiO2) %, ω(Al2O3) %, ω(FeOT) %, ω(CaO) %, ω(P2O5) %, ω(MgO) %.

Figure 9

Figure 10. Diagrams of (a) TiO2 (wt%) versus Nb/Ta ratios, (b) P2O5 (wt%) versus REE (ppm) (after Wang et al.2017), (c) Sr (ppm) versus Rb (ppm), (d) Sr (ppm) versus Ba (ppm), (e) Eu (ppm) versus Rb/Ba ratios and (F) Ba (ppm) versus Rb/Sr ratios, showing the fractional crystallization of different minerals. See Supplementary Table S7 for the partition of those minerals. Symbols are: Mt = magnetite; Sp = spinel; Hb = hornblende; Pl = plagioclase; Kf = K- feldspar; BI = biotite; Ap = apatite.

Figure 10

Figure 11. (a) Variation of Sr/Y vs. Y. Fields for adakites and island arcs are from Defant and Drummond (1990). (b)Variation of (La/Yb)N vs. YbN. Data are normalized (N) to C1 (chondrite) values of Sun & McDonough (1989). Fields for adakites and island arc magma are from Martin (1999).

Figure 11

Figure 12. Fe2+-Fe3+-Mg2+ ternary diagram (after Wones, 1989) of the biotite from the BZGs.

Supplementary material: File

Zhao et al. supplementary material

Zhao et al. supplementary material

Download Zhao et al. supplementary material(File)
File 550.9 KB