Next Article in Journal
Prediction of Au-Associated Minerals in Eastern Thailand Based on Stream Sediment Geochemical Data Analysis by S-A Multifractal Model
Next Article in Special Issue
Metallogenic Mechanism of Carlin-Type Gold Deposit in Zhen’an-Xunyang Basin, in the South Qinling of China: Constraints of In Situ Trace Elements and S Isotopes from Newly Discovered Wangzhuang Gold Deposit
Previous Article in Journal
Particle Classification in the Enhanced Gravity Field Using the Knelson Concentrator
Previous Article in Special Issue
Two-Stage Superimposed Gold Mineralization in the Xiejiagou Gold Deposit, Shandong Province: Insights from Fluid Inclusions, H-O-S Isotopes, and Trace Elements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ore Mineralogy, Fluid Inclusion Geochemistry, and Zircon U-Pb Geochronology of the Nanmingshui Gold Deposit in East Junggar, Xinjiang, Northwest China: Implications for Ore Genesis

1
School of Earth Sciences and Resources, China University of Geosciences, Beijing 100083, China
2
Xi’an Center of Mineral Resources Survey, China Geological Survey, Xi’an 710100, China
3
State Key Laboratory of Geological Processes and Mineral Resources, China University of Geosciences, Beijing 100083, China
4
Kunming Comprehensive Natural Resources Survey Center, China Geological Survey, Kunming 650111, China
*
Authors to whom correspondence should be addressed.
Minerals 2023, 13(10), 1296; https://doi.org/10.3390/min13101296
Submission received: 13 August 2023 / Revised: 1 October 2023 / Accepted: 3 October 2023 / Published: 6 October 2023

Abstract

:
The Nanmingshui gold deposit, located in the eastern segment of the Kalamaili gold belt (KGB), is hosted by the sub-greenschist facies rocks of the Lower Carboniferous Jiangbasitao Formation. The genesis of this deposit, however, has been debated for decades because of controversial constraints on the P-T-X conditions and origins of hydrothermal fluid and mineralization age. In this study, we present gold-bearing sulfide compositions, fluid inclusions, H-O isotopes, and the results of hydrothermal zircon U-Pb dating to provide new insights into the genesis of the gold deposit. Three gold mineralization stages are recognized: quartz–pyrite–minor native gold veins (early), quartz–tourmaline–arsenopyrite–pyrite–gold–polymetallic sulfide veins (middle), and quartz–calcite veinlets (late). Gold predominantly occurs as native gold with high fineness ranging from 941 to 944 in sulfides and quartz, and some as solid solutions (Au+) within the lattice of pyrite and arsenopyrite. Three types of primary fluid inclusions are identified in hydrothermal quartz: CO2-H2O (C-type), aqueous (W-type), and pure CO2 (PC-type) inclusions. The early-stage quartz mainly contains C-type and minor W-type inclusions, with total homogenization temperatures (Th) of 220–339 °C, salinities of 0.4–3.7 wt.% NaCl eqv., and bulk densities of 0.66–1.01 g/cm3. All three types of inclusions are observed in the middle-stage quartz, of which the C- and W-type inclusions yield Th values of 190–361 °C, with salinities of 0.4–6.0 wt.% NaCl eqv. and bulk densities of 0.69–0.99 g/cm3. The late-stage quartz contains only W-type inclusions that have lower Th values of 172–287 °C, higher salinities of 1.4–6.9 wt.% NaCl eqv., and bulk densities of 0.79–0.95 g/cm3. Trapping pressures estimated from C-type inclusions in the early and middle stages cluster at 280–340 MPa and 220–310 MPa, respectively, corresponding to metallogenic depths of 10–13 km and 8–11 km. The H-O isotopic compositions (δ18Owater = 1.8–10.9‰, δD = −99 to −62.9‰) and microthermometric data indicate that the ore-forming fluids belong to medium–high-temperature, low-salinity, medium-density, and CO2-rich-H2O-NaCl ± CH4 ± N2 systems, probably originating from metamorphic water. Fluid immiscibility is a crucial mechanism for gold precipitation. Additionally, the U-Pb dating of hydrothermal zircons, from the auriferous quartz–tourmaline vein, yield a weighted mean 206Pb/238U age of 314.6 ± 9.6 Ma. Taking all of the above, the Nanmingshui deposit can be reasonably classed as a typical mesozonal orogenic gold deposit in the KGB, which was formed in a Late Carboniferous tectonic transition from syn-collision between the Jiangjunmiao accretionary complex and Yemaquan arc to post-collision in the East Junggar Orogen. Our results serve to better understand the gold mineralization and genesis of the Late Paleozoic orogenic system in the Kalamaili area, Xinjiang.

1. Introduction

Orogenic gold deposits, a distinctive class of epigenetic mineral deposits, are currently the major source of global gold production [1,2]. They typically form at a 5 to 18 km depth under greenschist–amphibolite facies conditions, via deformation-induced fluid circulation preferentially driven along ductile–brittle transcurrent structures deeply rooted into the lithosphere at Archean-to-Phanerozoic-convergent tectonic margins [1,3]. Since the term “Orogenic gold deposits” was proposed by Groves et al. [3], numerous world-class and even several giant orogenic gold deposits have been reported worldwide, e.g., the largest Archean Golden Mile and Mt Charlotte gold deposits at Kalgoorlie in the Yilgarn Craton, west Australia [4]; the Obuasi, Damang, and Bogoso–Prestea gold deposits in the Paleoproterozoic Ashanti greenstone belt, southwest Ghana [5,6]; the Gara, Yalea, and Gounkoto gold deposits in the world-class Loulo mining district, Mali [7,8]; and the Jiaojia, Sanshandao, and Xincheng gold deposits in the Jiaodong Peninsula, China [9,10,11]. As one of the principal types of gold deposits in the metamorphic belt, orogenic gold deposits have been relatively well documented in tectonic and geodynamic settings, geology, ore fluids and isotopic chemistry, and genetic models [12,13,14,15,16]. Nevertheless, there remains an ongoing and spirited debate on the fluid sources of orogenic gold deposits, and various models have been proposed, such as the shallow crustal meteoric-fluid [17], magmatic–hydrothermal [18,19], and supracrustal metamorphic models [20]. The latter emphasizes a metamorphic fluid source resulting from the metamorphic devolatilization of supracrustal rocks under greenschist- to amphibolite-facies conditions and is widely accepted [1]. However, significant magmatic fluids are now input to explain the enrichment of tellurides in individual orogenic gold deposits [19,21]. Essentially, the spatiotemporal and genetic associations between gold mineralization and magmatism in an orogeny are still subjects in the study of orogenic gold deposits.
The Central Asian Orogenic Belt (CAOB) or Altaids represents the complex evolution of the Neoproterozoic–Paleozoic orogenic belt, which is situated between the Siberian craton to the north and the Tarim and North China (Sino-Korean) craton to the south (Figure 1A) [22]. Based on metallogenic models applied for either accretionary orogeny or collisional orogeny, the CAOB could contain orogenic deposits formed during the processes of oceanic plate subduction, terrane accretion, and continent–continent collision. Indeed, numerous world-class super-large and large orogenic gold deposits have been discovered in the Southern Tianshan area, the Chinese Tianshan area, Altay, and the West Junggar Orogen, e.g., the Muruntau (5246 t Au), Kumtor (570 t Au), Sawayaerdun (>100 t Au), and Hatu (56 t Au) gold deposits [23,24,25]. As an important part of the CAOB, however, orogenic deposits are rarely reported in the East Junggar Orogen, North Xinjiang, even though a large number of lode gold deposits and occurrences have been detected and worked by local farmers in the Kalamaili Gold Belt (KGB) in past decades [26]. The majority of these lode deposits, represented by the Shuangquan (24 t Au), Jinshuiquan (10 t Au), Nanmingshui (10 t Au), and Sujiquandong (7 t Au) gold deposits, are confined to a narrow KGB bounded by the NW-NWW-trending Qingshui–Sujiquan and Kalamaili faults (Figure 1C), and are similar to those of Phanerozoic gold deposits in geology and geochemistry [27]. Most studies in this area mainly focused on Paleozoic tectonic evolution, post-collisional magmatism, and intra-continental deformation [28,29,30]. To date, systematic studies on ore geology, mineralogy, fluids and isotope geochemistry, and geochronology are still scarce or remain equivocal in the literature, which limits the understanding of the genetic types and metallogenic regularities of the orogenic gold system in the Kalamaili area, East Junggar.
The Nanmingshui gold deposit, situated in the eastern part of the KGB, is a typical medium-sized gold deposit [31], and was discovered in the 1980s with an average grade of 6.1 g/t. Owing to its complex geological features and mineral associations, nearly all published data on this deposit are contentious. Two different views have emerged among geologists regarding the genetic types and metallogenic ages of the Nanmingshui gold deposit. Some advocate a magmatic–hydrothermal model, which is linked to contemporaneous granitic magmatism in the Early Carboniferous (337.5 ± 3.9 Ma) [32]. On the other hand, others propose a metamorphic–hydrothermal model associated with the tectonic transition that occurred in the Kalamaili area during the Late Carboniferous (312.9 ± 4.3 Ma) [23,27]. In this paper, we report new data obtained from field geology, electron microprobe analysis (EMPA) of gold-bearing sulfides, fluid inclusions, H-O isotopic compositions, and the LA-ICP-MS U-Pb dating of hydrothermal zircon in auriferous quartz–tourmaline veins from the Nanmingshui gold deposit. The aims of this study are to examine and define the characteristics and origins of ore fluids, the mechanisms and timing of gold mineralization, and ore genesis in the Nanmingshui gold deposit. Our results provide insights to better understand the gold mineralization and genesis of the Late Paleozoic orogenic gold system in Kalamaili.
Figure 1. (A) Sketch map of the Central Asian Orogenic Belt, showing the location of the study area (modified from [33]); (B) simplified regional geological map of the Junggar terrane in north Xinjiang (modified from [34,35]); (C) simplified regional geological map of the Kalamaili area, East Junggar (modified from [28]). Abbreviations: LYQ, Laoyaquan; BLKDK, Beilekuduke; HYS, Huangyangshan.
Figure 1. (A) Sketch map of the Central Asian Orogenic Belt, showing the location of the study area (modified from [33]); (B) simplified regional geological map of the Junggar terrane in north Xinjiang (modified from [34,35]); (C) simplified regional geological map of the Kalamaili area, East Junggar (modified from [28]). Abbreviations: LYQ, Laoyaquan; BLKDK, Beilekuduke; HYS, Huangyangshan.
Minerals 13 01296 g001

2. Regional Geology

The East Junggar Orogen, Xinjiang, NW China, lying at the junction zone of the Siberian and Kazakhstan–Junggar plates, is part of the CAOB and sandwiched between the Zaisan-Erqis (Irtysh) fault to the north and the Kalamaili fault to the south. The tectonic units in this region consist, from north to south, of the Dulate–Baytag arc, Yemaquan arc, and Jiangjunmiao accretionary complex (Figure 1B), which are separated from each other by two NW-SE-trending highly deformed and dismembered ophiolite belts (Zhaheba-Aermantai and Kalamaili) [34]. The Kalamili area lies on the southwestern edge of the Paleozoic East Junggar Orogen, which incorporates three distinct tectonic units from the Late Paleozoic, namely, from north to south, the Yemaquan arc, Kalamaili ophiolite belt (KOB; so-called Kalamaili shear zone or Kalamaili fault zone), and Jiangjunmiao accretionary complex (Figure 1C).
North of the KOB, the Yemaquan arc mainly comprises Devonian to Carboniferous volcanic and volcanosedimentary rocks, and Late Paleozoic post-collisional Alkaline granite plutons [29]. However, several Early Paleozoic intermediate-acid intrusive rocks have been reported in the eastern part of the Yemaquan arc, with zircon LA-ICP-MS U-Pb ages ranging from 427.4 to 418.5 Ma [36]. In the south of the Yemaquan arc, a number of A-type monzogranites, biotite granites, alkali granites, and alkali–feldspar granites are distributed along north of the KOB, the largest of which is the Laoyaquan–Huangyangshan batholith. The zircon U-Pb dating of these plutons, including the Laoyaquan, Beilekuduke, Huangyangshan, and Sujiquan granites, have ages between 311 and 283 Ma, which reveals the end of orogeny and the beginning of intraplate evolution in the Late Carboniferous in the Kalamali area, East Junggar [37,38,39]. South of the KOB, the Jiangjunmiao accretionary complex is mainly composed of a well-developed succession of Late Carboniferous clastic sedimentary rocks, and sporadic marine–terrestrial volcanic and carbonate rocks from the Late Silurian.
The NWW-trending KOB, approximately 5–15 km wide, is exposed over a length of nearly 150 km in the Kamst–Hongliugou–Nanmingshui areas to the west and the Zhifang–Aletunkunduo areas to the east [40]. It is primarily composed of fragmented ophiolite mélanges and a heavily deformed volcanic–sedimentary matrix. The rock lithology includes serpentinite, serpentinized peridotite, lherzolite, diabase, gabbro, basalt, plagiogranite, chert, and mafic lava [41], indicating fragments of a paleo-oceanic crust. Recent works gave wide ophiolite ages ranging from 416.7 to 329.9 Ma, with zircon LA-ICP-MS U-Pb ages of gabbro (406 ± 1.8 Ma [42]; 329.9 ± 1.6 Ma [43]) and diabase (416.7 ± 3.2 Ma [44]), and zircon SHRIMP U-Pb ages of plagiogranite (373 ± 10 Ma [45]; 351 ± 6 Ma [46]). The tuff in the Heishantou Formation, unconformably lying over ophiolite mélanges, yielded a zircon LA-ICP-MS U-Pb age of 343.0 ± 5.0 Ma [44], which constrained the upper limit of the emplacement age of the Kalamaili ophiolite. Therefore, the Kalamaili ophiolite, representing one suture between the Kazakhstan–Junggar and Siberian plates, were formed in the period between the Early Devonian and the Early Carboniferous [40].
The KGB investigated in our research is located in the KOB, south of the East Junggar Orogen. Actually, it is an Au-Cu polymetallic belt bounded by the NW-NWW-trending Qingshui–Sujiquan fault to the north and the Kalamaili fault to the south (Figure 1C). The Paleozoic lithologies in this area consist of fore-arc sedimentary rocks of the Devonian to Early Carboniferous, cherts on the ocean floor of the Early Carboniferous, and ophiolite mélanges [28]. From bottom to top, the dominant strata in the KGB include the Middle Devonian Beitashan and Yundukala Formations, and the Lower Carboniferous Jiangbasitao Formation. All of these strata strike northwest and dip to the northeast, and were metamorphosed to zeolite to sub-greenschist facies. As one of two backbone structures, the Kalamaili fault is a high-angle thrust fault that dips to the northeast with angles of 75–80°. From east to west, the spatial distribution of this fault is transferred from NWW- to NW-trending and intermittently extends ~300 km with a width of 100–300 m (Figure 1C) [47]. At the northern boundary of the KGB, the Qingshui–Sujiquan fault originates from Qingshuiquan in the west and extends eastwards to the south of Liushuquan via Sujiquan. This NW-trending thrust fault strikes 280–300° and dips to the north with angles of 60–80°, and exceeds 100 km in length and nearly 100 m in width.
More than 100 lode gold deposits, occurrences, and a few placer gold deposits occur in the KGB, most of which (e.g., the Shuangquan, Jinshuiquan, and Nanmingshui lode gold deposits) stretch along the northeastern side of the Kalamaili fault for approximately 400 km [23]. Almost all auriferous quartz ± tourmaline veins are controlled by the subsidiary NW-NWW- and/or NE-NEE-trending ductile–brittle faults of the Kalamaili shear zone.

3. Deposit Geology

The Nanmingshui gold deposit is located in the eastern segment of the KGB (Figure 1C), several kilometers south of the Qingshui–Sujiquan fault. The exposed lithologies in the mining district are lower-grade metamorphic rocks of the Jiangbasitao Formation that are intensely sheared and foliated, including tuffs, tuffaceous sandstone, basalt, and basaltic andesite (Figure 2). There are 19 auriferous quartz veins identified by the Xinjiang Geology and Mineral Bureau, with average gold grades ranging from 4.12 to 24.19 g/t, which are mined via open pitting and underground operations [31]. Compared with other lode gold deposits in the KGB, tourmaline occurs as a major or minor gangue mineral in most of the gold-bearing veins at Nanmingshui.

3.1. Features of the Ore-Hosting Faults

The spatial distribution of auriferous quartz veins at Nanmingshui is strictly controlled by the NW- to NWW-trending Kalamaili shear zone and its subsidiary ductile–brittle reverse fault systems (Figure 2). On the basis of different orientations, three groups of the subsidiary faults can be identified in the deposit. The first group, consisting of the NW- to NWW-trending faults, strikes 275–330° and dips to the north or southwest with angles of 42–85° (Figure 3A–C). These faults are nearly parallel to the regional faults of the KGB, which are generally 40–400 m in length and 0.3–4.0 m in width. It is noteworthy that the majority of auriferous quartz–tourmaline veins at Nanmingshui occur in reverse faults. The faults in the second group strike 40–75° and dip to the north, south, or east with angles of 35–82°, few of which are exposed in the deposit, with 42–115 m in length and 0.2–6.0 m in width. Only five gold-bearing veins exist in these NE-NEE-trending faults. Both of the groups above belong to pre-metallogenic structures. The nearly EW-trending faults (F1, F2, and F3; Figure 2) of the third group are relatively small in size, and are all post-metallogenic and have no effects on the gold-bearing veins occurring in other faults.

3.2. Characteristics of the Orebodies

The dominant mineralization in the Nanmingshui gold deposit consists of auriferous quartz veins, quartz–tourmaline veins, and subordinately altered rocks (Figure 3A–F). The orebodies typically occur as veins and veinlets, which are lenticular and agglomerated in shape. Among the 19 gold-bearing quartz ± tourmaline veins, most of them (e.g., Nos. 1, 7, 8, 9, and 13) strike 275–310° and dip to the north or south with variable angles of 21–85° (mostly 40–75°), elongating by 5–145 m with a thickness of 0.15–1.50 m. Other veins (including Nos. 4, 10, 12, and 18) strike 40–70° and dip to the north or south with angles of 42–80°, extending by 5–62 m with a thickness of 0.10–0.45 m (Figure 2). The largest orebody is at the curved part of No. 9, and is mainly composed of quartz–tourmaline veins and altered tuffaceous sandstone (Figure 3B,E). The orebody is 34 m long with an average width of 2 m and a thickness of 1.83 m, and its ore grades vary from 1.55 to 26.24 g/t (mean of 5.40 g/t).
The gold grades of different veins in the Nanmingshui gold deposit vary significantly and mostly range between 4 and 15 g/t. The majority of the orebodies are hosted in high-angle (mainly 40–75°) reverse faults with ductile–brittle deformation (Figure 3A,C). However, all high-grade orebodies are concentrated at sites of structural bends, strike changes, and splays, indicating an important role of vein interconnections for fluid flow and gold precipitation at Nanmingshui [48]. Brecciated and lamellar textures can be found in some quartz–tourmaline veins (Figure 3F,K), which probably show that abrupt fluid-pressure fluctuations from a fault-valve behavior occurred in the process of gold mineralization [49,50].

3.3. Ore Types and Mineral Assemblages

According to the characteristics of mineralization and mineral assemblages, three types of gold ores can be detected in the Nanmingshui gold deposit, including quartz-pyrite ore (Figure 3G), high-grade quartz–tourmaline–gold–polymetallic sulfide ore (Figure 3H), and low-grade altered wall-rock ore (Figure 3I).
The total contents of metallic minerals in different gold ores are commonly less than 5%. The major ore minerals are arsenopyrite and pyrite, while other subordinate species include native gold, chalcopyrite, tetrahedrite, and hematite, containing minor pyrrhotite, galena, and limonite (Figure 4A–I). The non-metallic minerals consist primarily of quartz, tourmaline, sericite, and calcite, with minor amounts of muscovite, chlorite, and epidote (Figure 4J–L). On the basis of morphology, texture, and paragenesis, two generations of pyrite are identified. The early generation of pyrite (Py1) is dominated by micro-grained (0.01–0.10 mm in diameter), euhedral grains exhibiting cubic forms, which occur in isolation in arsenopyrite (Figure 4D,E) or in aggregations in milky quartz veins (Figure 4I). The late generation of pyrite (Py2) mainly consists of fine- to medium-grained (0.2–1.0 mm in diameter) euhedral to subhedral grains, most of which have pentagonal–dodecahedral forms and coexist with arsenopyrite in smoky-gray quartz–tourmaline veins (Figure 4A,B). Many grains of Py2 and coexisting arsenopyrite are fractured or brecciated, showing foam and cataclastic textures with micro-fissures (Figure 4E,F). These textures provide highly favorable conditions for the occurrence of native gold and gold minerals.
The ores are mainly characterized by disseminated, vein-stockwork, brecciated, or veinlet structures (Figure 3G–K). Minerals occur as euhedral to subhedral or anhedral crystals. Metasomatic, cataclastic, and pressure-shadow textures are commonly observed, indicating that the Nanmingshui gold deposit was formed as a result of hydrothermal replacement (Figure 4).

3.4. Mineralization Stages and Wall-Rock Alteration

Based on the crosscutting relationships and mineral assemblages, gold mineralization can be divided into three stages. These are the quartz–pyrite–minor native gold (early) stage, the quartz–tourmaline–arsenopyrite–pyrite–gold–polymetallic sulfide (middle or main) stage, and the quartz–calcite (late) stage. The paragenetic sequence is illustrated in Figure 5.
The early-stage veins primarily consist of quartz, pyrite, and minor native gold, with widths ranging from less than 10 cm to 2 m. Quartz commonly coexists with py1 and minor fissure-filled native gold in the milky or whitish veins (Figure 3G), which are always intersected by the auriferous quartz–tourmaline veins (Figure 3J). In the middle-stage veins, hydrothermal minerals dominantly include quartz, tourmaline, native gold, arsenopyrite, and pyrite, with minor chalcopyrite, tetrahedrite, pyrrhotite, and galena. Quartz of this stage is typically smoky gray in color, and occurs as veins with a width of 0.2–2.0 m. It is commonly coherent with euhedral to subhedral arsenopyrite and pentagonal-dodecahedral pyrite, as well as native gold, chalcopyrite, tetrahedrite, and other sulfides. Tourmaline is a major or minor gangue mineral in gold-bearing veins, and occurs in the form of acicular crystals and veinlets and is visible in hand specimens (Figure 3H). Polymetallic sulfides generally occur as dissemination (Figure 3I) or veinlets (Figure 4I), while the native gold is exclusively confined to granular inclusions and fissure-filled gold (Figure 4C) in the auriferous veins. Late-stage veinlets mainly consist of white quartz and calcite, without any gold minerals observed, and are generally 2–5 cm in width. These barren quartz–calcite veinlets are always filled in the fractures of altered wall-rocks (Figure 3L) and crosscut the earlier gold-bearing veins.
The extensive wall-rock alteration at Nanmingshui is characterized by the hydrothermal minerals of quartz, sulfides, sericite, calcite, chlorite, and epidote in the alteration zone. Based on our field investigations, an intense lateral alteration is symmetrically distributed at the point of contact between gold-bearing veins and wall-rocks. From the center to the outside, the alteration zonation is as follows: (1) auriferous quartz–tourmaline-polymetallic sulfide veins; (2) altered wall-rocks with silicification, arsenopyritization, pyritization, and sericitization that occur in the proximal zone (approximately 5–20 m in width); (3) altered wall-rocks with alteration types of sericitization, carbonatization, chloritization, and rare epidotization in the distal zone; and (4) unaltered wall-rocks on the outermost side.

4. Samples and Analytical Methods

In this study, more than 50 representative samples were collected from field outcrops and mine dumps at Nanmingshui. Samples for electron probe microanalysis (EPMA), hydrogen and oxygen (H-O) isotopes, and zircon U-Pb dating were collected from the middle-stage quartz–tourmaline–gold–polymetallic sulfide veins. Others for the fluid inclusion study were selected from the auriferous and barren veins of the three different stages.

4.1. EPMA and Mapping

Twelve samples were double-side-polished as thin sections of 0.1 mm thickness for optical microscopy and EPMA. Backscattered electron (BSE) images were taken and the chemical compositions of the minerals (native gold and sulfides) were determined using a JEOL JXA-8230 Electron Probe Micro-Analyzer under the operation conditions of 15 kV, 20 nA, a beam size of 2 μm in diameter, and count times of 10−40 s (peak) and 5−20 s (backgrounds) at the Shandong Analysis Center of the China Metallurgical Geology Bureau (SAC-CMGB). The characteristic X-rays used were Fe(Kα), S (Kα), As (Lα), Au (Mα), Ag (Lα), Ni (Kα), Co (Kα), Cu (Kα), Zn (Kα), Sb (Lα), Se (Lα), Te (Lα), Cd (Lα), Pb (Mα), Hg (Mα), and Bi (Mα). The standard materials for each element were as follows: native metal standards for Au, Ag, Ni, Cd, and Bi; pyrite (FeS2) for Fe and S; cuprite (Cu2O) for Cu; galena (PbS) for Pb; sphalerite (ZnS) for Zn; bismuth selenide (Bi2Se3) for Se; modderite (CoAs) for Co and As; and stibnite (Sb2S3) for Sb. The analysis yielded an accuracy of 1–5%. The average detection limits obtained in this study were 184 ppm for Fe, 98 ppm for S, 142 ppm for As, 157 ppm for Au, 79 ppm for Ag, 116 ppm for Ni, 133 ppm for Co, 266 ppm for Cu, 185 ppm for Zn, 101 ppm for Sb, 113 ppm for Se, 115 pm for Te, 149 ppm for Cd, 304 ppm for Pb, and 252 ppm for Bi. Matrix effects were corrected using the ZAF (Atomic-Number–Absorption–Fluorescence) correction program.
Trace element mapping was performed at the Research Center of Genetic Mineralogy, China University of Geosciences, Beijing (RCGM-CUGB), using Tescan MIRA3-XMU scanning electron microscope (SEM) equipped with an X-ray energy-dispersive spectrometer (EDS) of Oxford INCA Energy X-Max 20 . The EDS X-ray elemental maps were collected at 20 kV with a working distance of 15 mm, a probe current of 83 μA, and a dwell time of 500 μs for acquisition at each point. Maps are up to 1024 × 1024 pixels in size.

4.2. Fluid Inclusion Study

Double-polished thin sections (0.2 mm thick) were made from more than 20 quartz samples for petrographic observation, which were used to identify the genesis, types, and vapor/liquid ratios of the fluid inclusions (FIs). Fourteen representative samples were selected for microthermometry and laser Raman spectroscopy analyses.
Microthermometric measurements of the FIs were performed using a Linkam THMSG-600 heating/freezing stage (from −196 to +600 °C) at the Ore Deposit Geochemistry Microanalysis Laboratory, State Key Laboratory of Geological Processes and Mineral Resources, China University of Geosciences, Beijing (ODGML-SKL-CUGB). The stage was calibrated by measuring the triple point of pure CO2 (−56.6 °C) and the freezing point of pure water (0 °C) in synthetic FIs. The estimated accuracy of the measurement was about ±0.1 °C during the cooling runs and ±2 °C during the heating runs, relative to the standard material. The heating and cooling rates were generally less than 20 °C/min, and reduced to 0.2 °C/min near the phase transitions. FIs were first quickly cooled to/below −100 °C, and then, heated progressively to obtain the temperatures of the phase transitions, including the melting temperature of solid CO2 (Tm-CO2), the temperature of final ice melting (Tm-ice), the temperature of CO2-clathrate dissociation (Tm-cla), the partial homogenization temperature of the CO2 phase (Th-CO2), and the total homogenization temperature (Th). The salinities of aqueous FIs were calculated using the equations for the NaCl-H2O system [51]. The W-ρ-X-P parameters (salinity–density–mole fraction–pressure) were calculated from microthermometric data using the Flincor program [52,53].
The gas compositions in a single fluid inclusion were analyzed using a LabRAM HR800 Raman spectrometer (Horiba Jobin Yvon) at the Analytical Laboratory of the Beijing Research Institute of Uranium Geology (BRIUG). The detecting conditions were as follows: a tunable argon ion laser with a wavelength of 532 nm, laser power up to 50 mW, a spectral resolution of 1.5 cm−1, and a spatial resolution of 1 μm. The spectral range fell between 100 and 4300 cm−1 for the analysis of CO2, CH4, N2, and H2O in the vapor and liquid phases.

4.3. H-O Isotope Analyses

The H-O isotope compositions of FIs in auriferous quartz were analyzed using a Finnigan MAT253-type mass spectrometer at the Analytical Laboratory of BRIUG, following the methods in reference [54]. Oxygen gas was liberated from the samples via a quantitative reaction with BrF5 in externally heated nickel vessels. Hydrogen was measured by reacting the H2O extracted from the FIs with zinc at 550 °C. The Standard Mean Ocean Water (SMOW) was adopted as the standard for oxygen and hydrogen, with total uncertainties of ±0.2‰ for δ18O and ±1‰ for δD. The δ18Owater values were calculated using the fractionation formula in reference [55].

4.4. Zircon U-Pb Dating

Approximately 20 kg fresh sample was collected from the auriferous quartz–tourmaline vein for zircon U-Pb dating. Zircon grains were separated using conventional crushing, heavy liquid, and magnetic separation techniques, and then, hand-picked under a binocular microscope. They were mounted in epoxy resin and polished to approximately half thickness. Representative zircon grains were prepared for cathodoluminescence (CL) investigations to expose the internal textures and guide U-Pb dating.
Zircon U-Pb dating and rare earth element (REE) analyses were conducted using a Thermo Fisher’s X-series II ICP-MS instrument connected to a Coherent GeoLasPro-193 nm laser (LA-ICP-MS) at the ODGML-SKL-CUGB. Helium was applied as the carrier gas and mixed with argon prior to entering the ICP-MS torch. The diameter of the laser spot was 32 μm with a 6 Hz laser pulse repetition rate. Zircon 91500 and NIST610 were used as the external standards for the U-Th-Pb isotopic ratios and rare earth element concentration, respectively. Meanwhile, zircon Plešovice was used as a monitoring standard for each analysis [56]. Time-dependent drifts of U-Th-Pb isotopic ratios were corrected using linear interpolation over time for every five analyses, according to the variations in the 91500 standard. For each analysis, 20 s of background was acquired for each analysis followed by 50 s of sample signal. Quantitative calibration for rare earth element concentrations and zircon U-Pb ages was carried out using ICPMSDataCal 10.7 software [57]. Data reduction and Concordia diagrams were processed using the ISOPLOT 4.15 program [58].

5. Results

5.1. Chemical Compositions of Native Gold and Sulfides

The EPMA data and mapping of ore minerals from the Nanmingshui gold deposit are shown in Table 1 and Figure 6, respectively. The gold and silver contents at the three points of gold grains are 89.35–90.51 wt.% (mean of 89.82 wt.%) and 2.81–3.08 wt.% (mean of 2.95 wt.%), respectively, similar to the chemical compositions of native gold (Au > 80 wt.%) [59]. The calculated gold fineness, expressed as 1000 Au/(Au + Ag + Hg), has a narrow range from 941 to 944 (mean of 943). However, high Hg (up to 2.56 wt.%) and low Cu (0.03–0.06 wt.%) contents of native gold are typically consistent with hydrothermal α-phase Au-Ag-Hg alloys [60]. The average contents of the trace elements Fe, S, and Bi in these gold minerals are 0.77 wt.%, 0.01 wt.%, and 0.97 wt.%, respectively, whereas the contents of other minor elements (As, Pb, Sb, Zn, Ni, Co, and Te) are rarely detected or below the detection limit.
The EDS maps of most elements, such as Fe, S, As, Au, Ag, Co, and Sb, in arsenopyrite display a nearly homogeneous distribution, with depletion in Au, Ag, and Sb and relative enrichment in Fe, S, As, and Co (Figure 6). Both the major and trace element compositions are slightly variable in arsenopyrite. Compared to the theoretical compositions (34.30 wt.% Fe, 46.01 wt.% As, and 19.69 wt.% S), eight arsenopyrite grains are intensely depleted in As and enriched in Fe and S, with the As, Fe, and S contents in the ranges of 39.84–41.38 wt.% (mean of 40.45 wt.%), 35.53–36.45 wt.% (mean of 35.93 wt.%), and 21.21–22.39 wt.% (mean of 22.02 wt.%), respectively. Only minor amounts of trace elements such as Se (0.20–0.27 wt.%) and Co (0.03–0.06 wt.%) are present in the analyzed arsenopyrite grains. However, the majority of the contents of other elements, such as Au, Ag, Cu, Pb, Sb, Zn, and Ni, vary from below the limit of detection (bdl) to less than 0.03 wt.%. The calculated formula of arsenopyrite is Fe1.03As0.86S1.10.
In the EDS maps of Py2, most elements (e.g., Fe, S, Au, Ag, and Sb) show a relatively homogeneous distribution, whereas As and Pb are slightly richer in the rims than in the cores. The Pb and S maps indicate that some galena inclusions occur in Py2 (Figure 6). For Py2, the Fe and S contents are 45.64–46.66 wt.% (mean of 46.05 wt.%) and 52.34–53.68 wt.% (mean of 52.78 wt.%), respectively. The As contents in these pyrites vary between 0.51 and 1.96 wt.% with an average of 1.21 wt.%, whereas the corresponding Au contents vary from bdl to 0.08 wt.%, which are similar to the compositions of gold-bearing arsenian pyrite [61]. The contents of Pb (bdl to 0.12 wt.%), Co (bdl to 0.15 wt.%), and Ni (bdl to 0.28 wt.%) in Py2 are higher than the detection limit and vary greatly, but the contents of other trace elements (Se, Cd, Ag, Cu, Sb, and Zn) are relatively low (Figure 6). The calculated formula of Py2 (Fe0.99S1.98) is near the theoretical composition (Table 1).
The contents of major elements in chalcopyrite vary more greatly than those in other sulfides, with Cu, Fe, and S contents of 32.06–35.93 wt.% (mean of 34.45 wt.%), 27.27–31.05 wt.% (mean of 29.24 wt.%), and 33.54–34.73 wt.% (mean of 34.30 wt.%). The contents of trace elements are below 0.05 wt.%, except for Pb (bdl to 0.10 wt.%), Sb (bdl to 0.12 wt.%), Zn (0.01 to 1.52 wt.%), and Co (0.04 to 0.07 wt.%). In particular, the Au contents at the 10 analyzed points range from bdl to 0.05 wt.%. Based on the EPMA data (Table 1), the calculated chalcopyrite formula is Cu1.01Fe0.98S2.00, which is the same as the theoretical composition.

5.2. Fluid Inclusion Geochemistry

5.2.1. Petrography and Classification

The FIs in quartz at the three stages of the veins at Nanmingshui are relatively large in size (3–22 μm) and rich in numbers. The vast majority are primary inclusions; however, pseudosecondary FIs are visible along the healing fractures that do not cut across the quartz crystal boundaries. Through phase relationships at room temperature (25 °C), phase transitions during heating and cooling, and compositional data from the Raman spectroscopic analysis, three types of FIs were identified: H2O-CO2 or aqueous–carbonic (C-type), aqueous (W-type), and pure CO2 or carbonic (PC-type) FIs (Figure 7), in decreasing order of their abundance.
C-type (H2O-CO2-NaCl ± CH4 ± N2) is the most predominant FI in the early- and middle-stage quartz, accounting for as much as 60% of the total FIs. These FIs consist of two (liquid CO2 + liquid H2O) or three phases (vapor CO2 + liquid CO2 + liquid H2O) at room temperature, and comprise CO2-H2O-NaCl FIs (type Ia) with the CO2 phase occupying 20–45% (mostly 25–40%) of the total inclusion volume (Figure 7A,D–G) and CO2-rich inclusions (type Ib) with the CO2 phase accounting for 50–80 vol.% (Figure 7H,I). When the temperature is cooled down to 20 °C or below during the freezing run, vapor CO2 can also be observed in two phases of the C-type FIs. C-type inclusions usually occur as single isolated inclusions or as clusters along the growth zones of quartz grains, most of which have negative crystal, ellipsoidal, and a variety of polygonal or rare irregular shapes with sizes ranging from 4 to 20 μm. The coexistence of C- and PC- or W-type FIs can often be observed in the same crystal (Figure 7A,E,F).
W-type (H2O-NaCl) FIs are commonly captured in quartz of all three stages of veins, accounting for 30% of the total FI population, but their abundance at the early, middle, and late stages shows an increasing trend. Especially in the quartz crystals of late stage, it is the only type of FI that can be identified. W-type inclusions usually appear as two-phase (liquid H2O + vapor H2O) aqueous FIs without a visible CO2 phase (Figure 7C) at room temperature. The vapor phase occupies  15–70 vol.% (mostly 15–30 vol.%) of the total inclusion volume, but some inclusions occur as monophase (liquid H2O or vapor H2O) aqueous inclusions individually (Figure 7B). The majority of them occur in isolation or as clusters or trails within quartz crystals, and generally show negative crystal or subrounded-oval shapes with sizes of 3–15 μm. Sometimes, W-type FIs coexist with C- or PC-type inclusions in quartz crystals of the early and middle stages (Figure 7A,F).
PC-type FIs can only be observed in quartz of the early and middle stages and account for less than 10% of the total FIs. All of these FIs consist of almost pure carbonic fluid at room temperature, including two-phase CO2 (vapor CO2 + liquid CO2) and monophase CO2 (liquid CO2). Only when the temperature is cooled down to 15 °C can vapor CO2 in the monophase CO2 of PC-type FIs be visible. They generally appear as isolated singles and show oval, negative crystal, or irregular shapes with sizes varying from 4 to 8 μm. PC-type always coexist with C-type FIs in the same field of petrographic view (Figure 7A,E), and rarely with the W-type FIs (Figure 7B).

5.2.2. Microthermometry

The microthermometric data of all three types of FIs from the early-, middle-, and late-stage veins at Nanmingshui are shown in Table 2 and Figure 8 and Figure 9.
Early stage: C-type FIs with a small number of W-type inclusions were observed as primary inclusions in the early-stage veins. The C-type FIs freeze to a solid phase below −100 °C during cooling runs. Their melting temperatures of solid CO2 (Tm-CO2) occur between −58.2 and −57.4 °C (mean of −57.8 °C), lower than the triple point of pure CO2 (−56.6 °C; Figure 9A), implying the presence of minor additional gases (mostly like CH4 and/or N2). The melting temperatures of CO2 clathrate (Tm-cla) are between 8.9 and 9.8 °C (mean of 9.5 °C; Figure 9B), and these values are representative of lower salinities varying from 0.4 to 2.2 wt.% NaCl eqv. (mean of 1.1 wt.% NaCl eqv.). The partial homogenization temperatures of the CO2 phases (Th-CO2) range from 18.8 to 29.9 °C (mostly between 22.0 and 24.0 °C and with a mean of 23.6 °C; Figure 9C) and all of them are homogenized to liquid CO2. The C-type inclusions are totally homogenized (Th) to aqueous phases at temperatures from 257 to 339 °C, peaking at 280–320 °C (mean of 298 °C). However, a few inclusions decrepitate prior to final homogenization at 238–334 °C. The calculated molar fractions of CO2 (XCO2) range from 0.11 to 0.31 (mean of 0.15), while the densities of the CO2 (ρCO2) phase range from 0.60 to 0.79 g/cm3 (mean of 0.73 g/cm3) and the bulk densities (ρtotal) range from 0.80 to 1.01 g/cm3 (mean of 0.91 g/cm3). The microthermometric analyses of W-type FIs show that the final ice-melting temperatures (Tm-ice) occur between −2.2 and −1.5 °C (mean of −1.9 °C), equivalent to salinities of 2.6 to 3.7 wt.% NaCl eqv. (mean of 3.2 wt.% NaCl eqv.). The total homogenization of W-type inclusions (to liquid H2O) shows a wide temperature range from 220 to 338 °C, with the majority clustering at 260–320 °C (mean of 276 °C). Their ρtotal values are within the range of 0.66–0.87 g/cm3 (mean of 0.77 g/cm3).
Middle stage: All three types of FIs were identified in the middle-stage quartz. For the most abundant C-type FIs, the Tm-CO2 values during the heating runs range from −62.2 to −56.0 °C (mean of −58.3 °C; Figure 9A), and are slightly below the triple point of pure CO2. The Tm-cla values of C-type inclusions are in a narrow range of 7.5–9.8 °C (mean of 9.0 °C; Figure 9B), corresponding to salinities between 0.4 and 4.8 wt.% NaCl eqv. (mean of 2.0 wt.% NaCl eqv.). The CO2 phases homogenize to a liquid at temperatures of 12.9–28.8 °C (mainly between 18.0 and 26.0 °C; Figure 9C). C-type inclusions in the middle-stage quartz contain a ρCO2 of 0.66–0.84 g/cm3 (mean of 0.74 g/cm3) and a ρtotal of 0.82–0.99 g/cm3 (mean of 0.91 g/cm3), with XCO2 ranging from 0.07 to 0.58 (mean of 0.16). These inclusions homogenize into both liquid and vapor states between 209 and 361 °C, with most clustering at 260–300 °C (mean of 277 °C). W-type inclusions are less abundant than C-type and yield Tm-ice values between −3.7 and −0.5 °C (mean of −2.2 °C), corresponding to salinities of 0.9−6.0 wt.% NaCl eqv. (mean of 3.6 wt.% NaCl eqv.) and a ρtotal of 0.69–0.91 g/cm3 (mean of 0.82 g/cm3). These inclusions finally homogenize into the liquid or vapor phases (due to the volume percentage of vapor H2O) with a temperature of 190–349 °C, clustering at 200–260 °C (mean of 249 °C). Occurring in very few numbers, the PC-type inclusions are generally observed coexisting with C-type inclusions. Their Tm-CO2 values range from −61.5 to −57.2 °C (mean of −59.0 °C; Figure 9A), indicating CH4 and/or N2 in the CO2 phases as additional gas species. The Th-CO2 (to liquid phase) values are between 12.3 °C and 19.0 °C (mean of 16.1 °C; Figure 9C) with a calculated ρCO2 of 0.80–0.86 g/cm3 (mean of 0.82 g/cm3).
Late stage: Only the W-type inclusions were observed in late-stage quartz crystals. The Tm-ice values, ranging from −4.3 to −0.8 °C (mean of −1.8 °C), are indicative of a narrow salinity range of 1.4–6.9 wt.% NaCl eqv. (mean of 3.0 wt.% NaCl eqv.). The W-type FIs are all homogenized into the liquid phase at temperatures of 172–287 °C (mean of 213 °C) with the majority clustering at 180–220 °C. The calculated ρtotal values range from 0.79 to 0.95 g/cm3 (mean of 0.88 g/cm3).

5.2.3. Metallogenic Pressure and Depth

C-type FIs are common in the gold-bearing quartz veins of the early and middle stages. The ore-forming fluids at Nanmingshui are approximately medium–high-temperature, low-salinity (<7.0 wt.% NaCl eqv.), and CO2-rich-H2O-NaCl ± CH4 ± N2 systems. Therefore, we chose C-type FIs to estimate trapping pressure by using the Flincor program for the CO2-H2O-NaCl system [53]. The results are shown in Figure 10. Based on the microthermometric data and estimated volume fractions of the CO2 phase, the calculated trapping pressures of the C-type inclusions are 208–367 MPa in the early stage and 206–392 MPa in the milled stage, with the majority clustering at 280–340 MPa and 220–310 MPa (Figure 10), respectively. Given 2.7 g/cm3 as an average density of the upper crust rocks, the metallogenic depths of the Nanmingshui gold deposit are estimated to be 10–13 km in the early stage and 8–11 km in the middle stage, assuming lithostatic conditions. Because Au mineralization mainly occurred in the middle stage, it can be concluded that the Nanmingshui gold deposit was formed at depth of ca. 8–11 km. These paleodepths are nearly consistent with the brittle–ductile transition zone, and are the favored depths (typically 5–18 km) of the most orogenic gold deposits in the world [3,62].
When neglecting the tectonic overpressure, the gold orebodies at Nanminghui experienced a minimum 3 km uplift from the early to middle stages, accompanied by a simultaneous reduction in trapping pressures of approximately 30–60 MPa. This phenomenon is common in orogenic deposits, and can be interpreted as the promotion of seismic activity related to uplift and exhumation during orogenic processes [49,63].

5.2.4. Laser Raman Spectroscopy

Representative FIs were measured using Raman spectroscopy to reveal their gas compositions. The analysis of C-type (including type Ia and Ib) inclusions from early- and middle-stage quartz shows that the vapor or liquid CO2 phases are dominated by CO2, but contain minor amounts of CH4 and/or N2 (Figure 11A–D), which are consistent with the Tm-CO2 being lower than −56.6 °C in the microthermometric process. Although no CO2 phase can be observed during the cooling–heating runs, the vapor bubbles of W-type FIs in middle-stage quartz still contain minor amounts of additional CO2 and CH4 other than H2O (Figure 11E). However, the existence of CO2 in late-stage W-type inclusions was not detected by either the microthermometric measurement or the laser Raman spectroscopic analysis (Figure 11F), implying that CO2-based gas loss or phase separation occurred during the evolution of ore-forming fluid systems from CO2-rich aqueous solution in the early and middle stages to CO2-poor aqueous solution in the late stage. The petrographic evidence of PC-type inclusions coexisting with W-type inclusions can also strongly support this fact (Figure 7A,B).

5.3. H-O Isotopic Compositions

The δ18O values of the middle-stage auriferous quartz show restricted variation between 13.3 and 19.7‰ with an average value of 14.9‰ (Table 3). Based on the equation of oxygen isotopic fractionation between quartz and water (1000lnαquartz-water = 3.38 × 106 T−2–3.4) [55], the calculated fluid δ18Owater values at Nanmingshui vary from 1.8 to 10.9‰ (mean of 6.1‰). Except for sample L1506, almost all the δ18Owater and δD values (−99.0 to −62.9‰) of ore-forming fluids in the deposit are consistent with those of gold deposits (Jinshuiquan, Shuangquan, Sujiquandong, etc.) in the KGB (δ18Owater = 4–12‰, δD = −100 to −60‰) [26,31,64].

5.4. Zircon U-Pb Geochronology

The LA-ICP-MS analytical results of U-Pb dating and REE concentrations for 30 zircon grains separated from the No. 3 auriferous quartz–tourmaline vein are listed in Tables S1 and S2 (see the Supplementary Materials), and illustrated in Figure 12 and Figure 13. According to the crystal morphology, internal structures, and dating results, two distinct types of zircons are recognized.
Type-1 zircons are generally beige-to-yellowish in color and rounded-to-subrounded in shape, and most are euhedral-to-subhedral crystals with diameters of 40–110 μm and aspect ratios of 1:1–3:1. More significantly, zircon crystals usually have bright CL responses and exhibit oscillatory zoning (Figure 12A), similar to igneous zircons crystallized during a single petrogenetic event [65]. The Th and U contents of type-1 zircons vary between 28–504 ppm and 48–840 ppm, respectively, with Th/U ratios of 0.13–1.17 (mostly >0.4). Their chondrite-normalized REE patterns from La to Lu are characterized by heavy (H) REE enrichment ((Lu/Sm)N = 47–433) and light (L) REE depletion ((Sm/La)N = 18–1572) with pronounced positive Ce anomalies (δCe = 4.77–182.43) and negative Eu anomalies (δEu = 0.03–0.65; Figure 13A), which are typical for magmatic zircon [66]. Furthermore, all data points of the type-1 zircons are plotted within or near the magmatic field in Figure 13B. The LA-ICP-MS U-Pb ages of 28 type-1 zircons fall into two populations (Figure 12A,B). One cluster of 13 analyzed points yields 206Pb/238U ages ranging from 426 to 410 Ma with a weighted mean of 416.6 ± 3.7 Ma (MSWD = 0.79, Figure 12C). The other cluster of nine analyzed points yields 206Pb/238U ages ranging from 404 to 385 Ma with a weighted mean of 396.3 ± 4.2 Ma (MSWD = 1.13, Figure 12D). These two weighted mean ages can be interpreted as xenocrysts incorporated from the Kalamaili ophiolite [44]. In addition, three analyses have a weighted mean 206Pb/238U age of 500 ± 11 Ma (MSWD = 0.56, Figure 12B), which is consistent with the early SHRIMP zircon U-Pb dates for plagiogranite (497 ± 12 Ma) from the KOB. Two zircon grains yield a weighted mean 206Pb/238U age of 962 ± 19 Ma (MSWD = 0.34, Figure 12A). One grain gives a Neoarchean 207Pb/206Pb age of 2611 ± 59 Ma (Figure 12 A), suggesting that it is a zircon xenocryst from the hidden Archean crust (ca. 3.8–2.55 Ga) beneath the Junggar terrane [67].
Two grains of type-2 zircon are smaller in size, with diameters of 58–62 μm and aspect ratios of 1:1–3:2. In the CL images, both have no or extremely low CL responses and show internally structureless or ambiguous oscillatory zoning (Figure 12A). Type-2 zircons have relatively higher Th contents (100–272 ppm, mean of 186 ppm) and lower U contents (171–248 ppm, mean of 210 ppm) than type-1 zircons (mean of 129 ppm Th and 304 ppm U), with Th/U ratios of 0.58–1.09. The lower total REE (∑REE = 827–1177 ppm vs. 415–4142 ppm for type-1 zircons) and HREE concentrations (HREE = 753–989 ppm vs. 405–4078 ppm for igneous zircons) and higher LREE concentrations (LREE = 74–188 ppm vs. 6–64 ppm for type-1 ones) of type-2 zircons, combined with their flatter or less steep chondrite-normalized REE patterns ((Sm/La)N = 0.62–0.92 and δCe = 2.37–2.85; Figure 13A), are obviously distinct from those of type-1 magmatic zircons. In particular, in the (Sm/La)N versus La discriminant diagram [68,69], type-2 zircons are plotted around the hydrothermal zircon field (Figure 13B), suggesting a hydrothermal origin. Two hydrothermal zircons yield 206Pb/238U ages that vary from 313 to 316 Ma with a Late Carboniferous weighted mean age of 314.6 ± 9.6 Ma (MSWD = 0.071, Figure 12B).
Figure 13. (A) Chondrite-normalized REE patterns (normalization values from [70]) and (B) discriminant diagrams of (Sm/La)N versus La for zircons from the auriferous quartz–tourmaline vein in the Nanmingshui gold deposit (after [68,69]).
Figure 13. (A) Chondrite-normalized REE patterns (normalization values from [70]) and (B) discriminant diagrams of (Sm/La)N versus La for zircons from the auriferous quartz–tourmaline vein in the Nanmingshui gold deposit (after [68,69]).
Minerals 13 01296 g013

6. Discussion

6.1. Gold Occurrences in Pyrite and Arsenopyrite

The presence of gold in natural ores, generally occurring as discrete gold minerals (native gold, Au-Ag tellurides, aurostibite, maldonite, auricupride, etc.) and submicroscopic or invisible (so-called refractory) gold [71,72], is a potential indicator for identifying Au-bearing deposit types [73]. It is also crucial for the recovery of gold during mineral processing [74]. In a great variety of gold deposit types (orogenic, Carlin, epithermal, intrusion-related, etc.), either native gold or invisible gold is principally hosted by (arsenian) pyrite and arsenopyrite. In particular, the concentrations of refractory gold can vary from a few tens of ppb to more than 1.0 wt.% of its hosts [61,75,76].
In the Nanmingshui gold deposit, four kinds of occurrence states for native gold (sizes of 1–60 μm) are observed in high-grade ores, including microscopic inclusions in arsenopyrite (60%, Figure 14A), intergranular grains in contact with the boundary of pyrite and arsenopyrite (30%, Figure 14B), elongated fissure-filled grains within the pseudomorphs of hematite after pyrite (8%, Figure 14C), and less common disseminated or isolated grains in quartz (2%, Figure 14D). The high gold fineness (941–944) and Hg contents (2.42–2.56 wt.%, mean of 2.50 wt.%) of native gold from the Nanmingshui gold deposit, together with gold textures, are almost consistent with the orogenic gold deposits [62,73]. On the basis of the EPMA data in this study (Table 1), both arsenian pyrite2 and arsenopyrite incorporate minor amounts of refractory gold, with Au values ranging from bdl to 0.08 wt.% (mean of 0.02 wt.%, n = 10) and bdl to 0.01 wt.% (mean of 0.004 wt.%, n = 8), respectively. The results of EDS mapping reveal that the gold-hosted minerals have homogeneous distributions of gold without any Au-rich points. Combined with the Au/As molar ratios of 0–0.015 (mostly 0–0.004) in arsenian pyrite and arsenopyrite falling below the maximum solid-solution Au lines (Figure 15) [75,77], it can be inferred that invisible gold may occur as solid solutions (Au+) in the crystal lattice [78] rather than as discrete microparticles and nanoparticles (Au0) [79].

6.2. Nature and Evolution of Ore-forming Fluids

The measured δ18O values of auriferous quartz from the Nanmingshui gold deposit show restricted variations between 13.3 and 19.7‰ (mean of 14.9‰; Table 3), which are typical of orogenic gold deposits from the Neoarchean to Cenozoic (δ18O = 10–18‰) [80,81,82]. Except for individual δ18Owater values, the majority fall into the range of most lode gold deposits (δ18Owater = 5–11‰) [12,13,83]. The δD values range from –99.0 to −62.9‰ with an average of -83.3‰ and overlap with those of the Juneau gold belts, Alaska (Figure 16) [84]. In Figure 16, all values are plotted away from the meteoric water line and below or in the metamorphic and primary magmatic H2O boxes, suggesting that the metallogenetic fluids are likely to be derived from metamorphic or magmatic H2O rather than meteoric water. Furthermore, the δ18Owater values of samples NMS4-30 and NMS7-7 are both 10.9‰ (Table 3), which are obviously higher than the maximum δ18O value (9‰) of magmatic water [85] and the δ18O values (5.4–10.35‰) of A-type granites in the north of the KOB [86]. Considering the geological fact that there are no reported granitic intrusions at the mining area and the fluid P-T-X conditions (Table 2), the ore-forming fluids at Nanmingshui probably originated from the metamorphic devolatilization of δ18O-enriched rocks instead of being granitic or meteoric H2O in origin.
In cases where gold particles in or around the FIs and temporal relationships between the ore and gangue minerals are not observed, relying on the inclusion data to constrain the geological environment or metallogenic processes remains inconclusive [88,89]. However, fluid inclusions, at least the primary FIs in the earliest-stage minerals, are potential indicators for the genesis of hydrothermal mineral systems [90]. Hydrothermal quartz at the Nanmingshui deposit generally coexists with gold-bearing minerals (Figure 14) and contains various types of FIs, which can be used to trace the nature and evolution of ore-forming fluids.
As shown in Figure 8, the Th values of FIs decrease from 220–339 °C (peaking at 260–320 °C) during the early stage to 190–361 °C (peaking at 220–280 °C) during middle stage, and to 172–287 °C (peaking at 180–220 °C) during late stage. Meanwhile, the salinities of the FIs increase slightly from 0.4–3.7 wt.% NaCl eqv. in the early stage to 0.4–6.0 wt.% NaCl eqv. in the middle stage, and to 1.4–6.9 wt.% NaCl eqv. in the late stage. The occurrence of numerous C- and W-type FIs in the early-stage quartz indicates that the initial fluids are most likely low-salinity and carbonic–aqueous. Furthermore, the coexistence of the C-, W-, and PC-type FIs with variable salinities and divergent homogenization behaviors implies that fluid immiscibility occurred in the early and middle stages, whereas the late-stage quartz only contains W-type inclusions without any CO2 detected, suggesting a CO2-poor aqueous solution. Thus, from the early to late stages, it can be concluded that the ore fluids evolved from CO2-rich to CO2-poor solutions through fluid immiscibility and CO2 escape.
In conclusion, the ore-forming fluids at the Nanmingshui deposit are characterized by medium–high-temperature, low-salinity (<7.0 wt.% NaCl eqv.), and CO2-rich-H2O-NaCl ± CH4 ± N2 systems with metamorphic water at the origin, similar to those of the orogenic gold deposits around the world [1,3,62].

6.3. Fluid Immiscibility and Gold Deposition

Fluid immiscibility is a process frequently invoked as the cause for gold deposition in many mesothermal lode gold deposits [89,91]. Most mesothermal veins typically formed below the levels at which water could boil, and the presence of CO2 in FIs allowed fluid immiscibility to occur at much greater depths [92]. Following the criteria outlined by Ramboz et al. [93], we found that fluid immiscibility can be a crucial mechanism for gold deposition in the Nanmingshui gold deposit. Our evidence is as follows: (1) The PC-type and/or C-type Ib, and W-type FIs, in many cases, coexist in the same clusters or trails of the same quartz crystal (Figure 7A,B,G). The petrographic, microthermometric, and laser Raman data collected from these inclusions imply that they are accidentally trapped in two end-member fluids (a CO2-rich fluid and a H2O-rich aqueous solution), resulting from the unmixing of a low-salinity H2O-CO2 parent fluid [94]. (2) The coexistence of these three-types of inclusions (PC-type, C-type Ib, and W-type) with rare C-type Ia inclusions, containing highly variable XCO2 concentrations (0.07–0.58; Table 2), can be explained by the heterogeneous entrapment of the carbonic and aqueous end-member fluids [8,91]. Heterogeneous trapping represents simple but solid evidence for fluid immiscibility. (3) The Th values for C-type inclusions are commonly higher than those for cogenetic W-type inclusions (Figure 7, Figure 17). This phenomenon may be due to the heterogeneous trapping of a small amount of H2O-rich fluid within CO2-rich inclusions [88,95], which is also consistent with fluid immiscibility [96]. (4) Compared with C-type inclusions in the same stage, W-type FIs tend to have higher salinities (Figure 17), as the salt is preferentially fractionated into the aqueous phase rather than the CO2-rich phase [92]. The relatively high salinity of H2O-rich fluid can be produced via unmixing of the low-salinity H2O-CO2 parent fluid. (5) Despite no aqueous phases being observed in PC-type inclusions at room temperature, the small size of these inclusions may contain microscopically undetectable H2O (up to 10–20 vol.%) [97,98], implying that the occurrence of PC-type FIs may be due to phase separation from a CO2-H2O-NaCl fluid.
The fluid immiscibility at Nanmingshui is likely to be induced by sudden drops in fluid pressure, rather than the gradual cooling of ore fluids. This process is strongly supported by the 30–60 MPa decreases in trapping pressure from the early to middle stages (Figure 10) and the partial overlapping of Th values between each stage (Figure 17).
On the basis of previous experiments on Au+ complexing in aqueous sulfide solutions [99,100], gold was probably transported as a bisulfide complex Au(HS)2 in lode-gold-only deposits like Nanmingshui [101]. Thus, the gold concentrations can be related to those of the sulfide species via the following: Au + H2S + HS = Au(HS)2 + 0.5H2, H2S = HS + H+. Two possible mechanisms can promote the deposition of gold at Nanmingshui by breaking down the bisulfide complexes. First, phase separation in the immiscible fluids can strongly fractionate H2S into the vapor phase, resulting in a pH increase in the liquid [27]. Second, the abundant CO2 and other additional volatiles (such as CH4 and N2) may play a critical role [102]. CO2 occurs as a pH buffer to maintain elevated concentrations of gold in ore fluids, and can be removed via phase separation, leading to the destruction of CO2 buffering pH [103]. Moreover, the existence of CH4 and N2 can push the solvus to higher P-T conditions, making phase separation easier [104].

6.4. Timing of Gold Mineralization

Zircon (ZrSiO4) is commonly considered to be an ideal mineral for geochronological studies, as a result of its stable physicochemical properties and high closure temperature (greater than 900 °C) for the U-Th-Pb isotopic system [105]. Hydrothermal zircon U-Pb geochronology has been widely used as a direct dating method to determine the metallogenic ages of distinct hydrothermal deposits, such as Archean massive sulfide deposits [106], orogenic Au [107,108,109], Fe skarn [110], Nb-Ta-U-REE [111,112,113], and intrusion-related W-Sn-Mo-Bi-Au deposits [114,115]. It has been demonstrated that the crystallization and growth of hydrothermal zircon are common in low-temperature environments, with P-T conditions of 260–380 °C and 7–200 MPa [107,108,115,116,117,118].
The P-T conditions above are similar to the ranges of the main ore fluid (220–280 °C and 220–310 MPa) in the Nanmingshui gold deposit, implying that hydrothermal zircon can crystallize from this mesothermal fluid. Indeed, two zircon grains from the auriferous quartz–tourmaline vein yield 206Pb/238U ages of 313–316 Ma, with a weighted mean age of 314.6 ± 9.6 Ma. Their CL images, U-Th-Pb concentrations, and REE compositions are more consistent with those of hydrothermal zircons from the Jack Hills, Australia, than those of magmatic zircons captured from the Archean crust and Kalamaili ophiolite (Table S1; Figure 13) [68]. Thus, the Late Carboniferous age of ca. 315 Ma can be recognized as the period of gold mineralization at Nanmingshui.
Besides the U-Pb dating results of hydrothermal zircon, three geological constrains may contribute to determining the mineralization age of the gold deposit: (1) The gold-bearing quartz veins are hosted in the rocks of the Upper Carboniferous Jiangbasitao Formation, whose zircon U-Pb ages are 339–336 Ma, giving a lower limit (336 Ma) of gold mineralization [119]. Both auriferous veins and wall-rocks were strongly deformed via foliation and mylonitization, while the A-type granitic intrusions (311–283 Ma) along the north of the KOB were not subjected to regional deformation and metamorphism [26]. Thus, the metallogenic age of the Nanmingshui gold deposit is reasonably restrained at 336–311 Ma. (2) Taking the latest research together, after the closure (ca. 340–330 Ma) of the Kalamaili Ocean and the collision (ca. 330–320 Ma) between the Jiangjunmiao accretionary complex and Yemaquan arc, a tectonic transition from syn-collisional compression to post-collisional extension (320–300 Ma) occurred during the Late Carboniferous [30,120]. This transition of tectono-thermal regimes is conducive to driving the fluid in deep and forming the hydrothermal deposits [121]. (3) Many gold deposits in the KGB were formed between 318 and 310 Ma, such as the Qingshui No. 48 (quartz Rb-Sr isochron age of 311 ± 46 Ma) [122], Huangyangshanxi (zircon U-Pb ages of 318–310 Ma) [123], and Jinshuiquan (hydrothermal zircon U-Pb age of 314.9 ± 3.2 Ma) [124] gold deposits, indicating that at least an essential hydrothermal event of gold mineralization occurred during the Late Carboniferous.
These facts correspond well with the hydrothermal zircon U-Pb ages in this study. Therefore, it is reasonable to deduce that the Nanmingshui gold deposit was formed in the Late Carboniferous (ca. 315 Ma) rather than Early Carboniferous (337.5 ± 3.9 Ma) [32], which marked a transitional setting from the syn-collision between the Jiangjunmiao accretionary complex and Yemaquan arc to post-collision in the Kalamaili area, East Junggar.

6.5. Ore Genetic Type

Many of the geological and geochemical features exhibited by Nanmingshui are similar to those of typical orogenic gold deposits [7,8,62,109,125]: (1) The deposit was formed in the Late Carboniferous, with an age of ca. 315 Ma, which was in a transitional period after the collision between the Jiangjunmiao accretionary complex and the Yemaquan arc in the East Junggar Orogen. (2) The gold-bearing veins, mainly controlled by the NW- to NWW-trending brittle–ductile faults, are hosted in the sub-greenschist facies rocks of the Lower Carboniferous Jiangbasitao Formation. (3) Metallic minerals in ores consist of arsenopyrite and (arsenian) pyrite and minor pyrrhotite, chalcopyrite, sphalerite, galena, tetrahedrite, and native gold, with total contents < 5%. Most of the gold occurs as native gold in sulfides and quartz, with high gold fineness of 941–944. (4) Extensive wall-rock alteration and lateral zoning can be observed. The alteration types primarily consist of silicification, arsenopyritization, pyritization, sericitization, carbonatization, chloritization, and epidotization. (5) The main ore fluid belongs to a mesothermal (220–280 °C), low-salinity (<7.0 wt.% NaCl eqv.), and CO2-rich-H2O-NaCl ± CH4 ± N2 system, with calculated pressures of 220–310 MPa and estimated depths of 8–11 km. (6) The δ18Owater (1.8–10.9‰) and δD values (−99.0 to −62.9‰) for the metallogenic fluids were probably derived from the metamorphic H2O. (7) Fluid immiscibility is one of the most important mechanisms for gold deposition.
Taking all of the above together, here, we propose that the Nanmingshui gold deposit can be classed as an orogenic gold deposit as defined by Groves et al. [3].

7. Conclusions

In this paper, a systemic study of mineralogy, geochemistry, and geochronology was conducted, aiming to constrain the gold mineralization and genesis of the Nanmingshui gold deposit in the KGB, Xinjiang, NW China. Our research draws the following conclusions:
(1) Three metallogenic stages are recognized at Nanmingshui, marked by the mineral assemblages of early-stage quartz–pyrite–minor native gold veins, middle-stage quartz–tourmaline–arsenopyrite–pyrite–gold–polymetallic sulfide veins, and late-stage quartz–calcite veinlets. Gold in ores mainly occurs as native gold with high fineness (941–944) in sulfides and quartz, and some as solid solutions (Au+) in the lattice of pyrite and arsenopyrite.
(2) The ore-forming fluids at Nanmingshui are charactered by medium–high-temperature, low-salinity, medium-density, and CO2-rich-H2O-NaCl ± CH4 ± N2 systems, probably derived from metamorphic water. Fluid immiscibility is a key mechanism for gold precipitation.
(3) The Nanmingshui deposit is a typical mesozonal (metallogenic depths of ca. 8–11 km) orogenic gold deposit in the KGB. It was formed in a Late Carboniferous tectonic transition from syn-collision between the Jiangjunmiao accretionary complex and Yemaquan arc to post-collision in the East Junggar Orogen.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min13101296/s1, Table S1: LA-ICP-MS U-Pb data of zircons from the auriferous quartz–tourmaline vein in the Nanmingshui gold deposit; Table S2: REE compositions (ppm) of zircons from the auriferous quartz–tourmaline vein in the Nanmingshui gold deposit.

Author Contributions

Conceptualization, Z.G., X.G. and Y.Z. (Yongmei Zhang); field investigation, Z.G., X.G., Y.Z. (Yongmei Zhang) and W.C.; experiments, Z.G. and W.C.; data analysis, Z.G., X.G., Y.Z. (Yongmei Zhang), D.H., X.Z., L.W. and Y.Z. (Yanrong, Zheng); writing—original draft, Z.G.; writing—review and editing, Z.G., X.G., Y.Z. (Yongmei Zhang), C.M. and M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 42130804), the Natural Science Basic Research Program of Shaanxi (No. 2023-JC-QN-0284), and the Geological Survey projects of the China Geological Survey (Nos. ZD20220306, ZD2023040308, and DD20230377).

Data Availability Statement

Not applicable.

Acknowledgments

We give special thanks to Ruiping Liu from the Beijing Research Institute of Uranium Geology for her laboratory help, as well as Xiaoyong Nie for his contributions to the experiments and Figure 11. We also thank the anonymous reviewers for their constructive comments and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Groves, D.I.; Santosh, M.; Deng, J.; Wang, Q.F.; Yang, L.Q.; Zhang, L. A holistic model for the origin of orogenic gold deposits and its implications for exploration. Miner. Depos. 2020, 55, 275–292. [Google Scholar] [CrossRef]
  2. Hou, Z.Q.; Wang, Q.F.; Zhang, H.J.; Yu, N.; Wang, R.; Groves, D.I.; Zheng, Y.C.; Han, S.C.; Gao, L.; Yang, L. Lithosphere architecture characterized by crust-mantle decoupling controls the formation of orogenic gold deposits. Natl. Sci. Rev. 2023, 10, nwac257. [Google Scholar] [CrossRef] [PubMed]
  3. Groves, D.I.; Goldfarb, R.J.; Gebre-Mariam, M.; Hagemann, S.G.; Robert, F. Orogenic gold deposits: A proposed classification in the context of their crustal distribution and relationship to other gold deposit types. Ore Geol. Rev. 1998, 13, 7–27. [Google Scholar] [CrossRef]
  4. Mueller, A.G.; Hagemann, S.G. A thematic issue dedicated to the giant Golden Mile Au-Te deposit, Kalgoorlie, Western Australia. Miner. Depos. 2020, 55, 597–600. [Google Scholar] [CrossRef]
  5. Chudasama, B.; Porwal, A.; Kreuzer, O.P.; Butera, K. Geology, geodynamics and orogenic gold prospectivity modelling of the Paleoproterozoic Kumasi Basin, Ghana, West Africa. Ore Geol. Rev. 2016, 78, 692–711. [Google Scholar] [CrossRef]
  6. White, A.J.R.; Waters, D.J.; Robb, L.J. Exhumation-driven devolatilization as a fluid source for orogenic gold mineralization at the Damang deposit, Ghana. Econ. Geol. 2015, 110, 1009–1025. [Google Scholar] [CrossRef]
  7. Lawrence, D.M.; Treloar, P.J.; Rankin, A.H.; Harbidge, P.; Holliday, J. The geology and mineralogy of the Loulo mining district, West Africa: Evidence for two distinct styles of orogenic gold mineralization. Econ. Geol. 2013, 108, 199–227. [Google Scholar] [CrossRef]
  8. Lawrence, D.M.; Treloar, P.J.; Rankin, A.H.; Boyce, A.; Harbidge, P. A fluid inclusion and stable isotope study at the Loulo Mining District, Mali, West Africa: Implications for multifluid sources in the generation of orogenic gold deposits. Econ. Geol. 2013, 108, 229–257. [Google Scholar] [CrossRef]
  9. Qiu, Y.M.; Groves, D.I.; McNaughton, N.J.; Wang, L.G.; Zhou, T.H. Nature, age, and tectonic setting of granitoid-hosted, orogenic gold deposits of the Jiaodong Peninsula, eastern North China craton, China. Miner. Depos. 2002, 37, 283–305. [Google Scholar] [CrossRef]
  10. Zhang, L.; Weinberg, R.F.; Yang, L.Q.; Groves, D.I.; Sai, S.X.; Matchan, E.; Phillips, D.; Kohn, B.P.; Miggins, D.P.; Liu, Y.; et al. Mesozoic orogenic gold mineralization in the Jiaodong Peninsula, China: A focused event at 120 ± 2 Ma during cooling of pregold granite intrusions. Econ. Geol. 2020, 115, 415–441. [Google Scholar] [CrossRef]
  11. Qiu, K.F.; Deng, J.; Laflamme, C.; Long, Z.Y.; Wan, R.Q.; Moynier, F.; Yu, H.C.; Zhang, J.Y.; Ding, Z.J.; Goldfarb, R. Giant Mesozoic gold ores derived from subducted oceanic slab and overlying sediments. Geochim. Cosmochim. Acta 2023, 343, 133–141. [Google Scholar] [CrossRef]
  12. McCuaig, T.C.; Kerrich, R. P-T-t deformation-fluid characteristics of lode gold deposits: Evidence from alteration systematics. Ore Geol. Rev. 1998, 12, 381–454. [Google Scholar] [CrossRef]
  13. Kerrich, R.; Goldfarb, R.J.; Groves, D.; Garwin, S.; Jia, Y.F. The characteristics, origins, and geodynamic settings of supergiant gold metallogenic provinces. Sci. China Ser. D 2000, 43, 1–68. [Google Scholar] [CrossRef]
  14. Goldfarb, R.J.; Groves, D.I.; Gardoll, S. Orogenic gold and geologic time: A global synthesis. Ore Geol. Rev. 2001, 18, 1–75. [Google Scholar] [CrossRef]
  15. Groves, D.I.; Santosh, M. The giant Jiaodong gold province: The key to a unified model for orogenic gold deposits? Geosci. Front. 2016, 7, 409–417. [Google Scholar] [CrossRef]
  16. Groves, D.I.; Santosh, M.; Goldfarb, R.J.; Zhang, L. Structural geometry of orogenic gold deposits: Implications for exploration of world-class and giant deposits. Geosci. Front. 2018, 9, 1163–1177. [Google Scholar] [CrossRef]
  17. Nesbitt, B.E. Gold deposit continuum: A genetic model for lode Au mineralization in the continental crust. Geology 1988, 6, 1044–1048. [Google Scholar] [CrossRef]
  18. Neumayr, P.; Walshe, J.; Hagemann, S.; Petersen, K.; Roache, A.; Frikken, P.; Horn, L.; Halley, S. Oxidized and reduced mineral assemblages in greenstone belt rocks of the St. Ives gold camp, Western Australia: Vectors to high-grade ore bodies in Archaean gold deposits? Miner. Depos. 2008, 43, 363–371. [Google Scholar] [CrossRef]
  19. Helt, K.M.; Williams-Jones, A.E.; Clark, J.R.; Wing, B.A.; Wares, R.P. Constraints on the genesis of the Archean oxidized, intrusion-related Canadian Malartic gold deposit, Quebec, Canada. Econ. Geol. 2014, 109, 713–735. [Google Scholar] [CrossRef]
  20. Goldfarb, R.J.; Groves, D.I. Orogenic gold: Common or evolving fluid and metal sources through time. Lithos 2015, 233, 2–26. [Google Scholar] [CrossRef]
  21. Spence-Jones, C.P.; Jenkin, G.R.; Boyce, A.J.; Hill, N.J.; Sangster, C.J.S. Tellurium, magmatic fluids and orogenic gold: An early magmatic fluid pulse at Cononish gold deposit, Scotland. Ore Geol. Rev. 2018, 102, 894–905. [Google Scholar] [CrossRef]
  22. Windley, B.F.; Alexeiev, D.; Xiao, W.J.; Kröner, A.; Badarch, G. Tectonic models for accretion of the Central Asian orogenic belt. J. Geol. Soc. Lond. 2007, 164, 31–47. [Google Scholar] [CrossRef]
  23. Rui, Z.Y.; Goldfarb, R.J.; Qiu, Y.M.; Zhou, T.H.; Chen, R.Y.; Pirajno, F.; Yun, G. Paleozoic-early Mesozoic gold deposits of the Xinjiang Autonomous Region, northwestern China. Miner. Depos. 2002, 37, 393–418. [Google Scholar] [CrossRef]
  24. Goldfarb, R.J.; Taylor, R.D.; Collins, G.S.; Goryachev, N.A.; Orlandini, O.F. Phanerozoic continental growth and gold metallogeny of Asia. Gondwana Rese. 2014, 25, 48–102. [Google Scholar] [CrossRef]
  25. Zheng, J.H.; Shen, P.; Feng, W.Y. Hydrothermal apatite record of ore-forming processes in the Hatu orogenic gold deposit, West Junggar, Northwest China. Contrib. Mineral. Petrol. 2022, 177, 27. [Google Scholar] [CrossRef]
  26. Gu, X.X.; Zhang, Y.M.; Ge, Z.L.; Chen, W.Z.; Xu, J.C.; Huang, G.; Tao, W. The orogenic Au mineralization system and regional tectonic evolution in the Kalamaili area, East Junggar, Xinjiang. Earth Sci. Front. 2020, 27, 254–275. (In Chinese) [Google Scholar]
  27. Gu, X.X.; Zhang, Y.M.; Ge, Z.L.; Chen, W.Z. Mineralization and genesis of the orogenic gold system in the Kalamaili area, East Junggar, Xinjiang, northwestern China. GSA Bull. 2023, 6. [Google Scholar] [CrossRef]
  28. Li, J.Y.; Yang, T.N.; LI, Y.P.; Zhu, Z.X. Geological features of the Karamaili faulting belt, eastern Junggar region, Xinjiang, China and its constraints on the reconstruction of Late Paleozoic ocean-continental framework of the Central Asian region. Geol. Bull. China 2009, 28, 1817–1826. (In Chinese) [Google Scholar]
  29. Zhang, Y.Y.; Pe-Piper, G.; Piper, D.J.W.; Guo, Z.J. Early Carboniferous collision of the Kalamaili orogenic belt, North Xinjiang, and its implications: Evidence from molasse deposits. GSA Bull. 2013, 125, 932–944. [Google Scholar] [CrossRef]
  30. Song, P.; Wang, T.; Tong, Y.; Zhang, J.J.; Huang, H. Late Carboniferous intrusions along the Kalamaili suture zone, southwestern Central Asian Orogenic Belt (CAOB): Implications for a tectonic switch from subduction to collision. Int. Geol. Rev. 2023, 65, 1601–1621. [Google Scholar] [CrossRef]
  31. Xu, B.; Lu, Y.M.; Gu, X.X.; Zhang, W.Z. Characteristics of ore-forming fluids and O, H, S isotopic geochemistry of the gold deposits in the Kalamaili Area, Xinjiang, China. Earth Sci. Front. 2010, 17, 227–240. (In Chinese) [Google Scholar]
  32. Nie, X.Y. Carboniferous Granitic Magma Activity and Gold Mineralization in the Kalamaily Region, Northern Xinjiang. Ph.D. Thesis, China University of Geosciences, Beijing, China, 2017. (In Chinese). [Google Scholar]
  33. Xiao, W.J.; Windley, B.F.; Yuan, C.; Sun, M.; Han, C.M.; Lin, S.F.; Chen, H.L.; Yan, Q.R.; Liu, D.Y.; Qin, K.Z.; et al. Paleozoic multiple subduction-accretion processes of the southern Altaids. Am. J. Sci. 2009, 309, 221–270. [Google Scholar] [CrossRef]
  34. Jahn, B.M.; Wu, F.Y.; Chen, B. Granitoids of the Central Asian Orogenic Belt and continental growth in the Phanerozoic. Earth Environ. Sci. Trans. R. Soc. Edinb. 2000, 91, 181–193. [Google Scholar]
  35. Chen, B.; Jahn, B.M. Genesis of post-collisional granitoids and basement nature of the Junggar Terrane, NW China: Nd-Sr isotope and trace element evidence. J. Asian Earth Sci. 2004, 23, 691–703. [Google Scholar] [CrossRef]
  36. Guo, L.S.; Zhang, R.; Liu, Y.L.; Xu, F.J.; Su, L. Zircon U-Pb age of Tonghualing intermediate-acid intrusive rocks, Eastern Junggar, Xinjiang. Acta Sci. Nat. Univ. Pekin. 2009, 1, 22–27. (In Chinese) [Google Scholar]
  37. Han, B.F.; Ji, J.Q.; Song, B.; Chen, L.H.; Zhang, L. Late Paleozoic vertical growth of continental crust around the Junggar Basin, Xinjiang, China (Part I): Timing of post-collisional plutonism. Acta Petrol. Sin. 2006, 22, 1077–1086. (In Chinese) [Google Scholar]
  38. Yang, G.X.; Li, Y.J.; Wu, H.E.; Zhong, X.; Yang, B.K.; Yan, C.X.; Yan, J.; Si, G.H. Geochronological and geochemical constrains on petrogenesis of the Huangyangshan A-type granite from the East Junggar, Xinjiang, NW China. J. Asian Earth Sci. 2011, 40, 722–736. [Google Scholar] [CrossRef]
  39. Liu, W.; Liu, X.J.; Liu, L.J. Underplating generated A-and I-type granitoids of the East Junggar from the lower and the upper oceanic crust with mixing of mafic magma: Insights from integrated zircon U-Pb ages, petrography, geochemistry and Nd-Sr-Hf isotopes. Lithos 2013, 179, 293–319. [Google Scholar] [CrossRef]
  40. Han, Y.G.; Zhao, G.C. Final amalgamation of the Tianshan and Junggar orogenic collage in the southwestern Central Asian Orogenic Belt: Constraints on the closure of the Paleo-Asian Ocean. Earth Sci. Rev. 2018, 186, 129–152. [Google Scholar] [CrossRef]
  41. Li, J.Y. Main Characteristics and emplacement process of the East Junggar ophiolites, Xinjiang, China. Acta Petrol. Sin. 1995, 11, 73–84. [Google Scholar]
  42. Fang, A.M.; Wang, S.G.; Zhang, J.M.; Zang, M.; Fang, J.H.; Hu, J.M. The U-Pb ages of zircons from the gabbro in the Kalamaili ophiolite, North Xinjiang and its tectonic significances. Chin. J. Geol. 2015, 50, 140–154. (In Chinese) [Google Scholar]
  43. Wang, B.Y.; Jiang, C.Y.; Li, Y.J.; Wu, H.E.; Xia, Z.D.; Lu, R.H. Geochemistry and tectonic implications of Karamaili ophiolite in East Junggar of Xinjiang. J. Mineral. Petrol. 2009, 29, 74–82. (In Chinese) [Google Scholar]
  44. Huang, G.; Niu, G.Z.; Wang, X.L.; Guo, J.; Yu, F. Formation and emplacement age of Karamaili ophiolite: LA-ICP-MS zircon U-Pb age evidence from the diabase and tuff in eastern Junggar, Xinjiang. Geol. Bull. China 2012, 31, 1267–1278. (In Chinese) [Google Scholar]
  45. Tang, H.F.; Su, Y.P.; Liu, C.Q.; Hou, G.S.; Wang, Y.B. Zircon U-Pb age of the plagiogranite in Kalamaili belt, northern Xinjiang and its tectonic implications. Geotecton. Metallog. 2007, 31, 110–117. (In Chinese) [Google Scholar]
  46. Qin, B. Time constraints and SHRIMP of zircon from plagioclase granite of ophiolitic in Aletunkunduo. Northwest. Geol. 2012, 45, 20–25. (In Chinese) [Google Scholar]
  47. Zhao, L.; Ji, J.Q.; Xu, Q.Q.; Gong, J.F.; Tu, J.Y.; Wang, J.D. Karamaili strike-slipping fault and deformational sequence since Late Paleozoic in the northern Xinjiang. Acta Petrol. Sin. 2010, 28, 2257–2268. (In Chinese) [Google Scholar]
  48. Garofalo, P.; Matthäi, S.K.; Heinrich, C.A. Three-dimensional geometry, ore distribution and time-integrated mass transfer through the quartz-tourmaline-gold vein network of the Sigma deposit (Abitibi belt, Canada). Geofluids 2002, 2, 217–232. [Google Scholar] [CrossRef]
  49. Sibson, R.H.; Robert, F.; Poulsen, K.H. High-angle reverse faults, fluid-pressure cycling, and mesothermal gold-quartz deposits. Geology 1988, 16, 551–555. [Google Scholar] [CrossRef]
  50. Cox, S.F.; Sun, S.S.; Etheridge, M.A.; Wall, V.J.; Potter, T.F. Structural and geochemical controls on the development of turbidite-hosted gold quartz vein deposits, Wattle Gully mine, central Victoria, Australia. Econ. Geol. 1995, 90, 1722–1746. [Google Scholar] [CrossRef]
  51. Bodnar, R.J. Revised equation and table for determining the freezing point depression of H2O-NaCl solutions. Geochim. Cosmochim. Acta 1993, 57, 683–684. [Google Scholar] [CrossRef]
  52. Brown, P.E.; Lamb, W.M. P-V-T properties of fluids in the system H2O±CO2±NaCl: New graphical presentations and implications for fluid inclusion studies. Geochim. Cosmochim. Acta 1989, 53, 1209–1221. [Google Scholar] [CrossRef]
  53. Brown, P.E. FLINCOR: A microcomputer program for the reduction and investigation of fluid-inclusion data. Am. Mineral. 1989, 74, 1390–1393. [Google Scholar]
  54. Clayton, R.N.; Mayeda, T.K. The use of bromine pentafluoride in the extraction of oxygen from oxides and silicates for isotopic analysis. Geochim. Cosmochim. Acta 1963, 27, 43–52. [Google Scholar] [CrossRef]
  55. Clayton, R.N.; O’Neil, J.R.; Mayeda, T.K. Oxygen isotope exchange between quartz and water. J. Geophys. Res. 1972, 77, 3057–3067. [Google Scholar] [CrossRef]
  56. Sláma, J.; Košler, J.; Condon, D.J.; Crowley, J.L.; Gerdes, A.; Hanchar, J.M.; Horstwood, M.S.A.; Morris, G.A.; Nasdala, L.; Norberg, N.; et al. Plešovice zircon—A new natural reference material for U-Pb and Hf isotopic microanalysis. Chem. Geol. 2008, 249, 1–35. [Google Scholar] [CrossRef]
  57. Liu, Y.S.; Hu, Z.C.; Gao, S.; Günther, D.; Xu, J.; Gao, C.G.; Chen, H.H. In situ analysis of major and trace elements of anhydrous minerals by LA-ICP-MS without applying an internal standard. Chem. Geol. 2008, 257, 34–43. [Google Scholar] [CrossRef]
  58. Ludwig, K.R. User’s Manual for Isoplot/ex Version 3.00: A Geochronological Toolkit for Microsoft Excel; Special Publication; Berkeley Geochronology Center: Berkeley, CA, USA, 2003. [Google Scholar]
  59. Boyle, R.W. The geochemistry of gold and its deposits. Geol. Surv. Can. Bull. 1979, 280, 1–584. [Google Scholar]
  60. Youngson, J.H.; Wopereis, P.; Kerr, L.C.; Craw, D. Au-Ag-Hg and Au-Ag alloys in Nokomai and Nevis valley placers, northern Southland and Central Otago, New Zealand, and their implications. N. Z. J. Geol. Geop. 2002, 45, 53–69. [Google Scholar] [CrossRef]
  61. Simon, G.; Kesler, S.E.; Chryssoulis, S. Geochemistry and textures of gold-bearing arsenian pyrite, Twin Creeks, Nevada: Implications for deposition of gold in Carlin-type deposits. Econ. Geol. 1999, 94, 405–422. [Google Scholar] [CrossRef]
  62. Goldfarb, R.J.; Baker, T.; Dubé, B.; Groves, D.I.; Hart, C.J.R.; Gosselin, P. Distribution, character, and genesis of gold deposits in metamorphic terranes. Econ. Geol. 2005, 100, 407–450. [Google Scholar]
  63. Zhou, Z.J.; Chen, Y.J.; Jiang, S.Y.; Zhao, H.X.; Qin, Y.; Hu, C.J. Geology, geochemistry and ore genesis of the Wenyu gold deposit, Xiaoqinling gold field, Qinling Orogen, southern margin of North China Craton. Ore Geol. Rev. 2014, 59, 1–20. [Google Scholar] [CrossRef]
  64. Gao, H.Z.; Zhang, W.S. Characteristics and metallogenic fluid and thermodynamic condition analysis of metallogenic system in highly strained structural belt, East Junggar. Earth Sci.-J. China Univ. Geosci. 2000, 25, 369–374. (In Chinese) [Google Scholar]
  65. Hoskin, P.W.O. Patterns of chaos: Fractal statistics and the oscillatory chemistry of zircon. Geochim. Cosmochim. Acta 2000, 64, 1905–1923. [Google Scholar] [CrossRef]
  66. Hoskin, P.W.O.; Schaltegger, U. The composition of zircon and igneous and metamorphic petrogenesis. Rev. Mineral. Geochem. 2003, 53, 27–62. [Google Scholar] [CrossRef]
  67. Wang, J.; Su, Y.P.; Zheng, J.P.; Belousova, E.A.; Griffin, W.L.; Zhou, X.; Dai, H.K. Hidden Eoarchean crust in the southwestern Central Asian Orogenic Belt. Lithos 2020, 360–361, 105437. [Google Scholar] [CrossRef]
  68. Hoskin, P.W.O. Trace-element composition of hydrothermal zircon and the alteration of Hadean zircon from the Jack Hills, Australia. Geochim. Cosmochim. Acta 2005, 69, 637–648. [Google Scholar] [CrossRef]
  69. Kirkland, C.L.; Whitehouse, M.J.; Slagstad, T. Fluid-assisted zircon and monazite growth within a shear zone: A case study from Finnmark, Arctic Norway. Contrib. Mineral. Petrol. 2009, 158, 637–657. [Google Scholar] [CrossRef]
  70. McDonough, W.F.; Sun, S.S. The composition of the Earth. Chem. Geol. 1995, 120, 223–253. [Google Scholar] [CrossRef]
  71. Cabri, L.J.; Newville, M.; Gordon, R.A.; Crozier, E.D.; Sutton, S.R.; McMahon, G.; Jiang, D.T. Chemical speciation of gold in arsenopyrite. Can. Mineral. 2000, 38, 1265–1281. [Google Scholar] [CrossRef]
  72. Chen, Y.L.; Li, H.; Zheng, C.Y.; Elatikpo, S.M.; Cheng, S.L.; Jiang, W.T. Ore-forming process of the Haigou gold deposit in the eastern Central Asian Orogenic Belt, NE China: Constrains from EPMA and LA-ICP-MS analysis of Au-bearing pyrite. Resour. Geol. 2022, 72, e12304. [Google Scholar] [CrossRef]
  73. Liu, H.M.; Beaudoin, G. Geochemical signatures in native gold derived from Au-bearing ore deposits. Ore Geol. Rev. 2021, 132, 104066. [Google Scholar] [CrossRef]
  74. Harris, D.C. The mineralogy of gold and its relevance to gold recoveries. Miner. Depos. 1990, 25, S3–S7. [Google Scholar] [CrossRef]
  75. Deditius, A.P.; Reich, M.; Kesler, S.E.; Utsunomiya, S.; Chryssoulis, S.L.; Walshe, J.; Ewing, R.C. The coupled geochemistry of Au and As in pyrite from hydrothermal ore deposits. Geochim. Cosmochim. Acta 2014, 140, 644–670. [Google Scholar] [CrossRef]
  76. Li, H.; Zhu, D.P.; Algeo, T.J.; Li, M.; Jiang, W.C.; Chen, S.F.; Elatikpo, S.M. Pyrite trace element and S-Pb isotopic evidence for contrasting sources of metals and ligands during superimposed hydrothermal events in the Dongping gold deposit, North China. Miner. Depos. 2023, 58, 337–358. [Google Scholar] [CrossRef]
  77. Reich, M.; Kesler, S.E.; Utsunomiya, S.; Palenik, C.S.; Chryssoulis, S.L.; Ewing, R.C. Solubility of gold in arsenian pyrite. Geochim. Cosmochim. Acta 2005, 69, 2781–2796. [Google Scholar] [CrossRef]
  78. Cabri, L.J.; Chryssoulis, S.L.; de Villiers, J.P.R.; Laflamme, J.G.; Buseck, P.R. The nature of “invisible” gold in arsenopyrite. Can. Mineral. 1989, 27, 353–362. [Google Scholar]
  79. Fougerouse, D.; Reddy, S.M.; Aylmore, M.; Yang, L.; Guagliardo, P.; Saxey, D.W.; Rickard, W.D.A.; Timms, N. A new kind of invisible gold in pyrite hosted in deformation-related dislocations. Geology 2021, 49, 1225–1229. [Google Scholar] [CrossRef]
  80. Böhlke, J.K.; Kistler, R.W. Rb-Sr, K-Ar, and stable isotope evidence for the ages and sources of fluid components of gold-bearing quartz veins in the northern Sierra Nevada foothills metamorphic belt, California. Econ. Geol. 1986, 81, 296–322. [Google Scholar] [CrossRef]
  81. Jia, Y.F.; Li, X.; Kerrich, R. Stable isotope (O, H, S, C, and N) systematics of quartz vein systems in the turbidite-hosted Central and North Deborah gold deposits of the Bendigo gold field, central Victoria, Australia: Constraints on the origin of ore-forming fluids. Econ. Geol. 2001, 96, 705–721. [Google Scholar] [CrossRef]
  82. Chen, H.Y.; Chen, Y.J.; Baker, M. Isotopic geochemistry of the Sawayaerdun orogenic-type gold deposit, Tianshan, northwest China: Implications for ore genesis and mineral exploration. Chem. Geol. 2012, 310-311, 1–11. [Google Scholar] [CrossRef]
  83. Ridley, J.R.; Diamond, L.W. Fluid chemistry of orogenic lode gold deposits and implications for genetic models. Rev. Econ. Geol. 2000, 13, 146–162. [Google Scholar]
  84. Goldfarb, R.J.; Newberry, R.J.; Pickthorn, W.J.; Gent, C.A. Oxygen, hydrogen, and sulfur isotope studies in the Juneau gold belt, southeastern Alaska: Constraints on the origin of hydrothermal fluids. Econ. Geol. 1991, 86, 66–80. [Google Scholar] [CrossRef]
  85. Zhang, J.; Chen, Y.J.; Pirajno, F.; Deng, J.; Chen, H.Y.; Wang, C.M. Geology, C-H-O-S-Pb isotope systematics and geochronology of the Yindongpo gold deposit, Tongbai Mountains, central China: Implication for ore genesis. Ore Geol. Rev. 2013, 53, 343–356. [Google Scholar] [CrossRef]
  86. Liu, J.Y.; Yu, H.X.; Wu, G.Q. Alkali granites and tin deposits of the Kalamaili area, northern Xinjiang. Geollogical. Explor. Non-Ferr. Met. 1997, 6, 129–135. (In Chinese) [Google Scholar]
  87. Taylor, H.P. The application of oxygen and hydrogen isotope studies to problems of hydrothermal alteration and ore deposition. Econ. Geol. 1974, 69, 843–883. [Google Scholar] [CrossRef]
  88. Wilkinson, J.J. Fluid inclusions in hydrothermal ore deposits. Lithos 2001, 55, 229–272. [Google Scholar] [CrossRef]
  89. Coulibaly, Y.; Boiron, M.C.; Cathelineau, M.; Kouamelan, A.N. Fluid immiscibility and gold deposition in the Birimian quartz veins of the Angovia deposit (Yaouré, Ivory Coast). J. Afr. Earth Sci. 2008, 50, 234–254. [Google Scholar] [CrossRef]
  90. Deng, X.H.; Chen, Y.J.; Santosh, M.; Yao, J.M. Genesis of the 1.76 Ga Zhaiwa Mo-Cu and its link with the Xiong’er volcanics in the North China Craton: Implications for accretionary growth along the margin of the Columbia supercontinent. Precambr. Res. 2013, 227, 337–348. [Google Scholar] [CrossRef]
  91. Craw, D.; Teagle, D.A.H.; Belocky, R. Fluid immiscibility in late-Alpine gold-bearing veins, Eastern and Northwestern European Alps. Miner. Depos. 1993, 28, 28–36. [Google Scholar] [CrossRef]
  92. Bowers, T.S.; Helgeson, H.C. Calculation of the thermodynamic and geochemical consequences of nonideal mixing in the system H2O-CO2-NaCl on phase relations in geologic systems: Metamorphic equilibria at high pressures and temperatures. Am. Mineral. 1983, 68, 1059–1075. [Google Scholar]
  93. Ramboz, C.; Pichavant, M.; Weisbrod, A. Fluid immiscibility in natural processes: Use and misuse of fluid inclusion data: II. Interpretation of fluid inclusion data in terms of Immiscibility. Chem. Geol. 1982, 37, 29–46. [Google Scholar] [CrossRef]
  94. Robert, F.; Kelly, W.C. Ore-forming fluids in Archean gold-bearing quartz veins at the Sigma Mine, Abitibi greenstone belt, Quebec, Canada. Econ. Geol. 1987, 82, 1464–1482. [Google Scholar] [CrossRef]
  95. Xu, Y.F.; Ni, P.; Wang, G.G.; Pan, J.Y.; Guan, S.J.; Chen, H.; Ding, J.Y.; Li, L. Geology, fluid inclusion and stable isotope study of the Huangshan orogenic gold deposit: Implications for future exploration along the Jiangshan-Shaoxing fault zone, South China. J. Geochem. Explor. 2016, 171, 37–54. [Google Scholar] [CrossRef]
  96. Roedder, E. Fluid inclusions: Reviews in Mineralogy and Geochemistry. Mineral. Soc. Am. 1984, 12, 1–644. [Google Scholar]
  97. Hollister, L.S. Enrichment of CO2 in fluid inclusions in quartz by removal of H2O during crystal-plastic deformation. J. Struct. Geol. 1990, 12, 895–901. [Google Scholar] [CrossRef]
  98. Hall, D.L.; Bodnar, R.J. Methane in fluid inclusions from granulites: A product of hydrogen diffusion? Geochim. Cosmochim. Acta 1990, 54, 641–651. [Google Scholar] [CrossRef]
  99. Benning, L.G.; Seward, T.M. Hydrosulphide complexing of Au(I) in hydrothermal solutions from 150–400 °C and 500-1500 bar. Geochim. Cosmochim. Acta 1996, 60, 1849–1871. [Google Scholar] [CrossRef]
  100. Stefánsson, A.; Seward, T.M. Gold(I) complexing in aqueous sulphide solutions to 500 °C at 500 bar. Geochim. Cosmochim. Acta 2004, 68, 4121–4143. [Google Scholar] [CrossRef]
  101. Phillips, G.N.; Powell, R. Formation of gold deposits: A metamorphic devolatilization model. J. Metamorph. Geol. 2010, 28, 689–718. [Google Scholar] [CrossRef]
  102. Phillips, G.N.; Evans, K.A. Role of CO2 in the formation of gold deposits. Nature 2004, 429, 860–863. [Google Scholar] [CrossRef]
  103. Neall, F.B.; Phillips, G.N. Fluid-wall rock interaction in an Archean hydrothermal gold deposit: A thermodynamic model for the Hunt Mine, Kambalda. Econ. Geol. 1987, 82, 1679–1694. [Google Scholar] [CrossRef]
  104. Naden, J.; Shepherd, T.J. Role of methane and carbon dioxide in gold deposition. Nature 1989, 342, 793–795. [Google Scholar] [CrossRef]
  105. Lee, J.K.W.; Williams, I.S.; Ellis, D.J. Pb, U and Th diffusion in natural zircon. Nature 1997, 390, 159–162. [Google Scholar] [CrossRef]
  106. Yeats, C.J.; McNaughton, N.J.; Groves, D.I. SHRIMP U-Pb geochronological constraints on Archean volcanic-hosted massive sulfide and lode gold mineralization at Mount Gibson, Yilgarn Craton, Western Australia. Econ. Geol. 1996, 91, 1354–1371. [Google Scholar] [CrossRef]
  107. Claoué-Long, J.C.; King, R.W.; Kerrich, R. Archaean hydrothermal zircon in the Abitibi greenstone belt: Constraints on the timing of gold mineralisation. Earth Planet. Sci. Lett. 1990, 98, 109–128. [Google Scholar] [CrossRef]
  108. Kerrich, R.; King, R. Hydrothermal zircon and baddeleyite in Val-d’Or Archean mesothermal gold deposits: Characteristics, compositions, and fluid-inclusion properties, with implications for timing of primary gold mineralization. Can. J. Earth Sci. 1993, 30, 2334–2351. [Google Scholar] [CrossRef]
  109. Xu, J.C.; Gu, X.X.; Zhang, Y.M.; Wang, J.L.; He, G.; Zhou, C.; Liu, R.P. Geology, fluid inclusions, HO isotope, and hydrothermal zircon U-Pb geochronology of the Daqingshan orogenic gold deposit in Beishan orogenic belt, Xinjiang, NW China. Miner. Depos. 2021, 56, 325–342. [Google Scholar] [CrossRef]
  110. Deng, X.D.; Li, J.W.; Wen, G. U-Pb geochronology of hydrothermal zircons from the early Cretaceous iron skarn deposits in the Handan-Xingtai district, North China craton. Econ. Geol. 2015, 110, 2159–2180. [Google Scholar] [CrossRef]
  111. Li, X.; Wang, K.Y.; Sun, G.S.; Sun, J.D.; Liu, G.Y.; Zhang, J.T.; Han, L.; Wang, G.W. Discussion on metallogenesis of Zhaojiagou Ta-Nb deposit in Inner Mongolia: Evidence from amazonitization and albitization granite geochronology and geochemistry. Acta Petrol. Sin. 2021, 37, 1765–1784. (In Chinese) [Google Scholar]
  112. Li, X.C.; Fan, H.R.; Zeng, X.; Yang, K.F.; Yang, Z.F.; Wang, Q.W.; Li, H.T. Identification of ~1.3 Ga hydrothermal zircon from the giant Bayan Obo REE deposit (China): Implication for dating geologically-complicated REE ore system. Ore Geol. Rev. 2021, 138, 104405. [Google Scholar] [CrossRef]
  113. Zhang, S.; Zhou, T.F.; Zhang, Z.Z.; Wu, M.A.; Wang, J.; Lü, Q.T. In-situ hydrothermal zircon U-Pb and phlogopite 40Ar-39Ar geochronology of uranium mineralisation in Luzong ore district scientific drilling (LTZK01), Anhui Province, SE China: Constraints on the mineralisation process. Ore Geol. Rev. 2021, 134, 104133. [Google Scholar] [CrossRef]
  114. Jiang, W.C.; Li, H.; Evans, N.J.; Wu, J.H. Zircon records multiple magmatic-hydrothermal processes at the giant Shizhuyuan W-Sn-Mo-Bi polymetallic deposit, South China. Ore Geol. Rev. 2019, 115, 103160. [Google Scholar] [CrossRef]
  115. Elatikpo, S.M.; Li, H.; Liu, B.; Zhang, W.D. Metallogenesis of the Bakoshi-Kundila gold deposit in northern West Nigerian Subshield: Insights from pyrite chemical and sulfur isotopic compositions and zircon U-Pb geochronology. Precambr. Res. 2022, 383, 106890. [Google Scholar] [CrossRef]
  116. Dubińska, E.; Bylina, P.; Kozlowski, A.; Dörr, W.; Nejbert, K.; Schastok, J.; Kulicki, C. U-Pb dating of serpentinization: Hydrothermal zircon from a metasomatic rodingite shell (Sudetic ophiolite, SW Poland). Chem. Geol. 2004, 203, 183–203. [Google Scholar] [CrossRef]
  117. Dempster, T.J.; Hay, D.C.; B1uck, B.J. Zircon growth in slate. Geology 2004, 32, 221–224. [Google Scholar] [CrossRef]
  118. Zhai, W.; Zhang, E.; Zheng, S.Q.; Santosh, M.; Sun, X.M.; Niu, H.C.; Fu, B.; Fu, Y.; Li, D.F.; Jiang, Y.H.; et al. Hydrothermal zircon: Characteristics, genesis and metallogenic implications. Ore Geol. Rev. 2022, 149, 105111. [Google Scholar] [CrossRef]
  119. Fan, T.T.; Zhou, X.H.; Liu, Y.Q.; Li, W.; Zheng, Z.Y.; Hu, T.; Liang, H. Crystal tuff zircon LA-ICP-MS U-Pb ages from the Lower Jiangbasitao Formation in the East Daheishan area, Xinjiang and their geological implications. Acta Sedimentol. Sin. 2011, 29, 312–320. (In Chinese) [Google Scholar]
  120. Wang, J.; Su, Y.P.; Zheng, J.P.; Belousova, E.A.; Chen, M.; Dai, H.K.; Zhou, L. Rapid transition from oceanic subduction to postcollisional extension revealed by Carboniferous magmatism in East Junggar (NW China), southwestern Central Asian orogenic belt. GSA Bull. 2022, 134, 1018–1042. [Google Scholar] [CrossRef]
  121. Xi, X.W.; Yang, L.Q.; Wang, Y.J.; Deng, J.; Lin, G.; Wang, J.P.; Lei, X.Q. Numerical modeling of the temperature effects of tectonic regimes transformation and their ore-forming dynamics. Earth Sci. Front. 2003, 10, 47–55. (In Chinese) [Google Scholar]
  122. Li, H.Q.; Chen, F.W. Regional Mineralization in Altay Area of Northern Xinjiang China; Geological Publishing House: Beijing, China, 2004; pp. 1–391. (In Chinese) [Google Scholar]
  123. Lu, Y.M.; Zhang, Y.J.; Pan, M.; Fan, J.J.; Liu, Y.F.; Zhang, D.; Chen, X.W.; Pan, A.J. Types and geological characteristics of gold deposits in East Junggar, Xinjiang. Acta Geosci. Sin. 2010, 31, 434–442. (In Chinese) [Google Scholar]
  124. Chen, W.Z.; Gu, X.X.; Zhang, Y.M.; He, G.; Huang, G.; Li, J.T. Geological features and geochronology of the Jinshuiquan gold deposit in the East Junggar, Xinjiang. Geol. Bull. China 2019, 38, 1240–1255. (In Chinese) [Google Scholar]
  125. Liu, J.; Bao, X.S.; Gao, Y.B.; Yang, K.; Kou, S.L.; Yang, W.; Wang, Z.B.; He, J.L.; Zhang, Y.X.; Chen, H. Composition and Geochemical Characteristics of Pyrite and Quartz: Constraints on the Origin of the Xinjiazui Gold Deposit, Northwestern Margin of the Yangtze Block, China. Minerals 2022, 12, 688. [Google Scholar] [CrossRef]
Figure 2. Geological map of the Nanmingshui gold deposit.
Figure 2. Geological map of the Nanmingshui gold deposit.
Minerals 13 01296 g002
Figure 3. Representative photos of orebodies and various ore types in the Nanmingshui gold deposit. (A) Quartz vein hosted in a high-angle reverse fault, with intensely altered wall-rocks; (B) quartz–tourmaline vein occurring in mylonitic tuffaceous sandstone; (C) auriferous quartz–tourmaline vein formed in a high-angle reverse fault; (D) limonitized gold-bearing quartz vein near the surface; (E) comb structures in the quartz–tourmaline vein; (F) laminated texture in the quartz–tourmaline vein; (G) scattered and fine-grained pyrite coexisting with arsenopyrite in early-stage milky quartz (Qz1); (H) laminated middle-stage quartz–tourmaline vein (Qz2) containing fine-grained arsenopyrite and pyrite; (I) coarse-grained, cubic or pentagonal–dodecahedral pyrite coexisting with arsenopyrite in the altered rocks; (J) early-stage milky quartz vein (Qz1) crosscut by middle-stage smoky-gray quartz–tourmaline vein (Qz2); (K) middle-stage quartz–tourmaline vein containing several brecciated altered wall-rocks; (L) late-stage quartz–calcite vein (Qz3) intruding in altered rocks. Abbreviations: Apy, arsenopyrite; Py, pyrite; Qz, quartz; Tur, tourmaline; Cal, calcite.
Figure 3. Representative photos of orebodies and various ore types in the Nanmingshui gold deposit. (A) Quartz vein hosted in a high-angle reverse fault, with intensely altered wall-rocks; (B) quartz–tourmaline vein occurring in mylonitic tuffaceous sandstone; (C) auriferous quartz–tourmaline vein formed in a high-angle reverse fault; (D) limonitized gold-bearing quartz vein near the surface; (E) comb structures in the quartz–tourmaline vein; (F) laminated texture in the quartz–tourmaline vein; (G) scattered and fine-grained pyrite coexisting with arsenopyrite in early-stage milky quartz (Qz1); (H) laminated middle-stage quartz–tourmaline vein (Qz2) containing fine-grained arsenopyrite and pyrite; (I) coarse-grained, cubic or pentagonal–dodecahedral pyrite coexisting with arsenopyrite in the altered rocks; (J) early-stage milky quartz vein (Qz1) crosscut by middle-stage smoky-gray quartz–tourmaline vein (Qz2); (K) middle-stage quartz–tourmaline vein containing several brecciated altered wall-rocks; (L) late-stage quartz–calcite vein (Qz3) intruding in altered rocks. Abbreviations: Apy, arsenopyrite; Py, pyrite; Qz, quartz; Tur, tourmaline; Cal, calcite.
Minerals 13 01296 g003
Figure 4. Photomicrographs showing mineral assemblages and textural features in the Nanmingshui gold deposit. (A) Galena and cube-shaped pyrite (Py1) occurring in arsenopyrite as inclusions (BSE); (B) pentagonal–dodecahedral pyrite (Py2) coexisting with fine-grained galena (BSE); (C) irregular native gold occurring in hematite as fissure-filled gold (whole view: BSE, amplified area: reflected light); (D) Py2 coexisting with euhedral arsenopyrite and py1 wrapped in arsenopyrite, displaying a poikilitic texture (reflected light); (E) arsenopyrite with cataclastic texture (reflected light); (F) Py2 occurring as pentagonal-dodecahedral crystals in auriferous quartz–tourmaline veins (reflected light); (G) coexistence of arsenopyrite, chalcopyrite, and tetrahedrite (reflected light); (H) Py2 replaced with pyrrhotine, forming the replacement residual texture (reflected light); (I) Py1 veinlet in quartz vein (reflected light); (J) quartz coexisting with sericite and calcite (transmitted light); (K) fibrous sericite and quartz around Py2, forming a pressure shadow structure (transmitted light); (L) acicular tourmaline coexisting with smoky-gray quartz (transmitted light). Abbreviations: Ser, Sericite; Gn, galena; Cpy, chalcopyrite; Thr, tetrahedrite; Po, pyrrhotite; Hem, hematite; Au, native gold. For other abbreviations, see Figure 3.
Figure 4. Photomicrographs showing mineral assemblages and textural features in the Nanmingshui gold deposit. (A) Galena and cube-shaped pyrite (Py1) occurring in arsenopyrite as inclusions (BSE); (B) pentagonal–dodecahedral pyrite (Py2) coexisting with fine-grained galena (BSE); (C) irregular native gold occurring in hematite as fissure-filled gold (whole view: BSE, amplified area: reflected light); (D) Py2 coexisting with euhedral arsenopyrite and py1 wrapped in arsenopyrite, displaying a poikilitic texture (reflected light); (E) arsenopyrite with cataclastic texture (reflected light); (F) Py2 occurring as pentagonal-dodecahedral crystals in auriferous quartz–tourmaline veins (reflected light); (G) coexistence of arsenopyrite, chalcopyrite, and tetrahedrite (reflected light); (H) Py2 replaced with pyrrhotine, forming the replacement residual texture (reflected light); (I) Py1 veinlet in quartz vein (reflected light); (J) quartz coexisting with sericite and calcite (transmitted light); (K) fibrous sericite and quartz around Py2, forming a pressure shadow structure (transmitted light); (L) acicular tourmaline coexisting with smoky-gray quartz (transmitted light). Abbreviations: Ser, Sericite; Gn, galena; Cpy, chalcopyrite; Thr, tetrahedrite; Po, pyrrhotite; Hem, hematite; Au, native gold. For other abbreviations, see Figure 3.
Minerals 13 01296 g004
Figure 5. Mineral paragenesis of the Nanmingshui gold deposit. The dotted line represents trace amounts of that mineral and the thickness of the solid line represents the amount of that mineral.
Figure 5. Mineral paragenesis of the Nanmingshui gold deposit. The dotted line represents trace amounts of that mineral and the thickness of the solid line represents the amount of that mineral.
Minerals 13 01296 g005
Figure 6. BSE images along with EDS X-ray compositional maps for Apy and Py2 in the auriferous quartz veins of the middle stage at Nanmingshui. The X-ray lines used were Kα for As and Lα for Pb, respectively.
Figure 6. BSE images along with EDS X-ray compositional maps for Apy and Py2 in the auriferous quartz veins of the middle stage at Nanmingshui. The X-ray lines used were Kα for As and Lα for Pb, respectively.
Minerals 13 01296 g006
Figure 7. Microphotographs showing the major fluid inclusion types in quartz from the Nanmingshui gold deposit. (A) Coexistence of the C-type, W-type, and PC-type FIs in the early-stage quartz; (B) coexistence of the PC-type and W-type FIs in the early-stage quartz; (C) W-type inclusion in the late-stage quartz; (D) C-type inclusion in the middle-stage quartz; (E) coexistence of the C-type and PC-type FIs in the middle-stage quartz; (F) coexistence of the C-type and W-type fluid inclusions in the middle-stage quartz; (G) coexistence of the type Ia and Ib FIs in the middle-stage quartz; (H) Ib-type FIs in the middle-stage quartz; (I) C-type FIs in the middle-stage quartz with negative crystal shape. Abbreviations: VCO2, vapor CO2; LCO2, liquid CO2; VH2O, vapor H2O; LH2O, liquid H2O.
Figure 7. Microphotographs showing the major fluid inclusion types in quartz from the Nanmingshui gold deposit. (A) Coexistence of the C-type, W-type, and PC-type FIs in the early-stage quartz; (B) coexistence of the PC-type and W-type FIs in the early-stage quartz; (C) W-type inclusion in the late-stage quartz; (D) C-type inclusion in the middle-stage quartz; (E) coexistence of the C-type and PC-type FIs in the middle-stage quartz; (F) coexistence of the C-type and W-type fluid inclusions in the middle-stage quartz; (G) coexistence of the type Ia and Ib FIs in the middle-stage quartz; (H) Ib-type FIs in the middle-stage quartz; (I) C-type FIs in the middle-stage quartz with negative crystal shape. Abbreviations: VCO2, vapor CO2; LCO2, liquid CO2; VH2O, vapor H2O; LH2O, liquid H2O.
Minerals 13 01296 g007
Figure 8. Histograms of total homogenization temperatures (Th) and salinities of C- and W-type FIs in the Nanmingshui gold deposit.
Figure 8. Histograms of total homogenization temperatures (Th) and salinities of C- and W-type FIs in the Nanmingshui gold deposit.
Minerals 13 01296 g008
Figure 9. Tm-CO2 (A), Tm-cla (B), and Th-CO2 (C) histograms for C- and PC-type FIs in the Nanmingshui gold deposit.
Figure 9. Tm-CO2 (A), Tm-cla (B), and Th-CO2 (C) histograms for C- and PC-type FIs in the Nanmingshui gold deposit.
Minerals 13 01296 g009
Figure 10. Histograms of trapping pressures of C-type FIs from the Nanmingshui gold deposit.
Figure 10. Histograms of trapping pressures of C-type FIs from the Nanmingshui gold deposit.
Minerals 13 01296 g010
Figure 11. Representative Raman spectra of FIs in the Nanmingshui gold deposit. (A) Spectrum for liquid CO2 phase of two-phase C-type inclusion in early-stage quartz, showing variable contents of CH4 and N2, in addition to CO2; (B) spectrum for liquid CO2 phase of three-phase C-type inclusion in middle-stage quartz, containing minor amounts of N2, CH4, and H2O, in addition to CO2; (C) spectrum for three-phase C-type inclusion in middle-stage quartz, showing minor contents of N2 in vapor CO2; (D) spectrum for liquid CO2 phase of CO2-rich inclusion in middle-stage quartz, containing variable N2 and CH4, in addition to CO2; (E) spectrum for vapor bubbles of W-type inclusion in middle-stage quartz, containing variable CO2 and CH4; (F) spectrum for vapor bubbles of W-type inclusion in late-stage quartz, containing variable CH4.
Figure 11. Representative Raman spectra of FIs in the Nanmingshui gold deposit. (A) Spectrum for liquid CO2 phase of two-phase C-type inclusion in early-stage quartz, showing variable contents of CH4 and N2, in addition to CO2; (B) spectrum for liquid CO2 phase of three-phase C-type inclusion in middle-stage quartz, containing minor amounts of N2, CH4, and H2O, in addition to CO2; (C) spectrum for three-phase C-type inclusion in middle-stage quartz, showing minor contents of N2 in vapor CO2; (D) spectrum for liquid CO2 phase of CO2-rich inclusion in middle-stage quartz, containing variable N2 and CH4, in addition to CO2; (E) spectrum for vapor bubbles of W-type inclusion in middle-stage quartz, containing variable CO2 and CH4; (F) spectrum for vapor bubbles of W-type inclusion in late-stage quartz, containing variable CH4.
Minerals 13 01296 g011
Figure 12. Representative CL images, U-Pb Concordia diagrams, and weighted mean ages of magmatic and hydrothermal zircons from the auriferous quartz–tourmaline vein in the Nanmingshui gold deposit. (A,B) U-Pb Concordia diagrams for all analyzed zircons; (C,D) U-Pb Concordia diagrams and weighted mean ages for type-1 zircons (magmatic).
Figure 12. Representative CL images, U-Pb Concordia diagrams, and weighted mean ages of magmatic and hydrothermal zircons from the auriferous quartz–tourmaline vein in the Nanmingshui gold deposit. (A,B) U-Pb Concordia diagrams for all analyzed zircons; (C,D) U-Pb Concordia diagrams and weighted mean ages for type-1 zircons (magmatic).
Minerals 13 01296 g012
Figure 14. Microphotographs of native gold in the Nanmingshui gold deposit. (A) Native gold occurring in arsenopyrite as anhedral inclusions (reflected light); (B) native gold occurring as intergranular grains in contact with the margin of arsenopyrite (reflected light); (C) elongated gold filling in fissures within the pseudomorph of hematite after pyrite (reflected light); (D) native gold occurring in quartz as isolated grains (reflected light). For mineral abbreviations, see Figure 4.
Figure 14. Microphotographs of native gold in the Nanmingshui gold deposit. (A) Native gold occurring in arsenopyrite as anhedral inclusions (reflected light); (B) native gold occurring as intergranular grains in contact with the margin of arsenopyrite (reflected light); (C) elongated gold filling in fissures within the pseudomorph of hematite after pyrite (reflected light); (D) native gold occurring in quartz as isolated grains (reflected light). For mineral abbreviations, see Figure 4.
Minerals 13 01296 g014
Figure 15. Au-As plot showing EPMA analyses of arsenian pyrite and arsenopyrite from the Nanmingshui gold deposit. Solid line represents the Au saturation line for Carlin-type gold deposits [77], and dashed line represents that for orogenic gold deposits [75]. For mineral abbreviations, see Figure 3.
Figure 15. Au-As plot showing EPMA analyses of arsenian pyrite and arsenopyrite from the Nanmingshui gold deposit. Solid line represents the Au saturation line for Carlin-type gold deposits [77], and dashed line represents that for orogenic gold deposits [75]. For mineral abbreviations, see Figure 3.
Minerals 13 01296 g015
Figure 16. Calculated δD and δ18O of the ore-forming fluids at Nanmingshui. Fields for primary magmatic and metamorphic waters are cited from reference [87], and domains for other deposits are given in the text.
Figure 16. Calculated δD and δ18O of the ore-forming fluids at Nanmingshui. Fields for primary magmatic and metamorphic waters are cited from reference [87], and domains for other deposits are given in the text.
Minerals 13 01296 g016
Figure 17. Plot of Th vs. salinity for fluid inclusions in quartz from the Nanmingshui gold deposit.
Figure 17. Plot of Th vs. salinity for fluid inclusions in quartz from the Nanmingshui gold deposit.
Minerals 13 01296 g017
Table 1. Electron probe data of native gold and sulfides in the Nanmingshui gold deposit.
Table 1. Electron probe data of native gold and sulfides in the Nanmingshui gold deposit.
MineralSeCdAuAgAsFeSCuPbSbZnNiCoHgTeBiTotal
Ng (n = 3)
Minwt.% 89.352.81bdl0.490.010.03bdlbdlbdlbdlbdl2.42bdl0.7896.89
Max 90.513.080.011.300.020.06bdl0.020.02bdl0.022.560.041.2997.26
Avg 89.822.950.010.770.010.04bdl0.010.01bdl0.012.500.010.9797.10
Minat.% 0.870.05bdl0.020.000.00bdlbdlbdlbdlbdl0.02bdl0.011.00
Max 0.900.050.000.040.000.00bdl0.000.00bdl0.000.020.000.011.00
Avg 0.880.050.000.030.000.00bdl0.000.00bdl0.000.020.000.011.00
Apy (n = 8)
Minwt.%0.20bdlbdlbdl39.8435.5321.21bdlbdlbdlbdlbdl0.03 98.26
Max0.270.030.010.0241.3836.4522.390.010.080.080.030.020.06 99.35
Avg0.230.010.000.0140.4535.9322.020.000.020.020.010.010.05 98.75
Minat.%0.00bdlbdlbdl0.851.021.07bdlbdlbdlbdlbdl0.00 3.00
Max0.010.000.000.000.891.041.120.000.000.000.000.000.00 3.00
Avg0.000.000.000.000.861.031.100.000.000.000.000.000.00 3.00
FormulaFe1.03As0.86S1.10 (to 3 atoms p.f.u.)
Py2 (n = 10)
Minwt.%bdlbdlbdl bdl0.5145.6452.34bdlbdlbdlbdlbdlbdl 99.44
Max0.010.020.080.011.9646.6653.680.050.120.020.030.280.15 101.11
Avg0.000.010.020.001.2146.0552.780.010.040.000.010.050.07 100.25
Minat.%bdlbdlbdlbdl0.010.991.97bdlbdlbdlbdlbdlbdl 3.00
Max0.000.000.000.000.031.002.000.000.000.000.000.010.00 3.00
Avg0.000.000.000.000.020.991.980.000.000.000.000.000.00 3.00
FormulaFe0.99S1.98 (to 3 atoms p.f.u.)
Cpy (n = 10)
Minwt.%bdlbdlbdl0.00bdl27.2733.5432.060.00bdl0.01bdl0.04 97.90
Max0.050.030.050.040.0331.0534.7335.930.100.121.520.010.07 99.48
Avg0.010.010.010.020.0129.2434.3034.450.040.050.380.000.05 98.57
Minat.%bdlbdlbdl0.00bdl0.921.980.940.00bdl0.00bdl0.00 4.00
Max0.000.000.000.000.001.042.021.070.000.000.040.000.00 4.00
Avg0.000.000.000.000.000.982.001.010.000.000.010.000.00 4.00
FormulaCu1.01Fe0.98S2.00 (to 4 atoms p.f.u.)
Note: bdl denotes the detection limit. Abbreviations: Ng, native gold; Min, minimum; Max, maximum; Avg, average. For other abbreviations, see Figure 3.
Table 2. Microthermometric data of fluid inclusions in quartz from the Nanmingshui gold deposit.
Table 2. Microthermometric data of fluid inclusions in quartz from the Nanmingshui gold deposit.
Stage/TypeRangeVapor
(vol.%)
XCO2Tm-CO2
(°C)
Tm-cla
(°C)
Th-CO2
(°C)
Tm-ice
(°C)
Th
(°C)
Salinity
(wt.% NaCl eqv.)
ρCO2
(g/cm3)
ρtotal
(g/cm3)
Early stage
C-typeMin300.11−58.28.918.8 2570.40.600.80
Max600.31−57.49.829.9 3392.20.791.01
Avg360.15−57.89.523.6 2981.10.730.91
n1918101518 19151818
W-typeMin20 −2.22202.6 0.66
Max30 −1.53383.7 0.87
Avg26 −1.92763.2 0.77
n9 797 7
Middle stage
C-typeMin200.07−62.27.512.9 2090.40.660.82
Max800.58−56.09.828.8 3614.80.840.99
Avg380.16−58.39.022.5 2772.00.740.91
n6969366469 69646969
W-typeMin15 −3.71900.9 0.69
Max70 −0.53496.0 0.91
Avg25 −2.22493.6 0.82
n125 9512595 95
PC-typeMin10 −61.5 12.3 0.80
Max30 −57.2 19.0 0.86
Avg21 −59.0 16.1 0.82
n7 7 7 7
Late stage
W-typeMin15 −4.31721.4 0.79
Max35 −0.82876.9 0.95
Avg23 −1.82133.0 0.88
n51 315131 31
Table 3. H-O isotopic compositions of auriferous quartz from the Nanmingshui gold deposit (‰).
Table 3. H-O isotopic compositions of auriferous quartz from the Nanmingshui gold deposit (‰).
Sample No.MineralδDwaterδ18OmineralT (°C)δ18OwaterReference
L1506Quartz−80.111.62331.8This study
L15114Quartz−68.714.02615.5
L15197Quartz−70.713.52373.9
NMS4-29Quartz−95.014.12696.0[26]
NMS4-30Quartz−92.019.026910.9
NMS7-6Quartz−82.014.82536.0
NMS7-7Quartz−99.019.725310.9
SQD-84Quartz−99.013.32564.6[31]
6N-53Quartz−62.913.92555.2[64]
Average −83.314.9 6.1
Note: T-average homogenization temperatures of fluid inclusions for each sample.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ge, Z.; Gu, X.; Zhang, Y.; Ma, C.; Hao, D.; Zheng, Y.; Zhang, X.; Wang, L.; Liu, M.; Chen, W. Ore Mineralogy, Fluid Inclusion Geochemistry, and Zircon U-Pb Geochronology of the Nanmingshui Gold Deposit in East Junggar, Xinjiang, Northwest China: Implications for Ore Genesis. Minerals 2023, 13, 1296. https://doi.org/10.3390/min13101296

AMA Style

Ge Z, Gu X, Zhang Y, Ma C, Hao D, Zheng Y, Zhang X, Wang L, Liu M, Chen W. Ore Mineralogy, Fluid Inclusion Geochemistry, and Zircon U-Pb Geochronology of the Nanmingshui Gold Deposit in East Junggar, Xinjiang, Northwest China: Implications for Ore Genesis. Minerals. 2023; 13(10):1296. https://doi.org/10.3390/min13101296

Chicago/Turabian Style

Ge, Zhanlin, Xuexiang Gu, Yongmei Zhang, Cheng Ma, Di Hao, Yanrong Zheng, Xiaoxing Zhang, Luzhi Wang, Ming Liu, and Weizhi Chen. 2023. "Ore Mineralogy, Fluid Inclusion Geochemistry, and Zircon U-Pb Geochronology of the Nanmingshui Gold Deposit in East Junggar, Xinjiang, Northwest China: Implications for Ore Genesis" Minerals 13, no. 10: 1296. https://doi.org/10.3390/min13101296

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop