Next Article in Journal
A Study of Thermal Stability of Hydroxyapatite
Next Article in Special Issue
Metallogenesis of the Bawang Sn-Zn Polymetallic Deposit, Wuxu Ore Field, Guangxi, South China: U-Pb Dating and C-O-S-Pb Isotopic Constraints
Previous Article in Journal
Characterization of Natural Consolidated Halloysite Nanotube Structures
Previous Article in Special Issue
Texture and Geochemistry of Scheelites in the Tongshankou Deposit in Daye, Hubei, China: Implication for REE Substitution Mechanism and Multistage W Mineralization Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluid Evolution and Ore Genesis of the Juyuan Tungsten Deposit, Beishan, NW China

1
Institute of Geomechanics, Chinese Academy of Geological Sciences, Beijing 100081, China
2
Key Laboratory of Paleomagnetism and Tectonic Reconstruction, Ministry of Natural Resources, Beijing 100081, China
3
Xinjiang Institute of Geological Survey, Urumqi 830000, China
*
Author to whom correspondence should be addressed.
Minerals 2021, 11(12), 1309; https://doi.org/10.3390/min11121309
Submission received: 6 October 2021 / Revised: 10 November 2021 / Accepted: 15 November 2021 / Published: 24 November 2021

Abstract

:
The newly discovered Juyuan tungsten deposit is hosted in Triassic granite in the Beishan Orogen, NW China. The tungsten mineralization occurred as quartz veins, and the main ore minerals included wolframite and scheelite. The age, origin, and tectonic setting of the Juyuan tungsten deposit, however, remain poorly understood. According to the mineralogical assemblages and crosscutting relationships, three hydrothermal stages can be identified, i.e., the early stage of quartz veins with scheelite and wolframite, the intermediate stage of quartz veinlets with sulfides, and the late stage of carbonate-quartz veinlets with tungsten being mainly introduced in the early stage. Quartz formed in the two earlier stages contained four compositional types of fluid inclusions, i.e., pure CO2, CO2-H2O, daughter mineral-bearing, and NaCl-H2O, but the late-stage quartz only contained the NaCl-H2O inclusions. The inclusions in quartz formed in the early, intermediate, and late stages had total homogenization temperatures of 230–344 °C, 241−295 °C, and 184−234 °C, respectively, with salinities no higher than 7.2 wt.% NaCl equiv (equivalent). Trapping pressures estimated from the CO2-H2O inclusions were 33−256 MPa and 36−214 MPa in the early and intermediate stages, corresponding to mineralization depths of 3–8 km. Fluid boiling and mixing caused rapid precipitation of wolframite, scheelite, and sulfides. Through boiling and inflow of meteoric water, the ore-forming fluid system evolved from CO2-rich to CO2-poor in composition and from magmatic to meteoric, as indicated by decreasing δ18Owater values from early to late stages. The sulfur and lead isotope compositions in the intermediate-stage suggest that the Triassic granite was a significant source of ore metals. The biotite 40Ar/39Ar age from the W-bearing quartz shows that the Juyuan tungsten system was formed at 240.0 ± 1.0 Ma, coeval with the emplacement of granitic rocks at the deposit. Integrating the data obtained from the studies including regional geology, ore geology, biotite Ar-Ar geochronology, fluid inclusion, and C-H-O-S-Pb isotope geochemistry, we conclude that the Juyuan tungsten deposit was a quartz-vein type system that originated from the emplacement of the granites, which was induced by collision between the Tarim and Kazakhstan–Ili plates. A comparison of the characteristics of tungsten mineralization in East Tianshan and Beishan suggests that the Triassic tungsten metallogenic belt in East Tianshan extends to the Beishan orogenic belt and that the west of the orogenic belt also has potential for the discovery of further quartz-vein-type tungsten deposits.

1. Introduction

Tungsten deposits in China are widely distributed in the South China block, the Himalaya, Tibetan, Sanjiang, Kunlun, Qilian, Qinling, Dabie, and Sulu orogens, as well as the Central Asian orogenic belt. The tungsten resource occurs mainly as skarn (63%), quartz-vein (17%), porphyry (17%), and greisen (3%) deposits in China [1]. The quartz-vein tungsten deposits are concentrated mostly in the South China block and are associated with Jurassic granitic magmas [2,3,4,5,6,7,8,9]. However, in recent years, a cluster of tungsten deposits associated with Triassic intrusions was discovered in the East Tianshan-Beishan belt of the Central Asian orogenic belt (CAOB) [10,11]. These tungsten deposits, in which the orebodies are mainly quartz-vein and skarn type, are currently under exploration. Most previous studies of these tungsten deposits have focused on geological characteristics, metallogenic granite chronology and geochemistry, metallogenic chronology, and metal source [12,13,14,15,16,17,18,19,20,21]. These studies have confirmed the existence of a Triassic tungsten metallogenic belt in East Tianshan-Beishan, China.
The Juyuan W deposit, which is situated in the western part of the Shuangyingshan–Huaniushan arc, is a quartz-vein tungsten deposit and was discovered in 2018 by the Xinjiang Institute of Geological Survey. The ore bodies occur mainly in the form of quartz veins and are hosted in Triassic monzonitic granite and two-mica granite. Previous studies of the deposit have focused mainly on the ages of the host granite, the tectonic setting of the granite inferred from geochemical data, and the trends of quartz veins to determine the principal stress direction applied to the deposit [22,23]. However, few studies have focused on the origin and evolution of ore-forming fluids of the deposit, and there is a lack of systematic research on fluid inclusions (FIs).
In this paper, we report a systematic study of the Juyuan tungsten deposit based on FIs in quartz C-H-O isotopes from hydrothermal minerals and S-Pb isotopes of sulfide minerals to investigate the nature of ore-forming fluids and deposited metals, the evolution of the fluids, and the mechanism of wolframite and scheelite deposition. In addition, we used Ar-Ar dating of muscovite to establish the formation age of the deposit.

2. Regional Geology

The CAOB is one of the largest accretionary orogens in the world and was formed by multiple subduction-accretion and collisional events associated with the closure of the Paleo-Asian Ocean [24,25,26]. Tectonically, the CAOB is bounded by the Siberian and Eastern European cratons to the north and the Tarim and North China cratons to the south [27,28,29].
As a significant part of the southern CAOB, the Beishan orogenic belt is located between the Southern Tian Shan suture to the west and the Solonker suture to the east and is tectonically regarded as the eastern extension of the Chinese Tian Shan [27,30]. The Southern Tian Shan suture is on the west of the Beishan orogenic belt, and the Solonker suture is on the east of Beishan orogenic belt. They are bordered but the boundary is not clear. Geographically, Beishan is a mountainous area in western Gansu Province and to the east of Xinjiang Uygur Autonomous Region.
The Beishan orogenic belt comprises Precambrian crystalline basement and overlying sedimentary rocks (Figure 1). The ages of the constituent rocks range mainly from late Precambrian to Mesozoic, including Precambrian metasedimentary rocks and Paleozoic arcs, ophiolitic mélanges, accretionary complexes, and arc-related basin sediments [31,32,33].
The E-W-trending tectonic units of the Beishan orogenic belt are separated by E-W-trending strike-slip faults and ophiolitic mélanges and are cut by later NE-SW-trending strike-slip faults with complicated shear senses [27,28,33].
The orogenic evolution of the Beishan belt can be divided roughly into pre-orogenic (Precambrian–early Cambrian), ocean–continental (Ordovician–Permian), and post-orogenic (Triassic) evolutionary periods. Based on the geological, geochemical, tectonic, and geophysical data, the Beishan orogenic collage can be subdivided into several tectonic units, which from north to south are the Queershan arc, Hongshishan ophiolitic mélange, Heiyingshan-Hanshan arc, Shibanjing-Xiaohuangshan ophiolitic mélange, Mazongshan arc, Hongliuhe-Niujuanzi-Xichangjing ophiolitic mélange, Shuangyingshan-Huaniushan arc, Liuyuan ophiolitic mélanges, and Shibanshan arc. The Heiyingshan-Hanshan arc and the Mazongshan and Shuangyingshan-Huaniushan arcs constitute the Gongpoquan arc-accretionary system [18,27,28,33]. These terranes consolidated and evolved at various times from the Cambrian to the Permian and were affected by multiple periods of magmatism, resulting in large numbers of granitic intrusions and mafic volcanic rocks, accompanied by lesser numbers of mafic–ultramafic intrusions.
Two episodes of magmatism related to coeval tungsten mineralization have been recognized in the Beishan orogenic belt: (1) 424–314 Ma (e.g., Hongjianbingshan, Guoqing, Yingzuihongshan, and Baixianishan deposits); the related intrusions were derived from the partial melting of Precambrian metasedimentary rock or juvenile lower crust [12,15,16,21,34,35]; (2) 286–220 Ma (e.g., Xiaobaishitou, Liushashan and Yushan deposit); the granites were derived from the ancient crustal sediments [17,21,36,37,38,39,40]. These two magmatic events are mainly distributed in the terranes with Precambrian metasedimentary strata, Shuangyingshan-Huaniushan arc, and Hanshan block.

3. Deposit Geology

The representative strata in the Juyuan tungsten ore district comprise mainly the Pingtoushan Formation of the Jixian System and the Kuqa Formation of the Tertiary System (Figure 2). The Pingtoushan Formation is composed of fragmented grayish marble with well-developed schistosity and strong weathering. The Tertiary Kuqa Formation is mainly reddish argillaceous clay that weathers to reddish-brown loose soil.
The structure of the ore district is relatively simple (Figure 3). There are no obvious faults, and quartz-vein dislocation can be seen locally, with displacements of no more than 1 m. Intrusive rocks are common in the ore district and include monzogranite, granite, and numerous dikes. The shape of monzogranite and granite is plutonic rock, and the dikes are mainly diabase dikes, with a small amount of quartz porphyry. The main intrusive phase is represented by Triassic monzogranite (Figure 4a), which is located in the central and northeastern parts of the ore district. The zircon U-Pb age of the monzogranite is 240.7 ± 0.5 Ma (unpublished data). The monzogranite is red in color, medium grained, and composed of quartz, with equal proportions of K-feldspar and plagioclase, and biotite, with accessory zircon, magnetite, apatite, and other minerals. The monzogranite shows argillitization and sericitization. Granite is distributed mainly in the north of the ore district and has a medium-grained granitic texture and massive structure. The surface of the granite is mostly weathered into granules, and ferritization and kaolinization are locally developed. Quartz porphyry, located in the central part of the ore deposit and running through the monzogranite (Figure 4a), is porphyritic and composed of quartz and K-feldspar as phenocrysts in a groundmass of quartz, K-feldspar, and plagioclase. Phenocrysts of K-feldspar have undergone intense argillitization, and the fine groundmass displays sericitization. Diabase and diorite dikes are also developed in the ore district. Diabase dikes are mostly 100–500 m in length and show weak chloritization. Diorite dikes are widely distributed, with lengths of 80–300 m, cutting through diabase dikes locally. The main mineral composition of diabase are pyroxene and basic plagioclase, as well as a small amount of biotite, apatite, magnetite, etc. The mineral composition of diorite is mainly amphibole and plagioclase, and the accessory minerals are mainly apatite, magnetite, and sphene.
At present, 22 ore bodies can be identified in the ore district (Figure 2). The tungsten ore body is 80–420 m long and 0.4–2.8 m wide, with a strike of 124°–152° and an overall south dip of 65°–80° (Figure 3). The tungsten grade 0.065–0.14%. The orebodies occur mainly in quartz veins in monzogranite. Wolframite and scheelite are the major ore minerals of the Juyuan tungsten deposit (Figure 4e,f,h,i), with a small amount of galena, pyrite, and other sulfides. Wolframite is mostly euhedral, with grain sizes ranging from 1 mm to 15 mm. Scheelite occurs in two forms: (1) euhedral, with a grain size smaller than that of wolframite, up to 5 mm (Figure 4i), and (2) distributed in fissures, with a shape structure, which commonly cuts quartz and wolframite. The ore structure is star shaped, vein shaped, or sparsely disseminated.
According to the occurrence and cross-cutting relationships of quartz veins, combined with the characteristics of microstructure and mineral assemblages, the quartz veins in the ore district can be divided into three stages. Early-stage quartz veins are ore veins and are generally gray white in color. Veins are 0.5–2 m thick, and the metallic minerals in the veins appear mainly as wolframates (e.g., wolframite and scheelite). Intermediate-stage quartz veins are smokey gray or gray white in color, and sulfides (e.g., galena and pyrite) (Figure 4l) are more abundant in the veins compared with the early stage. The quartz veins of the intermediate stage cut through the early stage of mineralized quartz veins. Late-stage quartz veins (post-metallogenic quartz veins) are white and 0.1–0.2 m thin. Calcite is present in the quartz veins. These late veins cross-cut the smokey gray quartz veins of the intermediate ore-bearing hydrothermal stage and no obvious metallic minerals are observed in them.

4. Samples and Analytical Methods

4.1. Fluid Inclusion

We undertook a microthermometric FI study of quartz crystals associated with the three different stages of mineralization in quartz veins. More than 150 double-polished thin-sections (200 μm thick) were prepared for FI analysis. Fluid inclusions were carefully observed to identify their genetic types, vapor-liquid ratios, and spatial clustering. Nineteen typical samples were selected for microthermometric measurements and laser Raman spectroscopy analysis.
Microthermometric measurements were conducted at the Institute of Geomechanics, Chinese Academy of Geological Sciences (CAGS), Beijing, China, using a Linkam THMSG 600 heating-freezing stage with a temperature range of −195 °C to +600 °C, attached to a Leitz Ortholux transmitted-light microscope connected to a television camera and screen. The stage was calibrated using synthetic FIs made by Fluid Inc., Oakland, CA, USA. The precision of the measured temperature was ±0.1 °C in the range of −195 °C to 30 °C and ±1 °C in the range of 30 °C to 600 °C. The heating rate was kept at 0.2 °C to 5.0 °C per minute and when approaching the phase-change conditions, the heating rate was changed into 0.2 °C per minute. To avoid inclusion decrepitation, the freezing processes were operated first. The salinities for aqueous fluids were calculated using the equations of [41] and for aqueous-carbonic fluids using the equations of [42,43]. According to the microthermometic data [44,45], the CO2-H2O-NaCl system was defined, and the densities and pressures were determined using the Flincor software package.
Laser Raman spectroscopic analysis of single fluids including vapor and liquid phases was done at the Raman Laboratory of the Beijing Institute of Uranium Geology, Beijing, using a LABHR-VIS LabRAM HR800 Laser Raman microspectrometer, (Paris, France). The analysis used an argon atom laser with a wavelength of 532 nm. The spectrum range was set between 100 cm−1 and 4200 cm−1. The accumulation time for each scan was 5 s, and the spectrum resolution was 2 cm−1.

4.2. Isotope Analysis

The samples for isotope analyses were obtained from different mineralization stages, confirmed by microscopy observation. The minerals were separated by handpicking under a binocular microscope to achieve >99% purity. Analyses of isotopes of O, H, C, and S were performed on a Finnigan MAT253 mass spectrometer, (Waltham, MA, USA) at the Analytical Testing Center of the Beijing Research Institute of Uranium Geology, using the methods outlined by [46].
For O isotope analyses of quartz, the quartz was ground to 200 mesh and then dried. Oxygen gas was liberated by reaction with BrF5 [47] in externally heated nickel vessels and converted to CO2 on a carbon rod. Isotope ratios (δ18O) were determined, with all values being reported relative to the V-SMOW standard, with an uncertainty of ±0.2‰.
Hydrogen isotope analyses were conducted on the same samples as investigated for fluid inclusions hosted in the quartz. The selected quartz samples were dried at a low temperature to remove the adsorbed water on the surface of the mineral, and secondary inclusions were removed by gradual heating under vacuum. The fluid inclusions decrepitated when the temperature was increased to 600 °C. Hydrogen in the liberated water was replaced by zinc and released for mass spectrometry analysis. The values are reported relative to the V-SMOW standard, and the uncertainty was better than ±1‰.
The method of analysis of C isotopes was similar to that of H isotopes. Quartz samples were dried and placed in a vacuum system to remove the gas generated by the burst of secondary inclusions during the process of heating. After heating to 600 ℃, the fluid inclusions were burst, and CO2 was liberated. CO2 was separated by combined cold traps and then sent for mass spectrometry analysis. We reported the C isotope data per mil relative to the Peedee Belemnite (PDB) standard, with an uncertainty of ±0.2‰.
Four galena and four pyrite samples were analyzed for S and Pb isotopic compositions at the Analytical Testing Center of the Beijing Research Institute of Uranium Geology.
For S isotope analysis, each sample was weighed to 15 mg, mixed with about 150 mg of CuO powder, ground to 200 mesh, and placed in a vacuum quartz tube. The powder was heated to 980 °C when the vacuum reached 2.0 × 10−2 Pa, and then SO2 was produced. After purification, SO2 was transferred to the sample tube, and the 34S/32S ratio was measured with a Finnigan MAT 251 mass spectrometer. The isotope data were reported relative to the Canyon Diablo Troilite (CDT) standard, and the uncertainty of δ34S values is ±0.2‰.
The O isotope ratios of water (δ18OW) were calculated using oxygen-isotope fractionation equilibrium formulas reported by [47].

4.3. Muscovite 40Ar/39Ar Dating

The muscovite grains were hand-picked carefully under a binocular microscope, followed by soaking in 5% dilute nitric acid. The separates were then washed repeatedly in the ultrasonic bath using acetone and deionized water for 15 min and placed in an oven to dry at 60 °C. Aliquots of ~10.8 mg were wrapped in aluminum foil and sealed in a quartz tube along with several neutron fluence monitors (ZBH-25 biotite) for calculation of the fluence gradient. The quartz tube was then vacuum-sealed and irradiated for 24 h at the HFETR facility of the Northwest Institute of Nuclear Technology, Beijing, China. Step-heating 40Ar/39Ar analyses were conducted using an Argus VI noble gas mass spectrometer in the Argon Laboratory of the Beijing Research Institute of Uranium Geology. The irradiated separates were incrementally heated by a double-vacuum tantulum furnace from 800 °C to total fusion at 1300 °C at a rate of 40 °C to 200 °C every 10 min in eight temperature steps to 1300 °C. Then, the gas released from the sample was purified by a stainless U-type trap with liquid nitrogen, and by two Zr-Al getters, with one at 450 ℃ and the other at room temperature, leaving only the noble gases, which entered the noble gas mass spectrometer for Ar isotope analysis. The interference correction values of the reactor, which were derived from irradiated CaF2 and K2SO4, are (36Ar/37Ar)Ca = 0.0003011, (39Ar/37Ar)Ca = 0.0007694, and (40Ar/39Ar)K = 0.0020826. J values for each irradiation position were calculated from the 40Ar/39Ar ratios by total fusion of the ZBH-25 biotite, using an assumed age of 132.9 ± 1.3 Ma [48]. The mass discrimination factor was calculated according to the atmospheric Ar ratio of 295.5 ± 0.5 [49], and the decay constant for 40K used in the calculation was 5.543 × 10−10 year−1 [50]. The data-processing software used was ArArCALC 2.4 [51].

5. Fluid Inclusion Geochemistry

5.1. Fluid Inclusion Petrography

Through phase relationships, relative phases present at room temperature (21 °C), phase changes during freezing processes, and laser Raman spectroscopic measurements [52], four types of FI were identified (Figure 5), namely, CO2-H2O or carbonic-aqueous (C-type), pure carbonic or CO2 (PC-type), aqueous (W-type), and fluid inclusion with daughter minerals (S-type).
Petrographic criteria suggested by [52] were used to distinguish primary and secondary FIs. Fluid inclusions occurring in growth zones or as isolated clusters in crystal cores are regarded as primary.
C-type were present in quartz from early- and intermediate stages. They usually occurred as two or three phases at room temperature (liquid CO2 + liquid H2O or vapor CO2 + liquid CO2 + liquid H2O); Figure 5a–e), with the CO2 phase volumetrically occupying typically 10–95 vol.%. When the temperature dropped to ≤10 °C, the vapor CO2 was observed in two-phase. The C-type FIs were commonly isolated individuals or appeared as clusters of randomly arrangement inclusions. They were variable in shape, including subrounded shapes and polygonal forms, with sizes ranging from 4 μm to 27 μm. According to the vapor/liquid ratio by volume, C-type FIs can be further divided into liquid-rich fluids (vapor/liquid < 50 vol.%; Figure 5a) and the vapor-rich fluids(vapor/liquid > 50 vol.%; Figure 5a). The liquid-rich fluids were more common and accounted for more than 90% of the total fluid (Figure 5).
PC-type FIs were rare and showed generally oval to subrounded shapes with a uniform grain size about 6–7 um. They appeared as monophase (liquid CO2) at room temperature (Figure 5d). During cooling, the monophase inclusions developed a CO2-rich vapor bubble.
W-type FIs occurred in all three stages, and at room temperature they usually appeared as two-phase (liquid H2O + vapor H2O) fluid inclusions (Figure 5d–g), with a vapor/liquid ratio of about 10–45 vol.%. Morphologically, the fluid inclusions (5–13 μm) showed subrounded, oval, or irregular forms. A small amount of CO2 was identified in the vapor bubbles by laser Raman spectroscopy, although the CO2 phase was not observed during heating or cooling runs. The primary W-type FIs occurred as isolated individuals or as clusters, but the secondary W-type FIs often occurred along healed fractures, showing irregular forms.
S-type FIs were scarce and occurred only in a few cases in early-stage quartz samples, appearing as three-phase inclusions (liquid CO2 + liquid H2O + rectangular-shaped daughter minerals; Figure 5b). The content of daughter minerals in the S-type FIs was low and unknown. The low-temperature phase transformation behavior of the inclusions was similar to that of W-type FIs. The daughter minerals did not melt at the bursting temperature, indicating that daughter minerals may have been captured accidentally during the formation of S-type FIs.

5.2. Microthermometry

5.2.1. Early-Stage Fluid Inclusions

The early-stage quartz included all four types of FIs identified in the Juyuan tungsten deposit (Table 1), with C-type FIs being the most common. The melting temperatures of solid CO2 (Tm-CO2) were between −69.6 °C and −56.7 °C, below the triple point of pure CO2 (−56.6 °C), implying the presence of minor amounts of dissolved components in the carbonic phase [52,53]. The melting temperatures of clathrates (Tm-cl) ranged from 4.0 °C to 10.0 °C. The carbonic phase (Th-CO2) in liquid-rich FIs was partially homogenized to liquid at 15.9 °C to 31.1 °C, and the vapor-rich FIs to vapor at 17.2 °C to 31.1 °C. The C-type FIs were totally homogenized to liquid or vapor at temperatures ranging from 230 °C to 344 °C (Th), and mostly in the range of 280–300 °C (Figure 6), although many vapor-rich inclusions decrepitated prior to final homogenization at about 300 °C. The CO2 densities calculated using the formula of [44] ranged from 0.52 g/cm3 to 0.81 g/cm3 and bulk densities from 0.61 g/cm3 to 0.99 g/cm3. Raman analysis indicated that CO2 and H2O were the principal species, with small amounts of CH4 (Figure 7a).
The PC-type FIs displayed two phases and were frozen to solid below −100 °C. The solid part melted at −60.2 °C to −63.2 °C. The Th-CO2 values of CO2 (L + V to L) were between 15.9 °C and 29.1 °C, with corresponding densities of 0.63 g/cm3 to 0.81 g/cm3.
The W-type FIs in early-stage quartz yielded ice-melting temperatures, measured at −2.8 °C to −7.4 °C, with salinities ranging from 4.7 to 11.0 wt.% NaCl equiv. They were homogenized to the liquid phase at 230 °C to 282 °C (Figure 6).
An S-type FI with an unidentified solid was observed, with an ice-melting temperature of −6.8 °C, a salinity of 10.2 wt.% NaCl equiv., and homogenization to the liquid phase at a temperature of 246 °C. The daughter mineral did not dissolve during heating to 500 °C, so its homogenization temperature was not observed.
As described above, coexisting C-type inclusions with different vapor/liquid ratios in the early-stage quartz had similar homogenization temperatures but commonly had different salinities, implying that fluid immiscibility occurred during the early stage of mineralization of the Juyuan deposit or that the FIs were trapped at the same time and thus represent an FIA.

5.2.2. Intermediate-Stage Fluid Inclusions

The intermediate-stage quartz, which was closely associated with polymetallic sulfide, contained many FIs. The FIs in this stage quartz were dominantly C-type with lesser PC-type. The Tm-CO2 values of C-type FIs were measured at −75.5 °C to −57.2 °C, below the triple point of pure CO2 (−56.6 °C). The CO2 clathrate (Tm-cl) melted at a temperature of 6.3 °C to 8.7 °C, and the corresponding salinities reached 6.9 wt.% NaCl equiv. The CO2 phases were totally homogenized to liquid at 30.0–31.1 °C (Th-CO2). The CO2 densities were calculated at 0.60 g/cm3 to 0.67 g/cm3 and bulk densities at 0.76 g/cm3 to 0.96 g/cm3. They were generally homogenized to liquid or vapor at 241–295 °C, peaking at 260–280 °C.
PC-type FIs were rarely observed in quartz of the intermediate stage. They yielded Tm-CO2 values of −76.3 to −62.0 °C and homogenized to liquid at 24.7 °C to 30.0 °C (Th-CO2). Raman analysis shows that the component of these FIs was dominated by CO2 and H2O (Figure 7c).

5.2.3. Late-Stage Fluid Inclusions

Only W-type FIs were observed in late-stage quartz. Their Tm-ice values were observed at −3.2 °C to −8.5 °C (Table 1), and corresponding salinities at 4.6 to 12.3 wt.% NaCl equiv. They were totally homogenized to liquid at 184–234 °C (Figure 6). The marked decrease in the abundance of C-type FIs from the intermediate to late stage indicates that the fluid system transformed from a relatively closed state to a relatively open state, which enabled a large amount of CO2 gas to escape and allowed a large volume of meteoric water to enter and mix with the ore-forming fluid.

5.3. Trapping Pressure and Mineralization Depth

Measurements of the partial and total homogenization temperatures and the ratios of vapor CO2 and liquid CO2 of C-type FIs allowed the minimum trapping pressures of FIs to be estimated using the Flincor program and the formula of [44,45] for the H2O-CO2-NaCl system. The calculated minimum trapping pressures of FIs of the early stage were estimated as 33–256 MPa, and those of the intermediate stage as 36–214 MPa, with the CO2 densities and bulk densities varying from 0.61 g/cm3 to 0.63 g/cm3 and 0.87 g/cm3 to 0.89 g/cm3, respectively.
Therefore, the minimum pressures estimated for the early-stage corresponded to a hydrostatic depth of about 3 km. The maximum pressure corresponded to a lithostatic depth of about 8 km, given that the density of the overburden rocks of the Triassic granite was ca. 2.90 g/cm3. Hence, the W mineralization depth of Juyuan deposit was about 3–8 km.

5.4. Laser Raman Spectroscopy

Phases in representative FIs were measured using laser Raman microspectroscopy to constrain their compositions. Results of the analyses show that vapor bubbles in FIs in early-stage quartz contained CO2, H2O, and CH4 (Figure 7a) and the liquid phases contained CO2 and H2O (Figure 7b). This is consistent with the microthermometric analyses results, which showed that the melting temperatures of solid CO2 were lower than −56.6 °C.
Vapor bubbles of C-type FIs in intermediate-stage quartz were dominated by H2O and CO2 with no CH4 detected by laser Roman spectroscopy (Figure 7c). Raman spectrum for the liquid and vapor of W-type inclusions in late-stage quartz was almost entirely H2O. However, the microthermometric measurement and the Laser Raman spectroscopic detection showed the absence of CO2 in any of the late-stage fluid inclusions (Figure 7), indicating that the fluid system evolved from CO2-rich to CO2-absent aqueous solution.

6. Isotope Geochemistry

6.1. Hydrogen and Oxygen

The H and O isotope compositions of selected minerals from different ore-forming stages were analyzed by determining the source of the mineralizing fluids. The O isotope data of fluids in equilibrium with quartz were calculated using the quartz-water equation of [47]. The analytical data and the calculated δ18Owater values of ore fluids are listed in Table 2. Measured δ18O values from the early stage ranged from 12.2‰ to 13.6‰. Using the equation, and corresponding fluid inclusion homogenization temperature data, the δ18O values for fluids from early-stage quartz were calculated to range from 4.8‰ to 6.0‰. The same calculation method applied to intermediate- and late-stage quartz yielded δ18O values of 4.2‰ to 6.0‰ and 1.3‰ to 0.5‰, respectively.
As the homogenization temperature of FIs provide only a lower limit of the trapping temperature, unless the FIs were entrapped from an immiscible or boiling fluid [47], the δ18O values of late-stage fluids calculated using FI homogenization temperatures may have been underestimated.
In a δD–δ18O diagram (Figure 8), the data from early-stage quartz and intermediate-stage approached those of the magmatic water box, with the δ18O values of the late-stage deviating towards the meteoric water line, suggesting that fluid in the Juyuan tungsten deposit was likely of magmatic origin, with possible later mixing of meteoric water.

6.2. Carbon

Fluid inclusions in different-stages quartz from the Juyuan deposit yielded δ13CCO2 values of −14.1‰ to −1.5‰, with a mean of −6.3‰ (Table 2). Comparing the values of various available carbon reservoirs, such as organic matter (−27‰; [48]), atmospheric CO2 (−8‰; [55]), dissolved CO2 in fresh water (−9‰ to −20‰; [56]), continental crust (−7‰; [57]), mantle (−5‰ to −7‰; [56]), and igneous rocks (−3‰ to −30‰; [56]), we concluded that the carbon of the fluid could not have been supplied by any of the above carbon reservoirs alone, instead showing a value range characteristic of mixed sources. Two carbon values (−2.3‰ to −1.5‰) of samples from the Juyuan deposit early-stage are higher than those of other carbon reservoirs but fall in the range of marine carbonate (−3‰ to 2‰ [56]; Figure 9). The higher δ13CO2 values may be due to diagenesis and dehydration of marine carbonates. Therefore, at least a portion of CO2 in ore fluids of the Juyuan deposit probably originated from Pingtoushan formation marine carbonates.

6.3. Sulfur

Sulfides formed in different geochemical systems have different isotopic compositions. For determining the sources of hydrothermal deposits, the sulfur isotopes of sulfides are known to be a key tracer [63,64,65,66,67]. Sulfur isotope compositions were obtained for intermediate-stage pyrite and galenite. Values of the eight δ34SV-CDT determinations made in this study are listed in Table 3 and presented in Figure 10. The results of the isotope analyses of pyrite were remarkably homogeneous, ranging from +6.2‰ to +6.3‰ (n = 4). Galenite separates were also fairly homogeneous, with δ34SV-CDT values ranging from +3.7‰ to +4.4‰ and a mean of 4.0‰ (n = 4).
Sulfur isotope was much more enriched in pyrite than in galena, suggesting sulfur isotope equilibrium between the coexisting pyrite and galena in intermediate-stage quartz [63]. The mineral assemblage in this stage suggest a reducing environment, and it is agreement with the fact that no sulfate mineral was observed in our study. This means that the sulfur isotopes of these sulfides should reflect the sulfur isotopic signatures of the hydrothemal fluids [63].
Sulfides from the intermediate stage had reasonably homogeneous δ34S values (+3.7‰ to +6.3‰), suggesting that the ore-forming fluids were derived mainly from crustal rocks. These values are higher than the average values of magmatic sources (mainly −3‰ to 3‰ δ34S) and the deep mantle (~0‰; [63]), and fall within the range of crust-derived granite (−4.0‰ to 9.0‰; [62]), indicating a crustal source (Figure 10). During the process of sulfide crystallization, sulfides underwent a degree of exchange reaction with the surrounding rock.

6.4. Lead

Lead isotope data for selected intermediate-stage pyrite and galena from the Juyuan deposit had 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb values of 18.476–18.538, 15.640–15.706, and 38.437–38.648, with mean values of 18.509, 15.673, and 38.541 (Table 4), respectively. In Figure 11, the Pb isotope data plot between the orogen and lower crust fields or orogen and upper crust fields showed the complexity of lead sources. The sulfide Pb isotopes yielded two-stage model ages of 209–161 Ma, suggesting that the metal was likely sourced from Late Triassic–Middle Jurassic strata (Figure 11). The two-stage model ages of sulfides are younger than the timing of tungsten mineralization (240.0 ± 1.0 Ma), suggesting that the formation of sulfides postdated the tungsten deposition at Juyuan.

6.5. Muscovite 40Ar/39Ar Dating

Step-heating 40Ar-39Ar analytical results for muscovite (Sample D18105-5) are reported in Table 5 and displayed in Figure 12. Uncertainties of individual step-heating analyses and on the plateau age are reported at the 1σ level. The age spectrum shows a flat plateau that accounted for 100% of the 39Ar released, indicating that K and radiogenic 40Ar were distributed homogeneously and that the K-Ar isotopic system remained closed and away from heating disturbance during the geological history of this sample. Nine continuous steps (800–1300 °C) for the muscovite sample yielded a well-defined weighted plateau age of 240.0 ± 1.0 Ma (1σ; MSWD = 2.2; Figure 12a), an isochron age of 238.3 ± 1.1 Ma (MSWD = 1.3) with an initial 40Ar/36Ar ratio of 440 ± 53 (Figure 12b), and an inverse isochron age of 238.3 ± 1.1 Ma (MSWD = 1.2) with an initial 40Ar/36Ar ratio of 441 ± 51 (Figure 12c). The 240.0 ± 1.0 Ma age is considered to be a reliable estimate for the crystallization age of muscovite in this sample from the Juyuan deposit.

7. Discussion

7.1. Fluid Nature, Evlution, and Tungsten Deposition

Fluid boiling is an important mechanism for ore-metal deposition in many hydrothermal deposits [71,72,73,74]. Fluid boiling occurred at the Juyuan deposit. Four different types of FIs coexisted in the early-stage quartz. The coexisting vapor-rich and liquid-rich FIs homogenized divergently to respective vapor and liquid phases at around the same temperature range (Table 1 and Figure 5a). It can be concluded that the boiling of fluids resulted in the early-stage precipitation of wolframite and scheelite in the Juyuan deposit, given that fluid boiling is inferred to have occurred in early-stage quartz. Fluid inclusions of the late-stage showed low homogenization temperatures, indicating cooling of the fluids, most likely reflecting the progressive influx and mixing of meteoric water. Of note, the late-stage fluids had relatively high salinity. A decrease in temperature from the early to late stage resembles the fluid-mixing trend proposed, but the slightly higher salinity in the late-stage may be due to high-salinity fluid that was carried by meteoric water during its descent (Figure 13). Furthermore, H-O isotope data (Figure 8) for the intermediate- and late-stage fluid systems plotted towards the meteoric water line, indicating the increasing input of meteoric water. This fluid mixing and/or inflow of meteoric fluids would have resulted in rapid deposition of polymetallic sulfides and gangue minerals.
In summary, boiling of the ore-forming fluids, mixing of the fluids with meteoric water, further pH increase, and decreasing temperature were likely the most important mechanisms that led to tungsten precipitation in the Juyuan deposit.
Various types of FI are present in quartz of different stages, providing insights into the nature and evolution of the fluid system of the Juyuan deposit. From the early to late stages, some regularities of fluid evolution are summarized below. The total homogenization temperatures of fluid inclusions decreased from 230–344 °C (mostly 280–300 °C) during the early-stage, through 241–295 °C (mostly 260–280 °C) during the intermediate-stage, and to 184–234 °C during the late-stage. Fluid cooling and influx of meteoric water occurred during the late-stage.
The C-, W-, S-, and PC-type inclusions in early-stage quartz indicate that the initial fluid system was carbonic and of low salinity. The existence of W-, C-, and PC-type inclusions in early- and intermediate-stage quartz show that the fluid system underwent boiling during these two stages.
It is common for C- and W-type inclusions to coexist in the same FIA [75], indicating that they were captured simultaneously. Microscope measurement results show that the C- and W-type FIs differed in their homogenization style. The W-type FIs totally homogenized to liquid, but the C-type FIs with greater vapor/liquid ratios (>50%) homogenized to vapor. The C-type FIs had higher homogenization temperatures than those of the W-type FIs, but lower salinity. These features indicate that the fluid became immiscible after being captured. The higher salinity of W-type inclusions compared with C-type inclusions can be explained by fluid immiscibility causing the CO2 phase that was dissolved in the fluid under higher pressure and temperature to become separated as a result of the decrease in pressure and temperature, leading to an increase in the salinity of the remaining fluid owing to the loss of gas.
Laser Raman spectroscopy analyses identified CO2 and H2O without CH4 in fluid inclusions from the intermediate-stage. We propose that an increase in oxygen fugacity, which occurred owing to inflow of oxidized meteoric waters, can explain the absence of CH4 (transition CH4 → CO2) [63]. Only W-type inclusions were observed in late-stage quartz, indicating that the late-stage fluid was a CO2-poor aqueous solution fluid.
Data points for δDwater–δ18Owater from the early-to late-stage shifted from the magmatic water field towards the meteoric water line (Figure 8), and a consistent feature of fluid inclusions is some anomalously low-δD fluids. As mentioned above, boiling of the early-stage fluid occurred, and vapor separation of ore-forming fluids by this boiling may explain the anomalously low-δD fluids [66]. Experimental studies have suggested that D fractionates preferentially into the vapor phase relative to H during fluid boiling, reducing the δDfluid values of the residual fluid by ~28‰ while scarcely changing the δ18Ofluid values [76,77].
It can be concluded that the fluid system evolved from a CH4-rich and CO2-rich system to one that was CO2 rich but lacked CH4. The system then evolved further into a CO2-poor and CH4-poor aqueous fluid system through CO2 escape, fluid boiling, and inflow of meteoric water. This conclusion is supported by H-O isotope signatures that indicate that the ore-forming fluids evolved from almost primary magmatic water to meteoric.

7.2. Timing and Tectonic Setting of Formation of the Juyuan Deposit

A series of Triassic tungsten deposits are found in the East Tianshan Mountains, including the Shadong (239 Ma) [18], Jiangshan (230 Ma) [18], and Xiaobaishitou (245 Ma) [40] tungsten deposits. These and other deposits form a Triassic tungsten metallogenic belt in the East Tianshan Mountains, whose metallogenic tectonic setting was an intra-plate extensional regime. Two episodes of magmatism related to coeval tungsten mineralization have been recognized in the Beishan orogenic belt: (1) 424–314 Ma; for example, the Hongjianbingshan, Guoqing, Yingzuihongshan, and Baixianishan deposits. The related intrusions were derived from the partial melting of Precambrian metasedimentary strata or juvenile lower crust [12,15,16,34,46]; and (2) 286–220 Ma; for example, the Liushashan (262–260 Ma; [36,37]), Huaheitan (226 Ma; [78]), and Yushan deposits [79]. The granites were derived from ancient crustal sediment [17,21,36,37,38,39,40,79]. Rocks of these two magmatic events are distributed mainly in terranes with Precambrian metasedimentary strata, the Shuangyingshan-Huaniushan arc, and the Hanshan block.
Structural data from Permian to Triassic rocks in regions adjacent to the Beishan belt indicate the occurrence of extensive thrusting, strike-slip faulting, and possibly also extension associated with the amalgamation of different blocks [80,81,82,83,84,85,86,87,88]. Furthermore, paleomagnetic, geochronological, and stratigraphic data show that the convergence of the archipelago and syn-collisional crustal shortening had ended by the end-Carboniferous to Permian, following which post-collisional extension began [28]. The extensional environment is inferred from the widespread post-collisional mafic-ultramafic rocks and A-type granitic stocks that occur throughout the Beishan orogenic belt [78,79,87,89], as well as from the widespread N-W-trending ultrabasic dikes near the Juyuan deposit (Figure 2). The post-collisional extension stage is interpreted as being responsible for the widespread emplacement of 286–232 Ma granites and associated contemporaneous W-(Mo) deposits.
To discuss the tectonic setting, we used the available granite U-Pb geochronological and geochemical data from the Xinjiang Institute of Geological Survey. The Juyuan granite had low A/CNK ratios (<1.1), high Th and Y contents that increase with increasing Rb content, and a negative relationship between P2O5 and SiO2, indicating an I-type granite affinity [90,91]. The granite also had high K, Rb, and Th values that are consistent with intra-plate magmatism. Isotopes of O and S support the interpretation that the hydrothermal fluids that formed the Juyuan deposit had a magmatic origin.
The orebodies of the Juyuan deposit run through the host granite, with mineralization styles dominated by quartz veins. The hydrothermal alteration consists of greisenization, sulfidization, and carbonation. The hydrothermal fluids were low in salinity and CO2 rich. These features, as well as the isotope signatures discussed above, suggest that Juyuan is a quartz-vein-type tungsten deposit. The studied muscovite from the Juyuan deposit yielded an Ar-Ar plateau age of 240.0 ± 1.0 Ma (1σ; Figure 12a), which is consistent with zircon 206Pb/238U ages of 240.9 ± 0.8 Ma and 240.7 ± 0.5 Ma for monzonitic and two-mica granites from the deposit, respectively (unpublished data). As muscovite generally coexists with scheelite at Juyuan, the excellent agreement between the U-Pb and 40Ar-39Ar chronometers confirms that the emplacement of the monzonitic granite and the tungsten mineralization in the Juyuan deposit formed contemporaneously during the Middle Triassic (240 Ma).
On the basis of the regional geology, characteristics of magmatic rocks, the formation of the Juyuan tungsten deposit and other similar Triassic deposits is considered to have taken place in an intra-continental plate setting, with the magmatism and metallogenesis being associated with regional extension. The discovery of the Juyuan tungsten deposit extends the known metallogenic belt eastward from East Tianshan to the Beishan structural belt, where quartz-vein-type tungsten deposits that formed in the same intra-plate extensional environment might be found.

8. Conclusions

We conducted a systematic geological and geochemical study of the evolution of mineralizing fluids and ore genesis of the Juyuan tungsten deposit, Beishan, NW China. Our main conclusions are as follows:
(1) The Juyuan tungsten deposit is hosted in Triassic granite. The vein-type mineralization can be divided into three stages, namely, early-stage quartz-oxide, intermediate-stage quartz-sulfide, and late-stage carbonate-quartz.
(2) Quartz in the Juyuan deposit contained CO2-H2O, pure CO2, NaCl-H2O, and daughter-mineral fluid inclusions. Fluid inclusion analysis reveals that the fluid system evolved from CO2-rich and CH4-rich fluid to CO2-rich and CH4-absent fluid, then further evolved into CO2-poor and CH4-poor aqueous fluid through CO2 escape, fluid boiling, and inflow/mixing of meteoric water. Fluid boiling and mixing were major processes leading to tungsten mineralization at Juyuan.
(3) Sulfides from the intermediate-stage had reasonably homogeneous δ34S values (+3.7‰ to +6.3‰), indicating the ore-forming fluids were derived mainly from granitic crust. Lead isotope data suggest that the formation of sulfides postdated tungsten deposition at Juyuan.
(4) Zircon U-Pb and muscovite 40Ar-39Ar geochronology confirms that the tungsten mineralization of the Juyuan deposit formed contemporaneously with the host granites during the Triassic (240.0 ± 1.0 Ma). The formation of the Juyuan tungsten deposit and other similar Triassic deposits is considered to have taken place in an intra-continental plate setting, with the magmatism and metallogenesis being associated with regional extension.
(5) A comparison of the tungsten deposits in East Tianshan and Beishan suggests that a single Triassic tungsten metallogenic belt extends from East Tianshan to the Beishan structural belt.

Author Contributions

Conceptualization, Q.Y. and Z.C.; sample collection, Q.Y., B.L., J.L., T.Z., Q.H. and W.Z. methodology, Q.Y. and Z.Z.; software, Z.Z.; article writing and figure drawing, Q.Y. and W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was jointly supported by the National Natural Science Foundation of China (grant no. 41902214), the National Key Technology Research and Development Programmer of the Ministry of Science and Technology of China (2018YFC0604005 and 2015BAB05B04), and Chinese Geological Survey Project (DD20190161).

Data Availability Statement

Not applicable.

Acknowledgments

We sincerely thank the two anonymous reviewers who have provided significant comments and suggestions which have greatly improved this paper.

Conflicts of Interest

The authors declare no conflict interest.

References

  1. Mao, J.W.; Ouyang, H.G.; Song, S.W.; Santosh, M.; Yuan, S.D.; Zhou, Z.H.; Zheng, W.; Liu, H.; Liu, P.; Cheng, Y.B.; et al. Geology and metallogeny of tungsten and tin deposits in China. SEG Spec. Publ. 2019, 22, 411–482. [Google Scholar]
  2. Huang, L.C.; Jiang, S.Y. Highly fractionated S-type granites from the giant Dahutang tungsten deposit in Jiangnan Orogen, Southeast China: Geochronology, petrogenesis and their relationship with W mineralization. Lithos 2014, 202–203, 207–226. [Google Scholar] [CrossRef]
  3. Mao, Z.H.; Liu, J.J.; Mao, J.W.; Deng, J.; Zhang, F.; Meng, X.Y.; Xiong, B.K.; Xiang, X.K.; Luo, X.H. Geochronology and geochemistry of granitoids related to the giant Dahutang tungsten deposit, middle Yangtze River region, China: Implications for petrogenesis, geodynamic setting, and mineralization. Gondwana Res. 2015, 28, 816–836. [Google Scholar] [CrossRef]
  4. Pan, X.; Hou, Z.; Zhao, M.; Chen, G.; Rao, J.; Li, Y.; Wei, J.; Ouyang, Y. Geochronology and geochemistry of the granites from the Zhuxi W–Cu ore deposit in South China: Implications for petrogenesis, geodynamical setting and mineralization. Lithos 2018, 304–307, 155–179. [Google Scholar] [CrossRef]
  5. Song, W.L.; Yao, J.M.; Chen, H.Y.; Sun, W.D.; Ding, J.Y.; Xiang, X.K.; Zuo, Q.S.; Lai, C.K. Mineral paragenesis, fluid inclusions, H–O isotopes and ore-forming processes of the giant Dahutang W–Cu–Mo deposit, South China. Ore Geol. Rev. 2018, 99, 116–150. [Google Scholar] [CrossRef]
  6. Su, Q.W.; Mao, J.W.; Wu, S.H.; Zhang, Z.C.; Xu, S.F. Geochronology and geochemistry of the granitoids and ore-forming age in the Xiaoyao tungsten polymetallic skarn deposit in the Jiangnan Massif tungsten belt, China: Implications for their petrogenesis, geodynamic setting, and mineralization. Lithos 2018, 296–299, 365–381. [Google Scholar] [CrossRef]
  7. Li, H.; Palinkaš, L.A.; Evans, N.J.; Watanabe, K. Genesis of the Huangshaping W–Mo–Cu–Pb–Zn deposit, South China: Role of magmatic water, metasomatized fluids, and basinal brines during intra-continental extension. Geol. J. 2020, 55, 1409–1430. [Google Scholar] [CrossRef]
  8. Li, H.; Kong, H.; Guo, B.Y.; Tamehe, L.S.; Zhang, Q.; Wu, Q.H.; Xi, X.S. Fluid inclusion, H–O–S isotope and rare earth element constraints on the mineralization of the Dong’an Sb deposit, South China. Ore Geol. Rev. 2020, 126, 103759. [Google Scholar] [CrossRef]
  9. Li, H.; Cao, J.; Algeo, T.J.; Jiang, W.; Liu, B.; Wu, Q. Zircons reveal multi-stage genesis of the Xiangdong (Dengfuxian) tungsten deposit, South China. Ore Geol. Rev. 2019, 111, 102979. [Google Scholar] [CrossRef]
  10. Qin, K.Z.; Fang, T.H.; Wang, S.L.; Zhu, B.Q.; Feng, Y.M.; Yu, H.F.; Xiu, Q.Y. Plate tectonics division, evolution and metallogenic settings in eastern Tianshan mountains, NW China. Xinjiang. Geology 2002, 20, 302–308. (In Chinese) [Google Scholar]
  11. Zhai, Y.S.; Yao, S.Z.; Cai, K.Q. Mineral Deposits, 3rd ed.; Geological Publishing House: Beijing, China, 2011; pp. 1–280. (In Chinese) [Google Scholar]
  12. Nie, F.J.; Jiang, S.H.; Hu, P.; Zhang, Y. Geological features and ore-forming material sources of Hongjianbingshan tungsten deposit in Beishan mountains, Gansu province. Min. Depos. 2004, 23, 11–19. (In Chinese) [Google Scholar]
  13. Jiang, S.H.; Nie, F.J. 40Ar–39Ar geochronology of Hongjianbingshan tungsten deposit in Beishan mountain, Gansu province, China. Min. Depos. 2006, 25, 89–94. (In Chinese) [Google Scholar]
  14. Ma, D.C.; Xi, Z. The ore-controlling factors of Yushan tungsten deposit. Northwest. Geol. 2012, 45, 57–60. (In Chinese) [Google Scholar]
  15. Yang, H.Q.; Li, Y.; Zhao, G.B.; Yang, J.G.; Li, W.M.; Yang, L.H. Research results report on prospecting problems in the Beishan metallogenic belt. 2008; 1–321, Unpublished Report. (In Chinese) [Google Scholar]
  16. Yang, H.Q.; Zhao, G.B.; Li, W.M.; Yang, J.G.; Li, Y.; Wang, X.H.; Jiang, H.B.; Tan, W.J. Formation age and source tracing of the tungsten-bearing granite belt in the Guoqing–Yingzuihongshan area, Inner Mongolia. Geol. Explor. 2010, 46, 353–370. (In Chinese) [Google Scholar]
  17. Zhang, X.H.; Su, L.; Cui, X.J.; Ding, S.H.; Zhao, J.G.; Chen, L. The age and controlling factors of Yushan tungsten deposit, Beishan Orogen, Gansu Province. Sci. China 2008, 53, 1077. (In Chinese) [Google Scholar]
  18. Zhao, G.B.; Li, W.M.; Yang, H.Q.; Tian, Y.T.; Quan, S.C.; Qiao, G.B.; Lei, Y.X.; Du, T.; Gao, Y.W.; Yang, T.; et al. Geological and geochemical characteristics of the Guoqing tungsten deposit in Beishan Orogen, Inner Mongolia. Min. Depos. 2010, 29, 341–342. (In Chinese) [Google Scholar]
  19. Zhao, G.B.; Li, W.M.; Yang, H.Q.; Tian, Y.T.; Gao, Y.W.; Du, T.; Quan, S.C.; Qiao, G.B.; Lei, Y.X.; Yang, T.; et al. Geological and geochemical characteristics of the Guoqing tungsten-bearing granite mass in Inner Mongolia and its genesis. Geol. Explor. 2011, 47, 828–836. (In Chinese) [Google Scholar]
  20. Zhao, G.B.; Li, W.M.; Yang, H.Q.; Gao, Y.W.; Yang, T.; Jia, J.; Quan, S.C.; Qiao, G.B.; Lei, Y.X. Mineralization and prospecting characteristics of the Guoqing tungsten-bearing granite mass in Inner Mongolia. Northwest. Geol. 2012, 45, 68–71. (In Chinese) [Google Scholar]
  21. Ding, J.X.; Han, C.M.; Xiao, W.J.; Wang, Z.M.; Yang, X.M. Geochemistry and U–Pb geochronology of tungsten deposit of Huaniushan island arc in the Beishan Orogenic Belt, and its geodynamic background. Acta Petrol. Sin. 2015, 31, 594–616. (In Chinese) [Google Scholar]
  22. Cao, J.; Xia, D.; Zheng, J.; Sun, Y.; Liu, Y.; Liu, Y. Geological characteristics of Juyuan gold–tungsten polymetallic deposit in Hami, Xinjiang. Xinjiang Geol. Surv. 2020, 38, 203–205. [Google Scholar]
  23. Sun, J.G.; Hu, S.X.; Shen, K.; Yao, F.L. Research on C, O isotopic geochemistry of intermediate–basic and intermediate–acid dykes in gold fields of Jiaodong Peninsula. Acta Petrol. Mineral. 2001, 30, 47–56. [Google Scholar]
  24. Sengör, A.M.C.; Natalin, B.A.; Buttman, U.S. Evolution of the Altaid tectonic collage and Paleozoic crustal growth in Eurasia. Nature 1993, 364, 299–307. [Google Scholar] [CrossRef]
  25. Jahn, B.M.; Wu, F.Y.; Chen, B. Massive granitoid generation in central Asia: Nd isotopic evidence and implication for continental growth in the Phanerozoic. Episodes 2000, 23, 82–92. [Google Scholar] [CrossRef] [Green Version]
  26. Windley, B.F.; Alexeiev, D.; Xiao, W.; Kröner, A.; Badarch, G. Tectonic models for accretion of the Central Asian Orogenic Belt. J. Geol. Soc. 2007, 164, 31–47. [Google Scholar] [CrossRef] [Green Version]
  27. Xiao, W.J.; Mao, Q.G.; Windley, B.F.; Han, C.M.; Qu, J.F.; Zhang, J.E.; Ao, S.J.; Guo, Q.Q.; Cleven, N.T.; Lin, S.F. Paleozoic multiple accretionary and collisional processes of the Beishan orogenic collage. Am. J. Sci. 2010, 310, 1553–1594. [Google Scholar] [CrossRef]
  28. Tian, Z.H.; Xiao, W.J.; Shan, Y.H.; Windley, B.; Han, C.M.; Zhang, J.E.; Wan, B.; Ao, S.J.; Song, D.F.; Feng, J.Y. Structures, ages and tectonic development of the Huoshishan–Niujuanzi ophiolitic mélange, Beishan, southernmost Altaids. Gondwana Res. 2014, 25, 820–841. [Google Scholar] [CrossRef]
  29. Xiao, W.J.; Santosh, M. The western Central Asian orogenic belt: A window to accretionary orogenesis and continental growth. Gondwana Res. 2014, 25, 1429–1444. [Google Scholar] [CrossRef]
  30. Li, P. Geological Characteristics and Genesis of Hongjianbingshan Tungsten Deposit, Beishan Mountain, Gansu Province. Master’s Thesis, China University of Geosciences, Wuhan, China, 2011; pp. 1–127. (In Chinese). [Google Scholar]
  31. Mao, Q.G. Paleozoic to Early Mesozoic Accretionary and Collisional Tectonics of the Beishan and Adjacent Area, Northwest China. Ph.D. Thesis, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing, China, 2008; pp. 1–228. (In Chinese). [Google Scholar]
  32. Guo, Q.Q.; Xiao, W.J.; Windley, B.F.; Mao, Q.G.; Han, C.M.; Qu, J.F.; Ao, S.J.; Li, J.L.; Song, D.F.; Yong, Y. Provenance and tectonic settings of Permian turbidites from the Beishan Mountains, NW China: Implications for the Late Paleozoic accretionary tectonics of the southern Altaids. J. Asian Earth Sci. 2012, 49, 54–68. [Google Scholar] [CrossRef]
  33. Tian, Z.H.; Xiao, W.J.; Shan, Y.H.; Windley, B.; Han, C.M.; Zhang, J.E.; Song, D.F. Mega-fold interference patterns in the Beishan orogen (NW China) created by change in plate configuration during Permo-Triassic termination of the Altaids. J. Struct. Geol. 2013, 52, 119–135. [Google Scholar] [CrossRef]
  34. Yang, H.Q.; Li, Y.; Li, W.M.; Yang, J.G.; Zhao, G.B.; Sun, N.Y.; Wang, X.H.; Tan, W.J. General discussion on metallogenetic tectonic setting of Beishan Mountain, northwestern China. Northwest. Geol. 2008, 41, 22–28. (In Chinese) [Google Scholar]
  35. Li, P.; Lv, X.B.; Chen, C.; Cao, X.F.; Mayila, A.; Su, Y.Y. Chronology and geochemical characteristics of K-feldspar granite in the Guoqing tungsten deposit and their geological implications. Acta Petrol. Mineral. 2011, 30, 13–24. (In Chinese) [Google Scholar]
  36. Nie, F.J.; Jiang, S.H.; Zhao, S.M.; Bai, D.M.; Liu, Y.; Zhao, Y.M.; Wang, X.L.; Su, X.X. Geological characteristics and types of Liushashan Au (Mo) deposit, Inner Mongolia. Geol. Geochem. 2002, 30, 1–7. (In Chinese) [Google Scholar]
  37. Nie, F.J.; Jiang, S.H.; Bai, D.M.; Zhang, Y.; Zhao, Y.M.; Wang, X.L. Type and temporal–spatial distribution of metal deposits in the Beishan Mountains, Inner Mongolia and its Neighboring Region. Acta Geol. Sin. 2003, 77, 367–378. (In Chinese) [Google Scholar]
  38. Li, H.Q.; Chen, F.W. Isotope Geochronology of Regional Mineralization in Xinjiang, NW China; Geological Publishing House: Beijing, China, 2004; pp. 1–391. (In Chinese) [Google Scholar]
  39. Li, H.Q.; Wu, H.; Chen, F.W.; Deng, G.; Yang, H.M.; Yang, Z.F.; Mei, Y.P.; Guo, J. New chronological evidence for Indosinian diagenetic mineralization in eastern Xinjiang, NW China. Acta Geol. Sin. 2005, 79, 264–275. (In Chinese) [Google Scholar]
  40. Li, N.; Yang, F.Q.; Zhang, Z.X.; Yang, C.D. Geochemistry and chronology of the biotite granite in the Xiaobaishitou W–(Mo) deposit, eastern Tianshan, China: Petrogenesis and tectonic implications. Ore Geol. Rev. 2019, 107, 999–1019. [Google Scholar] [CrossRef]
  41. 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]
  42. Collins, P.L.F. Gas hydrates in CO2-bearing fluid inclusions and use of freezing data for estimation of salinity. Econ. Geol. 1979, 74, 1435–1444. [Google Scholar] [CrossRef]
  43. Robert, S.D. An extended equation to calculate NaCl contents from final clathrate melting temperatures in H2O-CO2-NaCl fluid inclusions: Implications for P-T isochore location. Geochim. Cosmochim. Acta 1991, 55, 3869–3871. [Google Scholar]
  44. 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 geological systems:Equation of statemperature for H2O-CO2-NaCl fluids at high pressures and temperatures. Geochim. Cosmochim. Acta 1983, 47, 1247–1275. [Google Scholar] [CrossRef]
  45. 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]
  46. Ding, T.; Jiang, S.; Wan, D.; Li, Y.; Li, J.; Song, H.; Liu, Z.; Yao, X. Silicon Isotope Geochemistry; Geological Publishing House: Beijing, China, 1996; pp. 1–122. [Google Scholar]
  47. 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]
  48. Sang, H.Q.; Wang, F.; He, H.Y.; Wang, Y.L.; Yang, L.K.; Zhu, R.X. Intercalibration of ZBH-25 biotite reference material utilized for K–Ar and 40Ar–39Ar age determination. Acta Petrol. Sin. 2006, 22, 3059–3078. [Google Scholar]
  49. Nier, A.O. A redetermination of the relative abundances of isotopes of carbon, nitrogen, oxygen, argon and potassium. Phys. Rev. 1950, 77, 789–793. [Google Scholar] [CrossRef]
  50. Steiger, R.H.; Jäger, E. Subcommission on geochronology: Convention on the use of decay constants in geo- and cosmochronology. Earth Planet. Sci. Lett. 1977, 36, 359–362. [Google Scholar] [CrossRef]
  51. Koppers, A.A.P. Ar-Ar CALC-software for 40Ar/39Ar age calculations. Comput. Geosci. 2002, 28, 605–619. [Google Scholar] [CrossRef]
  52. Roedder, E. Fluid inclusions. Rev. Mineral. 1984, 12, 1–644. [Google Scholar]
  53. Zhou, Z.-J.; Chen, Y.-J.; Jiang, S.-Y.; Hu, C.-J.; Qin, Y.; Zhao, H.-X. Isotope and fluid inclusion geochemistry and genesis of the Qiangma gold deposit, Xiaoqinling gold field, Qinling Orogen, China. Ore Geol. Rev. 2015, 66, 47–64. [Google Scholar] [CrossRef]
  54. Taylor, H.P. Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits. In Geochemistry of Hydrothermal Ore Deposits; Barnes, H.L., Ed.; John Wiley and Sons: New York, NY, USA, 1997; pp. 229–302. [Google Scholar]
  55. Schwinn, G.; Wagner, T.; Baatartsogt, B.; Markl, G. Quantification of mixing processes in ore-forming hydrothermal systems by combination of stable isotope and fluid inclusion analyses. Geochim. Cosmochim. Acta 2006, 70, 965–982. [Google Scholar] [CrossRef]
  56. Hoefs, J. Stable Isotope Geochemistry, 3rd ed.; Springer: Berlin/Heidelberg, Germany, 1997; pp. 1–250. [Google Scholar]
  57. Faure, G. Principles of Isotope Geology, 2nd ed.; John Wiley and Sons: New York, NY, USA, 1986; pp. 1–608. [Google Scholar]
  58. Nelson, D.R.; Chivas, A.R.; Chappell, B.W.; McCulloch, M.T. Geochemical and isotopic systematics in carbonatites and implications for the evolution of ocean-island sources. Geochim. Cosmochim. Acta 1998, 52, 1–17. [Google Scholar] [CrossRef]
  59. Liu, J.M.; Liu, J.J. Basin fluid genetic model of sediment-hosted micro-disseminated gold deposits in the gold-triangle area between Guizhou, Guangxi and Yunnan. Acta Mineral. Sin. 1997, 17, 448–456. (In Chinese) [Google Scholar]
  60. Ray, J.S.; Ramesh, R.; Pande, K.; Trivedi, J.R.; Shukla, P.N.; Patel, P.P. Isotope and rare earth element chemistry of carbonatite–alkaline complexes of Deccan volcanic province: Implications to magmatic and alteration processes. J. Asian Earth Sci. 2000, 18, 177–194. [Google Scholar] [CrossRef]
  61. Liu, J.J.; He, M.Q.; Li, Z.M.; Liu, Y.P.; Li, C.Y.; Zhang, Q.; Yang, W.G.; Yang, A.P. Oxygen and carbon isotopic geochemistry of Baiyangping silver–copper polymetallic ore concentration area in Lanping Basin of Yunnan Province and its significance. Miner. Depos. 2004, 23, 1–10. (In Chinese) [Google Scholar]
  62. Hoefs, J. Stable Isotope Geochemistry, 7th ed.; Springer: Berlin/Heidelberg, Germany, 2015; pp. 1–389. [Google Scholar]
  63. Ohmoto, H.; Goldhaber, M.B. Sulphur and carbon isotopes. In Geochemistry of Hydrothermal Ore Deposits; Barnes, H.L., Ed.; John Wiley and Sons: New York, NY, USA, 1997; pp. 517–612. [Google Scholar]
  64. Jin, C.H.; Li, K.; Huang, L.; Zhang, Y.; Shen, Z.W. Characteristics of sulfur and lead isotope composition and metallogenic material source of the Nayongzhi Pb–Zn deposit, northwestern Guizhou Province. J. Mineral. Petrol. 2015, 35, 81–88. [Google Scholar]
  65. LaFlamme, C.; Sugiono, D.; Thebaud, N.; Caruso, S.; Fiorentini, M.; Selvaraja, V.; Jeon, H.; Voute, F.; Martin, L. Multiple sulfur isotopes monitor fluid evolution of an Archean orogenic gold deposit. Geochim. Cosmochim. Acta 2018, 222, 436–446. [Google Scholar] [CrossRef]
  66. Peng, N.J.; Jiang, S.Y.; Xiong, S.F.; Pi, D.H. Fluid evolution and ore genesis of the Dalingshang deposit, Dahutang W–Cu ore field, northern Jiangxi Province, South China. Miner. Depos. 2018, 53, 1079–1094. [Google Scholar] [CrossRef]
  67. Tsang, H.Y.; Cao, J.Y.; Yang, X.Y. Source of the Chaoyangzhai gold deposit, Southeast Guizhou: Constraints from LA–ICP–MS zircon U–Pb dating, whole-rock geochemistry and in situ sulfur isotopes. Minerals 2019, 9, 235. [Google Scholar] [CrossRef] [Green Version]
  68. Zheng, Y.; Chen, J. Stable Isotope Geochemistry; Science Press: Beijing, China, 2000; pp. 1–316. [Google Scholar]
  69. Ohmoto, H.; Rye, R.O. Isotopes of sulfur and carbon. In Geochemistry of Hydrothermal Ore Deposits, 2nd ed.; Barnes, H.L., Ed.; John Wiley and Sons: New York, NY, USA, 1979; pp. 509–567. [Google Scholar]
  70. Zartman, R.E.; Doe, B.R. Plumbotectonics—The model. Tectonophysics 1981, 75, 135–162. [Google Scholar] [CrossRef]
  71. Fan, H.R.; Xie, Y.H.; Zhao, R.; Wang, Y.L. Dual origins of Xiaoqinling gold-bearing quartz veins: Fluid inclusion evidences. Chin. Sci. Bull. 2000, 45, 1424–1430. [Google Scholar] [CrossRef]
  72. Fan, H.R.; Xie, Y.H.; Zhai, M.G.; Jin, C.W. A three stages fluid flow model for Xiaoqinling lode gold metallogenesis in the Henan and Shaanxi provinces, central China. Acta Petrol. Sin. 2003, 19, 260–266. (In Chinese) [Google Scholar]
  73. Hagemann, S.G.; Luders, V. P–T–X conditions of hydrothermal fluids and precipitation mechanism of stibnite–gold mineralization at the Wiluna lode-gold deposits, Western Australia: Conventional and infrared microthermometric constraints. Miner. Depos. 2003, 38, 936–952. [Google Scholar] [CrossRef]
  74. Coulibaly, Y.; Boiron, M.C.; Cathelineau, M.; Kouamelan, A.N. Fluid immiscibility and gold deposition in the Birimian quartz veins of the Angovia deposit (Yaoure, Ivory Coast). J. Afr. Earth Sci. 2008, 50, 234–254. [Google Scholar] [CrossRef]
  75. Chi, G.; Lu, H. Validation and representation of fluid inclusion microthermometric data using the fluid inclusion assemblage (FIA) concept. Acta Petrol. Sin. 2008, 24, 1945–1953. [Google Scholar]
  76. Shmulovich, K.I.; Landwehr, D.; Simon, K.; Heinrich, W. Stable isotope fractionation between liquid and vapour in water–salt systems up to 600 °C. Chem. Geol. 1999, 157, 343–354. [Google Scholar] [CrossRef]
  77. Driesner, T.; Seward, T.M. Experimental and simulation study of salt effects and pressure/density effects on oxygen and hydrogen stable isotope liquid–vapor fractionation for 4–5 molal aqueous NaCl and KCl solutions to 400 °C. Geochim. Cosmochim. Acta 2000, 64, 1773–1784. [Google Scholar] [CrossRef]
  78. Zhu, J.; Lü, X.B.; Cao, X.F.; Mo, Y.; Chen, C. U–Pb zircon geochronology, geochemistry and kinetics of the Huaniushan A-type granite in Northwest China. Chin. J. Geochem. 2012, 31, 85–94. [Google Scholar] [CrossRef]
  79. Ding, J.X.; Han, C.M.; Xiao, W.J.; Wang, Z.M.; Song, D.F. Geochronology, geochemistry and Sr-Nd isotopes of the granitic rocks associated with tungsten deposits in Beishan district, NW China, Central Asian Orogenic Belt: Petrogenesis, metallogenic and tectonic implications. Ore Geol. Rev. 2017, 89, 441–462. [Google Scholar] [CrossRef]
  80. Laurent-Charvet, S.; Charvet, J.; Monie, P.; Shu, L. Late Paleozoic strike-slip shear zones in eastern central Asia (NW China): New structural and geochronological data. Tectonics 2003, 22, 1009. [Google Scholar] [CrossRef] [Green Version]
  81. Lamb, M.A.; Badarch, G.; Navratil, T.; Poier, R. Structural and geochronologic data from the Shin Jinst area, eastern Gobi Altai, Mongolia: Implications for Phanerozoic intracontinental deformation in Asia. Tectonophysics 2008, 451, 312–330. [Google Scholar] [CrossRef]
  82. Wang, T.; Jahn, B.M.; Kovach, V.P.; Tong, Y. Mesozoic anorogenic granitic magmatism in the Altai Paleozoic accretionary orogen, NW China: Implications for crustal architecture and growth. Am. J. Sci. 2014, 314, 1–42. [Google Scholar] [CrossRef]
  83. Wang, Y.; Sun, G.; Li, J. U–Pb (SHRIMP) and 40Ar/39Ar geochronological constraints on the evolution of the Xingxingxia shear zone, NW China: A Triassic segment of the Altyn Tagh fault system. Geol. Soc. Am. Bull. 2010, 122, 487–505. [Google Scholar] [CrossRef]
  84. Xiao, W.J.; Windley, B.F.; Huang, B.C.; Han, C.M.; Yuan, C.; Chen, H.L.; Sun, M.; Sun, S.; Li, J.L. End-Permian to mid-Triassic termination of the accretionary processes of the southern Altaids: Implications for the geodynamic evolution, Phanerozoic continental growth, and metallogeny of Central Asia. Int. J. Earth Sci. 2009, 98, 1189–1287. [Google Scholar] [CrossRef]
  85. Ye, D.J.; Zhan, J.B.; Liang, M.H.; Xiao, L. Formational and Tectonic Map of the Mazongshan Region, Gansu, China, 1:250,000; Institute of Geological Survey: Lanzhou, China, 2008. [Google Scholar]
  86. Xiao, W.J.; Huang, B.C.; Han, C.M.; Sun, S.; Li, J.L. A review of the western part of the Altaids: A key to understanding the architecture of accretionary orogens. Gondwana Res. 2010, 18, 253–273. [Google Scholar] [CrossRef]
  87. Wu, F.Y.; Sun, D.Y.; Ge, W.C.; Zhang, Y.B.; Grant, M.L.; Wilde, S.A.; Jahn, B.M. Geochronology of the Phanerozoic granitoids in northeastern China. J. Asian Earth Sci. 2011, 41, 1–30. [Google Scholar] [CrossRef] [Green Version]
  88. Li, N.; Yang, F.; Zhang, Z. Fluid inclusions and isotope (C, H, O, S, He, and Ar) study of the Xiaobaishitou skarn W–(Mo) deposit, East Tianshan, NW China. Ore Geol. Rev. 2020, 122, 103520. [Google Scholar] [CrossRef]
  89. Pirajno, F.; Ernst, R.E.; Borisenko, A.S.; Fedoseev, G.; Naumov, E. Intraplate magmatism in Central Asia and China and associated metallogeny. Ore Geol. Rev. 2009, 35, 114–136. [Google Scholar] [CrossRef]
  90. Chappell, B.W. Aluminium saturation in I- and S-type granites and the characterization of fractionated hapogranites. Lithos 1999, 46, 535–551. [Google Scholar] [CrossRef]
  91. Chappell, B.W.; White, A.J.R. I- and S-type granites in the Lachlan Fold Belt. Trans. R. Soc. Edinb. Earth Sci. 1992, 83, 1–26. [Google Scholar]
Figure 1. (A) Location of the study area within the Central Asian Orogenic Belt (modified after [24]). Abbreviations: CAOB= Central Asian orogenic belt, NCC = North China Craton; (B) Simplified tectonic map of the Beishan orogenic collage and its adjacent area showing the tectonic subdivisions (modified after [27,28]).
Figure 1. (A) Location of the study area within the Central Asian Orogenic Belt (modified after [24]). Abbreviations: CAOB= Central Asian orogenic belt, NCC = North China Craton; (B) Simplified tectonic map of the Beishan orogenic collage and its adjacent area showing the tectonic subdivisions (modified after [27,28]).
Minerals 11 01309 g001
Figure 2. General geologic map of the Juyuan tungsten ore district.
Figure 2. General geologic map of the Juyuan tungsten ore district.
Minerals 11 01309 g002
Figure 3. Geologic section of No. 18 prospecting lines at the Juyuan tungsten deposit.
Figure 3. Geologic section of No. 18 prospecting lines at the Juyuan tungsten deposit.
Minerals 11 01309 g003
Figure 4. Representative photos and photomicrographs of the Juyuan ore deposit. (a) Distribution characteristics of the granite, quartz porphyry, and quartz vein in Juyuan mining area; (b) Late-stage quartz veinlets in monzogranite; (c) Early-stage quartz veins; (d) Intermediate ore-bearing hydrothermal quartz veins; (e) Wolframite in quartz vein; (f) Scheelite ore fluorescent under a tungsten lamp; (g) Microscopic characteristics of monzogranite; (h) Microscopic characteristics of wolframite; (i) The interspersed relationship between wolframite and scheelite under microscope; (j) Sericitized hand specimen; (k) Microscopic characteristics of Sericitized; (l) Galena and pyrite in the intermediate-stage of mineralization.
Figure 4. Representative photos and photomicrographs of the Juyuan ore deposit. (a) Distribution characteristics of the granite, quartz porphyry, and quartz vein in Juyuan mining area; (b) Late-stage quartz veinlets in monzogranite; (c) Early-stage quartz veins; (d) Intermediate ore-bearing hydrothermal quartz veins; (e) Wolframite in quartz vein; (f) Scheelite ore fluorescent under a tungsten lamp; (g) Microscopic characteristics of monzogranite; (h) Microscopic characteristics of wolframite; (i) The interspersed relationship between wolframite and scheelite under microscope; (j) Sericitized hand specimen; (k) Microscopic characteristics of Sericitized; (l) Galena and pyrite in the intermediate-stage of mineralization.
Minerals 11 01309 g004
Figure 5. Photomicrographs of fluid inclusions in quartz from the Juyuan tungsten deposit. (a) C-type fluid inclusions with various vapor/liquid in early-stage quartz; (b) S-type fluid inclusion in early-stage quartz; (c) C-type fluid inclusion in early-stage quartz; (d) coexisting W-type and PC-type fluid inclusions in intermediate-stage quartz; (e) coexisting W-type and C-type fluid inclusions in intermediate-stage quartz; (f,g) W-type fluid inclusions in late-stage quartz. Abbreviation: VCO2, vapor CO2; LCO2, liquid CO2; VH2O, vapor H2O; LH2O, liquid H2O; S, daughter mineral.
Figure 5. Photomicrographs of fluid inclusions in quartz from the Juyuan tungsten deposit. (a) C-type fluid inclusions with various vapor/liquid in early-stage quartz; (b) S-type fluid inclusion in early-stage quartz; (c) C-type fluid inclusion in early-stage quartz; (d) coexisting W-type and PC-type fluid inclusions in intermediate-stage quartz; (e) coexisting W-type and C-type fluid inclusions in intermediate-stage quartz; (f,g) W-type fluid inclusions in late-stage quartz. Abbreviation: VCO2, vapor CO2; LCO2, liquid CO2; VH2O, vapor H2O; LH2O, liquid H2O; S, daughter mineral.
Minerals 11 01309 g005
Figure 6. Histograms of homogenization temperatures and the salinities of fluid inclusions.
Figure 6. Histograms of homogenization temperatures and the salinities of fluid inclusions.
Minerals 11 01309 g006
Figure 7. Representative Laser Raman spectra for fluid inclusions from the Juyuan tungsten deposit (a) Spectrum for vapor bubbles of CO2-H2O inclusions in early-stage quartz, containing variable CO2 and CH4; (b) Liquid phase of CO2-H2O inclusions in early-stage quartz; (c) Spectrum for vapor bubbles of CO2-H2O inclusions in intermediate-stage quartz, containing variable CO2 and H2O; (d) Aqueous inclusions in late-stage quartz containing water only.
Figure 7. Representative Laser Raman spectra for fluid inclusions from the Juyuan tungsten deposit (a) Spectrum for vapor bubbles of CO2-H2O inclusions in early-stage quartz, containing variable CO2 and CH4; (b) Liquid phase of CO2-H2O inclusions in early-stage quartz; (c) Spectrum for vapor bubbles of CO2-H2O inclusions in intermediate-stage quartz, containing variable CO2 and H2O; (d) Aqueous inclusions in late-stage quartz containing water only.
Minerals 11 01309 g007
Figure 8. δ18O-δD plots of the ore fluids at the Juyuan tungsten deposit. Domains for metamorphic and magmatic fluids were obtained from [54].
Figure 8. δ18O-δD plots of the ore fluids at the Juyuan tungsten deposit. Domains for metamorphic and magmatic fluids were obtained from [54].
Minerals 11 01309 g008
Figure 9. Plot of δ13CPDB for fluid inclusions in quartz versus δ18OSMOW for the quartz from the Juyuan tungsten deposit. Data for the C-O reservoirs and geological processes reasonable for the C-O isolate variations were obtained from [23,58,59,60,61,62].
Figure 9. Plot of δ13CPDB for fluid inclusions in quartz versus δ18OSMOW for the quartz from the Juyuan tungsten deposit. Data for the C-O reservoirs and geological processes reasonable for the C-O isolate variations were obtained from [23,58,59,60,61,62].
Minerals 11 01309 g009
Figure 10. δ34S values of sulfides from the Juyuan deposit. Data of other deposits are presented in [63,66,68,69].
Figure 10. δ34S values of sulfides from the Juyuan deposit. Data of other deposits are presented in [63,66,68,69].
Minerals 11 01309 g010
Figure 11. Lead isotopic compositions of the Juyuan deposit. Base map from [70].
Figure 11. Lead isotopic compositions of the Juyuan deposit. Base map from [70].
Minerals 11 01309 g011
Figure 12. (a) Plateau, (b) isochron, and (c) inverse isochron 40Ar–39Ar ages of muscovite (sample D18105-5; 1σ) from the Juyuan tungsten deposit.
Figure 12. (a) Plateau, (b) isochron, and (c) inverse isochron 40Ar–39Ar ages of muscovite (sample D18105-5; 1σ) from the Juyuan tungsten deposit.
Minerals 11 01309 g012
Figure 13. Plot of homogenization temperature (Th) vs. salinity of FIs in quartz from the Juyuan deposit.
Figure 13. Plot of homogenization temperature (Th) vs. salinity of FIs in quartz from the Juyuan deposit.
Minerals 11 01309 g013
Table 1. Microthermometric data for fluid inclusions in quartz from the Juyuan tungsten deposit.
Table 1. Microthermometric data for fluid inclusions in quartz from the Juyuan tungsten deposit.
StageSample No.TypeNumberSize
(μm)
Vapor
(vol.%)
Tm-CO2
(°C)
Tm-cl
(°C)
Th-CO2
(°C)
Tm-ice
(°C)
Th-tot
(°C)
Salinity
(wt.% NaCl equiv.)
CO2 Density
(g/cm3)
Bulk Density
(g/cm3)
EarlyD1936-1C135–925–85−60.9 to −68.85.4–8.326.9–31.1 (L/V) 261–343 (L/V)3.3–8.40.54–0.690.69–0.92
PC25–720–30−62.5 to −63.25.4–8.328.3–29.1 (L) 0.63–0.65
D1938-3aC126–2715–50−58.2 to −69.66.2–8.727.0–31.1 (L) 266–327 (L)2.6–7.10.59–0.690.83–0.95
W26–920–30 −2.8 to −5.8275–282 (L)4.7–9.0
S11515 −6.8246 (L)10.2
D1958-2-2C84–1220–75−59.0 to −67.14.4–8.617.2–29.6 (L/V) 253–318 (L/V)2.8–9.90.65–0.810.79–0.97
PC1640−60.2 26.5 (L) 0.69
D1958-4aC25–715–20−61.0 to −61.96.0–6.728.0–30.6 (L) 249–321 (L)6.2–7.40.62–0.680.95–0.99
D1965-1C176–1125–95−58.2 to −66.76.0–8.728.8–31.1 (L/V) 249–333 (L/V)2.6–7.40.52–0.640.69–0.92
D1966-1C115–1015–45−60.5 to −67.07.2–10.027.2–30.7 (L) 265–336 (L)0.02–5.30.62–0.700.83–0.97
W1715 −5.5280 (L)8.6
D1967-1C194–915–90−59.8 to −63.95.5–8.526.1–30.6 (L/V) 240–340 (L/V)3.0–8.20.62–0.710.71–0.96
PC25–725–30−61.9 to −62.6 15.9–25.7 (L) 0.70–0.81
D2016-1C85–930–40−56.7 to −62.76.1–9.724.5–30.4 (L) 248–290 (L)0.6–7.20.62–0.740.86–0.95
D2017-1C55–730–35−57.4 to −64.36.9–8.829.7–30.8 (L) 260–324 (L)2.4–5.90.61–0.670.88–0.92
W1530 −6.5232 (L)9.9
D2022-1C65–630–40−56.7 to −59.76.9–9.227.7–29.8 (L) 241–340 (L)1.6–5.90.64–0.690.87–0.91
W25–635–40 −3.1 to −7.4230–278 (L)5.1–11
D2032-1C75–725–40−58.0 to −64.76.8–9.729.5–31.0 (L) 257–338 (L)0.6–6.00.60–0.670.86–0.92
D2033-1C325–1120–40−56.9 to −67.34.0–9.325.8–31.0 (L) 230–309 (L)1.4–10.50.60–0.720.84–0.96
W11015 −5.8 264 (L)9.0
D2045-1C105–1130–45−57.1 to −63.18.6–10.028.9–30.8 (L) 275–334 (L)0.02–2.770.61–0.660.83–0.90
D2048-1C85–830–45−57.2 to −65.46.7–9.325.7–30.9 (L) 257–344 (L)1.4–6.20.61–0.650.85–0.90
Intermediate D1935-1C85–1315–40−58.8 to −68.77.4–8.530.0–31.1 (L) 270–290 (L)3.0–5.00.61–0.640.86–0.96
D1938-1bC115–1720–70−60.2 to −75.56.8–8.729.3–31.1 (L/V) 241–286 (L/V)1.8–6.00.60–0.650.76–0.92
PC28–1225–45−62.0 to −76.3 24.7–30.0 (L) 0.60–0.72
D2070-1C66–720–40−57.2 to −62.16.3–8.428.7–30.9 (L) 275–295 (L)3.2–6.90.61–0.670.89–0.95
LateD1936-4W135–1310–25 −3.2 to −8.5184–234 (L)4.6–12.3
D1959-2aW1610 −8.1196 (L)11.8
1 Fluid inclusion type: C = CO2-rich inclusions; W = liquid-rich two-phase inclusions; PC = pure CO2 inclusions; S = liquid-rich three-phase inclusions with one daughter mineral; 2 Tm-CO2—Final solid CO2 melting temperature. 3 Tm-cl—Final CO2 clathrate dissociation temperature. 4 Th-CO2—CO2 phase homogenization temperature. 5 Th-tot—Total homogenization temperature. 6 Tm-ice—Final ice melting temperature.
Table 2. The δ18O, δD, and δ13C ratios (‰) of the Juyuan tungsten deposit.
Table 2. The δ18O, δD, and δ13C ratios (‰) of the Juyuan tungsten deposit.
No.Sample Mineralδ18Omδ18Ow δDδ13CCO2 T(°C)StageReference
1D1934-1aQuartz12.2 4.8 −86 −7.8 287 EarlyThis study
2D1935-5Quartz13.6 6.0 −82 −7.7 280 EarlyThis study
3D1937-5Quartz12.3 4.9 −98 −14.1 287 EarlyThis study
4D1937-6Quartz12.9 5.5 −94 −7.4 287 EarlyThis study
5D1938-3aQuartz13.2 5.7 −91 −3.3 285 EarlyThis study
6D1958-2-1Quartz12.9 5.3 −95 −4.0 282 EarlyThis study
7D1958-4aQuartz12.9 5.3 −88 −3.4 282 EarlyThis study
8D1960-1Quartz12.9 5.5 −91 −6.2 287 Early This study
9D1963-1Quartz12.9 5.5 −88 −2.3 287 EarlyThis study
10D1965-1Quartz12.4 5.3 −85 −1.5 294 EarlyThis study
11JYQuartz12.7 5.3 −72 −9.0 287 EarlyThis study
12D1935-2Quartz13.1 4.6 −92 −7.4 260 Intermediate This study
13D1938-1aQuartz12.7 4.2 −93 −6.6 260 IntermediateThis study
14D1936-4Quartz13.0 1.3 −91 −6.5 200 LateThis study
15D1959-2aQuartz12.2 0.5 −89 −7.5 200 LateThis study
Table 3. The δ34S values of ores and rocks at the Juyuan tungsten deposit.
Table 3. The δ34S values of ores and rocks at the Juyuan tungsten deposit.
No.Sample No.Sample DescriptionMineralδ34SStage
1D1935-6Sulfides -bearing Quartz VeinPyrite6.3Intermediate
2D1936-3Sulfides -bearing Quartz VeinPyrite6.2Intermediate
3D1936-4Sulfides -bearing Quartz VeinPyrite6.2Intermediate
4D1962-1Sulfides -bearing Quartz VeinPyrite6.3Intermediate
Average6.25
5D1935-6Sulfides -bearing Quartz VeinGalena4.4Intermediate
6D1936-3Sulfides -bearing Quartz VeinGalena4.0Intermediate
7D1936-4Sulfides -bearing Quartz VeinGalena3.7Intermediate
8D1962-1Sulfides -bearing Quartz VeinGalena3.7Intermediate
Average3.95
Table 4. Lead isotope ratios of ores and rocks at the Juyuan tungsten deposit.
Table 4. Lead isotope ratios of ores and rocks at the Juyuan tungsten deposit.
No.Sample No.Testing Object 206Pb/204Pb207Pb/204Pb208Pb/204Pb
1D1935-6Pyrite18.53815.70638.648
2D1936-3Galena18.47615.64138.437
3D1936-4Pyrite18.53415.70638.643
4D1936-4Galena18.51015.67238.534
5D1962-1Galena18.48715.64038.441
Average (n = 5) 18.50915.67338.541
Table 5. 40Ar-39Ar data for muscovite from the Juyuan tungsten deposit.
Table 5. 40Ar-39Ar data for muscovite from the Juyuan tungsten deposit.
T (°C)(40Ar/39Ar) m(36Ar/39Ar) m(37Ar/39Ar) m40Ar* (%)F (40Ar*/39Ar)39Ar (×10−14mol)39Ar (%)Age ± 1σ (Ma)
D18105-5, Sample weight = 10.8 mg, J = 0.006506
80022.73820.00240.018396.8822.029121.595.53243.9 ± 1.2
90021.89820.00080.009498.9321.664441.3210.59240.1 ± 1.1
95021.81630.00070.005399.0921.618452.2813.4239.6 ± 1.1
100021.73230.00050.001999.2621.572274.1419239.1 ± 1.1
105021.68910.00050.000799.2621.528879.6620.42238.7 ± 1.1
110021.67520.00070.003399.121.479760.8815.6238.2 ± 1.1
115021.78590.00070.003899.0421.576539.4810.12239.2 ± 1.1
120022.09630.00110.028198.4921.7627164.1241.1 ± 1.2
130022.48040.00270.074996.4121.67444.781.22240.2 ± 1.2
Notes: 40Ar* refer to radiogenic 40Ar.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, Q.; Chen, Z.; Zhou, Z.; Zhao, T.; Han, Q.; Li, J.; Liu, B.; Zhang, W. Fluid Evolution and Ore Genesis of the Juyuan Tungsten Deposit, Beishan, NW China. Minerals 2021, 11, 1309. https://doi.org/10.3390/min11121309

AMA Style

Yan Q, Chen Z, Zhou Z, Zhao T, Han Q, Li J, Liu B, Zhang W. Fluid Evolution and Ore Genesis of the Juyuan Tungsten Deposit, Beishan, NW China. Minerals. 2021; 11(12):1309. https://doi.org/10.3390/min11121309

Chicago/Turabian Style

Yan, Qiaojuan, Zhengle Chen, Zhenju Zhou, Tongyang Zhao, Qiong Han, Jilin Li, Bo Liu, and Wengao Zhang. 2021. "Fluid Evolution and Ore Genesis of the Juyuan Tungsten Deposit, Beishan, NW China" Minerals 11, no. 12: 1309. https://doi.org/10.3390/min11121309

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