Next Article in Journal
Zircon U-Pb Ages, Petrogenesis, and Tectono-Magmatic Evolution of Late Jurassic–Early Cretaceous S-Type Granitoid in Wujinshan Area of Northwestern Zhejiang, South China
Next Article in Special Issue
Ore Genesis and the Magmatism of the Yuhaixi Mo(Cu) Deposit in Eastern Tianshan, NW China: Constraints from Geology, Geochemistry, Zircon U-Pb and Molybdenite Re-Os Dating
Previous Article in Journal
Ceramic Grinding Kinetics of Fine Magnetite Ores in the Batch Ball Mill
Previous Article in Special Issue
In Situ Trace Element and Sulfur Isotope Composition of Pyrite from the Beiwagou Pb-Zn Deposit, Liaodong Peninsula, Northeast China: Implications for Ore Genesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ore Genesis of the Lower Urgen Porphyry Molybdenum Deposit in the Northern Great Xing’an Range, Northeast China: Constraints from Molybdenite Re-Os Dating, Fluid Inclusions, and H-O-S-Pb Isotopes

1
Key Laboratory of Ministry of Education on Safe Mining of Deep Metal Mines, Department of Geology, Northeastern University, Shenyang 110819, China
2
College of Earth Science and Engineering, Shandong University of Science and Technology, Qingdao 266590, China
3
Key Laboratory of Mineral Resources, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing 100029, China
4
Department of Earth Sciences, VU Amsterdam, De Boelelaan 1105, 1081 HV Amsterdam, The Netherlands
5
China Water Resources Bei Fang Investigation, Design & Research Co., Ltd., Tianjin 300222, China
6
China Railway Design Corporation, Tianjin 300308, China
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(9), 1189; https://doi.org/10.3390/min13091189
Submission received: 9 July 2023 / Revised: 17 August 2023 / Accepted: 8 September 2023 / Published: 10 September 2023

Abstract

:
The Lower Urgen molybdenum deposit (44,856 t Mo @ 0.141%), situated in the northern Great Xing’an Range, is a newly discovered porphyry molybdenum deposit. Mineralization is characterized by veinlet-disseminated- and vein-type quartz–sulfide orebodies primarily occurring in the cupola of the Early Cretaceous granite porphyry stock. In this study, we present a detailed description of the ore geology, molybdenite Re-Os dating, H-O-S-Pb isotopic compositions, and fluid inclusion (FI) analyses including petrography, laser Raman, and microthermometry to precisely constrain the timing of ore formation, the origin of ore-forming fluids and materials, as well as the metal precipitation mechanism. Molybdenite Re-Os dating yielded two model ages of 141.2 ± 1.5 and 147.7 ± 1.7 Ma, coeval with the regional Late Jurassic–Early Cretaceous molybdenum metallogenesis. The hydrothermal process can be divided into three stages: the quartz–molybdenite(–pyrite) stage, quartz–polymetallic sulfide stage, and quartz–carbonate stage. Four types of FIs were distinguished for quartz, including two-phase liquid-rich (L-type), saline (S-type), CO2-rich (C1-type), and CO2-bearing (C2-type) FIs. Microthermometric data showed that the homogenization temperatures and salinities from the early to late stages were 240–430 °C, 5.0–11.9, and 30.1–50.8 wt% NaCl equiv.; 180–280 °C and 3.0–9.1 wt% NaCl equiv.; and 120–220 °C and 0.2–7.9 wt% NaCl equiv., respectively, suggesting a decreasing trend. H-O isotopic compositions indicate that the ore-forming fluids were initially of magmatic origin with the increasing incorporation of meteoric water. S-Pb isotopic compositions indicate that the ore-forming materials originated from granitic magmas, and the mineralization is genetically related to the ore-bearing granite porphyry stock in the deposit. Fluid immiscibility and fluid–rock interaction are collectively responsible for the massive deposition of molybdenite in stage 1, whereas fluid mixing and immiscibility played a critical role in the deposition of polymetallic sulfide in stage 2.

1. Introduction

Porphyry deposits are commonly spatially, temporally, and genetically related to intermediate–acid hypabyssal porphyry intrusions [1], and provide more than 75%, 50%, and 20% of the world’s molybdenum, copper, and gold resources, respectively [2,3,4,5]. Mineralization mainly occurs as veinlets and disseminations concentrated in or near the cupolas of porphyry intrusions, and is related to the hydraulic fracturing resulting from fluid exsolution from magmas [3,6,7]. Taking the economic importance of porphyry molybdenum deposits into account, the study of the genesis and metallogenic models of these deposits has always been an essential subject in ore deposit research [8,9,10,11,12,13,14,15,16].
China has the largest molybdenum resources worldwide, accounting for about 54% of the world’s known economic molybdenum resources [17], which appear in six predominant molybdenum metallogenic provinces, i.e., NE China, Qinling-Dabie, Yanliao, Sanjiang, the middle-lower Yangtze River Valley, and South China [18]. NE China is situated in the eastern Central Asian Orogenic Belt (CAOB) tectonically. A long-term tectonomagmatic evolutionary history has made it a significant polymetallic metallogenic province around the world [19,20,21,22,23,24,25] (Figure 1). More than 11.4 Mt of molybdenum has now been discovered in NE China, including 4 super-large (>0.5 Mt Mo), 11 large (0.1–0.5 Mt Mo), 22 medium (0.01–0.1 Mt Mo), and 18 small (<0.01 Mt Mo) molybdenum deposits, marking NE China as the largest molybdenum metallogenic province in China [26] (Figure 1c).
The Lower Urgen deposit in Inner Mongolia, NE China, is a newly discovered porphyry molybdenum deposit, with estimated reserves of 44,856 t of molybdenum at an average grade of 0.141% [30]. The geochronology, petrogenesis, and tectonic context of the barren granites in the mining area have been the primary subjects of earlier investigations of the deposit [31]. Nonetheless, the direct timing of molybdenum mineralization, ore-forming fluids and the metal source, and metallogenic processes remains enigmatic. In accordance with a thorough field investigation and petrographic observations of thin sections, we present comprehensive descriptions of the geology, mineralization stages, and hydrothermal alteration in the Lower Urgen molybdenum deposit. The new molybdenite Re-Os geochronology, fluid inclusion (FI) study (petrography, laser Raman, and microthermometry analyses), and isotope geochemistry (H-O-S-Pb) for the deposit are investigated. By combining our results with available geochronology and geodynamic research in the region, we aim to (1) restrict the timing of molybdenum mineralization in the Lower Urgen molybdenum deposit; (2) determine ore-forming fluids and the metal source; and (3) decipher fluid evolution processes and ore precipitation mechanisms. This study not only contributes to the establishment of a plausible metallogenetic model for the Lower Urgen molybdenum deposit, but also has important implications for the comprehension of regional molybdenum metallogenesis and the future molybdenum prospecting and exploration.

2. Regional Geology

The tectonic components of NE China include the Erguna Block in the northwest, the Xing’an and Songliao-Xilinhot blocks in the central part, the Khanka-Jiamusi Block in the southeast, and the Liaoyuan Terrane in the south [28] (Figure 1). Major faults separating these tectonic units include the Hegenshan-Heihe, Solonker-Xar Moron-Changchun, Xinlin-Xiguitu, and Mudanjiang faults [28,29] (Figure 1b). NE China has experienced a multi-stage Phanerozoic tectonic evolutionary history, including the subduction and final closure of the Palaeo-Asian Ocean (PAO) during the Paleozoic to Early Mesozoic [28,32,33,34,35], the south-directed subduction and closure of the Mongol-Okhotsk Ocean (MOO), and the west-directed subduction and subsequent rollback of the Palaeo-Pacific Ocean (PPO) plate during the Mesozoic [29,36,37]. Due to the overprinting of multiple tectonic domains, NE China experienced extensive tectonomagmatic events, accompanied by intense metallogenesis [19,20,21,22,23,24,25]. Three episodes of Mesozoic molybdenum mineralization are recognized in the region, including the Triassic, Early–Middle Jurassic, and Late Jurassic–Early Cretaceous [18,24]. The major location of the Triassic molybdenum deposits is along the Solonker-Xar Moron-Changchun Fault (SXCF), with minor occurrences in the northern and central Great Xing’an Range (NCGXR). The Triassic molybdenum deposits near the SXCF were formed in a syn-collision compressional to post-collision extensional environment induced through the tectonic evolution of the PAO [18,38,39,40,41,42,43], whereas the Triassic molybdenum deposits in the NCGXR were formed in a subduction-dominated compressional environment induced through the SE-directed subduction of the MOO plate [44,45,46,47,48]. The Early–Middle Jurassic molybdenum deposits are concentrated in the Lesser Xing’an-Zhangguangcai Range (LXZR) and sparsely distributed in the NCGXR, which were formed in a compressional environment controlled by the overprinting influence of the SE-directed subduction of the MOO plate in the west and the NW-directed subduction of the PPO plate in the east [49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64]. Molybdenum deposits from the Late Jurassic to Early Cretaceous are mostly concentrated in the GXR, with sporadic occurrences in the LXZR, which were formed in a post-orogenic extensional environment triggered by the PPO and MOO tectonic domains [56,63,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88].
The NCGXR comprises two main units, i.e., the Erguna and Xing’an blocks (Figure 1c). The amalgamation of the Erguna and Xing’an blocks occurred along the Xinlin-Xiguitu Fault in the Early Palaeozoic induced by the closure of the PAO [28,89,90,91,92]. Volcanic-sedimentary rocks of the Xinghuadukou, Ergunahe, Jiageda, and Luomahu groups; Neoproterozoic magmatic rocks; and minor Palaeoproterozoic gneisses compose the basement of the Erguna Block [29,93,94]. Multiple Late Palaeozoic–Late Mesozoic episodes of accretion and collision occurred on the Erguna Block, resulting in the formation of tremendous intrusions, volcanics, and volcaniclastic rocks [95,96]. Widespread Mesozoic igneous rocks; Early Palaeozoic basement rocks composed of the Wolegen, Xinkailing, Zalantun, and Fengshuigouhe groups; together with the Xinhuadukou Complex and dispersed Palaeozoic plutons and sediments constitute the Xing’an Block, the primary component of the GXR [29]. Neoproterozoic granite is locally distributed in the western region [97]. Mesozoic volcanics are categorized by the Tamulangou, Manketouebo, Manitu, Baiyingaolao, and Meiletu Formations (Fms), from oldest to youngest [93]. The Palaeozoic strata were extensively emplaced through substantial Palaeozoic to Mesozoic granitic intrusions. The latest geochronological evidence indicates that the magmatism within the NCGXR took place in two periods: the Palaeozoic granites consisting of granite, monzogranite, granodiorite, quartz diorite, and syenogranite mainly distributed in the northeastern part with zircon U-Pb ages of 517–420 and 359–250 Ma; and the Mesozoic granites, comprising monzogranite, syenogranite, granodiorite, granite porphyry, quartz porphyry, and quartz monzogranite, which are more widespread than the Palaeozoic intrusions and yielded a zircon U-Pb age of 250–118 Ma [29,98,99,100,101,102,103,104] (Figure 1c).

3. Ore Deposit Geology

The Lower Urgen molybdenum deposit (120°27′30″ E, 50°22′42″ N), situated in the central NCGXR, discovered in 2017, is still in exploration and not in production (Figure 1c). The strata exposed in the mining area comprise the Tamulangou Fm of the Middle Jurassic, Manketouebo Fm of the Upper Jurassic, Baiyingaolao Fm of the Lower Cretaceous, and Quaternary [30] (Figure 2). The Tamulangou Fm comprises intermediate–basic volcanics including trachyandesite, andesite, and basaltic-trachyandesite. The Manketouebo Fm comprises felsic volcanics including volcanic breccia, rhyolite, and volcanic agglomerate. The Baiyingaolao Fm is locally dispersed in the northwestern part, which is composed of rhyolite, rhyolitic breccia lava, and tuff [30]. Mesozoic granitoids include quartz monzonite porphyry, syenite porphyry, and granite porphyry. The Middle Jurassic quartz monzonite porphyry outcrops are scattered in the western and central regions. The Early Cretaceous syenite porphyry locally occurs in the western region, while the Early Cretaceous granite porphyry outcrops are mostly distributed in the eastern region [30] (Figure 2). Zhang et al. [31] obtained precise zircon U-Pb ages of the barren quartz monzonite porphyry and syenite porphyry of 165.0 ± 2.5 Ma and 125.5 ± 1.9 Ma, respectively. Mineralization is characterized by concealed veinlet-disseminated- and vein-type quartz–sulfide orebodies mostly hosted in the cupola of the Early Cretaceous granite porphyry (Figure 2c and Figure 3). Ore minerals are mainly molybdenite and pyrite, as well as subordinate chalcopyrite, sphalerite, arsenopyrite, and galena (Figure 3). K-feldspar, quartz, and sericite are the predominant gangue minerals, accompanied by epidote, biotite, and calcite (Figure 3g and Figure 4). Molybdenite occurs as medium-coarse grained (1.2–7 mm) crystals with platy and leaf-like forms in ore-bearing quartz veins and granite porphyry (Figure 3a–c). Pyrite is present as euhedral to subhedral octahedral–dodecahedral crystals, and usually coexists with molybdenite, chalcopyrite, sphalerite, and galena in granular aggregate forms (Figure 3). Sphalerite occurs as medium-coarse grained (1.5–7 mm) anhedral crystals, and is often intergrown with galena, chalcopyrite, and pyrite (Figure 3e,f,i). Chalcopyrite appears as fine-medium grained (0.4–1.3 mm) irregular crystals. It often coexists with pyrite and occurs in small amounts with galena and sphalerite (Figure 3d,e,i). Arsenopyrite is observed as rhomboid euhedral crystals (Figure 3d). Galena is observed as fine-medium-grained (0.6–4.5 mm) irregular crystals, and is often coeval with sphalerite, pyrite, and chalcopyrite in quartz veins (Figure 3e,f,i). Superposition mineralization in the cupola of the granite porphyry is ubiquitous, and no obvious metallogenic zonation is observed.
Intensive hydrothermal alteration is a defining characteristic of the Lower Urgen molybdenum deposit. There are three alteration zones recognized according to field observations and thin-section petrography: an intense potassic alteration zone at depth, a phyllic alteration zone at mid-shallow depths, and a propylitic alteration zone near the surface and periphery of the deposit. The potassic and phyllic alteration zones are all distributed in the granite porphyry, while the propylitic alteration zone mainly occurs in the intermediate–mafic volcanics around the granite porphyry. The spatial relationship of the three alteration zones is delineated in Figure 2c. The potassic alteration has a high temperature and is featured by a mineral assemblage of K-feldspar, biotite, and quartz [1,3] (Figure 4a). Quartz–molybdenite veinlets are distributed in the granite porphyry with intense potassic alteration (Figure 4a). The phyllic alteration is distinguished by the replacement of primary K-feldspar and plagioclase in the granite porphyry with quartz and sericite (Figure 4e). Quartz–pyrite–chalcopyrite, quartz–galena–sphalerite, and quartz–molybdenite(–pyrite) veins are usually developed in the phyllic alteration zone (Figure 4b,c). The propylitic alteration generally represents low-temperature hydrothermal alteration, which is featured by plagioclase replaced by epidote [1,3] (Figure 4d,f).
There are three main stages of hydrothermal processes in the Lower Urgen molybdenum deposit according to crosscutting relationships and mineral paragenetic sequences, including the quartz–molybdenite(–pyrite) stage (stage 1), quartz–polymetallic sulfide stage (stage 2), and quartz–carbonate stage (stage 3).
Stage 1 is the main molybdenum mineralization stage. Molybdenite, pyrite, quartz, and sericite are developed in this stage. Various types of veins were formed, including quartz–molybdenite veins and veinlets (Figure 3a and Figure 4a), quartz–molybdenite stockworks (Figure 3f), and quartz–molybdenite–pyrite veins (Figure 3b and Figure 4b). Disseminated molybdenite ores can also be found (Figure 3c).
Stage 2 is featured by an assemblage of chalcopyrite, sphalerite, arsenopyrite, pyrite, galena, quartz, sericite, and epidote. The most common sulfide is pyrite, which appears as fine-medium-grained euhedral to subhedral crystals and aggregates, which often form quartz–sulfide veins with chalcopyrite, galena, sphalerite, and quartz, i.e., quartz–pyrite–chalcopyrite–arsenopyrite veins (Figure 3d), quartz–chalcopyrite–pyrite veins (Figure 4c), quartz–pyrite–chalcopyrite–sphalerite–galena veins (Figure 3e), and quartz–pyrite–sphalerite–galena veins (Figure 3f). Galena and sphalerite are often intergrown, and they form quartz–galena–sphalerite veins and veinlets (Figure 4b,d). The brecciated fragments of granite porphyry containing stage 1 quartz–molybdenite stockworks can be found in stage 2 quartz–pyrite–sphalerite–galena veins (Figure 3f). Additionally, stage 2 quartz–galena–sphalerite veins cut stage 1 quartz–molybdenite–pyrite veins (Figure 4b).
Stage 3 represents the termination of the hydrothermal activity in the Lower Urgen molybdenum deposit. During this stage, quartz and calcite make up the majority of the veins, and sulfides are no longer found in this veins’ stage (Figure 3g).

4. Sampling and Analytical Methods

4.1. Samples

Samples for this study were all collected from drillholes, containing three granite porphyry samples, twenty-nine sulfide samples, and twenty-eight quartz samples. The sample locations are indicated in Figure 2c. All granite porphyry samples (T0a1-1, T0a1-2, and T16-1) were fresh and used for whole-rock Pb isotope analyses. Twenty-nine sulfide samples comprised sixteen molybdenite samples, seven pyrite samples, four chalcopyrite samples, and two galena samples. Sixteen molybdenite samples were obtained from stage 1 molybdenite(–pyrite)-bearing quartz veins, while the other sulfide samples were all collected from stage 2 sulfide-bearing quartz veins. Two molybdenite samples (G5 and LE-1) were used for Re-Os isotope analyses. Six molybdenite samples (LU16-1, LU16-2, LU16-3, LU16-4, LU16-5, and LU16-6), two pyrite samples (LU16-7 and LU16-8), two chalcopyrite samples (LU16-9 and LU16-10), and one galena sample (LU16-11) were used for S isotope analyses. Eight molybdenite samples (D17-2, D17-3, D17-4, D17-5, D17-7, D0A1-1, D0A1-2, and D0a1-1), five pyrite samples (G10-2, G11-2, G13, G14, and D17-1), two chalcopyrite samples (D17-2 and G8-2), and one galena sample (G8-3) were used for Pb isotope analyses. The quartz samples were obtained from veins of all stages, including seven samples from stage 1 molybdenite(–pyrite)-bearing quartz veins, nine samples from stage 2 sulfide-bearing quartz veins, and five samples from stage 3 quartz–carbonate veins. Seven quartz samples of stage 1 veins (B16-1, B16-2, B16-3, B16-4, B16-5, B16-6, B16-7), seven quartz samples of stage 2 veins (D14-1, D16-1, D16-4, D16-6, D16-7, D16-8, and D16-9), and three quartz samples of stage 3 veins (G8-1, G10-1, and G11-1) were used for H-O isotope analyses. Seven quartz samples of stage 1 veins (BG16-1, BG16-2, BG16-3, BG16-4, BG16-5, BG16-6, and BG16-7), two quartz samples of stage 2 veins (BG8-1 and BG14-1), and two quartz samples of stage 3 veins (BG8-2 and BG8-3) were used for the FI study. For investigations of FIs in quartz cores, thirty-six double-polished thin sections (200 μm thick) were prepared.

4.2. Molybdenite Re-Os Isotope Dating

Re-Os isotope analyses were carried out at the State Key Laboratory of Ore Deposit Geochemistry, Institute of Geochemistry, Chinese Academy of Sciences in Guiyang, China. A TJA PQ ExCell ICP-MS was applied to measure the contents of Re, 187Re, and 187Os. We followed the analytical procedures reported in [105,106]. The Re-Os model ages were calculated using the formula t = 1/λ [ln (1 + 187Os/187Re)], where λ is the decay constant for 187Re (1.666 × 10−11 year−1) [107].

4.3. FIs Laser Raman Spectroscopy

A Renishaw inVia laser Raman probe at the Beijing Createch Testing Technology Co., Ltd. in Beijing, China, was applied to examine the vapor and liquid components of quartz-hosted FIs at room temperature. For detection, a laser power of 20 mW and an argon ion laser with a wavelength of 514.5 nm were used. The spectrum was measured from 1000 to 4000 cm−1 with a counting time of 20 s for each scan. Detailed analytical procedures are consistent with those in [108].

4.4. FIs Microthermometry

FI microthermometry was conducted using a Linkam THMS 600 heating-freezing stage with a stage of –195 to 600 °C mounted on a Zeiss microscope at the Beijing Createch Testing Technology Co., Ltd., Beijing, China. The accuracy of temperature observations is 0.2 °C through cooling, ±0.2 °C for 0–100 °C, and ±2 °C for 100–450 °C. To properly record the phase transformation process, the heating rate was decreased to 0.1 °C/min close to the freezing point and 0.2 °C/min close to the homogenization temperature. To prevent unintentional deviations, FI results were only collected on FI assemblages (FIAs) [109]. During heating, seven different temperature types were recorded, including CO2-ice melting temperatures (TmCO2), first ice melting temperatures (Tfmice), final ice melting temperatures (Tmice), final clathrate melting temperatures (Tmclath), CO2 homogenization temperatures (ThCO2), halite dissolution temperatures (TmH), and final homogenization temperatures (Th). Tmclath was used to determine the salinities of the fluid system composed of NaCl, CO2, and H2O [110]. Tmice and TmH were used to obtain the salinities of the fluid system composed of NaCl and H2O [111,112]. Using the Flincor program, the bulk molar volume of FIs, the density of the bulk fluids and carbonic liquid, and mole fractions of compositions were acquired [113,114,115].

4.5. Quartz H-O Isotopes

H-O isotope analyses were carried out using a MAT253 mass spectrometer at the Beijing Research Institute of Uranium Geology in Beijing, China. The accuracies of the H and O isotopic measurements are better than ±2‰ and ±0.2‰, respectively. The fractionation formula 1000 lnαquartz water = 3.38 × 106 T−1–3.40 was applied to calculate the δ18OH2O of fluids, where T is the Kelvin temperature [116]. The average FI homogenization temperature for each stage was used to determine the δ18OH2O values.

4.6. S-Pb Isotopes

The sulfide S-Pb isotope measurements were performed at the State Key Laboratory of Isotope Geochemistry, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences. A Delta V Plus mass spectrometer was used to analyze the S isotopic compositions of sulfides with an analytical accuracy of ±2‰. Sulfide δ34S values were calculated on sulfur dioxide produced via the traditional extraction method described in [117]. The δ34S results were recorded using the Canyon Diablo Troilite (CDT).
Thermal ionization mass spectrometry using an ISOPROBE–T mass spectrometer was used to analyze the Pb isotopic compositions of sulfides and granite porphyry. The accuracy for 1 μg of 208Pb/206Pb and 207Pb/206Pb was greater than 0.005% (2σ). The Pb isotopic analyses are reported with respect to Pb standard reference NBS-981 values, as follows: 206Pb/204Pb = 16.934 ± 0.007, 207Pb/204Pb = 15.486 ± 0.012, and 208Pb/204Pb = 36.673 ± 0.033 [118].

5. Results

5.1. Molybdenite Re-Os Ages

Table 1 lists the results of Re and Os isotope compositions of two molybdenite samples from the Lower Urgen deposit. The 187Re and 187Os contents of the samples vary from 19,468 to 27,503 ppb and 47.958 to 64.540 ppb, respectively. The Re contents of the samples are 31,202–43,934 ppb. The two Re-Os model ages are 141.2 ± 1.5 and 147.7 ± 1.7 Ma.

5.2. FI Petrography

Primary and secondary FIs were identified based on the criteria suggested by [109,119]. All descriptions strictly follow the concept of FIAs [109]. An FIA only applies to FI vacuoles that developed along the identical growth bands in an individual crystal, as well as within the same T-P circumstances. Nonetheless, extremely few FIs are dispersed within or along the growth bands, and the bulk is too small to be quantified (<5 μm). Most FIs that are suitable for measurement appear as isolated, clustered, or randomly distributed (Figure 5). As suggested by [109,119,120,121], FIs that are away from apparent cracks and appear as isolated, clustered, or randomly distributed forms are considered as primary FIs, and these FIs are selected for further investigation. In contrast, FIs along healed fractures that have apparent secondary causes are not considered. Four different types of FIs are recognized based on phase components at room temperature and phase changes due to variations in temperature, including the two-phase (L + V) liquid-rich (L-type), saline (S-type), CO2-rich (C1-type), and CO2-bearing (C2-type) FIs. The descriptions of the petrographic features of each FI type are as follows.
L-type FIs are ubiquitous in all stages of quartz. At room temperature, they have a vapor phase and a liquid phase with LH2O/(LH2O + VH2O) ratios of 65–90 vol% (Figure 5). With diameters ranging from 5 to 15 μm, these FIs can be ellipsoidal, rectangular, negative-crystal, or irregular in shape. L-type FIs occur alongside S-, C1-, and C2-type FIs in stage 1 as immiscibility FIAs (Figure 5a–c). They are also intergrown with C1- and C2-type FIs and form immiscibility FIAs in stage 2 quartz crystals (Figure 5g). In quartz veins of stages 1 and 2, L-type FIs are abundant and make up around 50% of all FIs; however, stage 3 veins solely contain L-type FIs.
Only stage 1 quartz veins have S-type FIs, which exhibit high salinities and densities. S-type FIs are composed of a cubic halite daughter mineral, a liquid phase, and a vapor bubble (3–10 vol%). They are elliptical, negative-crystal, or irregular in shape, with diameters of 7 to 15 μm. S-type FIs can exist alone or with L-, C1-, and C2-type FIs to form immiscibility FIAs in stage 1 (Figure 5b–d). In abundance, S-type FIs account for about 10% of all FIs in stage 1 veins.
C1-type FIs occur as ellipsoidal, circular, or negative-crystal shapes, and have a diameter of 5 to 13 μm. The CO2 phase content ranges from 60 to 85 vol%. Usually, 60 to 80 vol% of the entire CO2 phase is made up of vapor CO2. C1-type FIs commonly appear in the three-phase form (VCO2 + LCO2 + LH2O), with a small amount of the two-phase form (VCO2 + LH2O) (Figure 5a,b,e,g). The two-phase FIs are regarded as CO2-rich FIs as clathrates were observed during freezing runs. In abundance, C1-type FIs frequently coexist as immiscibility FIAs with L-, S-, and C2-type FIs (Figure 5a–c,g), and account for about 25 vol% and 30 vol% of all FIAs in veins of stages 1 and 2, respectively.
At room temperature, C2-type FIs exist in the three-phase (VCO2 + LCO2 + LH2O) state, and minor FIs display a two-phase (VCO2 + LH2O) form (Figure 5b,g). During freezing, clathrates are also observed in these two-phase C2-type FIs and can therefore be distinguished from L-type FIs. Typically, vapor CO2 makes about 60–80 vol% of the total CO2 phase, and the amount of CO2 phase varies from 15 to 45 vol%. These FIs are irregular or nearly ellipsoidal in shape with a diameter of 5–15 μm. C2-type FIs are intergrown with L- and C1-type FIs (Figure 5g), together with small quantities of S-type FIs (Figure 5b), and account for approximately 15 vol% and 20 vol% of all FIAs in veins of stages 1 and 2, respectively.

5.3. FI Laser Raman Spectroscopy

Figure 6 illustrates the components of typical FIs from quartz veins of all stages. A substantial volume of H2O and a negligible amount of CO2 (1385 cm−1, 1283 and 1388 cm−1, respectively) were detected in both the liquid and vapor phase in L-type FIs of stage 1 (Figure 6a). C1-type FIs of stage 1 contain CO2 as the only species in both the vapor and liquid CO2 phases (1284 and 1388 cm−1, 1282 and 1386 cm−1, respectively) (Figure 6b). C2-type FIs of stage 1 have a similar vapor and liquid phase composition to stage 1 C1-type FIs, with CO2 as the only component of the CO2 phase (1283 and 1388 cm−1, 1281 and 1386 cm−1, respectively) (Figure 6c). Analogous liquid and vapor phase components, which are predominately made of H2O and include a little amount of CO2 (1385 cm−1, 1284 and 1388 cm−1, respectively), are present in S-type FIs of stage 1 (Figure 6d). Consequently, a NaCl–H2O–CO2 system is represented by the fluids in the quartz–molybdenite(–pyrite) stage. C1-type FIs of stage 2 show typical CO2 spectra (1282 and 1387 cm−1) in the vapor CO2 phase (Figure 6e). The compositions of the vapor and liquid CO2 phase of C2-type FIs in stage 2 only contain CO2 (1281 and 1386 cm−1) (Figure 6f). H2O is the predominant gas species in the vapor phase of L-type FIs in stage 2 quartz, with only trace amounts of CO2 (1387 cm−1). In contrast, only H2O was detected in the liquid phase of L-type FIs in stage 2 quartz (Figure 6g). Thus, a NaCl–H2O–CO2 system is represented by the fluids in the quartz–polymetallic sulfide stage. L-type FIs of stage 3 lack CO2 detection and only have H2O for the compositions of the liquid and vapor phases, demonstrating the characteristic of a NaCl–H2O system (Figure 6h).

5.4. FI Microthermometry

Only primary FIs were chosen for microthermometry to prevent interpretation errors brought by the post-ore modification. Using the FIA concept, the reliability of microthermometric results has been assessed [109]. Large disparities in homogenization temperatures in spatially close FIs prevent them from being unequivocally classified as an FIA; instead, they are thought to be the result of necking, stretching, or heterogeneous entrapment [120]. The microthermometry data for isolated FIs are considered, barring a significant discrepancy with surrounding FIs [121]. If the homogenization temperature variation within a cluster of a single FI type is below 15 °C, the possibility of post-entrapment modification, generational overlap, or heterogeneous trapping of FIs is ruled out, and it is argued that these FIs represent a uniform fluid entrapment. We thus suggest that those results are trustworthy and that the overall average of a cluster is adequate for additional research [120]. An immiscibility FIA is a group of FIs of various types that possess comparable ranges of homogenization temperature (variation below 15 °C) but distinct homogenization processes. In this case, the complete FIA was represented for statistical analysis by the average of several types of FIs.
In this investigation, microthermometric results for 320 FIAs of three stages are listed in Table S1 (see the Supplementary Materials) and illustrated in Figure 7. The Tfmice for L-type FIs of all stages ranges from –21.6 to –20.8 °C, consistent with the eutectic temperature of the H2O–NaCl system, indicating that most of the dissolved salts are NaCl [111,122]. As mentioned in Section 5.3, the L- and S-type FIs sometimes display weak CO2 signals, but the CO2 characteristics cannot be observed during the petrography observation and microthermometry of these FIs. The detection limit of CO2 through laser Raman spectroscopy can be as low as 1 bar, while the phase transition of CO2 can be observed in microthermometry with a CO2 pressure of at least 10.4 bar (the melting of CO2 clathrate can be observed) or 45 bar (solid phase of CO2 melting can be observed) [123]. In summary, the amount of CO2 in the S-type FIs of stage 1 and L-type FIs of stages 1 and 2 is extremely low and almost negligible, and thus the salinities of these FIs were obtained using the TmH and Tmice, respectively.
The entrapped fluid characteristics are often dramatically changed through post-entrapment modification (dismemberment and/or migration) in quartz-hosted FIs [124,125]. Euhedral quartz crystals that are naturally formed frequently exhibit these modifications in addition to quartz phenocrysts and veins. Although not verified, a satisfactory FI preservation state is recommended if the FIs display constant phase ratios, develop in clear trails or growth bands, and provide densely concentrated microthermometric values [125]. On the contrary, regular negative-crystal forms are neither necessary nor indicative of the preservation of syn-entrapment characteristics of FIs. Unless they adhere to specified growth bands, clusters of randomly scattered FIs often reflect secondary or pseudosecondary FI trails that have dissolved. Audétat [124] demonstrated that the decline in FI volume caused by post-entrapment modification results in a decrease in the homogenization temperature, a steepening of the isochore, and an increase in fluid density. By contrast, in magmatic–hydrothermal systems, water loss from FIs seldom surpasses twenty percent; consequently, the fluid salinities that are determined should, in most instances, be almost accurate. In the current research, FIs inside or along the growth zones of euhedral quartz occur incredibly seldom, and the bulk of them are too small to investigate (<5 μm). Post-entrapment modification may have occurred to varied degrees in clustered, isolated, and randomly dispersed FIs, but it seems inescapable. However, to guarantee that the chosen FIAs only had slight effects of post-entrapment modification, FIAs that displayed obvious post-entrapment modification features, e.g., deformation, movement tracks, stretching, shrinkage, or deteriorating halos, have been excluded [119,124,125,126]. Nonetheless, the genuine densities and homogenization temperatures of the examined FIs are slightly lower and higher than the calculated data, respectively. Comparatively, FI salinities are frequently less influenced, and most of them tend to have maintained their original salinities. As a result, we primarily concentrate on the salinities of the examined FIs, while the homogenization temperatures are used as a reference.
Stage 1 veins contain L-, S-, C1-, and C2-type FIs. The Th of L-type FIs ranges from 240 to 400 °C, and the Tmice varies between –8.2 and –5.1 °C (corresponding to 8.0–11.9 wt% NaCl equiv.). The disappearance of the vapor bubbles indicates complete homogenization. The TmH of S-type FIs is between 160 and 428 °C (corresponding to 30.1–50.8 wt% NaCl equiv.). Halite dissolution or the disappearance of vapor bubbles is an indicator that S-type FIs homogenize between 245 and 428 °C. The TmCO2 of C1-type FIs ranges from –59.8 to –57.0 °C. C1-type FIs have a Tmclath of 5.9–7.4 °C and calculated salinities of 5.0–7.5 wt% NaCl equiv. At 14.5–21.5 °C, the CO2 phase homogenizes to the vapor CO2 phase. At 240–397 °C, the liquid H2O phase disappears, signifying the completion of the homogenization. C2-type FIs have a TmCO2 of –59.8 to –57.0 °C and Tmclath of 5.1–6.8 °C, corresponding to salinities of 6.0–8.8 wt% NaCl equiv. At 20.3–25.9 °C, CO2 homogenizes to the vapor CO2 phase, and at 241–395 °C, C2-type FIs finally homogenize to the liquid phase (Figure 7a,b).
L-, C1-, and C2-type FIs can be found in stage 2 quartz veins. Compared to stage 1 FIs, these FIs have a lower Th and lower salinities (Figure 7c,d). L-type FIs have a Tmice of –5.9 to –3.7 °C, corresponding to 6.0–9.1 wt% NaCl equiv. At 182–278 °C, L-type FIs homogenize and are characterized by the vapor bubbles’ disappearance. C1-type FIs yield a TmCO2 of –59.8 to –57.0 °C, Tmclath of 6.8–8.5 °C, and calculated salinities of 3.0–6.0 wt% NaCl equiv. At 14.6–22.5 °C, the CO2 phase homogenizes to the vapor CO2 phase, and at 181–277 °C, the liquid H2O phase is eliminated, indicating complete homogenization. C2-type FIs contain a TmCO2 that varies between –59.8 and –56.8 °C and a Tmclath of 6.2–7.9 °C (corresponding to 4.1–7.1 wt% NaCl equiv.). The homogeneity of the CO2 phase is featured by the elimination of the liquid CO2 phase, and the ThCO2 has a range of 21.9–25.9 °C. Total homogenization occurs when the vapor CO2 phase is eliminated, and the Th of C2-type FIs has a range of 180–277 °C (Figure 7c,d).
L-type FIs are the only FIs found in quartz–carbonate veins of stage 3, which have Tmice values between –5.0 and –0.1 °C and NaCl equivalent salinities between 0.2 and 7.9 wt%. At 120–218 °C, these FIs ultimately homogenize to the liquid phase (Figure 7e,f).

5.5. Quartz H-O Isotopic Compositions

The hydrogen and oxygen isotopic compositions of analyzed quartz samples are presented in Table 2 and shown in Figure 8. Seven quartz samples of stage 1 veins have δD and δ18O values ranging from –141.8 to –136.5‰ and 7.2–8.0‰, respectively. The δ18Ofluid values are determined to be between 1.1 and 1.9‰. The δD values of seven quartz samples collected from stage 2 veins are –144.1 to –129.9‰, the δ18O values range between 7.3 and 8.1‰, and the calculated δ18Ofluid values are –2.7 to –1.9‰. The δD and δ18O values in three quartz samples from stage 3 veins are –146.6 to –142.6‰ and 5.8–7.7‰, respectively, and the calculated δ18Ofluid values are between –8.0 and –6.1‰.

5.6. Sulfide S Isotopic Compositions

In Table 3 and Figure 9, the S isotopic compositions of molybdenite, pyrite, chalcopyrite, and galena collected from the Lower Urgen molybdenum deposit are displayed. Sulfide δ34S values range from 3.9 to 5.5‰, with an average value of 5.2‰. Six molybdenite samples contain δ34S values of 5.4–5.5‰. Two pyrite samples and two chalcopyrite samples have δ34S values of 5.4‰ and 4.8–5.4‰, respectively. One galena sample has a δ34S value of 3.9‰.

5.7. Sulfides and Whole-Rock Pb Isotopic Compositions

The Pb isotopic compositions of molybdenite, pyrite, chalcopyrite, galena, and granite porphyry are given in Table 4 and Figure 10. The 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios for eight molybdenite samples are 18.340–18.412, 15.540–15.546, and 38.130–38.212, respectively. Five pyrite samples contain 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios of 18.336–18.510, 15.542–15.569, and 38.131–38.312, respectively. The 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios of two chalcopyrite samples are 18.336–18.340, 15.541–15.547, and 38.131–38.143, respectively. One galena sample has 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios of 18.343, 15.550, and 38.149, respectively. The 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb ratios of three molybdenum-bearing granite porphyry samples range from 18.535 to 18.599, 15.571 to 15.574, and 38.419 to 38.583, respectively.

6. Discussion

6.1. Timing of Molybdenum Metallogenesis

From both scientific and economic perspectives, it is crucial to constrain the length and time of magmatic–hydrothermal systems [130]. Precise isotopic dating of hydrothermal minerals is an essential technique for determining the geochronology of hydrothermal processes in relation to intrusive magmatism and for constructing genetic models of hydrothermal deposits. As the primary product of hydrothermal mineralization, molybdenite obtained from ore-bearing quartz veins in molybdenum deposits is suggested as a Re-Os geochronometer for directly dating molybdenum mineralization events. In this contribution, two Re-Os model ages of 141.2 ± 1.5 and 147.7 ± 1.7 Ma were obtained. The wide range of Re-Os model ages could be attributed to Re and Os isotope decoupling, the considerable spatial Re heterogeneity, the post-ore deformation, and the analytical procedure [130,131,132,133,134,135,136]. For instance, many explanations based on post-crystallization diffusion of 187Os are proposed to answer the Re and Os decoupling mechanism. ReS2 inclusions or 3R polytype molybdenite formed during the crystallization of molybdenite would, over time, result in a chemical gradient for 187Os that may homogenize during a thermal event. Decoupling could also result from the exsolution of Os-rich sulfides or diffusion of Os to crystal dislocations within the molybdenite during metamorphism [133,134]. The heterogeneous distribution of Re in molybdenite samples could also lead to the discrepancy of Re-Os model ages, which may be related to the overall abundance of Re in the hydrothermal fluids [136], compositional changes in the hydrothermal fluid with time [132], or partitioning of Re between liquid and vapor [130]. Additionally, a relatively wide range of molybdenite Re-Os model ages has been reported in numerous magmatic–hydrothermal deposits, such as the Haigou Au deposit (311.3–306.7 Ma; [131]), Lala Fe-Cu deposit (1097–1081 Ma; [137]), Kenge Mo deposit (1957–1950 Ma; [138]), Porcupine Mo deposit (1896–1882 Ma; [138]), Kedengjian Cu deposit (99.1–95.7 Ma; [139]), Mada Mo deposit (150.9–136.9 Ma; [140]), Meiziwo W deposit (165.6–156.2 Ma; [141]), Yaoling W deposit (165.9–156.2 Ma; [141]), Banlashan Mo deposit (142.7–131.8 Ma; [26]), Huashandong W deposit (827.8–795.3 Ma; [142]), Jiguanshan Mo deposit (156.2–151.7 Ma; [143]), and Dongbeigou Mo deposit (132.6–127.1 Ma; [144]). Zhang et al. [31] reported that the barren quartz monzonite porphyry and syenite porphyry stocks were emplaced at 165.0 ± 2.5 and 125.5 ± 1.9 Ma, respectively. Our recent geochronology study yielded a zircon U-Pb age of 142.3 ± 0.5 Ma for molybdenum-bearing granite porphyry stock (unpublished data). In recent decades, advances in high-precision dating, numerical simulation, and diffusion modeling mostly argue for short-lived (<1 Ma) timescales for the lifetime of magmatic–hydrothermal systems [7,145,146,147,148,149,150,151,152]. The Re-Os model age of molybdenite sample G5 (141.2 ± 1.5 Ma) is highly consistent with the emplacement time of the molybdenum-bearing granite porphyry stock within the analytical uncertainty. Spatially, the molybdenum mineralization in the deposit is characterized by veinlet-disseminated- and vein-type quartz–sulfide orebodies hosted in the cupola of the ore-bearing granite porphyry stock (Figure 2c). Geochemically, the metal sulfides obtained in this study have similar lead isotopic compositions to those of the ore-bearing granite porphyry stock (discussed in Section 6.3). This evidence collectively demonstrates that the molybdenum mineralization at Lower Urgen was temporally, spatially, and genetically associated with the molybdenum-bearing granite porphyry stock, supporting an intrusion-related origin during the Early Cretaceous. In contrast, the barren quartz monzonite porphyry and syenite porphyry stocks record pre-ore and post-ore magmatic events, respectively. Our findings also indicate that accurate and precise molybdenite Re-Os ages representing the true timing of ore formation need an integrated combination of careful geology, geochemistry, and multiple dating methods.
Considering the Mesozoic intense and widespread molybdenum mineralization in NE China, molybdenum mineralization at Lower Urgen is not a singular instance. By collecting published molybdenite Re-Os ages for Mesozoic molybdenum deposits within NE China, we determine three periods of molybdenum metallogenesis events in the region: Middle–Late Triassic (245–225 Ma), Early–Middle Jurassic (200–165 Ma), and Late Jurassic–Early Cretaceous (160–130 Ma) (Table S2 and Figure 11). The Middle–Late Triassic molybdenum deposits are primarily located in the southern GXR (SGXR) and along the SXCF, with minor occurrences in the NCGXR, including the Laojiagou (234.9 Ma; [41]), Chehugou (245 Ma; [85]), Kulitu (236 Ma; [153]), and Badaguan deposits (228.8–225.1 Ma; [46,48,154]). Molybdenum deposits formed in the Early to Middle Jurassic are abundant within the LXZR, and scattered in the NCGXR, including the Daheishan (169.2–168.0 Ma; [52,155,156]), Xingshan (167.3 Ma; [53]), Jidetun (168.9–168.0 Ma; [54,58,157]), Fu’anpu (168.2–166.9 Ma; [57,158]), Houdaomu (167.5 Ma; [159]), Shuangshan (171.6 Ma; [160]), Jiapigou (196 Ma; [160]), Sancha (183.1 Ma; [160]), Dashihe (186.7 Ma; [161]), Liushengdian (169.3–168.7 Ma; [159,162]), Sifangdianzi (176.6 Ma; [59]), Luming (177.9–176.7 Ma; [61,62]), Huojihe (181.3 Ma; [64]), Xinhualong (171.3 Ma; [159]), Wunugetushan (180.5–177.6 Ma; [49,163]), and Taipingchuan deposits (200.1 Ma; [164]). Most of the molybdenum deposits dated between the Late Jurassic and Early Cretaceous are located in the GXR, including the Chalukou (150–148 Ma; [65,67]), Taipinggou (130.1–129.4 Ma; [72,165]), Jiazishan (147.4 Ma; [166]), Diyanqinamu (156.2 Ma; [167]), Hashitu (150 Ma; [75]), Aolunhua (132–129.4 Ma; [79,168,169]), Shabutai (135.3 Ma; [26]), Haisugou (136.4 Ma; [26]), Banlashan (143–136.1 Ma; [26,169]), Yangchang (138.5 Ma; [169]), Hongshanzi (138.2–137.2 Ma; [83]), Xiaodonggou (138.1–135.5 Ma; [84,170]), Nianzigou (154.3–153 Ma; [85,86,153]), and Jiguanshan deposits (155.4–151.1 Ma; [143,171]). The zircon U-Pb ages of ore-bearing granodiorites record the granitic magmatism in the Kanchuangou and Jinchanggou deposits at 111.1 [56] and 114 Ma [63], respectively. Unfortunately, no accurate molybdenite Re-Os ages of these two deposits have been reported, and the genetic relationship between molybdenum mineralization and host granodiorites remains to be further investigated. Hence, the molybdenum mineralization in these two deposits is not involved in the present study. The Early Cretaceous molybdenum mineralization in Lower Urgen is in accordance with the Late Jurassic–Early Cretaceous molybdenum metallogenesis in the NCGXR.

6.2. Source of Mineralizing Fluids

Stage 1 quartz has calculated δ18Ofluid values (1.1 to 1.9‰) that are lower than normal magmatic fluids (7–9‰; [172]) and substantially higher than normal Mesozoic meteoric water (–18‰; [173]). The δ18Ofluid values of the quartz samples from the late quartz–carbonate stage are more depleted (–8.0 to –6.1‰) and comparable to the values of Mesozoic meteoric water [127,173] (Figure 8). This δ18Ofluid variation trend implies the mixing of magmatic fluids with later meteoric water during the mineralization at Lower Urgen. Furthermore, in agreement with the documented δD values of Mesozoic meteoric water in the GXR (–140 to –90‰; [173]), the δD values of all stages (–146.6 to –129.9‰) are much lower than those of magmatic water (–50 to –85‰; [174]). The depletion of δD in the mineralizing fluids could be controlled through (1) fluid boiling or immiscibility [175], (2) magma degassing led by major meteoric water incursion [176], and (3) hydrogen isotope fractionation between the mineralizing fluids and some reduced species (e.g., CH4). As shown in Figure 5, the coexistence of multiple FI types in stages 1 and 2 indicates intense fluid immiscibility during ore precipitation. The decreasing trend of Th and salinity in the FIs from stage 1 to 3 indicates increasing meteoric water input toward the late fluid system (Figure 7). However, no reduced species such as CH4 have been detected in the H2O-dominated ore-forming fluids, which indicates that hydrogen isotope fractionation is likely negligible [127]. Hence, we preliminarily speculate that fluid immiscibility and meteoric water input could explain the δD depletion of ore-forming fluids at Lower Urgen. In the δD versus δ18Ofluid diagram, the calculated δ18Ofluid and δD values also show a clear evolution trend from stage 1 to 3, and the data are plotted between the magmatic and meteoric water fields (Figure 8). Therefore, we conclude that the mineralizing fluids of the Lower Urgen deposit were primarily of magmatic origin with progressive meteoric water input.

6.3. Source of Sulfur and Lead

S isotopes can be used to trace the source of sulfur in sulfide minerals, despite the fact that the entire S isotopic components could be influenced by pH, temperature, ion activity, and fO2 [177,178,179]. The δ34S data of sulfides can be utilized to determine their origin because no sulfate minerals have been found in the Lower Urgen district. The molybdenite, pyrite, chalcopyrite, and galena have δ34S values of 5.4–5.5‰, 5.4‰, 4.8–5.4‰, and 3.9‰, respectively, implying that these sulfides have reached sulfide equilibrium fractionation [179]. The sulfides δ34S values of the Lower Urgen deposit display a restricted range (3.9–5.5‰), which is commonly interpreted as a magma-derived sulfur [180] (Figure 9).
The source of the lead in the mineralizing fluids has been constrained using the lead isotope ratios of sulfides. The Pb isotopic ratios of sulfides can provide information about the source of Pb since sulfide minerals often have minimal radiogenic Pb isotopes and extremely low quantities of Th and U [129,181]. By contrasting the Pb isotope ratios of molybdenite, pyrite, chalcopyrite, galena, and molybdenum-bearing granite porphyry, the Pb source of the Lower Urgen molybdenum deposit can be determined. The sulfide samples have narrow ranges of 206Pb/204Pb (18.336–18.510), 207Pb/204Pb (15.540–15.569), and 208Pb/204Pb (38.130–38.312), implying an individual or successfully mixed lead origin. The Pb isotopic ratios of sulfides in the Lower Urgen deposit are distinct from those of the surrounding Mesozoic strata (e.g., Tamulagou and Manketouebo Fms) (206Pb/204Pb = 18.08–18.32, 207Pb/204Pb = 15.46–15.55, and 208Pb/204Pb = 37.62–38.10) [93,182], but similar to those of the ore-bearing granite porphyry stock in the deposit (206Pb/204Pb = 18.535–18.599, 207Pb/204Pb = 15.571–15.574, and 208Pb/204Pb = 38.419–38.583). Lead with high μ values (>9.58) was derived from the upper crust [129], whereas that with values of 8.92 was derived from the mantle [183]. The sulfides and granite porphyry stock from the Lower Urgen deposit yield μ values of 9.35–9.40 and 9.39–9.40, respectively, indicating that lead was basically from the crustal source and an extremely small amount of the mantle source. On the diagram of 207Pb/204Pb versus 206Pb/204Pb [129], the sulfide and ore-bearing granite porphyry samples are all located between the mantle and orogen evolution curves (Figure 10a). On the diagram of 208Pb/204Pb versus 206Pb/204Pb [129], most sulfide and ore-bearing granite porphyry samples are distributed on the orogen evolution curve (Figure 10b). These features imply that the metal sulfides and granite porphyry at Lower Urgen have the same lead source, which has been documented in many porphyry deposits [2,3,184], and that the granitic magmatism is probably largely responsible for the molybdenum mineralization in the Lower Urgen deposit.

6.4. Metal Precipitation Mechanisms

Fluid immiscibility, one of the most significant ore-forming mechanisms in magmatic–hydrothermal metallogenic systems, can change the temperature, pH, and salinity of the ore-bearing fluids, thereby reducing the solubility of metals in the fluids [14,185,186,187,188]. The occurrence of fluid immiscibility is supported by the following evidence from stage 1 quartz-hosted FIs: (1) L-, S-, C1-, and C2-type FIs have intimate spatial relationships (Figure 5a–c); (2) they exhibit distinct homogenization modes; and (3) these FIs show distinct salinities and comparable homogenization temperature ranges (Figure 7a,b). Similarly, at stage 2, L-, C1-, and C2-type FIs frequently coexist, indicating that they were simultaneously captured and showed different homogenization modes (Figure 5g and Figure 7c,d). Notably, in stages 1 and 2, C1- and C2-type FIs frequently show lower salinities but analogous homogenization temperatures than L-type FIs, which shows that fluid immiscibility occurred before their entrapment [189]. The increased salinity of the residual fluids caused by volatile separation from the ore-bearing fluids as temperature and pressure drops may be the cause of the higher salinities in L-type FIs than those in C1- and C2-type FIs [185,190]. The mineralizing fluids at Lower Urgen underwent fluid immiscibility processes during stages 1 and 2, as reflected by the diagram of salinity versus homogenization temperature (Figure 12). Field evidence, such as the widespread fracture and brecciation of granite porphyry cupola and the overlying intermediate–basic volcanic rocks at Lower Urgen, suggests a dramatic pressure drop that might result in fluid immiscibility, further supporting the existence of fluid immiscibility [119,191,192,193]. Previous research has demonstrated that a mafic magmatic source or the magmatic–hydrothermal formation procedure of a deep-seated magma chamber often contributes significant amounts of dissolved CO2 [119,194,195,196,197,198]. In combination with the H-O isotopic compositions of stage 1 quartz, we speculate that the original fluids of the Lower Urgen molybdenum deposit are CO2-rich magmatic–hydrothermal fluids. As the fluids rose, volatiles (such as CO2) started to separate from the fluids due to a decrease in pressure and temperature. The sudden supersaturation of the solution was caused by the release of volatiles, which also increased the pH of the solution and caused the decoupling of metal complexes. As a result of the procedure, significant amounts of molybdenite precipitated from stage 1 fluids, while chalcopyrite, pyrite, sphalerite, arsenopyrite, and galena were deposited from stage 2 fluids.
Hydrothermal mineralization frequently involves fluid–rock interaction, which is essential for the transport and deposition of metal ions, especially in an intrusive-related mineralization system [4,199,200,201]. Many investigations have demonstrated that changes in the physicochemical properties of fluids as a result of fluid–rock interactions affect metal solubility and mineral precipitation [202,203,204,205]. Different hydrothermal alteration types are characteristics of porphyry molybdenum mineralization, and phyllic alteration is frequently seen as a sign of prospecting [201,203,205,206], as indicated by the fact that molybdenum mineralization in the deposit is spatially related to phyllic alteration, and plagioclase and K-feldspar of the ore-bearing granite porphyry tend to be replaced by quartz and sericite in stage 1 (Figure 4b,e). Phyllic alteration lowered the H+ contents of the hydrothermal fluids and enriched the reduced sulfur (S2−), which mixed with Mo2+ to generate molybdenite. The following formulas reflect the pertinent reactions:
3KAlSi3O8 + 2H+ → KAl3Si3O10(OH)2 (sericite) + 6SiO2 (quartz) + 2K+
3NaAlSi3O8 + K+ + 2H+ → KAl3Si3O10(OH)2 (sericite) + 6SiO2 (quartz) + 3Na+
H2S → H+ + HS
HS → H+ + S2−
Therefore, we provide a preliminary argument that phyllic alteration could benefit the deposition of molybdenite at Lower Urgen. Nevertheless, a more detailed explanation to this conclusion is beyond the scope of our current study and additional alteration mineralogical and geochemical research is necessary before a full understanding can be attained.
Numerous studies have shown that fluid mixing is a crucial factor in the production of base metal sulfide deposits because it can dilute and cool hydrothermal solutions, which causes enormous sulfide mineral depositions [207,208,209,210,211]. Mineral deposition occurs when the solubility of metal ions in the fluids surpasses saturation, and substantial deposition occurs as the solubility is dramatically decreased [208]. In the Lower Urgen deposit, the continuous fluid immiscibility in stages 1 and 2 would have increased the permeability of the host granite porphyry stock, and further accelerated meteoric water circulation. Stage 2 fluids have much lower salinity and homogenization temperatures than stage 1 fluids (Figure 12). Additionally, stage 2 fluids have more depleted δ18Ofluid values than stage 1 fluids (Figure 8). These features collectively demonstrate the existence of fluid mixing processes in stage 2. To summarize, fluid mixing propelled the polymetallic sulfide mineralization in stage 2 at Lower Urgen. The progressive incursion of meteoric water into the magmatic–hydrothermal system might cause considerable cooling, the instability of metal chloride complexes, and sulfide deposition.
In conclusion, stage 1 molybdenite deposition is primarily controlled through fluid immiscibility and fluid–rock interaction. The increasing injection of meteoric water into the hydrothermal system and fluid immiscibility collectively controlled the massive deposition of polymetallic sulfides in stage 2.

6.5. Metallogenic Model for the Lower Urgen Magmatic–Hydrothermal System

Especially for intrusion-related magmatic–hydrothermal deposits, the development of a metallogenic model may successfully direct the investigation of the ore deposit and increase the successful rate of prospecting and prediction [4,212]. We construct the metallogenic model for the Lower Urgen molybdenum deposit upon the above discussion and previous research, as outlined below.
Due to the closure of the MOO, the surveyed region experienced an extensional environment in the Late Jurassic and Early Cretaceous, which promoted the partial melting of the lower crust and produced ore-bearing granitic magmas [25,50,213,214,215,216,217,218]. As a lithophile element, molybdenum tends to be abundant in the crust but deficient in the mantle [219]. Consequently, a potential source for molybdenum mineralization might be the lower crust under the Lower Urgen region. Metallic elements are partitioned into the melt through partial melting of the protolith. The ore-bearing granite porphyry, which represents the parental rock and served as a constant source of hot, ore-rich fluids for the generation of the hydrothermal mineralization system, was emplaced at about 142.3 Ma. The potassic and phyllic alteration zones were generated by these fluids accumulating at the cupola of the mineralization intrusion and interacting with cooling granite porphyry (Figure 2c). In the meantime, intense and widespread fluid immiscibility was created through the constant ascent and decompression of mineralizing fluids. Stage 1 fluids belong to a NaCl–H2O–CO2 system and contain bimodal salinities (5.0–11.9 and 30.1–50.8 wt% NaCl equiv.) and have medium–high temperatures (240–430 °C). Intense fluid immiscibility and fluid–rock interaction caused an increase in pH and a drop in temperature, which significantly reduced the solubility of molybdenum and aided in the precipitation of molybdenite. The molybdenum mineralization stage was characterized by widespread molybdenum-bearing quartz veinlets and disseminated molybdenite in the granite porphyry. In stage 2, massive meteoric water mixed with the evolved magmatic–hydrothermal fluids through localized regional hydraulic fractured and brecciated zones. Stage 2 fluids have salinities and homogenization temperatures of 3.0–9.1 wt% NaCl equiv. and 180–280 °C, respectively, and belong to a NaCl–H2O–CO2 system. The molybdenum mineralization came to an end with the extinction of veinlet-disseminated- and vein-type quartz–molybdenite(–pyrite) ores. The constant infiltration of meteoric water diluted ore-forming fluids and decreased their temperature. This process, combined with extensive fluid immiscibility, collectively led to the large-scale decoupling of base metal complexes and promoted the massive deposition of polymetallic sulfides. Stage 3 fluids belong to a NaCl–H2O system and have salinities of 0.2 to 7.9 wt% NaCl equiv. and homogenization temperatures of 120 to 220 °C, suggesting a hydrothermal environment with low salinity and temperature. The hydrothermal fluids were governed by meteoric water as the pressure of the system was dropped. The fluids produced quartz and calcite as precipitates. The absence of ore minerals in the veins of this stage suggests that the fluids are no longer useful for mineralization, signaling the termination of the hydrothermal process in the Lower Urgen region.

7. Conclusions

(1) Molybdenite Re-Os dating yields two model ages of 141.2 ± 1.5 and 147.7 ± 1.7 Ma for the Lower Urgen molybdenum deposit. The new geochronological data indicate that the molybdenum mineralization event in the Lower Urgen area is consistent with the intensive and extensive Late Jurassic–Early Cretaceous molybdenum mineralization events in NE China.
(2) Three hydrothermal stages of the Lower Urgen molybdenum deposit can be distinguished: (i) quartz–molybdenite(–pyrite) stage—ore-forming fluids have medium–high temperatures (240–430 °C) and bimodal salinities (5.0–11.9 and 30.1–50.8 wt% NaCl equiv.) and belong to a NaCl–H2O–CO2 system; (ii) quartz–polymetallic sulfide stage—ore-forming fluids have medium temperatures (180–280 °C) and low salinities (3.0–9.1 wt% NaCl equiv.) and belong to a NaCl–H2O–CO2 system; and (iii) quartz–carbonate stage—fluids have low temperatures (120–220 °C) and low salinities (0.2–7.9 wt% NaCl equiv.) and belong to a NaCl–H2O system.
(3) Hydrogen–oxygen isotope results indicate that the mineralizing fluids initially had a magmatic source with the progressive incursion of meteoric water. Sulfur–lead isotopic compositions indicate that the mineralizing materials have a genetic relationship with ore-bearing granite porphyry, and sulfur and lead are magmatically derived.
(4) Fluid immiscibility and fluid–rock interaction are the predominant mechanisms for molybdenite deposition in stage 1. Fluid mixing and immiscibility are cooperatively accountable for the massive deposition of polymetallic sulfides in stage 2.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/min13091189/s1, Table S1: Microthermometry data for fluid inclusions in quartz from the Lower Urgen molybdenum deposit; Table S2: Molybdenite Re-Os isochron ages for Mesozoic molybdenum deposits in NE China [26,41,43,46,48,49,52,53,54,57,59,61,62,64,65,67,72,73,75,79,81,83,84,85,86,143,153,154,155,156,158,159,160,161,162,163,164,165,167,168,169,170,171].

Author Contributions

Conceptualization, G.Z. and W.X.; methodology, G.Z., H.W. and K.Z.; formal analysis, W.X.; investigation, G.Z., W.X., S.W., H.W., K.Z., T.T. and P.M.; data curation, G.Z. and W.X.; writing—original draft preparation, G.Z. and W.X.; writing—review and editing, W.X., S.W., Q.Z. and L.Z.; visualization, W.X.; supervision, W.X.; funding acquisition, Q.Z. 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, grant number 91962104.

Data Availability Statement

The data presented in this study are available in the article and Supplementary Materials.

Acknowledgments

We are grateful to three anonymous reviewers for their constructive comments, which greatly help in improving our paper. We would like to thank the managers of the Baojin Mining Group Co., Ltd. for their support of our fieldwork.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kirkham, R.V. Porphyry deposits. In Report of Activities Part B; Blackadar, R.G., Ed.; Geological Survey of Canada 72-1b: Ottawa, ON, Canada, 1972; pp. 62–64. [Google Scholar]
  2. Bodnar, R.J. Fluid inclusion evidence for a magmatic source for metals in porphyry copper deposits. Magmas Fluids Ore Depos. Mineral. Assoc. Can. Short. Course 1995, 23, 139–152. [Google Scholar]
  3. Sinclair, W.D. Porphyry deposits. In Mineral Deposits of Canda: A Synthesis of Major Deposit-Types, District Metallogeny, The Evolution of Geological Provinces, and Exploration Methods; Special Publication 5; Geological Association of Canada, Mineral Deposits Division: St. John’s, NL, Canada, 2007. [Google Scholar]
  4. Sillitoe, R.H. Porphyry copper systems. Econ. Geol. 2010, 105, 3–41. [Google Scholar] [CrossRef]
  5. Chen, H.Y.; Wu, C. Metallogenesis and major challenges of porphyry copper systems above subduction zones. Sci. China Earth Sci. 2020, 63, 899–918. [Google Scholar] [CrossRef]
  6. Kooiman, G.J.A.; McLeod, M.J.; Sinclair, W.D. Porphyry tungsten-molybdenum orebodies, polymetallic veins and replacement bodies, and tin-bearing greisen zones in the Fire Tower Zone, Mount Pleasant, New Brunswick. Econ. Geol. 1986, 81, 1356–1373. [Google Scholar] [CrossRef]
  7. Weis, P.; Driesner, T.; Heinrich, C.A. Porphyry-copper ore shells form at stable pressure-temperature fronts within dynamic fluid plumes. Science 2012, 338, 1613–1616. [Google Scholar] [CrossRef]
  8. Audétat, A.; Pettke, T.; Heinrich, C.A.; Bodnar, R.J. The composition of magmatic-hydrothermal fluids in barren and mineralized intrusions. Econ. Geol. 2008, 103, 877–908. [Google Scholar] [CrossRef]
  9. Audétat, A. Source and evolution of molybdenum in the porphyry Mo (–Nb) deposit at Cave Peak, Texas. J. Petrol. 2010, 51, 1739–1760. [Google Scholar] [CrossRef]
  10. Klemm, L.M.; Pettke, T.; Heinrich, C.A.; Campos, E. Hydrothermal evolution of the El Teniente deposit, Chile: Porphyry Cu-Mo ore deposition from low-salinity magmatic fluids. Econ. Geol. 2007, 102, 1021–1045. [Google Scholar] [CrossRef]
  11. Klemm, L.M.; Pettke, T.; Heinrich, C.A. Fluid and source magma evolution of the Questa porphyry Mo deposit, New Mexico, USA. Miner. Depos. 2008, 43, 533. [Google Scholar] [CrossRef]
  12. Landtwing, M.R.; Furrer, C.; Redmond, P.B.; Pettke, T.; Guillong, M.; Heinrich, C.A. The Bingham Canyon porphyry Cu-Mo-Au deposit. III. Zoned copper-gold ore deposition by magmatic vapor expansion. Econ. Geol. 2010, 105, 91–118. [Google Scholar] [CrossRef]
  13. Rusk, B.G.; Reed, M.H.; Dilles, J.H.; Klemm, L.; Heinrich, C.A. Compositions of magmatic hydrothermal fluids determined by LA-ICP-MS of fluid inclusions from the porphyry copper-molybdenum deposit at Butte. Montana. Chem. Geol. 2004, 210, 173–199. [Google Scholar] [CrossRef]
  14. Rusk, B.G.; Reed, M.H.; Dilles, J.H. Fluid inclusion evidence for magmatic-hydrothermal fluid evolution in the porphyry copper-molybdenum deposit at Butte, Montana. Econ. Geol. 2008, 103, 307–334. [Google Scholar] [CrossRef]
  15. Ulrich, T.; Mavrogenes, J.A. An experimental study of the solubility of molybdenum in H2O and KCl–H2O solutions from 500 °C to 800 °C, and 150 to 300 MPa. Geochim. Cosmochimca Acta 2008, 72, 2316–2330. [Google Scholar] [CrossRef]
  16. Zhang, L.; Audétat, A.; Dolejs, D. Solubility of molybdenite (MoS2) in aqueous fluids at 600–800 °C, 200 MPa: A synthetic fluid inclusion study. Geochim. Cosmochimca Acta 2012, 77, 175–185. [Google Scholar] [CrossRef]
  17. Mao, J.W.; Pirajno, F.; Xiang, J.F.; Gao, J.J.; Ye, H.S.; Li, Y.F.; Guo, B.J. Mesozoic molybdenum deposits in the east Qinling-Dabie orogenic belt: Characteristics and tectonic settings. Ore Geol. Rev. 2011, 43, 264–293. [Google Scholar] [CrossRef]
  18. Zeng, Q.D.; Liu, J.M.; Qin, K.Z.; Fan, H.R.; Chu, S.X.; Wang, Y.B.; Zhou, L.L. Types, characteristics, and time–space distribution of molybdenum deposits in China. Int. Geol. Rev. 2013, 55, 1311–1358. [Google Scholar] [CrossRef]
  19. Zeng, Q.D.; Liu, J.M.; Yu, C.M.; Ye, J.; Liu, H.T. Metal deposits in the Da Hinggan Mountains, NE China: Styles, characteristics, and exploration potential. Int. Geol. Rev. 2011, 53, 846–878. [Google Scholar] [CrossRef]
  20. Zeng, Q.D.; Liu, J.M.; Chu, S.X.; Wang, Y.B.; Sun, Y.; Duan, X.X.; Zhou, L.L. Mesozoic molybdenum deposits in the East Xingmeng orogenic belt, northeast China: Characteristics and tectonic setting. Int. Geol. Rev. 2012, 54, 1843–1869. [Google Scholar] [CrossRef]
  21. Zeng, Q.D.; Qin, K.Z.; Liu, J.M.; Li, G.M.; Zhai, M.G.; Chu, S.X.; Guo, Y.P. Porphyry molybdenum deposits in the Tianshan–Xingmeng orogenic belt, northern China. Int. J. Earth Sci. 2015, 104, 991–1023. [Google Scholar] [CrossRef]
  22. Ouyang, H.G.; Mao, J.W.; Santosh, M.; Zhou, J.; Zhou, Z.H.; Wu, Y.; Hou, L. Geodynamic setting of Mesozoic magmatism in NE China and surrounding regions: Perspectives from spatio-temporal distribution patterns of ore deposits. J. Asian Earth Sci. 2013, 78, 222–236. [Google Scholar] [CrossRef]
  23. Ouyang, H.G.; Mao, J.W.; Zhou, Z.H.; Su, H.M. Late Mesozoic metallogeny and intracontinental magmatism, southern Great Xing’an Range, northeastern China. Gondwana Res. 2015, 27, 1153–1172. [Google Scholar] [CrossRef]
  24. Wang, R.L.; Zeng, Q.D.; Zhang, Z.C.; Zhou, L.L.; Qin, K.Z. Extensive mineralization in the eastern segment of the Xingmeng orogenic belt, NE China: A regional view. Ore Geol. Rev. 2021, 135, 104204. [Google Scholar] [CrossRef]
  25. Xie, W.; Zeng, Q.D.; Wang, R.L.; Wu, J.J.; Zhang, Z.M.; Li, F.C.; Zhang, Z. Spatial-temporal distribution and tectonic setting of Mesozoic W-mineralized granitoids in the Xing-Meng Orogenic Belt, NE China. Int. Geol. Rev. 2022, 64, 1845–1884. [Google Scholar] [CrossRef]
  26. Shu, Q.H.; Chang, Z.S.; Lai, Y.; Zhou, Y.T.; Sun, Y.; Yan, C. Regional metallogeny of Mo-bearing Deposits in Northeastern China, with new Re-Os Dates of porphyry Mo Deposits in the Northern Xilamulun District. Econ. Geol. 2016, 111, 1783–1798. [Google Scholar] [CrossRef]
  27. Feng, Z.Q.; Liu, Y.J.; Li, L.; Jin, W.; Jiang, L.W.; Li, W.M.; Wen, Q.B.; Zhao, Y.L. Geochemical and geochronological constraints on the tectonic setting of the Xinlin ophiolite, northern Great Xing’an Range, NE China. Lithos 2019, 326−327, 213–229. [Google Scholar] [CrossRef]
  28. Liu, Y.J.; Li, W.M.; Feng, Z.Q.; Wen, Q.B.; Neubauer, F.; Liang, C.Y. A review of the Paleozoic tectonics in the eastern part of Central Asian Orogenic Belt. Gondwana Res. 2017, 43, 123–148. [Google Scholar] [CrossRef]
  29. 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]
  30. Zhang, G.L. Metallogenic Regularity and Prospect Estimation of Urgen Au-Ag-Pb-Zn-Mo Polymetallic Ore Field in Erguna, Inner Mongolia. Ph.D. Thesis, Northeastern University, Shenyang, China, 2022. (In Chinese with English Abstract). [Google Scholar]
  31. Zhang, G.L.; Wen, S.Q.; Li, J.Y.; Xie, W. Petrogenesis and tectonic implications of alkali-rich porphyry in the Urgen ore field, Erguna Block, Inner Mongolia: Constraints from geochronology and geochemistry. Geotecton. Metallog. 2023, 47, 428–448, (In Chinese with English Abstract). [Google Scholar]
  32. Sengör, A.M.C.; Natal’in, B.A.; Burtman, V.S. Evolution of the Altaid tectonic collage and Paleozoic crustal growth in Eurasia. Nature 1993, 364, 299–307. [Google Scholar] [CrossRef]
  33. Eizenhöfer, P.R.; Zhao, G.C.; Zhang, J.; Sun, M. Final closure of the Paleo-Asian Ocean along the Solonker suture zone: Constraints from geochronological and geochemical data of Permian volcanic and sedimentary rocks. Tectonics 2014, 33, 441–463. [Google Scholar] [CrossRef]
  34. Xiao, W.J.; Windley, B.F.; Hao, J.; Zhai, M.G. Accretion leading to collision and the Permian Solonker suture, Inner Mongolia, China: Termination of the central Asian orogenic belt. Tectonics 2003, 22, 288–308. [Google Scholar] [CrossRef]
  35. 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]
  36. Xu, W.L.; Pei, F.P.; Wang, F.; Meng, E.; Ji, W.Q.; Yang, D.B.; Wang, W. Spatial-temporal relationships of Mesozoic volcanic rocks in NE China: Constraints on tectonic overprinting and transformations between multiple tectonic systems. J. Asian Earth Sci. 2013, 74, 167–193. [Google Scholar] [CrossRef]
  37. Wilde, S.A. Final amalgamation of the Central Asian Orogenic Belt in NE China: Paleo-Asian Ocean closure versus Paleo-Pacific plate subduction–A review of the evidence. Tectonophysics 2015, 662, 345–362. [Google Scholar] [CrossRef]
  38. Chen, P.W.; Zeng, Q.D.; Wang, Y.B.; Zhou, T.C.; Yu, B.; Chen, J.Q. Petrogenesis of the Dasuji porphyry Mo deposit at the northern margin of North China Craton: Constrains from geochronology, geochemistry and isotopes characteristics. Lithos 2018, 322, 87–103. [Google Scholar] [CrossRef]
  39. Chen, P.W.; Zeng, Q.D.; Zhou, T.C. Petrogenesis and Mo prospecting significance of Sadaigoumen granites on the northern margin of the North China Craton. J. Geochem. Explor. 2020, 214, 106536. [Google Scholar] [CrossRef]
  40. Tang, W.H.; Li, J.; Wang, K.Y.; Li, S.D.; Cai, W.Y.; Liu, H.L.; Wang, Y.C. Ore genesis and tectonic setting of the Laojiagou porphyry molybdenum deposit, Inner Mongolia, China: Evidence from geology, fluid inclusions, H–O isotopes, zircon U–Pb geochronology, and geochemistry. Arab. J. Geosci. 2019, 12, 731. [Google Scholar] [CrossRef]
  41. Duan, X.X.; Zeng, Q.D.; Yang, Y.H.; Liu, J.M.; Chu, S.X.; Sun, Y.; Zhang, Z.L. Triassic magmatism and Mo mineralization in Northeast China: Geochronological and isotopic constraints from the Laojiagou porphyry Mo deposit. Int. Geol. Rev. 2015, 57, 55–75. [Google Scholar] [CrossRef]
  42. Sun, Q.F.; Wang, K.Y.; Sun, F.Y.; Lai, C.K.; Zhang, M.; Zhao, C.G.; Sun, L.X. Superimposing porphyry Mo and vein-type Cu-Pb-Zn mineralization in the Panjiaduan deposit, Great Xing’an Range (NE China): Perspective from zircon U-Pb and sphalerite Rb-Sr dating, geochemistry and S-Pb isotopes. Ore Geol. Rev. 2021, 139, 104538. [Google Scholar] [CrossRef]
  43. Zeng, Q.D.; Yang, J.H.; Liu, J.M.; Chu, S.X.; Duan, X.X.; Zhang, Z.L.; Zhang, W.Q.; Zhang, S. Genesis of the Chehugou Mo-bearing granitic complex on the northern margin of the North China Craton: Geochemistry, zircon U-Pb age and Sr-Nd-Pb isotopes. Geol. Mag. 2011, 149, 753–767. [Google Scholar] [CrossRef]
  44. Chen, Z.G.; Zhang, L.C.; Lu, B.Z.; Li, Z.L.; Wu, H.Y.; Xiang, P.; Huang, S.W. Geochronology and geochemistry of the Taipingchuan copper-molybdenum deposit in Inner Mongolia, and its geological significances. Acta Petrol. Sin. 2010, 26, 1437–1449, (In Chinese with English Abstract). [Google Scholar] [CrossRef]
  45. Kang, Y.J.; She, H.Q.; Lai, Y.; Wang, Z.Q.; Li, J.W.; Zhang, Z.H.; Xiang, A.P.; Jiang, Z.S. Evolution of Middle-Late Triassic granitic intrusions from the Badaguan Cu-Mo deposit, Inner Mongolia: Constraints from zircon U-Pb dating, geochemistry and Hf isotopes. Ore Geol. Rev. 2018, 95, 195–215. [Google Scholar] [CrossRef]
  46. Kang, Y.J.; Wang, Y.J.; Huang, G.J.; She, H.Q.; Xiang, A.P.; Tian, J.; Guo, Z.J.; Dong, X.Z. Study of rock-forming and ore-forming ages of Badaguan porphyry Cu-Mo deposit in Inner Mongolia. Miner. Depos. 2014, 33, 795–806, (In Chinese with English Abstract). [Google Scholar]
  47. Mi, K.F.; Liu, Z.J.; Li, C.F.; Liu, R.B.; Wang, J.P.; Peng, R.M. Origin of the Badaguan porphyry Cu Mo deposit, Inner Mongolia, northeast China: Constraints from geology, isotope geochemistry and geochronology. Ore Geol. Rev. 2017, 81, 154–172. [Google Scholar] [CrossRef]
  48. Gao, B.Y.; Zhang, L.C.; Jin, X.D.; Li, W.J.; Chen, Z.G.; Zhu, M.T. Geochronology and geochemistry of the Badaguan porphyry Cu–Mo deposit in Derbugan metallogenic belt of the NE China, and their geological significances. Int. J. Earth Sci. 2015, 105, 507–519. [Google Scholar] [CrossRef]
  49. Wang, Y.H.; Zhao, C.B.; Zhang, F.F.; Liu, J.J.; Wang, J.P.; Peng, R.M.; Liu, B. SIMS zircon U–Pb and molybdenite Re–Os geochronology, Hf isotope, and whole-rock geochemistry of the Wunugetushan porphyry Cu–Mo deposit and granitoids in NE China and their geological significance. Gondwana Res. 2015, 28, 1228–1245. [Google Scholar] [CrossRef]
  50. Mi, K.F.; Liu, Z.J.; Liu, R.B.; Li, C.F.; Wang, J.P.; Peng, R.M. U–Pb zircon, geochemical and Sr–Nd–Hf isotopic constraints on age and origin of the intrusions from Wunugetushan porphyry deposit, Northeast China: Implication for Triassic–Jurassic Cu–Mo mineralization in Mongolia–Erguna metallogenic belt. Int. Geol. Rev. 2018, 60, 496–512. [Google Scholar] [CrossRef]
  51. Zhang, X.J.; Zhang, Y.Y.; Liu, K.; He, X.L.; Wang, S.X.; Jia, W.H.; Zhao, Z.N. Zircon U-Pb and Lu-Hf isotopic dating of magmatic rocks in the Wunugetushan porphyry copper-molybdenum deposit, Inner Mongolia. Rock. Miner. Anal. 2022, 41, 774–788, (In Chinese with English Abstract). [Google Scholar]
  52. Zheng, W.; Yu, X.F. Geochronological and Geochemical Constraints on the Petrogenesis and Geodynamic Setting of the Daheishan Porphyry Mo Deposit, Northeast China. Resour. Geol. 2018, 68, 1–21. [Google Scholar] [CrossRef]
  53. Zhou, L.L.; Zeng, Q.D.; Liu, J.M.; Friis, H.; Zhang, Z.L.; Duan, X.X. Geochronology of the Xingshan molybdenum deposit, Jilin Province, NE China, and its Hf isotope significance. J. Asian Earth Sci. 2013, 75, 58–70. [Google Scholar] [CrossRef]
  54. Wang, Z.G.; Wang, K.Y.; Wan, D.; Konare, Y.; Yang, T.N.; Liang, Y.H. Metallogenic age and hydrothermal evolution of the Jidetun Mo deposit in central Jilin Province, northeast China: Evidence from fluid inclusions, isotope systematics, and geochronology. Ore Geol. Rev. 2017, 89, 731–751. [Google Scholar] [CrossRef]
  55. Zhang, Y.; Sun, J.G.; Xing, S.W.; Zhao, K.Q.; Ma, Y.B.; Zhang, Z.J.; Wang, Y. Ore-forming granites from Jurassic porphyry Mo deposits, east–central Jilin Province, China: Geochemistry, geochronology, and petrogenesis. Int. Geol. Rev. 2016, 58, 1158–1174. [Google Scholar] [CrossRef]
  56. Hou, X.G.; Sun, D.Y.; Gou, J.; Yang, D.G. The origin of variable-δ18O zircons in Jurassic and Cretaceous Mo-bearing granitoids in the eastern Xing–Meng Orogenic Belt, Northeast China. Int. Geol. Rev. 2019, 61, 129–149. [Google Scholar] [CrossRef]
  57. Yu, X.F.; Hou, Z.Q.; Qian, Y.; Li, B.L. Ore–forming fluids, stable isotopes and metallogenic epoch of the Fu’anpu molybdenum deposit in Mid–East Jilin Province. Geol. Explor. 2012, 48, 1151–1162, (In Chinese with English Abstract). [Google Scholar]
  58. Zhang, Y.; Sun, J.G.; Chen, Y.J.; Zhao, K.Q.; Gu, A.L. Re–Os and U–Pb geochronology of porphyry Mo deposits in central Jilin Province: Mo ore-forming stages in northeast China. Int. Geol. Rev. 2013, 55, 1763–1785. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Xing, S.W.; Sun, J.G.; Zhao, K.Q.; Wang, Y. Re-Os and U-Pb ages of the Sifangdianzi molybdenite deposit in central Jilin Province, NE China. Acta Geol. Sin. Engl. Ed. 2014, 88, 1045–1046. [Google Scholar] [CrossRef] [PubMed]
  60. Yang, Y.C.; Han, S.J.; Sun, D.Y.; Guo, J.; Zhang, S.J. Geological and geochemical features and geochronology of porphyry molybdenum deposits in the Lesser Xing’an Range-Zhangguangcai Range metallogenic belt. Acta Petrol. Sin. 2012, 28, 379–390, (In Chinese with English Abstract). [Google Scholar]
  61. Hu, X.L.; Ding, Z.J.; He, M.C.; Yao, S.Z.; Zhu, B.P.; Shen, J.; Chen, B. A porphyry-skarn metallogenic system in the Lesser Xing’an Range, NE China: Implications from U–Pb and Re–Os geochronology and Sr–Nd–Hf isotopes of the Luming Mo and Xulaojiugou Pb–Zn deposits. J. Asian Earth Sci. 2014, 90, 88–100. [Google Scholar] [CrossRef]
  62. Cheng, G.H.; Wang, R.L.; Zeng, Q.D.; Guo, Y.P.; Duan, X.X.; Wei, J.J.; Zhang, J.S.; Gao, X.H. Zircon U-Pb ages, Hf isotopes of the granitoids and Re-Os ages of the molybdenites in Luming molybdenum ore area, Heilongjiang Province, and its geological significance. Acta Petrol. Sin. 2015, 31, 2450–2464, (In Chinese with English Abstract). [Google Scholar]
  63. Sun, J.G.; Zhang, Y.; Xing, S.W.; Zhao, K.Q.; Zhang, Z.J.; Bai, L.A.; Ma, Y.B.; Liu, Y.S. Genetic types, ore-forming age and geodynamic setting of endogenic molybdenum deposits in the eastern edge of Xing-Meng orogenic belt. Acta Petrol. Sin. 2012, 28, 1317–1332, (In Chinese with English Abstract). [Google Scholar]
  64. Hu, X.L.; Yao, S.Z.; Zeng, G.P.; Liu, W.H.; Zhang, Z.J. Multistage magmatism resulting in large-scale mineralizaion: A case from the Huojihe porphyry Mo deposit in NE China. Lithos 2019, 326-327, 397–414. [Google Scholar] [CrossRef]
  65. Zhang, C.; Li, N. Geochronology and zircon Hf isotope geochemistry of granites in the giant Chalukou Mo deposit, NE China: Implications for tectonic setting. Ore Geol. Rev. 2017, 81, 780–793. [Google Scholar] [CrossRef]
  66. Liu, Y.F.; Bagas, L.; Jiang, S.H.; Wang, F.X. The Chalukou deposit in the North Great Xing’an Range of China: A protracted porphyry Mo ore-forming system in a long-lived magmatic evolution cycle. Ore Geol. Rev. 2017, 89, 171–186. [Google Scholar] [CrossRef]
  67. Liu, J.; Mao, J.W.; Wu, G.; Wang, F.; Luo, D.F.; Hu, Y.Q. Zircon U-Pb and molybdenite Re-Os dating of the Chalukou porphyry Mo deposit in the northern Great Xing’an Range, China and its geological significance. J. Asian Earth Sci. 2014, 79, 696–709. [Google Scholar] [CrossRef]
  68. Li, Z.Z.; Qin, K.Z.; Li, G.M.; Ishihara, S.; Jin, L.Y.; Song, G.X.; Meng, Z.J. Formation of the giant Chalukou porphyry Mo deposit in northern Great Xing’an Range, NE China: Partial melting of the juvenile lower crust in intra-plate extensional environment. Lithos 2014, 202−203, 138–156. [Google Scholar] [CrossRef]
  69. Hu, X.L.; Yao, S.Z.; He, M.C.; Ding, Z.J.; Cui, Y.B.; Shen, J.; Chen, B.; Zhu, B.P. Geochemistry, U-Pb Geochronology and Hf Isotope Studies of the Daheishan Porphyry Mo Deposit in Heilongjiang Province, NE China. Resour. Geol. 2014, 64, 102–116. [Google Scholar] [CrossRef]
  70. Deng, C.Z.; Sun, D.Y.; Han, J.S.; Chen, H.Y.; Li, G.H.; Xiao, B.; Li, R.C.; Feng, Y.Z.; Li, C.L.; Lu, S. Late-stage southwards subduction of the Mongol-Okhotsk oceanic slab and implications for porphyry Cu Mo mineralization: Constraints from igneous rocks associated with the Fukeshan deposit, NE China. Lithos 2019, 326–327, 341–357. [Google Scholar] [CrossRef]
  71. Deng, C.Z.; Li, G.H. The Cu-Mo mineralization of the Late Jurassic Porphyry in the northern Great Xing’an Range: Constraints from zircon U-Pb ages of the ore-causative granites. Acta Geol. Sin. (Engl. Ed.) 2019, 93, 236–237. [Google Scholar] [CrossRef]
  72. Wang, S.W.; Wang, J.G.; Zhang, D.; Qi, X.J.; Wu, G.G.; Zhao, P.Z.; Yang, Z.F.; Liu, Y.B. Geochronological study on Taipinggou molybdenum deposit in Da Hinggan Mountain. Acta Petrol. Sin. 2009, 25, 2913–2923, (In Chinese with English Abstract). [Google Scholar]
  73. Wang, Z.L.; Yang, J.Z.; Zhao, K.Q.; Fu, Y.G.; Duan, Z.; Lin, L.J.; Xi, M.J.; Deng, Y.M.; Lu, G.F. Zircon U-Pb age, geochemical characteristics and geological significance of the granite of Jiazishan Mo(Ag) deposit, Inner Mongolia. Earth Sci. 2021. (In Chinese with English Abstract). [Google Scholar]
  74. Sun, H.R.; Huang, Z.L.; Li, W.B.; Leng, C.B.; Ma, D.Y.; Zhang, X.C. Chronology, geochemistry and Sr–Nd isotope studies of Jurassic intrusions in the Diyanqinamu porphyry Mo mine, central Inner Mongolia, China. J. Asian Earth Sci. 2014, 88, 85–97. [Google Scholar] [CrossRef]
  75. Zhai, D.G.; Liu, J.J.; Wang, J.P.; Yang, Y.Q.; Zhang, H.Y.; Wang, X.L.; Zhang, Q.B.; Wang, G.W.; Liu, Z.J. Zircon U–Pb and molybdenite Re–Os geochronology, and whole-rock geochemistry of the Hashitu molybdenum deposit and host granitoids, Inner Mongolia, NE China. J. Asian Earth Sci. 2014, 79, 144–160. [Google Scholar] [CrossRef]
  76. Zeng, Q.D.; Guo, W.K.; Chu, S.X.; Duan, X.X. Late Jurassic granitoids in the Xilamulun Mo belt, Northeastern China: Geochronology, geochemistry, and tectonic implications. Int. Geol. Rev. 2016, 58, 588–602. [Google Scholar] [CrossRef]
  77. Shang, Z.; Zhou, D.; Chen, Y.Q. Genesis, metallogenetic and tectonic significance of the A-type granites in Hashitu Mo deposit, southern Great Hinggan Range, NE China. Ore Geol. Rev. 2021, 138, 104388. [Google Scholar] [CrossRef]
  78. Wu, H.Y.; Zhang, L.C.; Wan, B.; Chen, Z.G.; Zhang, X.J.; Xiang, P. Geochronological and geochemical constraints on Aolunhua porphyry Mo–Cu deposit, northeast China, and its tectonic significance. Ore Geol. Rev. 2011, 43, 78–91. [Google Scholar] [CrossRef]
  79. Ma, X.H.; Chen, B.; Lai, Y.; Lu, Y.H. Petrogenesis and mineralization chronology study on the Aolunhua porphyry Mo deposit, Inner Mongolia, and its geological implications. Acta Petrol. Sin. 2009, 25, 2939–2950, (In Chinese with English Abstract). [Google Scholar]
  80. Shu, Q.H.; Lai, Y.; Zhou, Y.T.; Xu, J.J.; Wu, H.Y. Zircon U–Pb geochronology and Sr–Nd–Pb–Hf isotopic constraints on the timing and origin of Mesozoic granitoids hosting the Mo deposits in northern Xilamulun district, NE China. Lithos 2015, 238, 64–75. [Google Scholar] [CrossRef]
  81. Zeng, Q.D.; Liu, J.M. Zircon SHRIMP U-Pb dating and geological significance of the granite porphyry from Banlashan porphyry molybdenum deposit in Xilamulun molybdenum metallogenic belt. J. Jilin Univ. Earth Sci. Ed. 2010, 40, 828–834, (In Chinese with English Abstract). [Google Scholar]
  82. Zeng, Q.D.; Yang, J.H.; Zhang, Z.L.; Liu, J.M.; Duan, X.X. Petrogenesis of the Yangchang Mo-bearing granite in the Xilamulun metallogenic belt, NE China: Geochemistry, zircon U-Pb ages and Sr-Nd-Pb isotopes. Geol. J. 2014, 49, 1–14. [Google Scholar] [CrossRef]
  83. Ji, H.W.; Niu, Z.L.; Dong, Q.; Zhang, C. Zircon U-Pb and molybdenite Re-Os dating for Hongshanzi uranium-molybdenum deposit in the Inner Mongolia and its geological significance. Uranium Geol. 2021, 37, 810–822, (In Chinese with English Abstract). [Google Scholar]
  84. Zeng, Q.D.; Liu, J.M.; Qin, F.; Zhang, Z.L. Geochronology of the Xiaodonggou Porphyry Mo Deposit in Northern Margin of North China Craton. Resour. Geol. 2010, 60, 192–202. [Google Scholar] [CrossRef]
  85. Zeng, Q.D.; Liu, J.M.; Zhang, Z.L.; Chen, W.J.; Zhang, W.Q. Geology and geochronology of the Xilamulun molybdenum metallogenic belt in eastern Inner Mongolia, China. Int. J. Earth Sci. 2011, 100, 1791–1809. [Google Scholar] [CrossRef]
  86. Li, Y.; Yang, Y.Q.; Hua, L.; Leng, Y.X.; Xin, J. Geochronology and Tectonic Implications of the Nianzigou Granites and Associated Mo Deposit, Inner Mongolia. Minerals 2022, 12, 791. [Google Scholar] [CrossRef]
  87. Wu, H.Y.; Zhang, L.C.; Pirajno, F.; Xiang, P.; Wan, B.; Chen, Z.G.; Zhang, X.J. The Jiguanshan porphyry Mo deposit in the Xilamulun metallogenic belt, northern margin of the North China Craton, U–Pb geochronology, isotope systematics, geochemistry and fluid inclusion studies: Implications for a genetic model. Ore Geol. Rev. 2014, 56, 549–565. [Google Scholar] [CrossRef]
  88. Ge, W.C.; Wu, F.Y.; Zhou, C.Y.; Zhang, J.H. The ore-forming time of porphyry Cu and Mo deposits in eastern segment of the Xing-Meng orogen and its geodynamic significance. Chinese Sci. Bull. 2007, 52, 2407–2417. (In Chinese) [Google Scholar] [CrossRef]
  89. Ge, W.C.; Wu, F.Y.; Zhou, C.Y.; Abdel Rahman, A.A. Emplacement age of the Tahe granite and its constraints on the tectonic nature of the Ergun block in the northern part of the Da Hinggan Range. Chinese Sci. Bull. 2005, 50, 2097–2105. [Google Scholar] [CrossRef]
  90. Zhou, J.B.; Wang, B.; Wilde, S.A.; Zhao, G.C.; Cao, J.L.; Zheng, C.Q.; Zeng, W.S. Geochemistry and U–Pb zircon dating of the Toudaoqiao blueschists in the Great Xing’ an Range, northeast China, and tectonic implications. J. Asian Earth Sci. 2015, 97, 197–210. [Google Scholar] [CrossRef]
  91. Feng, Z.Q.; Liu, Y.J.; Jin, W.; Li, W.M.; Wen, Q.B. Is there the Paleoproterozoic komatiite related to mantle plume in the Jifeng area, Northern Great Xing’an Range, NE China? Int. J. Earth Sci. 2017, 106, 2225–2231. [Google Scholar] [CrossRef]
  92. Miao, L.C.; Zhang, F.; Jiao, S.J. Age, protoliths and tectonic implications of the Toudaoqiao blueschist, Inner Mongolia, China. J. Asian Earth Sci. 2015, 105, 360–373. [Google Scholar] [CrossRef]
  93. IMBGMR. Regional Geology of Inner Mongolia; Geological Publishing House: Beijing, China, 1991. (In Chinese) [Google Scholar]
  94. Zhao, S.; Xu, W.L.; Tang, J.; Li, Y.; Guo, P. Timing of formation and tectonic nature of the purportedly Neoproterozoic Jiageda formation of the Erguna Massif, NE China: Constraints from field geology and U-Pb geochronology of detrital and magmatic zircons. Precambrian Res. 2016, 281, 585–601. [Google Scholar] [CrossRef]
  95. Li, S.C.; Liu, Z.H.; Xu, Z.Y.; Li, G.; Zhang, C. Age and tectonic setting of volcanic rocks of the Tamulangou Formation in the Great Xing’an Range, NE China. J. Asian Earth Sci. 2015, 113, 471–480. [Google Scholar] [CrossRef]
  96. Ji, Z.; Ge, W.C.; Wang, Q.H.; Yang, H.; Zhao, G.C.; Bi, J.H.; Dong, Y. Petrogenesis of early Cretaceous volcanic rocks of the Manketouebo formation in the Wuchagou region, central great Xing’an range, NE China, and tectonic implications: Geochronological, geochemical, and Hf isotopic evidence. Int. Geol. Rev. 2016, 58, 556–573. [Google Scholar] [CrossRef]
  97. Badarch, G.; Cunningham, W.D.; Windley, B.F. A new terrane subdivision for Mongolia: Implications for the Phanerozoic crustal growth of Central Asia. J. Asian Earth Sci. 2002, 21, 87–110. [Google Scholar] [CrossRef]
  98. Li, S.; Wang, T.; Wilde, S.A.; Tong, Y. Evolution, source and tectonic significance of early Mesozoic granitoid magmatism in the central Asian Orogenic Belt (central segment). Earth-Sci. Rev. 2013, 126, 206–234. [Google Scholar] [CrossRef]
  99. Gou, J.; Sun, D.Y.; Ren, Y.S.; Liu, Y.J.; Zhang, S.Y.; Fu, C.L.; Wang, T.H.; Wu, P.F.; Liu, X.M. Petrogenesis and geodynamic setting of Neoproterozoic and Late Paleozoic magmatism in the Manzhouli-Erguna area of Inner Mongolia, China: Geochronological, geochemical and Hf isotopic evidence. J. Asian Earth Sci. 2013, 67–68, 114–137. [Google Scholar] [CrossRef]
  100. Gou, J.; Sun, D.Y.; Yang, D.G.; Tang, Z.Y.; Mao, A.Q. Late Paleozoic igneous rocks of the Great Xing’an Range, NE China: The Tayuan example. Int. Geol. Rev. 2017, 21, 314–340. [Google Scholar]
  101. Tang, J.; Xu, W.L.; Wang, F.; Wang, W.; Xu, M.J.; Zhang, Y.H. Geochronology and geochemistry of Early-Middle Triassic magmatism in the Erguna Massif, NE China: Constraints on the tectonic evolution of the Mongol-Okhotsk suture belt. Lithos 2014, 184–187, 1–16. [Google Scholar] [CrossRef]
  102. Sun, D.Y.; Gou, J.; Wang, T.H.; Ren, Y.S.; Liu, Y.J.; Gou, H.Y.; Liu, X.M.; Hu, Z.C. Geochronological and geochemical constraints on the Erguna massif basement, NE China subduction history of the Mongol-Okhotsk oceanic crust. Int. Geol. Rev. 2013, 55, 1801–1816. [Google Scholar] [CrossRef]
  103. Wang, W.; Tang, J.; Xu, W.L.; Wang, F. Geochronology and geochemistry of Early Jurassic volcanic rocks in the Erguna Massif, northeast China: Petrogenesis and implications for the tectonic evolution of the Mongol–Okhotsk suture belt. Lithos 2015, 218−219, 73–86. [Google Scholar] [CrossRef]
  104. Tang, J.; Xu, W.L.; Wang, F.; Zhao, S.; Wang, W. Early Mesozoic southward subduction history of the Mongol-Okhotsk oceanic plate: Evidence from geochronology and geochemistry of Early Mesozoic intrusive rocks in the Erguna Massif, NE China. Gondwana Res. 2016, 31, 218–240. [Google Scholar] [CrossRef]
  105. Du, A.D.; Wu, S.Q.; Sun, D.Z.; Wang, S.X.; Qu, W.J.; Markey, R.; Stein, H.; Morgan, J.; Malinovoskiy, D. Preparation and certification of Re-Os dating reference materials, molybdenite HLP and JDC. Geostand. Geoanalytical Res. 2004, 28, 41–52. [Google Scholar] [CrossRef]
  106. Shirey, S.B.; Walker, R.J. Carius tube digestion for low-blank Rhenium-Osmium analysis. Anal. Chem. 1995, 67, 2136–2141. [Google Scholar] [CrossRef]
  107. Smoliar, M.I.; Walker, R.J.; Morgan, J.W. Re–Os ages of group IIA, IIIA, IVA, and IVB iron meteorites. Science 1996, 271, 1099–1102. [Google Scholar] [CrossRef]
  108. Wopenka, B.; Pasteris, J.D.; Freeman, J.J. Analysis of individual fluid inclusions by Fourier transform infrared and Raman microspectroscopy. Geochim. Cosmochimca Acta 1990, 54, 519–533. [Google Scholar] [CrossRef]
  109. Goldstein, R.H.; Reynolds, T.J. Systematics of fluid inclusions in diagenetic minerals. Soc. Sediment. Geol. 1994, 31, 199. [Google Scholar]
  110. Collins, P.L.F. Gas hydrates in CO2-bearing fluid inclusions and the use of freezing data for estimation of salinity. Econ. Geol. 1979, 74, 1435–1444. [Google Scholar] [CrossRef]
  111. Bodnar, R.J. Revised equation and table for determining the freezing point depression of H2O-NaCl solutions. Geochim. Cosmochimca Acta 1993, 57, 683–684. [Google Scholar] [CrossRef]
  112. Bodnar, R.J.; Vityk, M.O. Interpretation of microthermometric data for H2O-NaCl fluid inclusions. In Fluid Inclusion in Minerals: Methods and Applications; Springer: Berlin/Heidelberg, Germany, 1994; pp. 117–130. [Google Scholar]
  113. Brown, P.E.; Lamb, W.M. P-V-T properties of fluids in the system H2O ± CO2 ± NaCl: New graphic presentations and implications for fluid inclusion studies. Geochim. Cosmochimca Acta 1989, 53, 1209–1221. [Google Scholar] [CrossRef]
  114. Duan, Z.; Moller, N.; Weare, J.H. An Equation of State for the CH4-CO2-H2O System: I. Pure Systems from 0 to 1000 °C and 0 to 8000 Bar. Geochim. Cosmochimca Acta 1992, 56, 2605–2617. [Google Scholar] [CrossRef]
  115. Duan, Z.; Moller, N.; Weare, J.H. An Equation of State for the CH4-CO2-H2O System: II. Mixtures from 50 to 1000 °C and 0 to 1000 bars. Geochim. Cosmochimca Acta 1992, 56, 2619–2631. [Google Scholar] [CrossRef]
  116. 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]
  117. Robinson, B.W.; Kusakabe, M. Quantitative preparation of sulfur dioxide, for 34S/32S analyses, from sulfides by combustion with cuprous oxide. Anal. Chem. 1975, 47, 1179–1181. [Google Scholar] [CrossRef]
  118. Todt, W.; Cliff, R.A.; Hanser, A.; Hofmann, A.W. Re-calibration of NBS lead standards using a 202Pb-205Pb double spike. Terra Abstr. 1993, 5, 396. [Google Scholar]
  119. Roedder, E. Fluid inclusions. Geol. Soc. Am. Bull. 1984, 12, 1–644. [Google Scholar]
  120. Chi, G.X.; Lu, H.Z. Validation and representation of fluid inclusion microthermometric data using the fluid inclusion assemblage (FIA) concept. Acta Petrol. Sin. 2008, 24, 1945–1953, (In Chinese with English Abstract). [Google Scholar]
  121. Chi, G.X.; Diamond, L.W.; Lu, H.Z.; Lai, J.Q.; Chu, H.X. Common Problems and Pitfalls in Fluid Inclusion Study: A Review and Discussion. Minerals 2021, 11, 7. [Google Scholar] [CrossRef]
  122. Davis, D.W.; Lowenstein, T.K.; Spencer, R.J. Melting behavior of fluid inclusions in laboratory-grown halite crystals in the systems NaCl-H2O, NaCl-KCl-H2O, NaCl-MgCl2-H2O, and NaCl-CaCl2-H2O. Geochim. Cosmochimca Acta 1990, 54, 591–601. [Google Scholar] [CrossRef]
  123. Rosso, K.M.; Bodnar, R.J. Microthermometric and Raman spectroscopic detection limits of CO2 in fluid inclusions and the Raman spectroscopic characterization of CO2. Geochim. Cosmochimca Acta 1995, 59, 3961–3975. [Google Scholar] [CrossRef]
  124. Audétat, A. A Plea for More Skepticism toward Fluid Inclusions: Part II. Homogenization via Halite Dissolution in Brine Inclusions from Magmatic-Hydrothermal Systems Is Commonly the Result of Postentrapment Modifications. Econ. Geol. 2023, 118, 43–55. [Google Scholar] [CrossRef]
  125. Zhang, D.H.; Audétat, A. A Plea for More Skepticism toward Fluid Inclusions: Part I. Postentrapment Changes in Fluid Density and Fluid Salinity Are Very Common. Econ. Geol. 2023, 118, 15–41. [Google Scholar] [CrossRef]
  126. Audétat, A.; Günther, D. Mobility and H2O loss from fluid inclusions in natural quartz crystals. Contrib. Miner. Petr. 1999, 137, 1–14. [Google Scholar] [CrossRef]
  127. 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]
  128. Zhang, L.G. Application of Stable Isotopes in Geological Sciences; Shanxi Science and Technology Press: Xi’an, China, 1985. (In Chinese) [Google Scholar]
  129. Zartman, R.E.; Doe, B.R. Plumbotectonics: The model. Tectonophysics 1981, 75, 135–162. [Google Scholar] [CrossRef]
  130. Stein, H.J. Dating and tracing the history of ore formation. In Treatise in Geochemistry, 2nd ed.; Turekian, K.K., Holland, H.D., Scott, S.D., Eds.; Elsevier-Pergamon: Oxford, UK, 2014; Volume 12. [Google Scholar]
  131. Zhai, D.G.; Williams-Jones, A.E.; Liu, J.J.; Selby, D.; Li, C.; Huang, X.W.; Qi, L.; Guo, D.H. Evaluating the use of the molybdenite Re-Os chronometer in dating gold mineralization: Evidence from the Haigou deposit, northeastern China. Econ. Geol. 2019, 114, 897–915. [Google Scholar] [CrossRef]
  132. Stein, H.J.; Markey, R.J.; Morgan, J.W.; Hannah, J.L.; Scherstén, A. The remarkable Re-Os chronometer in molybdenite: How and why it works. Terra Nova 2001, 13, 479–486. [Google Scholar] [CrossRef]
  133. Stein, H.; Schersten, A.; Hannah, J.; Markey, R. Subgrain-scale decoupling of Re and 187Os and assessment of laser ablation ICP-MS spot dating in molybdenite. Geochim. Cosmochim. Acta 2003, 67, 3673–3686. [Google Scholar] [CrossRef]
  134. Selby, D.; Creaser, R.A. Macroscale NTIMS and microscale LA-MC-ICP-MS Re-Os isotopic analysis of molybdenite: Testing spatial restrictions for reliable Re-Os age determinations, and implications for the decoupling of Re and Os within molybdenite. Geochim. Cosmochim. Acta 2004, 68, 3897–3908. [Google Scholar] [CrossRef]
  135. Porter, S.J.; Selby, D. Rhenium-Osmium (Re-Os) molybdenite systematics and geochronology of the Cruachan granite skarn mineralization, Etive Complex: Implications for emplacement chronology. Scott. J. Geol. 2010, 46, 17–21. [Google Scholar] [CrossRef]
  136. Rathkopf, C.; Mazdab, F.; Barton, I.; Barton, M.D. Grain-scale and deposit-scale heterogeneity of Re distribution in molybdenite at the Bagdad porphyry Cu-Mo deposit, Arizona. J. Geochem. Explor. 2017, 178, 45–54. [Google Scholar] [CrossRef]
  137. Chen, W.T.; Zhou, M.F. Paragenesis, stable isotopes, and molybdenite Re-Os isotope age of the Lala iron-copper deposit, southwest China. Econ. Geol. 2012, 107, 459–480. [Google Scholar] [CrossRef]
  138. Lawley, C.J.M.; Selby, D. Re-Os geochronology of quartz-enclosed ultrafine molybdenite: Implications for ore geochronology. Econ. Geol. 2012, 107, 1499–1505. [Google Scholar] [CrossRef]
  139. Li, J.; Xu, L.L.; Bi, X.W.; Tang, Y.Y.; Sheng, X.Y.; Yu, H.J.; Liu, G.; Ma, R. New titanite U–Pb and molybdenite Re–Os ages for a hydrothermal vein-type Cu deposit in the Lanping Basin, Yunnan, SW China: Constraints on regional metallogeny and implications for exploration. Miner. Depos. 2021, 56, 441–456. [Google Scholar] [CrossRef]
  140. Girei, M.B.; Li, H.; Vincent, V.I.; Algeo, T.J.; Elatikpo, S.M.; Bute, S.I.; Ahmed, H.A.; Amuda, A.K. Genesis and timing of Mo mineralization in the Mada Ring Complex, north-central Nigeria: Insights from whole-rock geochemistry, Nd-Sr isotopes, zircon U-Pb-Hf isotopes, and molybdenite Re-Os systematics. Miner. Depos. 2022, 57, 601–620. [Google Scholar] [CrossRef]
  141. Qi, H.W.; Hu, R.Z.; Wang, X.F.; Qu, W.J.; Bi, X.W.; Peng, J.T. Molybdenite Re–Os and muscovite 40Ar/39Ar dating of quartz vein-type W–Sn polymetallic deposits in Northern Guangdong, South China. Miner. Depos. 2012, 47, 607–622. [Google Scholar] [CrossRef]
  142. Huang, X.D.; Huang, D.; Lu, J.J.; Zhang, R.Q.; Ma, D.S.; Jiang, Y.H.; Chen, H.W.; Liu, J.X. Neoproterozoic tungsten mineralization: Geology, chronology, and genesis of the Huashandong W deposit in northwestern Jiangxi, South China. Miner. Depos. 2023, 58, 771–796. [Google Scholar] [CrossRef]
  143. Wu, H.Y.; Zhang, L.C.; Wan, B.; Chen, Z.G.; Xiang, P.; Pirajno, F.; Du, A.D.; Qu, W.J. Re–Os and 40Ar/39Ar ages of the Jiguanshan porphyry Mo deposit, Xilamulun metallogenic belt, NE China, and constraints on mineralization events. Miner. Depos. 2010, 46, 171–185. [Google Scholar] [CrossRef]
  144. Zhang, P.; Zhao, Y.; Kou, L.L.; Yang, H.Z.; Yang, F.C. Zircon U–Pb and molybdenite Re–Os geochronology of copper–molybdenum deposits in southeast Liaoning Province, China. Int. Geol. Rev. 2016, 58, 1481–1491. [Google Scholar] [CrossRef]
  145. Buret, Y.; von Quadt, A.; Heinrich, C.; Selby, D.; Walle, M.; Peytcheva, I. From a long-lived upper-crustal magma chamber to rapid porphyry copper-emplacement: Reading the geochemistry of zircon crystals at Bajo de la Alumbrera (NW Argentina). Earth Planet. Sci. Lett. 2016, 450, 120–131. [Google Scholar] [CrossRef]
  146. Cernuschi, F.; Dilles, J.H.; Grocke, S.B.; Valley, J.W.; Kitajima, K.; Tepley, F.J. Rapid formation of porphyry copper deposits evidenced by diffusion of oxygen and titanium in quartz. Geology 2018, 46, 611–614. [Google Scholar] [CrossRef]
  147. Li, Y.; Li, X.H.; Selby, D.; Li, J.W. Pulsed magmatic fluid release for the formation of porphyry deposits: Tracing fluid evolution in absolute time from the Tibetan Qulong Cu-Mo deposit. Geology 2018, 46, 7–10. [Google Scholar] [CrossRef]
  148. Li, Y.; Allen, M.B.; Li, X.H. Millennial pulses of ore formation and an extra-high Tibetan Plateau. Geology 2022, 50, 665–669. [Google Scholar] [CrossRef]
  149. Li, Y.; Zhang, R.Q.; He, S.; Chiaradia, M.; Li, X.H. Pulsed exsolution of magmatic ore-forming fluids in tin-tungsten systems: A SIMS cassiterite oxygen isotope record. Miner. Depos. 2022, 57, 343–352. [Google Scholar] [CrossRef]
  150. Von Quadt, A.; Erni, M.; Martinek, K.; Moll, M.; Peytcheva, I.; Heinrich, C.A. Zircon crystallization and the lifetimes of ore-forming magmatic-hydrothermal systems. Geology 2011, 39, 731–734. [Google Scholar] [CrossRef]
  151. Chiaradia, M.; Schaltegger, U.; Spikings, R.; Wotzlaw, J.F.; Ovtcharova, M. How accurately can we date the duration of magmatic-hydrothermal events in porphyry systems?—An invited paper. Econ. Geol. 2013, 108, 565–584. [Google Scholar] [CrossRef]
  152. Cathles, L.M.; Erendi, A.H.J.; Barrie, T. How long can a hydrothermal system be sustained by a single intrusive event? Econ. Geol. 1997, 92, 766–771. [Google Scholar] [CrossRef]
  153. Zhang, L.C.; Wu, H.Y.; Wan, B.; Chen, Z.G. Ages and geodynamic settings of Xilamulun Mo–Cu metallogenic belt in the northern part of the North China Craton. Gondwana Res. 2009, 16, 243–254. [Google Scholar] [CrossRef]
  154. Li, C.F.; Liu, Z.J.; Mi, K.F.; Wang, J.P.; Liu, R.B.; Zou, J.Y.; Wang, Z.B. Metallogenic age and ore genesis of the Badaguan porphyry copper-molybdenum deposit in Inner Mongolia. Geol. China 2014, 41, 1253–1269, (In Chinese with English Abstract). [Google Scholar]
  155. Han, C.M.; Xiao, W.J.; Windley, B.F.; Zhao, G.C.; Su, B.X.; Ao, S.J.; Zhang, J.E.; Zhang, Z.Y.; Wan, B.; Cui, B.; et al. Re-Os Age of Molybdenite from the Daheishan Mo Deposit in the Eastern Central Asian Orogenic Belt, NE China. Resour. Geol. 2014, 64, 379–386. [Google Scholar] [CrossRef]
  156. Wang, C.H.; Song, Q.H.; Wang, D.H.; Li, L.X.; Yu, C.; Wang, Z.G.; Qu, W.J.; Du, A.D.; Ying, L.J. Re-Os isotopic dating of molybdenite from the Daheishan molybdenum deposit of Jilin Province and its geological significance. Rock. Miner. Anal. 2009, 28, 269–273, (In Chinese with English Abstract). [Google Scholar]
  157. Wang, Z.G.; Wang, K.Y.; Konare, Y.; Yang, T.N.; Liang, Y.H. Metallogenesis and hydrothermal evolution of Middle Jurassic porphyry Mo deposits in the southern Zhangguangcai Range, NE China: Evidence from fluid inclusions, H–O–S isotopes, and Re–Os geochronology. Int. Geol. Rev. 2017, 59, 1391–1412. [Google Scholar] [CrossRef]
  158. Li, L.X.; Song, Q.H.; Wang, D.H.; Wang, C.H.; Qu, W.J.; Wang, Z.G.; Bi, S.Y.; Yu, C. Re-Os isotopic dating of molybdenite from the Fu’anpu molybdenum deposit of Jilin Province and discussion on its metallogenesis. Rock. Miner. Anal. 2009, 28, 283–287, (In Chinese with English Abstract). [Google Scholar]
  159. Zhang, Y. Research on Characteristics of Geology, Geochemistry and Metallogenic Mechanism of the Jurassic Molybdenum Deposits in the Mid-East Area of Jilin. Ph.D. Thesis, Jilin University, Changchun, China, 2013. (In Chinese with English Abstract). [Google Scholar]
  160. Zhang, Y.; Sun, J.G.; Xing, S.W.; Zhao, K.Q.; Ma, Y.B. Geochronology and metallogenesis of porphyry Mo deposits in east-central Jilin province, China: Constraints from molybdenite Re–Os isotope systematics. Ore Geol. Rev. 2015, 71, 363–372. [Google Scholar] [CrossRef]
  161. Ju, N.; Ren, Y.S.; Wang, C.; Wang, H.; Zhao, H.L.; Qu, W.J. Ore genesis and molybdenite Re-Os dating of Dashihe molybdenum deposit in Dunhua, Jilin. Glob. Geol. 2012, 31, 68–76, (In Chinese with English Abstract). [Google Scholar]
  162. Wang, H.; Ren, Y.S.; Zhao, H.L.; Ju, N.; Qu, W.J. Re-Os dating of molybdenite from the Liushengdian molybdenum deposit in Antun area of Jilin Province and its geological significance. Acta Geosci. Sin. 2011, 32, 707–715, (In Chinese with English Abstract). [Google Scholar]
  163. Chen, Z.G.; Zhang, L.C.; Wan, B.; Wu, H.Y.; Cleven, N. Geochronology and geochemistry of the Wunugetushan porphyry Cu-Mo deposit in NE China, and their geological significance. Ore Geol. Rev. 2011, 43, 92–105. [Google Scholar] [CrossRef]
  164. Zhang, L.C.; Gao, B.Y.; Li, W.J.; Chen, Z.G.; Sakyi, P.A.; Jin, X.D. Early Mesozoic tectono-magmatic activity and mineralization in Northeast China: Evidence from Re–Os to U–Pb studies of the Taipingchuan porphyry Cu–Mo deposit in the Derbugan metallogenic belt. Int. Geol. Rev. 2014, 56, 1837–1851. [Google Scholar] [CrossRef]
  165. Zhai, D.G.; Liu, J.J.; Wang, J.P.; Peng, R.M.; Wang, S.G.; Li, Y.N.; Chang, Z.Y. Re-Os isotopic chronology of molybdenite from the Taipinggou porphyry-type molybdenum deposit in Inner Mongolia and its geological significance. Geoscience 2009, 23, 262–268, (In Chinese with English Abstract). [Google Scholar]
  166. Wang, Z.L.; Zhao, K.Q.; Yang, J.Z.; Fu, Y.G.; Xi, M.J.; Li, C.; Deng, Y.M.; Lu, G.F. The molybdenite Re-Os dating of the Jiazishan Mo (Ag) deposit, Inner Mongolia and its geological significance. Acta Geol. Sin. 2021, 95, 3768–3777, (In Chinese with English Abstract). [Google Scholar]
  167. Leng, C.B.; Zhang, X.C.; Huang, Z.L.; Huang, Q.Y.; Wang, S.X.; Ma, D.Y.; Luo, T.Y.; Li, C.; Li, W.B. Geology, Re-Os ages, sulfur and lead isotopes of the Diyanqinamu porphyry Mo deposit, Inner Mongolia, NE China. Econ. Geol. 2015, 110, 557–574. [Google Scholar] [CrossRef]
  168. Shu, Q.H.; Jiang, L.; Lai, Y.; Lu, Y.H. Geochronology and fluid inclusion study of the Aolunhua porphyry Cu-Mo deposit in Arhorqin area, Inner Mongolia. Acta Petrol. Sin. 2009, 25, 2601–2614, (In Chinese with English Abstract). [Google Scholar]
  169. Zeng, Q.D.; Liu, J.M.; Zhang, Z.L. Re-Os geochronology of porphyry molybdenum deposit in south segment of Da Hinggan Mountains, Northeast China. J. Earth Sci.-China 2010, 21, 392–401. [Google Scholar] [CrossRef]
  170. Nie, F.J.; Zhang, W.Y.; Du, A.D.; Jiang, S.H.; Liu, Y. Re-Os isotopic dating on molybdenite separates from the Xiaodonggou porphyry Mo deposit, Hexigten Qi, Inner Mongolia. Acta Geol. Sin. 2007, 81, 898–905, (In Chinese with English Abstract). [Google Scholar]
  171. Chen, W.J.; Liu, J.M.; Liu, H.T.; Sun, X.G.; Zhang, R.B.; Zhang, Z.L.; Qin, F. Geochronology and fluid inclusion study of the Jiguanshan porphyry Mo deposit, Inner Mongolia. Acta Petrol. Sin. 2010, 26, 1423–1436, (In Chinese with English Abstract). [Google Scholar]
  172. Hedenquist, J.W.; Lowenstern, J.B. The role of magmas in the formation of hydrothermal ore deposits. Nature 1994, 370, 519–527. [Google Scholar] [CrossRef]
  173. Zhang, L.G. Theory of Diagenesis and Metallogenesis and Ore Prospecting; Beijing University of Technology Press: Beijing, China, 1989. (In Chinese) [Google Scholar]
  174. Barnes, H.L. Geochemistry of Hydrothermal Ore Deposits; Wiley: New York, NY, USA, 1979; p. 781. [Google Scholar]
  175. Kodera, P.; Lexa, J.; Rankin, A.H.; Fallick, A.E. Epithermal gold veins in a caldera setting: Banska Hodrusa, Slovakia. Miner. Depos. 2005, 39, 921–943. [Google Scholar] [CrossRef]
  176. Rye, R.O.; Ohmoto, H. Sufur and carbon isotopes and ore genesis: A review. Econ. Geol. 1974, 69, 826–842. [Google Scholar] [CrossRef]
  177. Seal, R.R. Sulfur Isotope Geochemistry of Sulfide Minerals. Rev. Mineral. Geochem. 2006, 61, 633–677. [Google Scholar] [CrossRef]
  178. Zhu, D.C.; Zhao, Z.D.; Niu, Y.L.; Mo, X.X.; Chung, S.L.; Hou, Z.Q.; Wang, L.Q.; Wu, F.Y. The Lhasa Terrane: Record of a microcontinent and its histories of drift and growth. Earth Planet. Sci. Lett. 2011, 301, 241–255. [Google Scholar] [CrossRef]
  179. Ohmoto, H.; Rye, R.O. Isotope 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. [Google Scholar]
  180. Ohmoto, H. Stable isotope geochemistry of ore deposits. Rev. Mineral. Geochem. 1986, 16, 491–559. [Google Scholar]
  181. Pass, H.E.; Cooke, D.R.; Davidson, G.; Maas, R.; Dipple, G.; Rees, C.; Ferreira, L.; Taylor, C.; Deyell, C.L. Isotope geochemistry of the Northeast Zone, Mount Polley Alkalic Cu-Au-Ag Porphyry Deposit, British Columbia: A case for carbonate assimilation. Econ. Geol. 2014, 109, 859–890. [Google Scholar] [CrossRef]
  182. Guo, F.; Fan, W.M.; Gao, X.F.; Li, C.W.; Miao, L.C.; Zhao, L.; Li, H.X. Sr-Nd-Pb isotope mapping of Mesozoic igneous rocks in NE China: Constraints on tectonic framework and Phanerozoic crustal growth. Lithos 2010, 120, 563–578. [Google Scholar] [CrossRef]
  183. Doe, B.R.; Zartman, R.E. Plumbotectonics I, the Phanerozoic. In Geochemistry of Hydrothermal Ore Deposits, 2nd ed.; Barnes, H.L., Ed.; John Wiley & Sons Inc.: New York, NY, USA, 1979; pp. 22–70. [Google Scholar]
  184. Cooke, D.R.; Hollings, P.; Walsh, J.L. Giant porphyry deposits: Characteristics, distribution, and tectonic controls. Econ. Geol. 2005, 100, 801–818. [Google Scholar] [CrossRef]
  185. Xiong, Y.Q.; Shao, Y.J.; Mao, J.W.; Wu, S.C.; Zhou, H.D.; Zheng, M.H. The polymetallic magmatic-hydrothermal Xiangdong and Dalong systems in the W–Sn–Cu–Pb–Zn–Ag Dengfuxian orefield, SE China: Constraints from geology, fluid inclusions, H–O–S–Pb isotopes, and sphalerite Rb–Sr geochronology. Miner. Depos. 2019, 54, 1101–1124. [Google Scholar] [CrossRef]
  186. Cline, J.S.; Bodnar, R.J. Direct evolution of brine from crystallizing silicic melt at the Questa, New Mexico, Molybdenum deposit. Econ. Geol. 1994, 89, 1780–1802. [Google Scholar] [CrossRef]
  187. Ulrich, T.; Günther, D.; Heinrich, C.A. The evolution of a porphyry Cu-Au deposit, based on LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera. Argentina. Econ. Geol. 2002, 97, 1889–1920. [Google Scholar] [CrossRef]
  188. Goldfarb, R.J.; Ayuso, R.A.; Miller, M.L.; Ebert, S.; Marsh, E.E.; Petsel, S.A. The late Cretaceous Donlin Creek gold deposit, southwestern Alaska: Controls on epizonal ore formation. Econ. Geol. 2004, 99, 643–671. [Google Scholar] [CrossRef]
  189. Shepherd, T.J.; Rankin, A.H.; Alderton, D.H.M. A Practical Guide to Fluid Inclusion Studies; Chapman & Hall: London, UK, 1985; p. 239. [Google Scholar]
  190. Wei, W.F.; Hu, R.Z.; Bi, X.W.; Peng, J.T.; Su, W.C.; Song, S.Q.; Shi, S.H. Infrared microthermometric and stable isotopic study of fluid inclusions in wolframite at the Xihuashan tungsten deposit, Jiangxi province, China. Miner. Depos. 2012, 47, 589–605. [Google Scholar] [CrossRef]
  191. Kwak, T.A.P. Fluid inclusions in skarns (carbonate replacement deposits). J. Metamorph. Geol. 1986, 4, 363–384. [Google Scholar] [CrossRef]
  192. Lu, H.Z.; Fan, H.R.; Ni, P.; Ou, G.X.; Shen, K.; Zhang, W.H. Fluid Inclusions; Science Press: Beijing, China, 2004; p. 444. (In Chinese) [Google Scholar]
  193. Roedder, E.; Bodnar, R.J. Geologic pressure determinations from fluid inclusion studies. Annu. Rev. Earth Planet. Sci. 1980, 8, 263–301. [Google Scholar] [CrossRef]
  194. Chen, Y.J.; Wang, P.; Li, N.; Yang, Y.F.; Pirajno, F. The collision-type porphyry Mo deposits in Dabie Shan. China. Ore Geol. Rev. 2017, 81, 405–430. [Google Scholar] [CrossRef]
  195. Hagemann, S.; Lüders, 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]
  196. Lowenstern, J.B. Carbon dioxide in magmas and implications for hydrothermal systems. Miner. Depos. 2001, 36, 490–502. [Google Scholar] [CrossRef]
  197. Soloviev, S.G.; Kryazhev, S.G.; Dvurechenskaya, S.S. Geology, mineralization, and fluid inclusion characteristics of the Meliksu reduced tungsten skarn deposit, Alai Tien Shan, Kyrgyzstan: Insights into conditions of formation and regional links to gold mineralization. Ore Geol. Rev. 2019, 111, 103003. [Google Scholar] [CrossRef]
  198. Xue, L.W.; Wang, G.W.; Tang, L.; Cao, Y.; Du, J.G.; Du, Y.S.; Luo, L.P.; Cheng, H.T. Genesis and hydrothermal evolution of the Zhazigou skarn W (Mo) deposit, East Qinling, China: Constraints from fluid inclusions and H-O-S-Pb isotopes. Ore Geol. Rev. 2021, 138, 104374. [Google Scholar] [CrossRef]
  199. Pirajno, F. Hydrothermal Mineral Deposits; Springer: Dordrecht, The Netherlands, 2009; p. 1250. [Google Scholar]
  200. Putnis, A. Mineral replacement reactions. Rev. Mineral. Geochem. 2009, 70, 87–124. [Google Scholar] [CrossRef]
  201. Chen, P.W.; Zeng, Q.D.; Zhou, T.C.; Wang, Y.B.; Yu, B.; Chen, J.Q. Evolution of fluids in the Dasuji porphyry Mo deposit on the northern margin of the North China Craton: Constraints from Microthermometric and LA-ICP-MS analyses of fluid inclusions. Ore Geol. Rev. 2019, 104, 26–45. [Google Scholar] [CrossRef]
  202. Guillemette, N.; Williams-Jones, A.E. Genesis of the Sb-W-Au deposits at Ixtahuacan, Guatemala: Evidence from fluid inclusions and stable isotopes. Miner. Depos. 1993, 28, 167–180. [Google Scholar] [CrossRef]
  203. Chen, P.W.; Zeng, Q.D.; Zhou, L.L.; Zhou, T.C. Fluid inclusion evidence for hydrothermal evolution of the Sadaigoumen porphyry Mo deposit on the northern margin of the north China Craton. Ore Geol. Rev. 2021, 134, 104145. [Google Scholar] [CrossRef]
  204. Zhao, X.Y.; Zhong, H.; Hu, R.Z.; Mao, W.; Bai, Z.J.; Lan, T.G.; Xue, K. Evolution of Multistage Hydrothermal Fluids in the Luoboling Porphyry Cu-Mo Deposit, Zijinshan Ore Field, Fujian Province, China: Insights from LA-ICP-MS Analyses of Fluid Inclusions. Econ. Geol. 2021, 116, 581–606. [Google Scholar] [CrossRef]
  205. Xie, W.; Zeng, Q.D.; Huang, L.L.; Zhou, L.L.; Fan, H.R.; Wu, J.J.; Wang, R.L.; Zhu, H.P. Composition and evolution of ore-forming fluids in the Sansheng porphyry W-Mo deposit, Inner Mongolia, NE China: Evidence from LA-ICP-MS analysis of fluid inclusions. Ore Geol. Rev. 2023, 158, 105481. [Google Scholar] [CrossRef]
  206. Hu, D.L.; Jiang, S.Y.; Xiong, S.F.; Dong, J.X.; Wang, K.X. Genesis of the Maogongdong deposit in the Dahutang W-Cu-(Mo) ore field of northern Jiangxi Province, South China: Constraints from mineralogy, fluid inclusions, and H-O-C-S isotopes. Miner. Depos. 2022, 57, 1449–1468. [Google Scholar] [CrossRef]
  207. Baker, T.; Van Achterberg, E.; Ryan, C.G.; Lang, J.R. Composition and evolution of ore fluids in a magmatic-hydrothermal skarn deposit. Geology 2004, 32, 117–120. [Google Scholar] [CrossRef]
  208. Seward, T.M.; Barnes, H.L. Geochemistry of Hydrothermal Ore Deposits; John Wiley & Sons: New York, NY, USA, 1997. [Google Scholar]
  209. Shu, Q.H.; Chang, Z.S.; Hammerli, J.; Lai, Y.; Huizenga, J.M. Composition and Evolution of Fluids Forming the Baiyinnuo’er Zn-Pb Skarn Deposit, Northeastern China: Insights from Laser Ablation ICP-MS Study of Fluid Inclusions. Econ. Geol. 2017, 112, 1441–1460. [Google Scholar] [CrossRef]
  210. Shu, Q.H.; Chang, Z.S.; Mavrogenes, J. Fluid compositions reveal fluid nature, metal deposition mechanisms, and mineralization potential: An example at the Haobugao Zn-Pb skarn, China. Geology 2021, 49, 473–477. [Google Scholar] [CrossRef]
  211. Stoffell, B.; Appold, M.S.; Wilkinson, J.J.; McClean, N.A.; Jeffries, T.E. Geochemistry and evolution of mississippi valley-type mineralizing brines from the tri-state and northern arkansas districts determined by LA-ICP-MS microanalysis of fluid inclusions. Econ. Geol. 2008, 103, 1411–1435. [Google Scholar] [CrossRef]
  212. Hedenquist, J.W.; Arribas, A.; Reynolds, T.J. Evolution of an intrusion centered hydrothermal system: Far Southest-Lepanto porphyry and epithermal Cu-Au deposit. Econ. Geol. 1998, 93, 373–404. [Google Scholar] [CrossRef]
  213. Fritzell, E.H.; Bull, A.L.; Shephard, G.E. Closure of the Mongol-Okhotsk Ocean: Insights from seismic tomography and numerical modelling. Earth Planet. Sci. Lett. 2016, 445, 1–12. [Google Scholar] [CrossRef]
  214. Wang, Z.H.; Ge, W.C.; Yang, H.; Zhang, Y.L.; Bi, J.H.; Tian, D.X.; Xu, W.L. Middle Jurassic oceanic island igneous rocks of the Raohe accretionary complex, northeastern China: Petrogenesis and tectonic implications. J. Asian Earth Sci. 2015, 111, 120–137. [Google Scholar] [CrossRef]
  215. Gou, J.; Sun, D.Y.; Qin, Z. Late Jurassic–Early Cretaceous tectonic evolution of the Great Xing’an Range: Geochronological and geochemical evidence from granitoids and volcanic rocks in the Erguna Block, NE China. Int. Geol. Rev. 2019, 61, 1842–1863. [Google Scholar] [CrossRef]
  216. Feng, Z.; Sun, D.Y.; Yue, Y.; Mao, A.Q.; Tian, L.; Sun, C.; Gou, J. Petrogenesis of highly differentiated I-type volcanic rocks: Reinjection of high-temperature magma-An example from Suolun silicic volcanic rocks, central Great Xing’an Range, China. Geol. J. 2020, 55, 6677–6695. [Google Scholar] [CrossRef]
  217. Zhang, G.L.; Xie, W.; Wen, S.Q.; Gong, E.P.; Guo, R.R.; Tang, T.Q. Petrogenesis and tectonic implications of Late Mesozoic volcanic rocks in the northern and central Great Xing’an Range, NE China: Constraints from geochronology and geochemistry. Geol. J. 2020, 55, 8282–8308. [Google Scholar] [CrossRef]
  218. Deng, C.Z.; Sun, D.Y.; Han, J.S.; Li, G.H.; Feng, Y.Z.; Xiao, B.; Li, R.C.; Shi, H.L.; Xu, G.Z.; Yang, D.G. Ages and petrogenesis of the Late Mesozoic igneous rocks associated with the Xiaokele porphyry Cu–Mo deposit, NE China and their geodynamic implications. Ore Geol. Rev. 2019, 107, 417–433. [Google Scholar] [CrossRef]
  219. Rudnick, R.L.; Gao, S. 3.01−Composition of the continental crust. Treatise Geochem. 2003, 3, 1–64. [Google Scholar]
Figure 1. (a) Simplified tectonic map of the Central Asian Orogenic Belt (modified from [27,28]). (b) Schematic tectonic map of NE China (modified from [29]). (c) Distribution of major molybdenum deposits in NE China (modified from [20]). Deposit locations: 1—Jiliuhe; 2—Fukeshan; 3—Taipingchuan; 4—Lower Urgen; 5—Badaguan; 6—Zishi; 7—Wunugetushan; 8—Chalukou; 9—Daheishan; 10—Yezhugou; 11—Taipinggou; 12—Baojiagou; 13—Nanxingan; 14—Jiazishan; 15–Diyanqinamu; 16—Aolunhua; 17—Banlashan; 18—Laojiagou; 19—Haolibao; 20—Hashitu; 21—Erbadi; 22—Yangchang; 23—Longtoushan; 24—Hongshanzi; 25—Xiaodonggou; 26—Liutiaogou; 27—Gangzi; 28—Tuohe; 29—Chehugou; 30—Nianzigou; 31—Yuanbaoshan; 32—Jiguanshan; 33—Baimashi; 34—Kulitu; 35—Huojihe; 36—Cuiling; 37—Luming; 38—Wudaoling; 39—Kanchuangou; 40—Jinchanggou; 41—Fuanpu; 42—Jidetun; 43—Dashihe; 44—Houdaomu; 45—Xingshan; 46—Daheishan; 47—Sangedingzi; 48—Sifangdianzi; 49—Xinhualong; 50—Liushengdian; 51—Sanchazi; 52—Dongfeng; 53—Shirengou.
Figure 1. (a) Simplified tectonic map of the Central Asian Orogenic Belt (modified from [27,28]). (b) Schematic tectonic map of NE China (modified from [29]). (c) Distribution of major molybdenum deposits in NE China (modified from [20]). Deposit locations: 1—Jiliuhe; 2—Fukeshan; 3—Taipingchuan; 4—Lower Urgen; 5—Badaguan; 6—Zishi; 7—Wunugetushan; 8—Chalukou; 9—Daheishan; 10—Yezhugou; 11—Taipinggou; 12—Baojiagou; 13—Nanxingan; 14—Jiazishan; 15–Diyanqinamu; 16—Aolunhua; 17—Banlashan; 18—Laojiagou; 19—Haolibao; 20—Hashitu; 21—Erbadi; 22—Yangchang; 23—Longtoushan; 24—Hongshanzi; 25—Xiaodonggou; 26—Liutiaogou; 27—Gangzi; 28—Tuohe; 29—Chehugou; 30—Nianzigou; 31—Yuanbaoshan; 32—Jiguanshan; 33—Baimashi; 34—Kulitu; 35—Huojihe; 36—Cuiling; 37—Luming; 38—Wudaoling; 39—Kanchuangou; 40—Jinchanggou; 41—Fuanpu; 42—Jidetun; 43—Dashihe; 44—Houdaomu; 45—Xingshan; 46—Daheishan; 47—Sangedingzi; 48—Sifangdianzi; 49—Xinhualong; 50—Liushengdian; 51—Sanchazi; 52—Dongfeng; 53—Shirengou.
Minerals 13 01189 g001
Figure 2. (a) Schematic regional geological map of the Lower Urgen molybdenum deposit. (b) Geological sketch map of the Lower Urgen molybdenum deposit. (c) Schematic geological cross-section of no. 0 exploration line (modified after [30]).
Figure 2. (a) Schematic regional geological map of the Lower Urgen molybdenum deposit. (b) Geological sketch map of the Lower Urgen molybdenum deposit. (c) Schematic geological cross-section of no. 0 exploration line (modified after [30]).
Minerals 13 01189 g002
Figure 3. Representative photographs (ag) and photomicrographs (h,i) of the mineralization in the Lower Urgen molybdenum deposit. (a) Quartz–molybdenite vein of stage 1 in the granite porphyry. (b) Quartz–molybdenite–pyrite vein of stage 1 in the granite porphyry. (c) Molybdenite occurs as disseminations. (d) Quartz–pyrite–chalcopyrite–arsenopyrite ore of stage 2. (e) Quartz–pyrite–chalcopyrite–galena–sphalerite ore of stage 2. (f) The brecciated fragments of granite porphyry containing quartz–molybdenite stockworks of stage 1 hosted by a quartz–pyrite–galena–sphalerite vein of stage 2. (g) Quartz–calcite vein of stage 3. (h) Molybdenite. (i) Galena intergrown with sphalerite. Sphalerite replacing chalcopyrite along crystal margins. Abbreviations: Mol, molybdenite; Py, pyrite; Ccp, chalcopyrite; Gn, galena; Sp, sphalerite; Apy, arsenopyrite; Qtz, quartz; Cal, calcite.
Figure 3. Representative photographs (ag) and photomicrographs (h,i) of the mineralization in the Lower Urgen molybdenum deposit. (a) Quartz–molybdenite vein of stage 1 in the granite porphyry. (b) Quartz–molybdenite–pyrite vein of stage 1 in the granite porphyry. (c) Molybdenite occurs as disseminations. (d) Quartz–pyrite–chalcopyrite–arsenopyrite ore of stage 2. (e) Quartz–pyrite–chalcopyrite–galena–sphalerite ore of stage 2. (f) The brecciated fragments of granite porphyry containing quartz–molybdenite stockworks of stage 1 hosted by a quartz–pyrite–galena–sphalerite vein of stage 2. (g) Quartz–calcite vein of stage 3. (h) Molybdenite. (i) Galena intergrown with sphalerite. Sphalerite replacing chalcopyrite along crystal margins. Abbreviations: Mol, molybdenite; Py, pyrite; Ccp, chalcopyrite; Gn, galena; Sp, sphalerite; Apy, arsenopyrite; Qtz, quartz; Cal, calcite.
Minerals 13 01189 g003
Figure 4. Representative hand specimen (ad) and photomicrographs (crossed-polarized light) (e,f) of gangue minerals from the Lower Urgen molybdenum deposit. (a) Quartz–molybdenite veinlets of stage 1 occur in the granite porphyry with potassic alteration. (b) Quartz–molybdenite–pyrite veins of stage 1 cut by quartz–galena–sphalerite of stage 2. These ore veins occur in the granite porphyry with phyllic alteration. (c) Quartz–pyrite–chalcopyrite vein of stage 2 occurs in the granite porphyry with phyllic alteration. (d) Quartz–galena–sphalerite veinlets of stage 2 are spatially related to silicification and propylitic alteration. (e) Phyllic-altered granite porphyry is characterized by quartz and sericite replacing plagioclase along fractures and crystal margins. (f) Propylitic-altered andesite is featured with epidote replacing plagioclase. Abbreviations: Qtz, quartz; Kfs, K-feldspar; Pl, plagioclase; Ser, sericite; Ep, epidote; Mol, molybdenite; Py, pyrite; Gn, galena; Sp, sphalerite; Ccp, chalcopyrite.
Figure 4. Representative hand specimen (ad) and photomicrographs (crossed-polarized light) (e,f) of gangue minerals from the Lower Urgen molybdenum deposit. (a) Quartz–molybdenite veinlets of stage 1 occur in the granite porphyry with potassic alteration. (b) Quartz–molybdenite–pyrite veins of stage 1 cut by quartz–galena–sphalerite of stage 2. These ore veins occur in the granite porphyry with phyllic alteration. (c) Quartz–pyrite–chalcopyrite vein of stage 2 occurs in the granite porphyry with phyllic alteration. (d) Quartz–galena–sphalerite veinlets of stage 2 are spatially related to silicification and propylitic alteration. (e) Phyllic-altered granite porphyry is characterized by quartz and sericite replacing plagioclase along fractures and crystal margins. (f) Propylitic-altered andesite is featured with epidote replacing plagioclase. Abbreviations: Qtz, quartz; Kfs, K-feldspar; Pl, plagioclase; Ser, sericite; Ep, epidote; Mol, molybdenite; Py, pyrite; Gn, galena; Sp, sphalerite; Ccp, chalcopyrite.
Minerals 13 01189 g004
Figure 5. Representative photomicrographs of FIs in quartz veins of different stages at Lower Urgen. (a) Phase separation recorded by L- and C1-type FIs at stage 1. (b) An immiscibility FIA consisting of L-, S-, C1-, and C2-type FIs at stage 1. (c) Coexistence of L- and S-type FIs at stage 1. (d) Isolate S-type FI at stage 1. (e) Isolate C1-type FI at stage 1. (f) An FIA containing many primary L-type FIs at stage 1. (g) Phase separation recorded by L-, C1-, and C2-type FIs at stage 2. (h) An FIA containing many primary L-type FIs at stage 2. (i) An FIA containing four L-type FIs at stage 3. Abbreviations: H, halite; LCO2, liquid CO2; VCO2, vapor CO2; LH2O, liquid H2O; VH2O, vapor H2O.
Figure 5. Representative photomicrographs of FIs in quartz veins of different stages at Lower Urgen. (a) Phase separation recorded by L- and C1-type FIs at stage 1. (b) An immiscibility FIA consisting of L-, S-, C1-, and C2-type FIs at stage 1. (c) Coexistence of L- and S-type FIs at stage 1. (d) Isolate S-type FI at stage 1. (e) Isolate C1-type FI at stage 1. (f) An FIA containing many primary L-type FIs at stage 1. (g) Phase separation recorded by L-, C1-, and C2-type FIs at stage 2. (h) An FIA containing many primary L-type FIs at stage 2. (i) An FIA containing four L-type FIs at stage 3. Abbreviations: H, halite; LCO2, liquid CO2; VCO2, vapor CO2; LH2O, liquid H2O; VH2O, vapor H2O.
Minerals 13 01189 g005
Figure 6. Representative laser Raman spectra for FIs from the Lower Urgen molybdenum deposit. (a) Raman spectra of stage 1 L-type FI reflect the presence of H2O and minor amounts of CO2. (b) Raman spectra of stage 1 C1-type FI indicate the presence of CO2 in the CO2 phase (VCO2 + LCO2). (c) Raman spectra of stage 1 C2-type FI reflect the existence of CO2 in the CO2 phase (VCO2 + LCO2). (d) Raman spectra of stage 1 S-type FI indicate the presence of H2O and minor amounts of CO2. (e) Raman spectra of stage 2 C1-type FI reflect the occurrence of CO2 in vapor CO2. (f) Raman spectra of stage 2 C2-type FI indicate the existence of CO2 in the CO2 phase (VCO2 + LCO2). (g) Raman spectra of stage 2 L-type FI reflect the presence of H2O and minor amounts of CO2. (h) Raman spectra of stage 3 L-type FI indicate the existence of H2O.
Figure 6. Representative laser Raman spectra for FIs from the Lower Urgen molybdenum deposit. (a) Raman spectra of stage 1 L-type FI reflect the presence of H2O and minor amounts of CO2. (b) Raman spectra of stage 1 C1-type FI indicate the presence of CO2 in the CO2 phase (VCO2 + LCO2). (c) Raman spectra of stage 1 C2-type FI reflect the existence of CO2 in the CO2 phase (VCO2 + LCO2). (d) Raman spectra of stage 1 S-type FI indicate the presence of H2O and minor amounts of CO2. (e) Raman spectra of stage 2 C1-type FI reflect the occurrence of CO2 in vapor CO2. (f) Raman spectra of stage 2 C2-type FI indicate the existence of CO2 in the CO2 phase (VCO2 + LCO2). (g) Raman spectra of stage 2 L-type FI reflect the presence of H2O and minor amounts of CO2. (h) Raman spectra of stage 3 L-type FI indicate the existence of H2O.
Minerals 13 01189 g006
Figure 7. Histograms of homogenization temperatures (a,c,e) and salinities (b,d,f) for FIs in quartz from the Lower Urgen molybdenum deposit according to the concept of FIAs.
Figure 7. Histograms of homogenization temperatures (a,c,e) and salinities (b,d,f) for FIs in quartz from the Lower Urgen molybdenum deposit according to the concept of FIAs.
Minerals 13 01189 g007
Figure 8. δD versus δ18Ofluid diagram for the Lower Urgen molybdenum deposit (base map after [127]). The magmatic water field of the Au-Cu-Mo series is from [128].
Figure 8. δD versus δ18Ofluid diagram for the Lower Urgen molybdenum deposit (base map after [127]). The magmatic water field of the Au-Cu-Mo series is from [128].
Minerals 13 01189 g008
Figure 9. Histogram of sulfur isotopic compositions of metal sulfides from the Lower Urgen molybdenum deposit.
Figure 9. Histogram of sulfur isotopic compositions of metal sulfides from the Lower Urgen molybdenum deposit.
Minerals 13 01189 g009
Figure 10. (a) 207Pb/204Pb versus 206Pb/204Pb diagram. (b) 208Pb/204Pb versus 206Pb/204Pb diagram (modified from [129]).
Figure 10. (a) 207Pb/204Pb versus 206Pb/204Pb diagram. (b) 208Pb/204Pb versus 206Pb/204Pb diagram (modified from [129]).
Minerals 13 01189 g010
Figure 11. Histogram of molybdenite Re-Os isochron ages for Mesozoic molybdenum deposits in NE China. Abbreviations: NCGXR, northern and central Great Xing’an Range; SGXR, southern Great Xing’an Range; LXZR, lesser Xing’an-Zhangguangcai Range. Data sources are listed in Table S2.
Figure 11. Histogram of molybdenite Re-Os isochron ages for Mesozoic molybdenum deposits in NE China. Abbreviations: NCGXR, northern and central Great Xing’an Range; SGXR, southern Great Xing’an Range; LXZR, lesser Xing’an-Zhangguangcai Range. Data sources are listed in Table S2.
Minerals 13 01189 g011
Figure 12. Homogenization temperature versus salinity of FIs from all stages in the Lower Urgen molybdenum deposit.
Figure 12. Homogenization temperature versus salinity of FIs from all stages in the Lower Urgen molybdenum deposit.
Minerals 13 01189 g012
Table 1. Re–Os isotopic data for molybdenite from the Lower Urgen molybdenum deposit.
Table 1. Re–Os isotopic data for molybdenite from the Lower Urgen molybdenum deposit.
Sample No.Re (ppb)Os (ppb)187Re (ppb)187Os (ppb)Model Age (Ma)
ReOs187Re187OsAge
G543,93415580.0830.00827,50397564.5400.698141.21.5
LE-131,2025280.1280.01519,46823147.9580.523147.71.7
Table 2. Hydrogen and oxygen isotopic compositions of quartz from the Lower Urgen molybdenum deposit.
Table 2. Hydrogen and oxygen isotopic compositions of quartz from the Lower Urgen molybdenum deposit.
Sample No.StageMineralδD (‰)δ18Oquartz (‰)T (°C)δ18Ofluid (‰)
B16-1Stage 1Quartz–140.27.63231.5
B16-2Stage 1Quartz–138.37.23231.1
B16-3Stage 1Quartz–136.57.43231.3
B16-4Stage 1Quartz–140.97.93231.8
B16-5Stage 1Quartz–139.08.03231.9
B16-6Stage 1Quartz–141.88.03231.9
B16-7Stage 1Quartz–139.17.63231.5
D14-1Stage 2Quartz–142.77.9229–2.1
D16-1Stage 2Quartz–135.57.6229–2.4
D16-4Stage 2Quartz–142.57.7229–2.3
D16-6Stage 2Quartz–135.47.9229–2.1
D16-7Stage 2Quartz–129.98.1229–1.9
D16-8Stage 2Quartz–142.77.6229–2.4
D16-9Stage 2Quartz–144.17.3229–2.7
G8-1Stage 3Quartz–146.67.7170–6.1
G10-1Stage 3Quartz–143.95.8170–8.0
G11-1Stage 3Quartz–142.67.5170–6.3
Table 3. Sulfur isotopic compositions of sulfides from the Lower Urgen molybdenum deposit.
Table 3. Sulfur isotopic compositions of sulfides from the Lower Urgen molybdenum deposit.
Sample No.StageMineralδ34S (‰)
LU16-1Stage 1Molybdenite5.5
LU16-2Stage 1Molybdenite5.4
LU16-3Stage 1Molybdenite5.5
LU16-4Stage 1Molybdenite5.5
LU16-5Stage 1Molybdenite5.5
LU16-6Stage 1Molybdenite5.5
LU16-7Stage 2Pyrite5.4
LU16-8Stage 2Pyrite5.4
LU16-9Stage 2Chalcopyrite5.4
LU16-10Stage 2Chalcopyrite4.8
LU16-11Stage 2Galena3.9
Table 4. Pb isotopic composition of sulfide and granite porphyry samples from the Lower Urgen molybdenum deposit.
Table 4. Pb isotopic composition of sulfide and granite porphyry samples from the Lower Urgen molybdenum deposit.
Sample No.StageMineral206Pb/204Pb207Pb/204Pb208Pb/204Pbμ
D17-21Molybdenite18.3400.000415.5420.000338.1370.00089.36
D17-31Molybdenite18.3400.000515.5430.000538.1530.00119.36
D17-41Molybdenite18.3550.000415.5440.000438.1640.00099.36
D17-51Molybdenite18.3430.000415.5400.000438.1350.00099.35
D17-71Molybdenite18.3400.000415.5400.000438.1300.00099.35
D0A1-11Molybdenite18.4120.000415.5460.000438.2120.00099.36
D0A1-21Molybdenite18.3870.000415.5440.000438.1830.00109.36
D0a1-11Molybdenite18.3630.000415.5430.000338.1700.00099.36
G10-22Pyrite18.3730.000515.5490.000438.1740.00109.37
G11-22Pyrite18.5100.000415.5570.000438.2490.00109.37
G132Pyrite18.4710.000415.5690.000338.3120.00089.40
G142Pyrite18.3410.000315.5490.000338.1480.00089.37
D17-12Pyrite18.3360.000315.5420.000238.1310.00069.36
D17-22Chalcopyrite18.3360.000515.5410.000438.1310.00119.36
G8-22Chalcopyrite18.3400.000415.5470.000338.1430.00089.37
G8-32Galena18.3430.000415.5500.000438.1490.00099.37
T0a1-1Granite porphyry18.5990.000515.5740.000538.5830.00139.40
T0a1-2Granite porphyry18.5650.000515.5730.000438.5150.00129.40
T16-1Granite porphyry18.5350.000415.5710.000438.4190.00109.39
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

Zhang, G.; Xie, W.; Wen, S.; Zeng, Q.; Zhou, L.; Wang, H.; Zhang, K.; Tang, T.; Ma, P. Ore Genesis of the Lower Urgen Porphyry Molybdenum Deposit in the Northern Great Xing’an Range, Northeast China: Constraints from Molybdenite Re-Os Dating, Fluid Inclusions, and H-O-S-Pb Isotopes. Minerals 2023, 13, 1189. https://doi.org/10.3390/min13091189

AMA Style

Zhang G, Xie W, Wen S, Zeng Q, Zhou L, Wang H, Zhang K, Tang T, Ma P. Ore Genesis of the Lower Urgen Porphyry Molybdenum Deposit in the Northern Great Xing’an Range, Northeast China: Constraints from Molybdenite Re-Os Dating, Fluid Inclusions, and H-O-S-Pb Isotopes. Minerals. 2023; 13(9):1189. https://doi.org/10.3390/min13091189

Chicago/Turabian Style

Zhang, Guangliang, Wei Xie, Shouqin Wen, Qingdong Zeng, Lingli Zhou, Hui Wang, Kailun Zhang, Tieqiao Tang, and Pengcheng Ma. 2023. "Ore Genesis of the Lower Urgen Porphyry Molybdenum Deposit in the Northern Great Xing’an Range, Northeast China: Constraints from Molybdenite Re-Os Dating, Fluid Inclusions, and H-O-S-Pb Isotopes" Minerals 13, no. 9: 1189. https://doi.org/10.3390/min13091189

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