Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Hindawi Geofluids Volume 2018, Article ID 2197891, 23 pages https://doi.org/10.1155/2018/2197891 Research Article The Fe-Zn Isotopic Characteristics and Fractionation Models: Implications for the Genesis of the Zhaxikang Sb-Pb-Zn-Ag Deposit in Southern Tibet Da Wang,1,2 Youye Zheng ,1,3 Ryan Mathur ,2 and Song Wu1 1 State Key Laboratory of Geological Processes and Mineral Resources, School of Earth Sciences and Resources, China University of Geosciences, Beijing 100083, China 2 Department of Geology, Juniata College, Huntingdon, PA 16652, USA 3 State Key Laboratory of Geological Processes and Mineral Resources and Faculty of Earth Resources, China University of Geosciences, Wuhan 430074, China Correspondence should be addressed to Youye Zheng; zhyouye@163.com and Ryan Mathur; MATHURR@juniata.edu Received 30 August 2017; Revised 17 December 2017; Accepted 21 January 2018; Published 20 March 2018 Academic Editor: Bin Chen Copyright © 2018 Da Wang et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. The genesis of the Zhaxikang Sb-Pb-Zn-Ag deposit remains controversial. Three different geological environments have been proposed to model mineralization: a hot spring, a magmatic-hydrothermal fluid, and a sedimentary exhalative (SEDEX) overprinted by a hot spring. Here, we present the electron probe microanalysis (EPMA) and Fe-Zn isotopic data (microsampled) of four samples from the first pulse of mineralization that show annular textures to constrain ore genesis. The Zn/Cd ratios from the EPMA data of sphalerite range from 296 to 399 and overlap the range of exhalative systems. The 𝛿56 Fe values of MnFe carbonate and 𝛿66 Zn values of sphalerite gradually decrease from early to late stages in three samples. A combination of the EPMA and isotopic data shows the Fe-Zn contents also have different correlations with 𝛿66 Zn values in sphalerite from these samples. Rayleigh distillation models this isotope and concentration data with the cause of fractionation related to vapour-liquid partitioning and mineral precipitation. In order to verify this Rayleigh distillation model, we combine our Fe-Zn isotopic data with those from previous studies to establish 12 Fe-Zn isotopic fractionation models. These fractionation models indicate the 𝛿56 Fe𝑖 and 𝛿66 Zn𝑖 values (initial Fe-Zn isotopic compositions) of the ore-forming system are in the range of −0.5‰∼−1‰ and −0.28‰∼0‰, respectively. To conclude, the EPMA data, Fe-Zn isotopic characteristics, and fractionation models support the SEDEX model for the first pulse of mineralization. 1. Introduction To date, the Zhaxikang Sb-Pb-Zn-Ag deposit is the only super large deposit that has been identified within the North Himalayan Polymetallic Metallogenic Belt (NHMB). Although basic research including the geology, petrography, geochronology, and geochemistry studies has been conducted (e.g., [1, 2]), the genesis of this deposit is still debated due to the complicated mineralogy and the presence of multiple stages of mineralization. The main viewpoints involve a hot spring [3, 4], two magmatichydrothermal fluids [5, 6], and a SEDEX overprinted by hot spring [7] genetic models. However, most of these genetic models are based on the S, C, O, H, and Si isotopic evidence, which cannot absolutely trace the metal source. The traditional light stable C, H, O, S, and N isotopes have been widely used to constrain fluid evolution and metal sources in ore deposit studies (e.g., [8, 9]). However, the evidences for metal source from these elements are always indirect and putative as they are not the metallogenic elements themselves [10]. For instance, these elements usually have different characteristics with the changing of tectonic settings, and sometimes they may even have different sources from the metallogenic elements [11]. However, the nontraditional transition metal stable isotopes (e.g., Fe, Zn, Cu, Cd, Mg, Cr, Sn, and Mo) are more precise tracers for the metal sources and ore-forming processes in metallogenic systems 2 [10, 12, 13]. The development of Multicollector-Inductively Coupled Plasma Mass Spectrometer (MC-ICP-MS) technology has greatly improved the precision of isotopic analyses [14, 15], which results in the wide application of the nontraditional transition metal stable isotopes in economic geology studies (e.g., [16, 17]). The Fe-Zn isotopes are two of the most representative isotopes applied in ore deposit studies. For example, Mason et al. [18] and Wilkinson et al. [19] both identified that the 𝛿66 Zn values of minerals precipitating from the same hydrothermal fluid become heavier over time by studying the Zn isotopic fractionation of the Alexandrinka volcanic hosted massive sulfide (VHMS) type deposit in Russia and Midlands Irishtype deposit in Ireland, respectively. The gradual increasing 𝛿66 Zn values both from early to late stages and from south to north within the Red Dog ore district in Alaska record the temporal-spatial evolution of the ore-forming fluid and constrain the SEDEX genesis with a single Zn source [20]. In addition, Fe isotopic studies of skarn Cu−Au±Fe deposits in South China excluded the possibility that Fe originated from sedimentary strata [21, 22]. These Fe isotopes matched the igneous source rocks and mineralization, and the 𝛿56 Fe values of sulfides gradually increase both from early to late stages and away from the ore-related igneous rocks. Wang et al. [21, 22] also revealed that the Fe isotopes fractionate during fluid exsolution and that the ore-forming fluid is enriched in light isotopes relative to ore-related igneous rocks. To the contrary, Wawryk and Foden [23] investigated the Fe-isotope fractionation in the Renison Sn-W deposit in Australia and discovered that Fe isotopic compositions of pyrite (0.61‰∼1.14‰), chalcopyrite (0.18‰∼0.71‰), and magnetite (0.50‰∼0.70‰) are isotopically heavier than Renison granite (0.27‰∼0.45‰) and thus hypothesized that a magmatic-hydrothermal fluid exsolved from an isotopically heavy reduced magma could deposit isotopically heavy ore minerals whereas oxidized magmas crystallise magmatic magnetite could result in an isotopically lighter melt and fluid. These studies demonstrate the potential of Fe-Zn isotopes to trace the metal source and provide insights into ore-forming evolution. With regard to the Zhaxikang deposit, Duan et al. [5] have investigated the Zn isotope of sphalerite, galena, FeMn carbonates, and igneous rocks and speculated that the 𝛿66 Zn values of the hydrothermal fluid are 0.39 ± 0.10‰. This value is consistent with those of basement rocks (average value of 0.36‰ ± 0.03‰) and Fe-Mn carbonates (average value of 0.27‰ ± 0.15‰), which is identified as the evidence for the magmatic origin. Meanwhile, the contribution of regional sedimentary rocks is conjectured by the Zn-Pb-S isotopes: (1) the Zn isotopic variation range of sulfides (−0.25‰∼0.03‰) is larger than basement rocks (0.05‰∼0.21‰); (2) the radiogenic Pb isotopic compositions of sulfides (e.g., 206 Pb/204 Pb = 18.727∼19.896) is higher than regional igneous rocks (206 Pb/204 Pb = 18.4∼19.2); (3) the 𝛿34 S values of sulfides (6‰∼12‰) are lighter than regional sedimentary wall rocks (10.94‰∼11.49‰) but higher than mantle value (0 ± 2‰). However, the Zn isotopic fractionation during the fluid exsolution and leaching process Geofluids [27, 28] has been ignored in Duan et al. [5]. In addition, Wang et al. [26] also studied the Fe-Zn isotopes of the pyrite, sphalerite, and Mn-Fe carbonate in Zhaxikang deposit, which successfully constrained the two pulses of mineralization by the temporally increasing 𝛿56 Fe and decreasing 𝛿66 Zn values recorded in the deposit that coincided with an increase in alteration. The Fe-Zn isotopic research also demonstrated the magmatic-hydrothermal fluid origin of the second pulse of mineralization by the heavier 𝛿56 Fe values of stage 3 pyrite and excluded the possibility that slate is the metal source by the similar 𝛿66 Zn values of slate and sphalerite. Nevertheless, the attempt to trace the metal source of the first pulse of mineralization failed in Wang et al. [26]. In this study, we present the Fe-Zn isotopic values and variations within four annular polished section samples from Zhaxikang deposit to provide more credible evidence for the primary and earlier stages of ore genesis. 2. Geological Setting 2.1. Regional Geology. The NHMB is in the eastern section of the North Himalayan Tectonic Belt (NH), in the Himalayan terrane. From north to south, the Himalayan terrane is divided into four tectonic belts: the North Himalayan Tethys Sedimentary Fold Belt (the North Himalayan Tectonic Belt; NH), the High Himalayan Crystalline Rock Belt (HH), the Low Himalayan Fold Belt (LH), and the Sub-Himalayan Tectonic Belt (SH; Figure 1(a)) [29–31]. These belts are separated by four nearly EW-trending faults, including the South Tibet Detachment System (STDS), the Main Central Thrust (MCT), the Main Boundary Thrust (MBT), and the Main Frontal Thrust (MFT; Figure 1(a)) [24, 32]. The NH, composed of a set of Palaeozoic marine sedimentary sequences that formed in a passive continental margin environment within northern India [33], is located between the Indus-Yarlung Zangbo Suture Zone (IYZS) and the HH (Figure 1(a)). The sedimentary sequence in the NH records Late Precambrian to Devonian prerift, Carboniferous to Early Jurassic syn-rift and Middle Jurassic to Cretaceous passive continental margin sediments (Figure 1(b)) [66–68]. These sediments crop out in an EW and NWW trend and are predominated by the Precambrian Laguigangri Group and a series of Upper Triassic, Jurassic, Lower Cretaceous, and Quaternary sediments. The Laguigangri Group crops out in the core of the Yelaxiangbo dome and is composed of schist, gneiss, and migmatite units (Figure 1(b)). A set of Late TriassicEarly Cretaceous flysch formations deposit in neritic-bathyal environments and crop out across the NH. This set of formations are dominated by turbidite deposits and host the majority of the Au-Sb-Pb-Zn-Ag deposits in the NH. The lithology of these Late Triassic-Early Cretaceous flysch formations is weak-metamorphic slate that is intercalated with metamorphosed fine-grained sandstone, argillaceous limestone, micrite, and siliceous rock that is intercalated with volcanic rocks [7]. Some quaternary sediments also occur in the central and southern area of the NH, which are composed of gravel, sand gravel, sandy loam, clay, and ice boulder. Geofluids 3 90∘ 91∘ 92∘ 93∘ N (b) IYZS (km) Lhasa Terrane Jiacha Ranba dome 29 Baijia Haweng ∘ Jiangzi Langkazi Kangbugunba Shalagang Qusong Wuladui Qiangtang Terrane Rongbu-Gudui Fault Jibupu (km) BNSZ Cheqiongzhubu Lhasa Terrane Lhasa IYZS STDS LH Cuomei Rangla Xiaba 30∘ Jiena NH Fig. (b) Keyue Longzi Zedang Jiangzhula Jisong Langzhang Zhaxikang D ST Cuona Bhutan Cretaceous: clastic and carbonate rocks, interlayered with volcanic rocks Jurassic: shelf-facies shale (slab), sandstone, limestone, interlayered with volcanic rocks Upper Triassic:deep-water shelf-facies sandstone shale (slab), interlayered with volcanic rocks Lower-middle Triassic: shallow-marine sandstone and shale (slab) Permian:conglomerate, sandstone, marble, and silty slate Gudui Luo zha faul t HH MCT MBT SH India continent MFT 88∘ 92∘ Mazhala 91 ∘ 29∘ Chalapu Zhegu Cimalong 200 0 Xigong Zhemulang Lazi-Qiongd uojiang fau lt Xiaqiong Kangma dome (a) Bangbu Xiangdala Quexiaxiongqu Yelaxiangbo dome Kaliepu Delong 25 0 28∘ 92 ∘ 93 ∘ Precambrian metamorphic rocks Late Jurassic-Early Cretaceous mafic rocks Thrust fault Cenozoic granitoids Fault Different grade fault Gological boundary Pb-Zn deposit Ductile shear zone Internation border Sb-Pb-Zn(Ag) deposit Detachment fault County town or city Au deposit Sb deposit Sb-Au deposit Figure 1: (a) Tectonic framework of the Himalayan Terrane (modified from Yin [24]). (b) Regional geological map of the North Himalayan Polymetallic Metallogenic Belt (modified from Zheng et al. [25]). BNSZ: the Bangong-Nujiang Suture Zone; IYZS: the Indus-Yarlung Zangbo Suture Zone; STDS: the South Tibet Detachment System; MCT: the Main Central Thrust Fault; MBT: the Main Boundary Thrust Fault; MFT: the Main Frontal Thrust; NH: the North Himalayan Tethys Sedimentary Fold Belt; HH: the High Himalayan Crystalline Rock Belt; LH: the Low Himalayan Fold Belt; SH: the Sub-Himalayan Tectonic Belt. The EW-trending and NS-trending faults, both with the several episodes of motion, are present in the NH. The EWtrending faults, controlling the distribution of intermediateacid magmatic rocks and ore deposits in the NH [7], are older and cover a larger area than the NS-trending faults. The representative EW-trending faults include the LaziQiongduojiang, Rongbu-Gudui, and Luozha faults as well as the STDS and numerous metamorphic core complexes (Figure 1(b)). A series of rifts that formed from 25 Ma to present are associated with these faults [29, 69–71], mainly including the Sangri-Cuona, Yadong-Gulu, Shenzha-Xietongmen, and Dangreyongcuo-Gucuo rift zones from east to west [72]. In addition, the NS-trending faults that are considered as the result of east-west extension of the Qinghai-Tibet Plateau [73, 74] also formed during this period especially from 18 to 4 Ma [75, 76]. These NS-trending faults are also the important ore-controlling structures in the NH [1]. Magmatism in the NH primarily includes the Mesozoic and Cenozoic magmatism. The Mesozoic magmatism generated multiple suites of mafic-intermediate igneous rocks between the Late Triassic and the Early Cretaceous, including basaltic volcanic interlayers, dyke swarms, and subvolcanic dykes. According to the previous geochronological data, the SHRIMP U-Pb ages of the basic dyke swarms from different area in the NH are 134.9 ± 1.8 Ma, 135.5 ± 2.1 Ma [77], and 138.0 ± 3.5 Ma [78], respectively. The SHRIMP U-Pb age of the gabbro is 155.8 Ma [79]. Tong et al. [78], Pan et al. [80], and Zhong et al. [81] regarded these basic dyke swarms as the result of late-stage massive expansion of Neo-Tethys Ocean under the structural environment of the Himalaya passive continental margin intensive stretching and breaking-off, lithosphere extension-thinning, and asthenosphere upwelling. On the contrary, Zhu et al. [82] and Qiu et al. [83] suggested that these basic dyke swarms are the result of interaction between mantle plume and lithospheric mantle material and form in the continental-rift environment. The Cenozoic magmatism is characterized by the formation of monzogranite, leucogranite, diorite, porphyritic diorite, and aplite units [84, 85]. These Cenozoic intermediate-acidic intrusive masses are distributed along the EW-trending faults and in the core of Ranba, Kangma, and Yelaxiangbo dome in the form of batholith, laccolith, and dykes (Figure 1(b)). This phenomenon is considered to be the result of crustal thickening [86] related to the collision of the India Plate and the Eurasia Plate during the postcollision stage (25 to 0 Ma) [87, 88]. 4 The NHMB contains many Sb, Au, Sb-Au, Pb-Zn, and Sb-Pb-Zn-Ag deposits, and the Zhaxikang Sb-Pb-Zn-Ag, the Mazhala Au-Sb, the Chalapu Au, the Bangbu Au, the Shalagang Sb, the Cheqiongzhuobu Sb deposits are representative (Figure 1(b)) [7, 25]. The geneses and metallogenic age of these deposits are controversial due to the limited research, the genetic models mainly include the SEDEX overprinted by hot spring, carlin and carlin-like, hot spring, subvolcanic magmatic-hydrothermal fluid, and orogenic types [25]. 2.2. Ore Deposit Geology. The Zhaxikang Sb-Pb-Zn-Ag deposit is located ∼48 km west from Longzi County Town within the southeastern Yangzuoyong-Nariyong composite syncline in the NH (Figure 1(b)). This deposit has a reserve of 1.268 Mt Pb-Zn with a 3.66% average Zn grade and a 2.45% average Pb grade, 0.138 Mt Sb with an average grade of 1.08%, 1800 t Ag with an average of 99.55 g/t, 3.9 t associated Au, 361 t associated Ga, and 20 Mt Mn-Fe carbonate ores with an average grade of 42% for Fe + Mn [89], which makes it the largest deposit within the NHMB. The majority of mineralization in the orefield is hosted by the Lower Jurassic Ridang formation that consists of epimetamorphic marine clastic rocks. This formation, dipping shallowly to the north and striking eastwest, is divided into five lithologic sections (Figure 2(a)). A few Upper Jurassic Weimei formations composed of fine-grained metamorphic quartzose sandstone, silty slate, and calcarenite as well as Quaternary sediments distributed along valleys also crop out in the orefield (Figure 2(a)) [7]. The Zhaxikang deposit developed extensive geological structures. A near northsouth striking fault system is prevalent in the orefield, which coexists with a group of northeaststriking faults and some folds. Engineering and geological mapping projects have identified 16 faults, the majority of which are steeply dipping normal faults associated with tensional stress and wrench faults associated with torsional stress. Faults F2, F4, F5, F6, F7, F13, F14, and F16 are the main ore-bearing faults, faults F1 and F10 are partly mineralized, fault F3 was associated with late-stage mineralization, faults F8 and F9 are wrench faults without any mineralization, and fault F12 is a nonmineralized regional fault (Figure 2(a)). The orebodies I–VI are hosted by nearly NS-striking faults and orebodies VII–IX are hosted by nearly NE-striking faults (Figure 2). Our samples in this study are all from the orebody V, which is the largest and richest one among these orebodies within the orefield and hosts more than 80% of the reserves. This orebody is >1400 m long, 1 to 30 m wide, and controlled by fault F7 (Figure 2). The magmatism in the orefield is associated with diabase, porphyritic rhyolite, basalt, and leucogranite units as well as some granite porphyry dykes that intruded into the porphyritic rhyolite (Figure 2(a)). The diabase is identified by drillholes and footrill in the central part of the orefield as dykes that emplaced into the Jurassic Ridang Formation and has been dated by zircon U-Pb methods to ∼133 Ma [7]. The rhyolite porphyry with the zircon SHRIMP U-Pb age of ∼135 Ma crops out in the western part of the orefield [90] and the leucogranites crop out in the southern part of the orefield over an area of <1 km2 . Additionally, the basalt usually occurs Geofluids near the orebody in the form of consequent layer or shear layer distributed in slate and the contact region of slate and diabase. Various types of alteration associated with mineralization have occurred in the orefield, including (1) the silicification that is associated with Sb mineralization and generally located in fault zones in the form of quartz veins, radiating quartz, and quartz druse; (2) the carbonatization that is associated with Pb-Zn mineralization in the form of MnFe carbonate veins and also formed the postmineralization calcite; (3) the chlorite alteration that is generally confined to massive and stellated aggregates of chlorite within diabase; (4) the weak sericite alteration that is associated with chlorite alteration and barren quartz; and (5) the clay alteration that developed along the edges of fracture-related crushed zones. Furthermore, the ore-forming elements display a vertical sequence that is zoned from a lowermost Zn (Pb + Ag) zone through a central Zn + Pb + Ag-(Sb) zone to an uppermost Pb + Zn + Sb + Ag zone, although no horizontal zoning is present [89]. 2.3. Ore Paragenetic Sequence. The paragenetic sequence in the Zhaxikang deposit is divided into six stages of ore formation based on the detailed hand specimen and microscopic observations. These six stages are assigned to two clear pulses: the first pulse consists of stages 1 to 2 and is characterized by the assemblages of Mn-Fe carbonates and sulfides, and the second pulse includes stages 3 to 6 and is primarily dominated by quartz, calcite, sulfosalt minerals, and sulfides (Figure 3). Stage 1, dominated by a Mn-Fe carbonate + sphalerite + pyrite + arsenopyrite assemblage, is the initial stage of ore formation in the Zhaxikang deposit. Majority of the fine-grained sphalerite, pyrite, and arsenopyrite are hosted by fine-grained Mn-Fe carbonate in the form of laminae (Figures 4(a)–4(c)), and a few sulfides occur within the MnFe carbonates as stellated aggregates (Figure 4(d)). The finegrained layered and colloform with synsedimentary features (Figure 4(b)). The laminae and Mn-Fe carbonates in some samples have been cut by later stage 4 quartz-boulangerite veins (Figure 4(c)) or have been affected by the stage 2 coarsegrained sphalerite (Figure 4(a)). Stage 2 hosts majority of the Pb-Zn mineralization in the Zhaxikang deposit and comprises an assemblage of MnFe carbonate + galena + sphalerite + pyrite ± arsenopyrite. The more abundant and coarser-grained sulfides are hosted by coarse-grained Mn-Fe carbonate and slate to form the banded (Figure 4(k)), net-veined (Figure 4(i)), massive (Figure 4(h)), concentric annular (Figures 4(g) and 5), and Dal Matianite (Figure 4(l)) ores. The Mn-Fe carbonate during this stage recrystallized to different degree (Figures 4(f), 4(g) and 4(i)–4(k)), some even formed the druse containing the idiomorphic columnar quartz, needle-like boulangerite, or valentinite (Figure 4(j)). We can also observe that the later sphalerite replaces the earlier pyrite containing automorphic stage 1 arsenopyrite to form a skeletal texture during this stage (Figure 4(w)). Zheng et al. [7] considered that the ore textures in stages 1 and 2 are similar to those of the Red Dog SEDEXtype ore district in Alaska. Geofluids 5 F16 91∘ 59 92∘ 00 IX 92∘ 01 N F7 28∘ 23 F5 ZK708 F1 F3 ZK705 ZK704 ZK701 ZK706 ZK721 ZK722 F6 ZK703 Line 7 ZK702 50 265 35 270 F4 28∘ 23 F2 62 270 58 270 37 270 61 270 III 42 300 IV V II VI 28∘ 22 28∘ 22 (m) 0 200 J1 r5 Sandstone inserted with slate J1 r4 Slate inserted with sandstone/basalt J1 r3 Sandstone J1 r2 Slate J1 r 1 Metamorphic quartz-sandstone Rhyolite porphyry Diabase vein Quaternary 28∘ 21 Ore body and number II F3 Geological boundary 42 300 Line 7 ZK701 56 315 Fault number >CJ ;HAF? >CJ 55 315 Exploration line and drill hole number 91∘ 59 VII 28∘ 21 F13 F14 VIII 92∘ 01 92∘ 00 (a) 93∘ 4883 4800 ZK708 ZK706 ZK705 ZK703 ZK704 ZK701 III 4700 200 (m) ZK702 ZK721 ZK722 VII VI 120.02 PD7 CM710 200 V Altitude: 4575m 4500 m .4 9 m IV 4600 (m) 0 307.43 m 312.16 m 323.58 m 4400 473.68 m 464.96 m 4300 4200 664.75 m 4100 4000 3950 883.7 m ZK701 Orebody Drill hole PD7 CM710 Trench Sampling location (b) Figure 2: (a) Geological map of the Zhaxikang Sb-Pb-Zn-Ag polymetallic deposit (modified from Zheng et al. [7]). (b) Cross-section along Exploration Line 7. 𝐽1 𝑟1 to 𝐽1 𝑟5 : the first to fifth lithologic section of the Lower Jurassic Ridang formation. Stage 3, characterized by the formation of a quartz ± calcite + pyrite + sphalerite + galena ± chalcopyrite ± arsenopyrite assemblage without Mn-Fe carbonate, is the earliest stage of the second pulse of mineralization. The massive, veined, net-veined, and brecciated sphalerite, galena, and pyrite occur in the quartz and calcite (Figures 4(l)–4(p), 4(x)), and most of the sulfides form by the modification of sulfides from earlier stages. Some chalcopyrite grains are distributed in sphalerite, galena, and pyrite as stellated aggregates (Figures 4(y) and 4(z)). Some of these sulfides have been cross-cut by the later quartz-boulangerite or quartzcalcite veins (Figure 4(o)). Stage 4 is marked by a mineral assemblage composed of quartz + antimony-lead-silver sulfosalt minerals (Figure 4(r)) that prevailingly include boulangerite and jamesonite, as well as minor bournonite, tetrahedrite, and andorite. This stage hosts the majority of the Sb and Ag mineralization and yields the ores with relatively high average Ag grades. The minerals 6 Geofluids Supergene Stage Stage 1 Stage 2 Stage 3 Stage 4 Stage 5 Stage 6 stage Mineral Mn-Fe carbonate Sphalerite Galena Pyrite Arsenopyrite Chalcopyrite Quartz Calcite Sericite Boulangerite Jamesonite Bournonite Zinckenite Freibergite Andorite Tetrahedrite Stibnite Cinnabar Ferrihydrite Smithsonite Sardinianite Valentinite Travertine Malachite Siliceous sinter Notes: Abundant Intermediate Minor Figure 3: Mineral paragenesis within the Zhaxikang Sb-Pb-Zn-Ag deposit (modified from Wang et al. [26]). formed in earlier stages are replaced and cross-cut by the quartz-boulangerite veins and boulangerite of this stage (Figures 4(c) and 4(aa)). Some samples also contain quartz druse filled with needle-like boulangerite (Figure 4(q)). Stage 5 is distinguished by the formation of a quartz + stibnite + cinnabar assemblage and hosts part of the Sb mineralization within the deposit. Elongate-radial stibnite and massive stibnite-cinnabar are hosted by quartz (Figures 4(s) and 4(t)). Some stibnite cross-cut the stage 4 boulangerite (Figure 4(aa)). Stage 6, representing the youngest stage of mineralization in the Zhaxikang deposit, is identified by a quartz ± calcite assemblage without sulfides. The quartz-calcite veins of this stage cross-cut the earlier formed minerals (Figures 4(o) and 4(ab)). Zheng et al. [7] regarded the ore textures in the second pulse of mineralization as typical hot spring type metallogenic features. Supergene stage in the Zhaxikang deposit consists of ferrihydrite, smithsonite, sardinianite, valentinite, travertine, malachite, and siliceous sinter (Figures 4(u) and 4(v)). 3. Sampling and Analytical Methods 3.1. Sampling. The sampling points for EPMA and Fe-Zn isotopic analyses (the powders are sampled by microdrill) are all in the annular polished section samples 9-3, 9-8, ZXK-1, and ZXK-2. The specific numbers, locations, and photomicrographs of these sampling points are given in Figures 5 and 6, respectively. These four samples are all from the first pulse of mineralization, only the sample ZXK-2 has been cut by a stage 3 quartz vein (Figure 5(d)). 3.2. EPMA. Chemical compositions of sulfide, Mn-Fe carbonate, and quartz were determined on a JEOL (Japan Electron Optics Laboratory) JXA-8100 electron microprobe (EMP) at the Second Institute of Oceanography, State Oceanic Administration of China. The accelerating voltage is 15 kV for Mn-Fe carbonate and quartz and 20 kV for sulfide, the beam current is 10 nA, the beam diameter is 1 𝜇m, the secondary electronic resolution is 6 nm with the operating distance of 11 mm, and the repeat accuracy of the sample stage is within 1 nm. The standards are natural minerals and synthetic oxides as those of Sun et al. [2]. The correction program supplied by the manufacturer is used for matrix corrections [91, 92]. 3.3. Fe-Zn Isotopic Analyses. Approximately 10–50 milligrams of sample powders was placed in 15 ml Teflon jars and the solids were dissolved in 4 ml of heated ultrapure aqua regia. The solutions were dried and then Fe and Zn were purified using the BioRad MP-1 anion exchange resin using the protocol from Maréchal et al. [14]. Yields from the columns were tested volumetrically on the ICP-OES at Pennsylvania State University and were all greater than 95%. Isotope values are reported in the traditional per mil values (‰). The Fe isotopes were measured on the Neptune MC-ICPMS at Pennsylvania State University. The instrument setup, sample introduction, and running conditions are discussed in greater detail in Yesavage et al. [93]. Samples were diluted to a 3 ppm Fe solution which produced approximately a 10 V signal on the shoulder to the argon interference peak (56 Fe and 40 Ar16 O). Sample intensities matched the intensity of the bracketing standard within 10%. Mass bias was corrected for by standard-sample-standard bracketing. In-house and international standards were measured throughout the sessions and yielded overlapping values of SRM-3126a 𝛿56 Fe = 0.33 ± 0.08‰, 𝑛 = 8 (accepted values 𝛿56 Fe = 0.34 ± 0.1‰ 2𝜎) [93], and HPS-WU 𝛿56 Fe = 0.62 ± 0.11‰, 𝑛 = 8 (accepted values 𝛿56 Fe = 0.60 ± 0.07‰ 2𝜎) [34]. The samples are reported relative to the international standard IRMM014 (𝛿56 Fe (‰) = [(56 Fe/54 Fe)sample /(56 Fe/54 Fe)IRMM–014 − 1] × 1000). Reported values are an average of two different measurements and the errors fall within the range 0.1‰ 2𝜎 of the standards. The Zn isotopes were measured on Neptune MC-ICP-MS at Rutgers University. Correction of mass bias for Zn using Cu (NIST 976) was employed for these samples as suggested in [94–97] and the corrected values were then bracketed by the standards. The samples are reported relative to the newly developed Zn isotope standard (AA-ETH; 𝛿66 Zn (‰) = [(66 Zn/64 Zn)sample /(66 Zn/64 Zn)AA–ETH − 1] × 1000) and all the quoted data from previous literatures in this paper are converted relative to the AA-ETH standard (𝛿66 ZnAA–ETH = 𝛿66 ZnJMC 3–0749 L − 0.28‰) [98]. We also compared the new Geofluids 7 Sp2 Sp2 Sp2 Mcar2 Py2 Mcar2 Apy1-Py1-Sp1 Apy1-Py1-Sp1 Gn2 Mcar1 Py2 Mcar1-Sp1 Sp2 Mcar1 Gn2 Apy1-Py1-Sp1 Qtz4-Blr4 Vein 2 =G 4 =G (a) 3 =G 3 =G (b) (c) (d) Py2 Mcar2 Mcar2 Py2 Sp2 Gn2 Py2 Sp2 Slate 4 =G 2 =G 3 =G (f) (e) 5 =G (h) (g) Mcar2 Py2 Sp2 Sp2 Py2 Py2 Mcar2 Mcar2 Sp2 Mcar2 Sp2 Sp2 Qtz5-valentinite Qtz3 4 =G 3 =G 2 =G (j) (i) (k) (l) Sp3 Slate Sp3 Cal3 Qtz6-Cal6 Qtz3-Cal3 Sp3 Sp3-Py3 Qtz4-Blr4 Gn3 Qtz3 2 =G 3 =G (m) 2 =G (o) (n) Blr4 (p) Qtz5 Qtz4 Stb5 Stb5 Qtz4 Blr4 2 =G Blr4 Ci5 5 =G 3 =G (q) 3 =G (r) (s) Qtz5 3 =G (t) Sp3 siliceous sinter sardinianite Lm Sp2 Qtz3 Py2 Apy1 2 =G (u) 5 =G (v) 100 G 100 G (w) Figure 4: Continued. Mcar2 (x) 8 Geofluids Blr4 Stb5 Sp3 Stb5 Ccp3 Ccp3 Qtz6 Sp3 Py3 Qtz5 20 G (y) 100 G 200 G (z) (aa) 200 G (ab) Figure 4: Hand specimen photographs and photomicrographs of representative samples from the Zhaxikang deposit. (a) Stage 1 lamellar sphalerite-pyrite-arsenopyrite and stage 2 massive sphalerite-pyrite hosted within fine-grained Mn-Fe carbonate. (b) Stage 1 lamellar and stage 2 banded Mn-Fe carbonate-sphalerite-galena ore with visible synsedimentary features including rhythmic sedimentation in the upper part and angular folding in the lower part of the sample. (c) Stage 1 lamellar sphalerite-pyrite-arsenopyrite and stage 2 massive and banded sphalerite-pyrite hosted by fine-grained Mn-Fe carbonate. The mineral assemblage is in turn cross-cut by stage 4 quartz-boulangerite veins. (d) Coarse-grained stage 2 Mn-Fe carbonate-sphalerite formed by the recrystallization of fine-grained stage 1 Mn-Fe carbonate-sphalerite. (e) Stage 2 massive coarse-grained pyrite hosted by slate. (f) Massive and brecciated stage 2 sphalerite hosted in stage 2 Mn-Fe carbonate. (g) Stage 2 massive, globular, and concentric annular sphalerite-pyrite hosted by coarse-grained Mn-Fe carbonate. (h) Stage 2 massive galena and pyrite. (i) Stage 2 massive and veined sphalerite-pyrite hosted by coarse-grained Mn-Fe carbonate. (j) Stage 1 lamellar sphalerite-pyrite-arsenopyrite and stage 2 massive sphalerite hosted by fine-grained Mn-Fe carbonate. The sample also contains a Mn-Fe carbonate druse dominated by idiomorphic columnar quartz and valentinite. (k) Stage 2 coarse-grained sphalerite-pyrite hosted by stage 2 Mn-Fe carbonate with banded textures. (l) Stage 2 sphalerite and Mn-Fe carbonate ore with typical Dal Matianite texture. (m) Disseminated stage 3 sphalerite and pyrite hosted in stage 3 quartz and calcite, and stage 3 quartz and calcite cut the slate. (n) Stage 3 brecciated sphalerite within stage 3 quartz-calcite. (o) Stage 3 sphalerite-galena veins cross-cut by stage 6 quartz-calcite veins. (p) Stage 3 sphalerite cross-cut by stage 4 quartz-boulangerite veins. (q) Stage 4 massive and needle-like boulangerite hosted by stage 4 quartz. (r) Stage 4 boulangerite-quartz. (s) Stage 5 elongate stibnite hosted by stage 5 quartz. (t) Stage 5 stibnite-cinnabar hosted by stage 5 quartz. (u) Siliceous sinter formed during the Supergene stage. (v) Ferrihydrite and sardinianite formed during the Supergene stage. (w) Stage 2 pyrite containing automorphic stage 1 arsenopyrite is replaced by later stage 2 sphalerite to form a skeletal texture. (x) Stage 3 sphalerite occurs in stage 3 quartz. (y) The emulsion-like and disseminated stage 3 chalcopyrite grains are dotted in the stage 3 sphalerite. (z) The stage 3 chalcopyrite grains are dotted among the grains of stage 3 pyrite. (aa) Stage 3 sphalerite replaced by stage 4 boulangerite that is in turn cross-cut by stage 5 stibnite. (ab) Stage 5 stibnite cross-cut by stage 6 quartz. Mcar1 = stage 1 fine-grained Mn-Fe carbonate; Apy1 = stage 1 lamellar arsenopyrite; Py1 = stage 1 lamellar pyrite; Sp1 = stage 1 lamellar sphalerite; Mcar2 = stage 2 coarse-grained Mn-Fe carbonate; Apy2 = stage 2 arsenopyrite; Py2 = stage 2 pyrite; Sp2 = stage 2 sphalerite; Gn2 = stage 2 coarse-grained galena; Py3 = stage 3 pyrite; Sp3 = stage 3 sphalerite; Gn3 = stage 3 galena; Ccp3 = stage 3 chalcopyrite; Qtz3 = stage 3 quartz; Cal3 = stage 3 calcite; Blr4 = stage 4 boulangerite; Qtz4 = stage 4 quartz; Stb5 = stage 5 stibnite; Ci5 = stage 5 cinnabar; Qtz5 = stage 5 quartz; Cal6 = stage 6 calcite; Qtz6 = stage 6 quartz; Lm = Supergene stage ferrihydrite. standard relative to IRMM 3702 and obtained a 𝛿66 Zn = 0.03‰, which is within the error reported in Archer et al. [98]. Solutions were kept at 100 ppb Cu and 150 ppb Zn which generated 63 Cu = 7 V and 66 Zn = 4 V. One block of 30 ratios is reported and the average error for the standard compared to itself throughout the session is 0.05‰ 2𝜎. 4. Results 4.1. EPMA. All the EPMA data are given in Tables 1 and 2. The Mn-Fe carbonate contains 23.562∼31.806 wt% Fe and 27.514∼ 32.232 wt% Mn, with a negative correlation between Fe and Mn contents (Figure 7(a)), which indicates that the Mn-Fe carbonates form by the isomorphic substitution of Fe2+ and Mn2+ ions and have a molecular formula of (Mn0.5 Fe0.5 )CO3 . The Fe contents are around 36 wt% for the arsenopyrite samples and 46 wt% for the pyrite samples. The sphalerite samples have 56.941∼60.552 wt% Zn and 5.375∼9.424 wt% Fe with a negative correlation between these two elements (Figure 7(b)). 4.2. Fe-Zn Isotopes. All the Fe-Zn isotopic data are given in Table 3. The annular polished section samples have 𝛿56 FeIRMM-014 of −1.95‰∼0.43‰, with an average of −0.50‰ ± 1.09‰ (2SD, 𝑛 = 19), and 𝛿66 ZnAA-ETH of −0.38‰∼0.07‰ with an average of −0.25‰ ± 0.19‰ (2SD, 𝑛 = 12). The Mn-Fe carbonate and pyrite show the 𝛿56 Fe values range from −1.95‰ to −0.59‰ (average value of −0.97‰ ± 0.86‰; 2SD, 𝑛 = 8) and from −0.26‰ to 0.23‰ (average value of −0.07‰ ± 0.35‰; 2SD, 𝑛 = 9), respectively. The two arsenopyrite samples exhibit the 𝛿56 Fe values of −1.01‰ (ZXK-1-3; Figure 5(c)) and −0.18‰ (ZXK-2-2; Figure 5(d)). In sample 9-3, the 9-3-1, 9-3-2 (Figure 6(a)), and 9-3-3 sphalerite have the 𝛿66 Zn values of −0.12‰, −0.23‰, and −0.31‰ with a gradually decreasing trend, the 9-3-4 and 9-35 Mn-Fe carbonate show the same trend with the 𝛿56 Fe values of −0.59‰ and −1.95‰, and the 𝛿56 Fe value of 9-3-7 pyrite is −0.26‰ (Figure 5(b)). Similarly, in sample 9-8, from 98-2 (−0.09‰; Figure 6(c)) through 9-8-5 (−0.23‰) to 9-83 (−0.35‰; Figure 6(d)), the 𝛿66 Zn values of sphalerite also present a gradually decreasing trend; meanwhile the 𝛿56 Fe values also decrease from −1.06‰ (9-8-8; Figure 6(b)) to −1.38‰ (9-8-9) for Mn-Fe carbonate and from 0.23‰ (9-84) to 0.09‰ (9-8-7) for pyrite, although the 9-8-1 sphalerite and 9-8-6 pyrite have the 𝛿66 Zn value of −0.17‰ and 𝛿56 Fe values of −0.26‰ (Figure 5(a)). In sample ZXK-1, the 𝛿66 Zn Geofluids Table 1: EPMA data for Mn-Fe carbonate and quartz from the Zhaxikang deposit. Sample number SiO2 (%) 9-3-4 0.015 9-3-5 0.024 9-8-8 0.002 9-8-9 – ZXK-1-6 0.001 ZXK-1-7 0.028 ZXK-2-7 0.005 ZXK-2-8 – ZXK-2-9 0.009 ZXK-2-10 0.015 ZXK-2-11 97.000 Na2 O (%) 0.005 0.109 0.031 0.025 0.028 0.011 0.039 0.051 0.049 – 0.026 K2 O (%) 0.007 0.027 0.007 – 0.001 0.019 – – 0.008 – 0.025 TiO2 (%) 0.011 – – – 0.022 0.036 0.055 0.012 – 0.023 – Al2 O3 (%) 0.003 0.010 – 0.008 – 0.014 – 0.010 0.037 0.012 0.219 MgO (%) 1.231 1.080 1.276 1.189 1.207 1.995 1.236 1.140 1.552 1.119 0.028 CaO (%) 2.061 2.151 1.978 1.144 5.178 2.940 5.542 4.298 1.946 4.849 0.018 FeO (%) 28.598 28.372 27.341 31.806 24.070 29.639 27.235 24.287 26.739 23.562 0.020 MnO (%) 29.840 30.582 31.338 27.514 30.717 27.943 27.604 32.694 30.968 32.232 — SrO (%) 0.007 – 0.010 – – – – – 0.007 – – BaO (%) 0.057 0.002 0.022 – – 0.017 – 0.008 0.013 0.003 – CO2 (%) 38.164 37.643 37.993 38.316 38.775 37.358 38.285 37.500 38.673 38.185 2.664 Total (%) 99.999 100.000 99.998 100.002 99.999 100.000 100.001 100.000 100.001 100.000 100.000 Mineral Mcar Mcar Mcar Mcar Mcar Mcar Mcar Mcar Mcar Mcar Qtz Notes. (1) Mcar = Mn-Fe carbonate; Qtz = quartz; (2) for the sampling points, please see Figure 5. 9 10 Table 2: EPMA data for sulfides from the Zhaxikang deposit. Sample number As (%) 9-3-1 – 9-3-2 – 9-3-3 – 9-3-6 0.784 9-3-7 2.244 9-8-1 0.002 9-8-2 – 9-8-3 – 9-8-5 0.009 9-8-6 0.006 9-8-7 0.015 ZXK-1-1 – ZXK-1-2 – ZXK-1-3(1) 39.757 ZXK-1-3(2) 2.435 ZXK-1-3(3) – ZXK-1-4 1.942 ZXK-1-5 2.056 ZXK-2-1 – ZXK-2-2 39.643 ZXK-2-3 – ZXK-2-4 39.444 ZXK-2-5 2.417 ZXK-2-6 0.031 S (%) 33.165 33.423 33.262 52.884 51.934 33.740 33.649 33.500 34.064 53.867 53.359 33.502 33.428 24.131 51.666 33.337 51.982 52.248 33.684 23.789 33.868 24.229 51.755 53.630 Se (%) – – – – – – – – – – – – – – – 0.002 0.001 – – – – – – – Cu (%) 0.133 0.015 – 0.005 0.002 0.009 0.011 0.006 0.089 0.033 0.471 0.012 0.036 0.006 0.018 0.014 0.008 0.008 0.061 0.002 0.025 – – 0.107 Co (%) 0.009 0.012 0.015 0.044 0.049 0.010 0.005 0.003 0.004 0.050 0.043 0.009 – 0.043 0.067 0.012 0.046 0.054 0.010 0.050 0.012 0.041 0.045 0.047 Zn (%) 59.095 58.224 57.114 – – 57.705 59.584 56.941 57.475 0.058 0.044 57.497 60.190 0.009 0.185 58.623 0.015 – 59.726 0.008 58.034 – 0.027 – Sb (%) – – – – – – – – – – – – 0.002 – – – – – – 0.196 – 0.173 – – Ag (%) – 0.002 – – – 0.004 0.005 – – 0.005 – – 0.002 – 0.007 0.015 – – – 0.006 0.001 – – – Au (%) – – – 0.016 0.004 – – – – 0.012 – – – 0.027 0.013 – 0.003 0.013 – – – 0.026 – 0.009 Cd (%) 0.171 0.151 0.159 – – 0.160 0.150 0.160 0.151 – – 0.172 0.201 – – 0.147 – – 0.170 – 0.196 – – – Bi (%) – – – – – – – – – – – – – – – – – – – – – – – – Pb (%) – – – 0.134 0.132 0.018 – – – 0.300 0.174 – – 0.063 0.113 – 0.196 0.137 – 0.081 – 0.061 0.131 0.168 NiO (%) – – – – – 0.008 – 0.002 – 0.002 – – – 0.002 0.004 0.003 – – 0.005 0.098 0.002 – – – FeO (%) 6.985 8.060 9.405 46.585 46.044 8.827 7.112 9.424 8.901 46.539 46.518 9.142 5.810 36.340 45.877 7.747 46.004 46.163 6.904 36.245 8.714 36.424 46.227 46.654 Total (%) 99.558 99.887 99.955 100.452 100.409 100.483 100.516 100.036 100.693 100.872 100.624 100.334 99.669 100.378 100.385 99.900 100.197 100.679 100.560 100.118 100.852 100.398 100.602 100.646 Mineral Sp Sp Sp Py Py Sp Sp Sp Sp Py Py Sp Sp Apy Py Sp Py Py Sp Apy Sp Apy Py Py Notes. (1) Py = pyrite, Sp = sphalerite, and Apy = arsenopyrite; (2) for the sampling points, please see Figure 5. Geofluids Geofluids 11 56 Fe values –0.26%₀ 0.23‰ 0.09‰ Pyrite 66 9-8  Zn values Gn 3 –0.35%₀ 5 Sp Sp –0.23%₀ 4 2 –0.09%₀ 1 –0.17%₀ 6 Slate Sphalerite Fe contents 9.424% 8.901% Zn contents 56.941% 57.475% 7.112% 8.827% 59.584% 57.705% 8 Sphalerite 66 Zn values −0.31%₀ Zn Fe contents contents 57.114% 9.405% 2 −0.23%₀ 58.224% 8.060% 1 −0.12%₀ 59.095% 6.985% 9-3 7 56 Fe values Mn-Fe Carbonate –0.78%₀ –0.92%₀ 9 Mcar 56 Fe values –0.26%₀ Pyrite Py 7 3 Mcar 56  Fe values Mn-Fe Carbonate −1.95%₀ 6 5 Py Sp 4 −0.59%₀ 3 cm 3 cm (a) (b) −0.17%₀ −0.07%₀ 56 Fe values Pyrite 56  Fe values −0.05%₀ −0.22%₀ 0.12‰ Pyrite ZXK-2 ZXK-1 Sp 3 11 9 6 56 56  Fe values Arsenopyrite −1.01%₀ 5 Py-Sp-Apy 3 2 Sp  Fe values Mn-Fe Carbonate −0.78%₀ 7 1 4 −0.66%₀ 6 Mcar Py Mcar 56 Fe values Arsenopyrite −0.18%₀ 2 5 Apy 1 Py 8 56 Fe values Mn-Fe Carbonate −1.04%₀ −1.06%₀ 7 3 cm 3 cm (c) 4 Sp 10 Gn Py −0.15%₀ −0.18%₀ −0.38%₀ 66 Zn values 60.190% 58.623% 57.497% Zn contents Sphalerite 5.810% 7.747% 9.142% Fe contents Qtz −0.28%₀ 58.034% 8.714% −0.32%₀ 66 Zn values 59.726% Zn contents Fe contents 6.904% Sphalerite (d) Figure 5: The sampling points and Fe-Zn isotopic-elemental variations in the annular polished section samples: (a) 9-8; (b) 9-3; (c) ZXK-1; (d) ZXK-2. Mcar = Mn-Fe carbonate; Apy = arsenopyrite; Py = pyrite; Sp = sphalerite; Qtz = quartz. values of ZXK-1-3 (−0.18‰; Figures 6(f) and 6(m)) and ZXK1-2 (−0.15‰; Figure 6(g)) sphalerite are almost the same and heavier than ZXK-1-1 (−0.38‰) sphalerite; the 𝛿56 Fe values for pyrite increase from ZXK-1-3 (−0.22‰) to ZXK1-4 (0.12‰; Figure 6(e)) and ZXK-1-5 (−0.05‰); and, as for Mn-Fe carbonate, the ZXK-1-6 (−0.66‰) has heavier 𝛿56 Fe values than ZXK-1-7 (−0.78‰; Figure 5(c)). By contrast, in sample ZXK-2, the ZXK-2-1 (−0.32‰; Figure 6(h)) and ZXK2-3 (−0.28‰; Figures 6(k) and 6(p)) sphalerite have almost the same 𝛿66 Zn values; the ZXK-2-7 (−1.06‰) and ZXK-28 (−1.04‰) Mn-Fe carbonate also yield approximate 𝛿56 Fe values; also the ZXK-2-5 (Figures 6(i) and 6(n)) and ZXK-26 pyrite show the similar 𝛿56 Fe values of −0.07‰ and −0.17‰ (Figure 5(d)). 5. Discussion 5.1. The Fe-Zn Isotopic and Elemental Variations. Sample 93 with typical concentric annular texture has the gradually decreasing 𝛿66 Zn values of sphalerite and 𝛿56 Fe values of MnFe carbonate from core to edge (early to late; Figure 5(b)). Similarly, sample 9-8 shows the same isotopic variation trend of pyrite, sphalerite, and Mn-Fe carbonate except the sulfides in the core (Figure 5(a)). As the core in 9-8 consists of crushed, earlier formed slate breccias and sulfides, the annular sulfides and Mn-Fe carbonate formed around the core, and thus this core is not included in the decreasing trend. In sample ZXK-1, as the laminae (ZXK-1-3; Figure 5(c)) is regarded as the earliest, the Fe-Zn isotopic values of sphalerite and Mn-Fe carbonate still present gradually decreasing trend from ZXK-1-3 (−0.18‰) to ZXK-1-1 (−0.38‰) and ZXK-1-6 (−0.66‰) to ZXK-1-7 (−0.78‰), respectively (Figure 5(c)), whereas the 𝛿56 Fe values of pyrite show the different variation, which might be related to the influence of Fe isotopic fractionation within pyrite-arsenopyrite-sphalerite mineral pair (Figures 6(f) and 6(m)). However, there is no similar decreasing Fe-Zn isotopic variation trend in sample ZXK-2, and the modification by the stage 3 quartz vein must be the main cause (Figures 5(d), 6(h)–6(l) and 6(n)–6(p)). Previous studies [12, 62, 99] have proposed a Rayleigh distillation model to explain an increasing trend in 𝛿66 Zn values within precipitates over time for the hydrothermal fluid. This Rayleigh distillation model is as follows: the oreforming materials derived from a single source would be subjected to kinetic Rayleigh fractionation that would lead to the early formed mineral precipitants being preferentially enriched in light isotopes, as well as residual fluids and later precipitants with heavier isotopic values, causing an increasing trend in isotopic values within precipitants over time. Several previous studies have used this distillation model to explain the Zn isotopic variation within different types of deposits (e.g., VHMS: [18]; Irish-type: [19, 100]; SEDEX: [20, 53]). Likewise, this Rayleigh distillation model is also applicable to the Fe isotopic variation in skarn deposits [21, 22]. However, the 𝛿66 Zn and 𝛿56 Fe values gradually decrease from early to late stages within Zhaxikang deposit, which cannot be explained by this distillation model. Another Rayleigh distillation mechanism models this 12 Geofluids Mcar Mcar Mcar Slate Sp Sp Sp Py Sp Sp 500 G 500 G (a) 200 G (b) 200 G (c) (d) Qtz Mcar Mcar Mcar Sp Sp Sp Sp-Py-Apy Py 500 G 500 G (e) 500 G 500 G (f) (g) (h) Mcar Mcar Mcar Qtz Qtz Sp Qtz Qtz Apy Py Apy Apy 500 G 500 G Mcar (i) 500 G 500 G (j) (k) (l) Qtz Sp Apy Qtz Py Mcar Sp Apy Py Py Apy Mcar Gn Mcar Mcar (m) (n) (o) Qtz Sp (p) Figure 6: The photomicrographs and electron probe micrographs of sampling area in the annular polished section samples for EPMA and Fe-Zn isotopic analyses. (a) 9-3-2; (b) 9-8-8; (c) 9-8-2; (d) 9-8-3 and 9-8-5; (e) ZXK-1-4; (f) ZXK-1-3; (g) ZXK-1-2; (h) ZXK-2-1; (i) ZXK-2-5; (j) ZXK-2-2; (k) ZXK-2-3; (l) ZXK-2-4; (m) ZXK-1-3; (n) ZXK-2-5; (o) ZXK-2-2; (p) ZXK-2-3. Abbreviations are as in Figure 5. decreasing trend: the metallogenic elements are transported by the ore-forming system consisting of vapour and liquid phases, and there is partitioning between vapour-liquid phases and the ratios change with the temperature decreasing. Then the minerals precipitate from the liquid phase of the ore-forming system. During this period, the vapour-liquid partitioning and mineral precipitation cause the Rayleigh fractionation, and this Rayleigh fractionation leads to the mineral precipitation being preferentially enriched in heavy isotopes relative to the ore-forming system. Thus, the isotopic values of subsequent minerals are lighter and lighter [101, 102]. This Rayleigh distillation model is supported by the following evidence: (1) The vapour-liquid partitioning and related isotopic fractionation for transition metal elements (e.g., Cu and Mo) have been confirmed by previous research in the Dahutang W-Cu-Mo ore field [103]; (2) minerals typically precipitate from the liquid phase; however, according to the previous literature [58], in unique cases the vapour phase containing metal can even directly condensate to form solid phases from high-temperature ore-forming system (e.g., VMS and volcano related system), which can demonstrate the existence of the vapour phase for metal elements. As for the Zhaxikang deposit, the theoretical calculated oreforming temperature from Fe isotopic data is 500∼800∘ C [26], and thus there should be a transitory high-temperature period, making the vapour-liquid partitioning possible; (3) the fluid inclusion data demonstrate that Mn-Fe carbonates and sulfides exist in three types of inclusions: A gas-liquid two-phase water inclusions (W type, more than 90%), B pure 13 34 61 32 60 Zn content (%) Mn content (%) Geofluids 30 y = −0.535x + 44.583 28 R2 = 0.5242 y = −0.909x + 65.701 R2 = 0.9624 58 57 26 24 59 56 26 28 30 Fe content (%) 32 34 5 6 7 (a) 10 (b) 0.05 0.05 0.00 0.00 −0.05 −0.10 66 Zn (% ₀) 66 Zn (% ₀) 9 Sphalerite Mn-Fe carbonate −0.15 −0.20 y = −0.0633x + 0.2947 −0.25 R2 = 0.6149 −0.30 −0.35 −0.40 5.00 8 Fe content (%) 6.00 7.00 8.00 Fe content (%) 9.00 10.00 Sphalerite −0.05 −0.10 −0.15 −0.20 −0.25 −0.30 −0.35 −0.40 56.00 y = 0.072x − 4.4143 R2 = 0.6496 57.00 58.00 59.00 Zn content (%) 60.00 61.00 Sphalerite (c) (d) Figure 7: (a) The negative correlation between Mn and Fe contents within Mn-Fe carbonate. (b) The negative correlation between Zn and Fe contents within sphalerite. (c) The negative correlation between 𝛿66 Zn values and Fe contents within sphalerite from 9-3, 9-8, and ZXK-1. (d) The positive correlation between 𝛿66 Zn values and Zn contents within sphalerite from 9-3, 9-8, and ZXK-1. liquid inclusions (L type), and C pure CO2 inclusions (PG type) [104] from Zhaxikang deposit. On the other hand, for sphalerite, there are positive correlations between Zn contents and 𝛿66 Zn values, with negative correlations between Fe contents and 𝛿66 Zn values in samples 9-8 (Figure 5(a)), 9-3 (Figure 5(b)), and ZXK-1 (Figure 5(c)), respectively. Moreover, plotting all the data from these three samples on a diagram together, the correlations are also good with 𝑅2 > 0.6 (Figures 7(c) and 7(d)). Zinc and iron usually show similar geochemical behaviour as both of them are highly mobile in chloride-bearing hydrothermal fluids [23, 100]. Thus, the Zn2+ and Fe2+ ions are preferentially enriched in the liquid phase relative to the vapour phase before precipitation [23, 100], which cause the decreasing Zn contents of sphalerite over time. As the total content of Zn2+ and Fe2+ is constant in sphalerite, the Fe contents of sphalerite gradually increase with the decreasing Zn contents. These correlations further support the hypothesis that the oreforming system is the mixture of vapour and liquid phases. In addition, the sample ZXK-2 cut by a later stage 3 quartz vein does not present the same correlations and variations, which is a new evidence for two pulses of mineralization proposed by Zheng et al. [7] and Wang et al. [26]. All of the evidence above reveals that the ore-forming elements are transported by the ore-forming system that consists of vapour and liquid phases. The vapour-liquid partitioning and mineral precipitation are the main cause of Fe-Zn isotopic and elemental variations. Afterwards, the overprint by the second pulse of mineralization has also partly modified the Fe-Zn isotopic and elemental compositions of some earlier samples (Figure 5(d)) [26]. 5.2. The Fe-Zn Isotopic Fractionation Models for the OreForming System. In order to verify the Rayleigh distillation model in Section 5.1 and obtain more information of the oreforming system, we use the following equations to establish Fe-Zn isotopic fractionation models for the ore-forming system and mineral precipitation: 𝛿Minerals (‰) = 𝛼 × (𝛿𝑖 + 1000) × 𝐹(𝛼𝑚 −1) 𝛼𝑖 − 1000, 1 × (𝛿𝑖 + 1000) × 𝐹(𝛼𝑚 −1) 𝛿Ore-forming System (‰) = 𝛼𝑖 − 1000. (1) 14 Geofluids Table 3: Fe-Zn isotopic data for annular polished section samples from the Zhaxikang deposit. Sample number 9-3-1 9-3-2 9-3-3 9-3-4 9-3-5 9-3-7 9-8-1 9-8-2 9-8-3 9-8-4 9-8-5 9-8-6 9-8-7 9-8-8 9-8-9 ZXK-1-1 ZXK-1-2 ZXK-1-3(1) ZXK-1-3(2) ZXK-1-3(3) ZXK-1-4 ZXK-1-5 ZXK-1-6 ZXK-1-7 ZXK-2-1 ZXK-2-2 ZXK-2-3 ZXK-2-5 ZXK-2-6 ZXK-2-7 ZXK-2-8 Mineral Sp Sp Sp Mcar Mcar Py Sp Sp Sp Py Sp Py Py Mcar Mcar Sp Sp Apy Py Sp Py Py Mcar Mcar Sp Apy Sp Py Py Mcar Mcar 𝛿56 FeIRMM–014 2𝜎 −0.59 −1.95 −0.26 0.08 0.08 0.08 0.23 0.08 −0.26 0.09 −0.78 −0.92 0.08 0.08 0.08 0.08 −1.01 −0.22 0.08 0.08 0.12 −0.05 −0.66 −0.78 0.08 0.08 0.09 0.08 −0.18 0.08 −0.07 −0.17 −1.06 −1.04 𝛿66 ZnAA–ETH −0.12 −0.23 −0.31 2𝜎 0.05 0.05 0.05 −0.17 −0.09 −0.35 0.05 0.05 0.05 −0.23 0.05 −0.38 −0.15 0.05 0.05 −0.18 0.05 −0.32 0.05 −0.28 0.05 0.08 0.08 0.08 0.08 Notes. (1) Abbreviations and sampling points are as in Tables 1 and 2; (2) 2𝜎 is two times the standard deviation. 𝛿Minerals , 𝛿Ore-forming System , and 𝛿𝑖 are the 𝛿56 Fe-𝛿66 Zn values of momentary mineral precipitation and momentary and initial ore-forming system, respectively; 𝛼, 𝛼𝑖 , and 𝛼𝑚 are the isotopic fractionation factors between ore-forming system and mineral precipitation that refer to the momentary condensation temperature 𝑇, the initial temperature of oreforming system 𝑇𝑖 , and (𝑇 + 𝑇𝑖 )/2, respectively; and 𝐹 is the fraction of remaining ore-forming system that consists of vapour and liquid phases [102, 105]. In addition, the equations for fractionation factors are approximately using ln 𝛼Fe = 0.4432 × 106 /𝑇2 for Fe [106] and ln 𝛼Zn = 0.2853 × 106 /𝑇2 + 0.0535 for Zn (𝑇 is absolute temperature in K) [107]. Wang et al. [26] have calculated the ore-forming temperature (500∼800∘ C) of the first pulse of mineralization in Zhaxikang deposit using the Fe isotopic fractionation between pyrite and Mn-Fe carbonate. Although this temperature range is a little wide, the highest temperature of such ore-forming system can reach around 500∘ C according to the previous ore deposit studies [6, 22, 23]; hence we regard 500∘ C as the initial temperature of the ore-forming system. The homogenization temperature (240∘ C) of the fluid inclusions [104] from the first pulse of mineralization in Zhaxikang deposit is considered as the momentary condensation temperature. Furthermore, for the purpose of making the fractionation models more comprehensive and exact, we also quote the Fe-Zn isotopic data of pyrite (𝛿56 Fe: stage 1: −0.33‰ to −0.09‰; stage 2: −0.30‰ to 0.19‰; stage 3: 0.16‰ to 0.43‰), sphalerite (𝛿66 Zn: −0.31‰ to 0.07‰), and Mn-Fe carbonate (𝛿56 Fe: −0.80‰ to −0.55‰; 𝛿66 Zn: −0.11‰ to 0.04‰) from Wang et al. [26], as well as the 𝛿66 Zn values of sphalerite (−0.25‰ to 0.03‰) and Mn-Fe carbonate (−0.01‰) from Duan et al. [5]. Finally, we set up 12 FeZn isotopic fractionation models for pyrite, sphalerite, and Mn-Fe carbonate (Figure 8) with different 𝛿56 Fe𝑖 values of 0‰ (mean value of magma) [51], −0.5‰, and −1‰, as well as 𝛿66 Zn𝑖 values of −0.28‰ (the lightest value of seafloor hydrothermal fluids) [62], 0‰ (mean value of bulk earth) 48.8% Ore-forming system F 1.00 Pyrite Ore-forming system −1.00 −2.00 48.7% 26.7% 0.07‰ −0.38%₀ Sphalerite Ore-forming system −3.00 −4.00 2.00 1.00 0.00 −1.00 −2.00 −3.00 −4.00 66 :Hi = 0.23%₀ Ore-forming system F F −1.95%₀ 7.1% Ore-forming system 10.9% 56 &?i = −0.5%₀ Mn-Fe carbonate −0.55%₀ 39.7% −1.95%₀ Ore-forming system F 1.00 0.00 −1.00 −2.00 −3.00 −4.00 −5.00 −6.00 −7.00 56 &?i = −1%₀ 63.3% Mn-Fe carbonate 17.3% Ore-forming system F F (i) (h) 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 Ore-forming system −1.00 66 :Hi = 0%₀ 46.8% 38.2% Mn-Fe carbonate 0.04‰ −0.11%₀ −2.00 −3.00 Ore-forming system −4.00 F (j) 2.00 1.00 0.00 −1.00 −2.00 −3.00 −4.00 66 :Hi = 0.23%₀ 34.8% Ore-forming system F (k) Mn-Fe carbonate 0.04‰ −0.11%₀ 28.5% 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 Mn-Fe carbonate 0.00 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 55.6% 1.00 0.04‰ −0.11%₀ 66 Zn (%₀ ) 66 :Hi = −0.28%₀ 68.1% −0.55%₀ −1.95%₀ 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 −0.55%₀ 1.00 0.00 −1.00 −2.00 −3.00 −4.00 −5.00 −6.00 −7.00 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 24.9% 56 Fe (%₀ ) Mn-Fe carbonate (f) 56 Fe (%₀ ) (e) 56 &?i = 0%₀ 0.07‰ −0.38%₀ 36.3% 19.8% Sphalerite 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 Ore-forming system 66 :Hi = 0%₀ 0.00 0.43‰ −0.33%₀ F 66 Zn (%₀ ) Sphalerite 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 56 Fe (%₀ ) 77.7% (c) 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 −0.38%₀ 38.7% 66 Zn (%₀ ) 0.07‰ 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 66 Zn (%₀ ) 66 :Hi = −0.28%₀ 70.9% (g) 66 Zn (%₀ ) 56 &?i = −1%₀ F F 0.50 0.00 −0.50 −1.00 −1.50 −2.00 −2.50 −3.00 1.00 0.00 −1.00 −2.00 −3.00 −4.00 −5.00 −6.00 −7.00 (b) (d) 2.00 1.00 0.00 −1.00 −2.00 −3.00 −4.00 −5.00 −6.00 0.43‰ −0.33%₀ Pyrite (a) 0.50 0.00 −0.50 −1.00 −1.50 −2.00 −2.50 −3.00 99.0% 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 Ore-forming system 56 &?i = −0.5%₀ 56 Fe (%₀ ) 30.1% Pyrite 1.00 0.00 −1.00 −2.00 −3.00 −4.00 −5.00 −6.00 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 62.5% 0.43‰ −0.33%₀ 56 Fe (%₀ ) 56 &?i = 0%₀ 66 Zn (%₀ ) 2.00 1.00 0.00 −1.00 −2.00 −3.00 −4.00 −5.00 −6.00 15 0.01 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 0.99 56 Fe (%₀ ) Geofluids F (l) Figure 8: The Fe-Zn isotopic fractionation models for the ore-forming system and minerals (pyrite, sphalerite, and Mn-Fe carbonate) with different 𝛿56 Fe𝑖 and 𝛿66 Zn𝑖 values. [54], and 0.23‰ (the mean value of deep sea water) [63, 108, 109]. These fractionation models show that the 𝐹 ranges for ore-forming system highly depend on the 𝛿𝑖 values (Figure 8). The details are as follows: the pyrite covers the 𝐹 ranges of 30.1%∼62.5% (𝛿56 Fe𝑖 = 0‰; Figure 8(a)), 48.8%∼99.0% (𝛿56 Fe𝑖 = −0.5‰; Figure 8(b)), and more than 77.7% (𝛿56 Fe𝑖 = −1‰; Figure 8(c)). The 𝐹 ranges for sphalerite are 38.7%∼70.9% (𝛿66 Zn𝑖 = −0.28‰; Figure 8(d)), 26.7%∼48.7% (𝛿66 Zn𝑖 = 0‰; Figure 8(e)), and 19.8%∼36.3% (𝛿66 Zn𝑖 = 0.23‰; Figure 8(f)), respectively. In comparison, the Mn-Fe carbonates has the 𝐹 ranges of 7.1%∼24.9% (𝛿56 Fe𝑖 = 0‰; Figure 8(g)), 10.9%∼39.7% (𝛿56 Fe𝑖 = −0.5‰; Figure 8(h)), and 17.3%∼63.3% (𝛿56 Fe𝑖 = −1‰; Figure 8(i)) for Fe isotope, as well as 55.6%∼68.1% (𝛿66 Zn𝑖 = −0.28‰; Figure 8(j)), 38.2%∼46.8% (𝛿66 Zn𝑖 = 0‰; Figure 8(k)), and 28.5%∼34.8% (𝛿66 Zn𝑖 = 0.23‰; Figure 8(l)) for Zn isotope. All of these results suggest that the Fe-Zn isotopic data of Zhaxikang deposit fit these Rayleigh fractionation models well. However, we need to consider the following facts in the Zhaxikang deposit: (1) the second pulse of mineralization has brought some Fe to form the stage 3 pyrite with heavier 𝛿56 Fe values (0.23‰∼0.43‰) [26]; thus the stage 3 pyrite does not fit the fractionation models; (2) as most of the sphalerite and pyrite are paragenetic during the first pulse of mineralization (Figures 4(g), 4(i), 4(k), 5(a)–5(c) and 6(c)), especially in the earliest lamina (Figures 4(a)–4(c), 5(c), 6(f) and 6(m)), the sphalerite and pyrite should overlap on 𝐹 16 ranges; (3) in theory, the Mn-Fe carbonates would have the same 𝐹 range for Fe-Zn isotopes. Nonetheless, in view of the tight Zn isotopic variation range, the 𝐹 range for Fe isotope should cover that for Zn isotope; (4) during the earlier period, as the ore-forming system consisting of vapour and liquid phases is dominant, 𝐹 values in fractionation models should be large. Taking all these facts and 12 fractionation models into consideration, the 𝛿56 Fe𝑖 value is supposed to be in the range of −0.5‰∼−1‰, and the 𝛿66 Zn𝑖 value should be between −0.28‰ and 0‰. 5.3. Implications for the Genesis of Zhaxikang Deposit 5.3.1. Excluding the Possibility of Hot Spring Genetic Model. The hot spring model predicts that metals (e.g., Zn, Pb, Sb, Ag, and Fe) are leached from sedimentary wall rocks, which is supported by the following evidence: (1) the 𝛿34 S values of the sulfides (4.5‰∼12‰) are similar to those of sedimentary wall rocks (4.93‰∼11.49‰); (2) the 𝛿30 Si values of quartz (−0.90‰∼−0.40‰) are the same as those of siliceous rocks with hot spring genesis; (3) the 𝛿DV–SMOW (−135‰∼−127‰) and 𝛿18 OH2 O values (−13.7‰∼12.4‰) of fluid inclusions trapped in quartz are similar to those of the south Tibetan hot spring; (4) the Pb isotopes (206 Pb/204 Pb: 18.474∼19.637; 207 Pb/204 Pb: 15.649∼ 15.774; 208 Pb/204 Pb: 39.660∼40.010) show the characteristics of radiogenic Pb; (5) the He-Ar isotopes demonstrate the contribution of crustal fluid and meteoric water [3, 4]. However, this genetic model is inconsistent with textural and Fe-Zn isotopic evidence presented here. Firstly, the primary sedimentary wall rocks in the orefield are slate. As suggested by the continuous batch experimental research of Fernandez and Borrok [27], the ore-forming fluid would preferentially leach out the heavy Zn isotopes. Likewise, Chen et al. [110] have analyzed the Zn isotopic compositions of samples from 8 hot springs, and the results show most of the hot springs have relatively constant and heavier 𝛿66 Zn values (approximately 0.42‰) than host rocks (−0.42‰ to 0.14‰). Therefore, if the metallogenic elements are leached from the slate by hot spring, there would be some sphalerite with heavier 𝛿66 Zn values than these slates in Zhaxikang deposit. Nevertheless, the 𝛿66 Zn values of the slate from Zhaxikang orefield range from −0.23‰ to 0.10‰ that are similar to sphalerite, especially the unmodified slate sample with the 𝛿66 Zn value of 0.10‰ that is even a little heavier than that of sphalerite (−0.38‰ to 0.07‰; Figure 9(b)) [26]. As the slate has heavier Zn isotopic compositions than those of host rocks from Chen et al. [110], the 𝛿66 Zn values of the hot spring in Zhaxikang orefield would even be heavier than 0.42‰, which are much heavier than the 𝛿66 Zn𝑖 values (−0.28‰∼0‰) of the fractionation models in Section 5.2. Secondly, Sharam et al. [46] have measured the 𝛿56 Fe values (−0.59‰ to −0.12‰) of hot springs in Juan de Fuca Ridge, which is much heavier than the 𝛿56 Fe𝑖 value (−0.5‰∼−1‰) gained from the fractionation models. Moreover, the marine fluids usually have lighter Fe isotopic compositions (Figure 9(a)); thus the hot springs in Tibet would have heavier 𝛿56 Fe values than those of Sharam et al. [46]. Geofluids The possibility of hot spring genesis for the first pulse of mineralization event can be excluded by Fe-Zn isotopic data. Evidence for the second pulse of mineralization is based on the fact that the Fe-Zn isotope values do not follow similar concentric patterns with the ore textures as seen in sample ZXK-2. Meanwhile, the evidence from Si-H-O isotopes demonstrates that the second pulse of mineralization may be related to hot spring. Additionally, the Fe-Zn isotopic data demonstrate that the sedimentary wall rocks have not provided significant amounts of metals, although the S-Pb isotopic data show that these wall rocks constitute some contribution for S-Pb [4, 5], whereas the similar Zn isotopic compositions of slate and sphalerite suggest that they share the same Zn origin. 5.3.2. Inconsistency with the Magmatic-Hydrothermal Fluid Genetic Models. There are two genetic models for the magmatic-hydrothermal fluid genesis. In the first model, Duan et al. [5] considered that the genesis of Zhaxikang deposit relates to the mid-low temperature magma-related hydrothermal activity and that the metallogenic elements are mainly sourced from the mixing of basement and the sedimentary wall rocks. The evidence is mainly from the ZnS-Pb isotopes that we mentioned in Section 1. Duan et al. [5] have analyzed the 𝛿66 Zn values of the sulfides (−0.25‰∼0.03‰) and basement rocks (0.05‰∼ 0.21‰). The dominating sedimentary wall rocks in the orefield are slate (𝛿66 Zn values: −0.23‰∼0.10‰) [26], which is formed by the epimetamorphism of shale and sandstone. Meanwhile, combining the data from Wang et al. [26] with this study, the sphalerite should have a range from −0.38‰ to 0.07‰ in Zhaxikang deposit. Both the basement and sedimentary wall rocks have heavier Zn isotopic compositions than sphalerite in Zhaxikang deposit. However, just like we discussed in Section 5.3.1, if the Zn is sourced from mixing of the basement and sedimentary wall rocks, there should be some sphalerite with heavier 𝛿66 Zn values than these rocks. This inference is also evidenced by the research of Zhou et al. [64]: the Paleozoic carbonate host rocks and Precambrian basements are considered to be the origin of metals, and these rocks have lighter 𝛿66 Zn values (−0.52‰ to 0.16‰) than the sphalerite from the Tianqiao (−0.54‰ to 0.30‰) and Bangbangqiao (−0.21‰ to 0.43‰) deposits in the Sichuan-Yunnan-Guizhou Pb-Zn metallogenic province (Figure 9(b)). Additionally, in respect of the Fe isotope, the Schwarzwald hydrothermal vein deposit in Germany can be used as an analogy [52]. Iron in this deposit originates from the basement consisting of granites and gneisses, as well as sedimentary rocks including shale and sandstone. The basement of Zhaxikang is composed of dolerite, quartz diorite, rhyolite porphyry, pyroclastics, and porphyritic monzogranite [5], and these crust-derived igneous rocks have the similar Fe isotopic composition with granites according to the data from previous studies [21–23, 28, 36–45]. Likewise, the 𝛿56 Fe values of shale and sandstone in the Schwarzwald deposit are −0.21‰, 0.03‰, and 0.22‰, all of which fall into the Fe isotopic variation range of shale (−0.39‰∼0.71‰) from Beard et al. [34] and Rouxel et al. [35]. Therefore, we Geofluids 17 56 F?i value −1%₀∼−0.5%₀ 66 ZHi value −0.28%₀∼0%₀ Earth 0.03 ± 0.09%₀ Earth 0 ± 0.05%₀ Deep sea water: 0.23‰ Zhaxikang Sb-Pb-Zn-Ag deposit Zhaxikang Sb-Pb-Zn-Ag deposit Dongshengmiao SEDEX-type deposit, China Dongshengmiao SEDEX-type deposit, China Red Dog SEDEX-type ore district, America Renison Sn-W deposit, Australia Alexandrinka VHMS-type deposit, Russia Schwarzwald hydrothermal vein deposit, Germany Cévennes MVT-type deposit, France Irish-type deposit, Ireland Fenghuangshan Skarn-type deposits in Tongling ore district Dongguashan China Xinqiao Basements and ore-hosted rocks Bangbangqiao Tianqiao Iron ores in Bayan Obo Fe-REE magmatic deposit,China Carbonated-hosted Pb-Zn Sulfide Deposit, Southwest China Skarn-type deposits in Tongling ore district, China Gorno and Raibl magmatic deposits, Italy Midoceanic ridges pyrite Seafloor hydrothermal fluid system Seafloor hydrothermal fluid system Deep-sea carbonate Igneous rock Igneous rock Shale –2.5 –2.0 –1.5 –1.0 –0.5 0.0 0.5 1.0 Sedimentary rock 1.5 2.0 –0.8 –0.6 –0.4 –0.2 Pyrrhotite Siderite Chalcopyrite Hematite 0.2 0.4 0.6 0.8 1.0 1.2  Zn (%₀ )  Fe (%₀ ) Pyrite Mn-Fe carbonate Arsenopyrite Sphalerite 0.0 66 56 Magnetite Bornite (a) Mn-Fe carbonate Sphalerite Slate (b) Figure 9: (a) Fe isotopic compositions of pyrite, Mn-Fe carbonate, and arsenopyrite in the Zhaxikang deposit (part of the data is quoted from Wang et al. [26]). Other Fe isotopic data for the Bulk Silicate Earth [10], shale [34, 35], igneous rocks [21–23, 28, 36–45], seafloor hydrothermal fluid system [34, 46–49], midoceanic ridges pyrite [48, 50], the Bayan Obo Fe-REE magmatic-type deposit in China [51], the skarn-type deposits in Tongling ore district in China [21, 22], the Schwarzwald hydrothermal vein deposit in Germany [52], the Renison Sn-W deposit in Australia [23], and the Dongshengmiao SEDEX-type deposit in China [53] are also plotted for comparison. (b) Zn isotopic compositions of sphalerite, Mn-Fe carbonate, and slate in the Zhaxikang deposit (part of the data is quoted from Wang et al. [26]), compared with the Bulk Silicate Earth [54], sedimentary rocks [55–57], igneous rocks [28, 54, 57–60], deep-sea carbonates [61], seafloor hydrothermal fluid system [62], deep sea water [51, 54, 63], and other deposits with different geneses: the Gorno and Raibl magmatic-type deposit in Italy [14], the skarn-type deposits in the Tongling ore district in China [12], the Tianqiao and Bangbangqiao carbonated-hosted Pb-Zn sulfide deposits in China [64], the Irish-type deposit in Ireland [19], the Cévennes MVT deposit in France [65], the Alexandrinka VHMS-type deposit in Russia [18], the Red Dog SEDEX-type ore district in Alaska [20], and the Dongshengmiao SEDEX-type deposit in China [53]. regard the fact that the metal-sourced rocks in these two deposits have the similar Fe isotopic compositions. And, yet, Markl et al. [52] suggested that fluid-rock interaction make the ore-forming fluid have the 𝛿56 Fe value of −0.5‰∼0‰, which is heavier than the 𝛿56 Fe𝑖 value (−0.5‰∼−1‰) of the fractionation models in Section 5.2. Furthermore, it is generally known that the equilibrium isotope fractionation is a function of temperature, with larger fractionation generated at lower temperatures [111]. Although the temperature of oreforming fluid (100∼200∘ C) [52] in the Schwarzwald deposit is lower than that in Zhaxikang deposit (∼250∘ C) [5], the lightest 𝛿56 Fe value (−1.36‰) of siderite from this deposit is much heavier than that of Mn-Fe carbonate (−1.95‰) from Zhaxikang deposit (Figure 9(a)). All of these inconsistencies from Fe-Zn isotopic data indicate that this genetic model is not appropriate for Zhaxikang deposit. In the second genetic model, Xie et al. [6] suggested the mineralization in Zhaxikang deposit was genetically related to the Miocene dome-related magmatism. This magmatism generated the pegmatite and two-mica granite in the core of the regional domes (Figure 1(b); e.g., Cuonadong, Yalaxiangbo, Ranba, and Kangma), and the high-temperature ore-forming fluids were derived from magmatic melts exsolution. This hypothesis is based on the research of field geology, petrography, melt and fluid inclusions, and C−H−O isotopes: (1) the 𝛿13 CV–PDB (−6.1‰∼−6.9‰) and 𝛿18 OV–SMOW (+9.9‰∼+11.8‰) values of rhodochrosite show the mantle origin; (2) the 𝛿13 DV–SMOW ( −144.8‰∼−110‰) and 𝛿18 OV–SMOW (−9.85‰∼−8.89‰) values, the low salinity (0.2∼7.9 wt.% NaCleqv), high temperature (298∼457∘ C), and rich CO2 content with minor CH4 , N2 , C2 H6 , C3 H8 , and C6 H6 of the melt and fluid inclusions trapped in quartz and beryl within pegmatite indicate the magmatic origin; (3) the 40 Ar-39 Ar plateau age of pegmatite (18.93 ± 0.27 Ma) is similar to stage 5 quartz-pyrite-stibnite (17.9 ± 0.5 Ma). The pegmatite and two-mica granite in the domes have the typical characteristics of S-type granite: (1) containing abundant Al-rich minerals, A/CNK: 1.07∼1.24; (2) w(SiO2 ): 73.26%∼74.33%, w(K2 O)/w(Na2 O): 1.1∼1.2, w(FeO)/w(Fe + Mn): 0.64∼0.76; (3) the content of corundum molecules > 1% in the CIPW standard minerals [112, 113]. Meanwhile, based on the facts that all the Fe-bearing minerals in Zhaxikang 18 deposit are ferrous, and the magmatic-hydrothermal fluid is CO2 -rich with minor CH4 , N2 , C2 H6 , C3 H8 , and C6 H6 [6], and the parent magma is most likely S-type reduced magma. In consideration of the situation in the Renison SnW deposit that we mentioned in Section 1, the ore-forming fluid is considered to exsolve from S-type reduced magma and the sulfides have heavier Fe isotopic compositions than ore-related igneous rocks [23]. The case in Zhaxikang deposit is contrary yet the pyrite from the first pulse of mineralization has the 𝛿56 Fe value of −0.33‰∼0.23‰ that is much lighter than those of granitoids (−0.08% to 0.59%) [22, 28]. Moreover, Heimann et al. [43] proposed that we would not expect Fe isotopic compositions of high F/Cl magmatic/hydrothermal systems to significantly deviate from the average of igneous rocks, hence as Xie et al. [6] considered that the Zhaxikang parent magma has high F contents, the magmatic-hydrothermal fluid in this genetic model would have the similar Fe isotopic compositions with granitoids, which is not in line with our fractionation models. Even if the parent magma in Zhaxikang orefield is oxidized-type and is similar to the magma related to I-type granitoids with 𝛿56 Fe values of −0.04%∼0.59% in Tongling ore district (Figure 9(a)) [21, 22], it is still hard to produce so light 𝛿56 Fe𝑖 value (−0.5‰∼−1‰). Similarity, Zn isotopic data do not support this genetic model, either. Chen et al. [54] studied the Zn isotopic fractionation during igneous process and suggested that the maximum Zn isotopic variation induced by high-temperature igneous processes is no larger than 𝛿66 Zn∼0.10%. Besides this, Telus et al. [28] measured the 𝛿66 Zn values of pegmatite (0.25‰∼0.59‰) and some other granitoids (−0.16‰∼0.21‰) and then found there is even no variation in 𝛿66 Zn values during fluid exsolution in some cases. As the granitoids in the regional domes are principally dominated by pegmatite and the magmatic fluids which have high temperature of 298∼457∘ C [6], it is also hard to generate 𝛿66 Zn𝑖 value between −0.28‰ and 0‰ as yielded from the Fe-Zn isotopic fractionation models. Consequently, the FeZn isotopic data are also not in favor of the second genetic model. On the other hand, neither of these two genetic models can explain the different Fe-Zn isotopic and elemental variations in sample ZXK-2, which is considered to result from the overprint by the second pulse of mineralization. Overall, all of the Fe-Zn isotopic and elemental evidences are inconsistent with both of these magmatic-hydrothermal fluid genetic models. Nevertheless, the evidence for these two genetic models may prove that the second pulse of mineralization is related to magmatic-hydrothermal fluids. 5.3.3. Constraints on SEDEX Modified by Hydrothermal Fluid Genetic Model. Zheng et al. [7, 25] considered that the first pulse of mineralization (Pb-Zn) has the SEDEX genesis, and the second pulse of mineralization (Sb-Ag) is related to hot spring that overprints the earlier mineralization. The previous evidences are mainly as follows: (1) the evidence from ore textures and obviously late Sb mineralization compared to Pb-Zn mineralization in Section 2.3; (2) the stage 1 and 2 ores have high Mn, Fe, Ba, and B contents, Ga ≫ In Geofluids (Ga/In: 1.49∼4.47), Pb + Zn ≫ Cu, and host rocks are exhalative rocks and Mn-Fe carbonates; (3) the 𝛿34 S values of sphalerite and galena (7.7‰∼12‰) are different from those of stibnite (4.5‰200B∼7.1‰); (4) the data from valentinite (𝛿30 Si values: −0.9‰∼−0.4‰) and fluid inclusions trapped in quartz (𝛿DV–SMOW : −162‰∼−142‰, 𝛿18 OH2 O : −12.9‰∼−1.9‰, homogenization and freezing temperatures: 164∼313∘ C and −3.2∼−0.7∘ C, salinity: 0.7%∼5.3%, and density: 0.74∼0.93 g/cm3 ) both demonstrate the Sb mineralization is related to the south Tibetan hot spring; (5) the Rb-Sr isotopic isochrone age of stage 2 sphalerite is 147.2 ± 3.2 Ma that is similar to the Jurassic sedimentary wall rocks, whereas a later stage quartz-pyrite-stibnite vein has the 40 Ar39 Ar plateau age of 17.9 ± 0.5 Ma. Wen et al. [17] investigated several Pb-Zn deposits with different geneses in China, and the results show that the Zn/Cd ratios of sphalerite vary with different geneses: (1) high-temperature systems including the porphyry, magmatic hydrothermal, skarn, and volcanic hosted massive sulfide (VMS)-type deposits: 155∼223; (2) low-temperature systems that include the Mississippi Valley-type (MVT) deposits: 17∼ 201; (3) SEDEX-type deposits of exhalative systems: 316∼ 368; (4) seafloor hydrothermal sulfides of exhalative systems: 211∼510. According to the EPMA data, the Zn/Cd ratios of sphalerite range from 296 to 399 in Zhaxikang deposit, which overlap the range of exhalative systems and much higher than those of high-temperature and low-temperature systems. The Fe-Zn isotopic data also conform to the SEDEX modified by hydrothermal fluid genetic model. Firstly, as discussed in Sections 5.3.1 and 5.3.2, neither the hot spring nor the magmatic-hydrothermal fluids can have the 𝛿56 Fe𝑖 (−0.5‰∼ − 1‰) and 𝛿66 Zn𝑖 (−0.28‰∼0‰) values to satisfy the Fe-Zn isotopic fractionation models. However, the FeZn isotopic compositions of seafloor hydrothermal fluid system covers the range of −1.79‰∼0.04‰ for Fe isotope and −0.28‰∼0.96‰ for Zn isotope according to previous studies (Figure 9) [38, 46–49, 62], which can generate the ore-forming system with 𝛿56 Fe𝑖 and 𝛿66 Zn𝑖 values to meet the fractionation models. Secondly, although there are large overlaps in Zn isotopic compositions among deposits with different geneses, the Zn isotopic compositions of Zhaxikang deposit is most similar to the Alexandrinka VHMS-type and Red Dog SEDEX-type deposits with marine origin, as well as obviously distinguishing from the narrow range of magmatic-hydrothermal deposits (Figure 9(b)). Thirdly, the lightest 𝛿56 Fe value in Zhaxikang deposit is −1.95‰, and only the minerals with marine origin (mid-oceanic ridges pyrite) or SEDEX genesis (sphalerite and pyrrhotite in Dongshengmiao SEDEX-type deposit) can have so light 𝛿56 Fe values (Figure 9(a)). Fourthly, under the conditions of high 𝑃CO2 and low PH (<8), the Zn precipitated as sulfides is isotopically nearly unfractionated with respect to the parent hydrothermal fluid, whereas, under the conditions of high 𝑃CO2 and high-PH (>9), negative 𝛿66 Zn values down to 0.6‰ can be expected in sulfides precipitated from hydrothermal fluid [107]. In Zhaxikang deposit, 𝑃CO2 would be high as there are plenty of Mn-Fe carbonates; meanwhile, owing to the facts that the modern seawater has the PH around 8 Geofluids and there is more CO2 in Jurassic atmosphere than present, the Jurassic seawater would have a lower PH than modern seawater. This can well explain that the 𝛿66 Zn values of sphalerite (−0.38‰∼0.07‰) slightly fractionate with the ore-forming system (𝛿66 Zn𝑖 value: −0.28‰∼0‰). Besides these, the overprinting of earlier ores by second pulse of mineralization is not only proved by the different Fe-Zn isotopic and elemental variations in sample ZXK-2 from the other 3 samples in this research (Figure 5) but also evidenced by the temporally increasing 𝛿56 Fe and decreasing 𝛿66 Zn values recorded in this deposit that coincided with an increase in alteration [26]. Nevertheless, further research is required to confirm whether this hydrothermal fluid is hot spring or magmatic-hydrothermal fluid. To sum up, among the various genetic models, the Fe-Zn isotopic and EPMA evidence indicate the SEDEX modified by hydrothermal fluid genetic model is the most plausible. Our research also demonstrates that the Fe-Zn isotopes have the potential to trace the metal source and provide insights into ore-forming processes. 6. Conclusions The EPMA and Fe-Zn isotopic data allow us to make the following conclusions: (1) The ore-forming elements are transported by the ore-forming system that is the mixture of vapour and liquid phases; the vapour-liquid partitioning and mineral precipitation are the main cause of Fe-Zn isotopic and elemental variations. (2) The Fe-Zn isotopic fractionation models demonstrate that the 𝛿56 Fe𝑖 and 𝛿66 Zn𝑖 values of the oreforming system are in the range of −0.5‰∼−1‰ and −0.28‰∼0‰, respectively. (3) Based on the evidence from the EPMA data, FeZn isotopic characteristics, and fractionation models, the SEDEX modified by hydrothermal fluid genetic model is most plausible for the Zhaxikang deposit. (4) There are two pulses of mineralization in the Zhaxikang deposit; the overprint by the second pulse of mineralization has also partly modified the Fe-Zn isotopic and elemental compositions of some earlier samples. Conflicts of Interest The authors declare that they do not have any commercial or associative interest that represents conflicts of interest in connection with the submitted work. Acknowledgments This work was carried out while Da Wang was a visiting student at the Juniata College. The authors would like to thank Matthew Gonzalez (Pennsylvania State University) and Linda Godfrey (Rutgers University) for aid in measuring and access to the Neptune instruments. They acknowledge 19 support from the Program for Changjiang Scholars and Innovative University Research Teams (IRT14R54, IRT1083), the Commonwealth Project from the Ministry of Land and Resources (201511015), and the Fundamental Research Funds for the Central Universities (2652015044 and 2652015354). References [1] W. Liang, Z. Q. Hou, Z. S. Yang et al., “Remobilization and overprinting in the Zhaxikang Pb-Zn-Ag-Sb polymetal ore deposit, Southern Tibet: Implications for its metallogenesis,” Acta Petrologica Sinica, vol. 29, no. 11, pp. 3828–3842, 2013 (Chinese). [2] X. Sun, Y. Zheng, F. Pirajno et al., “Geology, S–Pb isotopes, and 40Ar/39Ar geochronology of the Zhaxikang Sb–Pb–Zn–Ag deposit in Southern Tibet: implications for multiple mineralization events at Zhaxikang,” Mineralium Deposita, pp. 1–24, 2017. [3] X. J. Meng, Z. S. Yang, X. X. Qi, Z. Q. Hou, and Z. Q. Li, “Siliconoxygen-hydrogen isotopic compositions of Zhaxikang antimony polymetallic deposit in Southern Tibet and its responses to the ore-controlling structure,” Acta Petrologica Sinica, vol. 24, no. 7, pp. 1649–1655, 2008 (Chinese). [4] J. F. Zhang, Y. Y. Zheng, G. Y. Zhang et al., “Genesis of Zhaxikang Pb-Zn-Sb-Ag deposit in northern Himalaya: constraints from multi-isotope geochemistry,” Earth Science Journal of China University of Geosciences, vol. 35, pp. 1000–1010, 2010 (Chinese). [5] J. L. Duan, J. X. Tang, and B. Lin B, “Zinc and lead isotope signatures of the Zhaxikang Pb Zn deposit, South Tibet: Implications for the source of the ore-forming metals,” Ore Geology Review, vol. 78, pp. 58–68, 2016. [6] Y. Xie, L. Li, B. Wang et al., “Genesis of the Zhaxikang epithermal Pb-Zn-Sb deposit in southern Tibet, China: Evidence for a magmatic link,” Ore Geology Reviews, vol. 80, pp. 891–909, 2017. [7] Y. Y. Zheng, M. Y. Liu, X. Sun et al., “Type, discovery process and significance of Zhaxikang antimony polymetallic ore deposit, Tibet,” Earth Science Journal of China University of Geosciences, vol. 37, no. 5, pp. 1004–1014, 2012 (Chinese). [8] H. P. Taylor Jr., “The application of oxygen and hydrogen isotope studies to problems of hydrothermal alteration and ore deposition,” Economic Geology, vol. 69, no. 6, pp. 843–883, 1974. [9] M. J. Le Bas, B. Spiro, and Y. Xueming, “Oxygen, carbon and strontium isotope study of the carbonatitic dolomite host of the Bayan Obo Fe-Nb-REE deposit, Inner Mongolia, N China,” Mineralogical Magazine, vol. 61, no. 4, pp. 531–541, 1997. [10] Y. Wang and X.-K. Zhu, “Fe isotope systematics and its implications in ore deposit geology,” Acta Petrologica Sinica, vol. 28, no. 11, pp. 3638–3654, 2012. [11] J. W. Hedenquist and J. B. Lowenstern, “The role of magmas in the formation of hydrothermal ore deposits,” Nature, vol. 370, no. 6490, pp. 519–527, 1994. [12] Y. Wang and X. K. Zhu, “Application of Zn Isotopes to study of Mineral Deposits: A review,” Mineral Deposits, vol. 29, no. 5, pp. 843–852, 2010 (Chinese). [13] Y. Wang and X. K. Zhu, “Applications of Cu isotopes on studies of mineral deposits: A status report,” Journal of Jilin University (Earth Science Edition), vol. 40, no. 4, pp. 739–751, 2010. [14] C. N. Maréchal, P. Télouk, and F. Albarède, “Precise analysis of copper and zinc isotopic compositions by plasma-source mass spectrometry,” Chemical Geology, vol. 156, no. 1-4, pp. 251–273, 1999. 20 [15] N. S. Belshaw, X.-K. Zhu, Y. Guo, and R. K. O’Nions, “High precision measurement of iron isotopes by plasma source mass spectrometry,” International Journal of Mass Spectrometry, vol. 197, no. 1-3, pp. 191–195, 2000. [16] R. Mathur, S. Titley, F. Barra et al., “Exploration potential of Cu isotope fractionation in porphyry copper deposits,” Journal of Geochemical Exploration, vol. 102, no. 1, pp. 1–6, 2009. [17] H. J. Wen, C. W. Zhu, Y. X. Zhang, C. Cloquet, H. F. Fan, and S. H. Fu, “Zn/Cd ratios and cadmium isotope evidence for the classification of lead-zinc deposits,” Scientific Reports, vol. 6, p. 25273, 2016. [18] T. F. D. Mason, D. J. Weiss, J. B. Chapman et al., “Zn and Cu isotopic variability in the Alexandrinka volcanic-hosted massive sulphide (VHMS) ore deposit, Urals, Russia,” Chemical Geology, vol. 221, no. 3-4, pp. 170–187, 2005. [19] J. J. Wilkinson, D. J. Weiss, T. F. D. Mason, and B. J. Coles, “Zinc isotope variation in hydrothermal systems: Preliminary evidence from the Irish midlands ore field,” Economic Geology, vol. 100, no. 3, pp. 583–590, 2005. [20] K. D. Kelley, J. J. Wilkinson, J. B. Chapman, H. L. Crowther, and D. J. Weiss, “Zinc isotopes in sphalerite from base metal deposits in the Red Dog district, northern Alaska,” Economic Geology, vol. 104, no. 6, pp. 767–773, 2009. [21] Y. Wang, X.-K. Zhu, J.-W. Mao, Z.-H. Li, and Y.-B. Cheng, “Iron isotope fractionation during skarn-type metallogeny: A case study of Xinqiao Cu-S-Fe-Au deposit in the Middle-Lower Yangtze valley,” Ore Geology Reviews, vol. 43, no. 1, pp. 194–202, 2011. [22] Y. Wang, X.-K. Zhu, and Y. Cheng, “Fe isotope behaviours during sulfide-dominated skarn-type mineralisation,” Journal of Asian Earth Sciences, vol. 103, pp. 374–392, 2015. [23] C. M. Wawryk and J. D. Foden, “Fe-isotope fractionation in magmatic-hydrothermal mineral deposits: A case study from the Renison Sn-W deposit, Tasmania,” Geochimica et Cosmochimica Acta, vol. 150, pp. 285–298, 2015. [24] A. Yin, “Cenozoic tectonic evolution of the Himalayan orogen as constrained by along-strike variation of structural geometry, exhumation history, and foreland sedimentation,” Earth-Science Reviews, vol. 76, no. 1-2, pp. 1–131, 2006. [25] Y. Zheng, X. Sun, L. Tian et al., “Mineralization, deposit type and metallogenic age of the gold antimony polymetallic belt in the eastern part of North Himalayan,” Geotectonica et Metallogenia, vol. 38, no. 1, pp. 108–118, 2014. [26] D. Wang, X. Sun, Y. Zheng et al., “Two pulses of mineralization and genesis of the Zhaxikang Sb–Pb–Zn–Ag deposit in southern Tibet: Constraints from Fe–Zn isotopes,” Ore Geology Reviews, vol. 84, pp. 347–363, 2017. [27] A. Fernandez and D. M. Borrok, “Fractionation of Cu, Fe, and Zn isotopes during the oxidative weathering of sulfide-rich rocks,” Chemical Geology, vol. 264, no. 1-4, pp. 1–12, 2009. [28] M. Telus, N. Dauphas, F. Moynier et al., “Iron, zinc, magnesium and uranium isotopic fractionation during continental crust differentiation: The tale from migmatites, granitoids, and pegmatites,” Geochimica et Cosmochimica Acta, vol. 97, pp. 247–265, 2012. [29] T. M. Harrison, P. Copeland, W. S. F. Kidd, and A. Yin, “Raising tibet,” Science, vol. 255, no. 5052, pp. 1663–1670, 1992. [30] M. A. Murphy, A. Yin, P. Kapp et al., “Structural evolution of the Gurla Mandhata detachment system, southwest Tibet,” Geological Society of America Bulletin, vol. 114, no. 4, pp. 428– 447, 2002. Geofluids [31] G. T. Pan, X. X. Mo, Z. Q. Hou et al., “Spatial-temporal framework of the Gangdese Orogenic Belt and its evolution,” Acta Petrologica Sinica, vol. 22, pp. 521–533, 2006 (Chinese). [32] J. E. Spratt, A. G. Jones, K. D. Nelson et al., “Crustal structure of the India-Asia collision zone, southern Tibet, from INDEPTH MT investigations,” Physics of the Earth and Planetary Interiors, vol. 150, no. 1-3, pp. 227–237, 2005. [33] G. M. Yu and C. S. Wang, Sedimentary Geology of Tethys, Tibet, Geology Publishing House, Beijing, China, 1999. [34] B. L. Beard, C. M. Johnson, J. L. Skulan, K. H. Nealson, L. Cox, and H. Sun, “Application of Fe isotopes to tracing the geochemical and biological cycling of Fe,” Chemical Geology, vol. 195, no. 1-4, pp. 87–117, 2003. [35] O. J. Rouxel, A. Bekker, and K. J. Edwards, “Iron isotope constraints on the Archean and Paleoproterozoic ocean redox state,” Science, vol. 307, no. 5712, pp. 1088–1091, 2005. [36] B. L. Beard and C. M. Johnson, “High precision iron isotope measurements of terrestrial and lunar materials,” Geochimica et Cosmochimica Acta, vol. 63, no. 11-12, pp. 1653–1660, 1999. [37] X. K. Zhu, Y. Guo, R. J. P. Williams, R. K. O’Nions, A. Matthews, and N. S. Belshaw, “Mass fractionation processes of transition metal isotopes,” Earth and Planetary Science Letters, vol. 200, no. 1, pp. 47–62, 2002. [38] B. L. Beard, C. M. Johnson, K. L. Von Damm, and R. L. Poulson, “Iron isotope constraints on Fe cycling and mass balance in oxygenated Earth oceans,” Geology, vol. 31, no. 7, pp. 629–632, 2003. [39] F. Poitrasson, A. N. Halliday, D.-C. Lee, S. Levasseur, and N. Teutsch, “Iron isotope differences between Earth, Moon, Mars and Vesta as possible records of contrasted accretion mechanisms,” Earth and Planetary Science Letters, vol. 223, no. 3-4, pp. 253–266, 2004. [40] F. Poitrasson and R. Freydier, “Heavy iron isotope composition of granites determined by high resolution MC-ICP-MS,” Chemical Geology, vol. 222, no. 1-2, pp. 132–147, 2005. [41] N. Dauphas, P. E. Janney, R. A. Mendybaev et al., “Chromatographic separation and multicollection-ICPMS analysis of iron. Investigating mass-dependent and-independent isotope effects,” Analytical Chemistry, vol. 76, no. 19, pp. 5855–5863, 2004. [42] N. Dauphas, F.-Z. Teng, and N. T. Arndt, “Magnesium and iron isotopes in 2.7 Ga Alexo komatiites: Mantle signatures, no evidence for Soret diffusion, and identification of diffusive transport in zoned olivine,” Geochimica et Cosmochimica Acta, vol. 74, no. 11, pp. 3274–3291, 2010. [43] A. Heimann, B. L. Beard, and C. M. Johnson, ““The role of volatile exsolution and sub-solidus fluid/rock interactions in producing high 56 Fe/54 Fe ratios in siliceous igneous rocks,” Geochimica et Cosmochimica Acta, vol. 72, no. 17, pp. 4379–4396, 2008. [44] J. A. Schuessler, R. Schoenberg, and O. Sigmarsson, “Iron and lithium isotope systematics of the Hekla volcano, Iceland Evidence for Fe isotope fractionation during magma differentiation,” Chemical Geology, vol. 258, no. 1-2, pp. 78–91, 2009. [45] J. Foden, P. A. Sossi, and C. M. Wawryk, “Fe isotopes and the contrasting petrogenesis of A-, I-and S-type granite,” Lithos, vol. 212, pp. 32–44, 2015. [46] M. Sharam, M. Polizzotto, and A. D. Anbar, “Iron isotopes in hot springs along the Juan de Fuca Ridge,” Earth and Planetary Science Letters, vol. 194, no. 1, pp. 39–51, 2011. Geofluids [47] S. Severmann, C. M. Johnson, B. L. Beard et al., “The effect of plume processes on the Fe isotope composition of hydrothermally derived Fe in the deep ocean as inferred from the Rainbow vent site, Mid-Atlantic Ridge, 36∘14󸀠N,” Earth and Planetary Science Letters, vol. 225, no. 1-2, pp. 63–76, 2004. [48] O. Rouxel, W. C. Shanks III, W. Bach, and K. J. Edwards, “Integrated Fe- and S-isotope study of seafloor hydrothermal vents at East Pacific Rise 9-10∘N,” Chemical Geology, vol. 252, no. 3-4, pp. 214–227, 2008. [49] S. A. Bennett, O. Rouxel, K. Schmidt, D. Garbe-Schoznberg, P. J. Statham, and C. R. German, “Iron isotope fractionation in a buoyant hydrothermal plume, 5∘S Mid-Atlantic Ridge,” Geochimica et Cosmochimica Acta, vol. 73, no. 19, pp. 5619–5634, 2009. [50] O. Rouxel, Y. Fouquet, and J. N. Ludden, “Subsurface processes at the lucky strike hydrothermal field, Mid-Atlantic ridge: Evidence from sulfur, selenium, and iron isotopes,” Geochimica et Cosmochimica Acta, vol. 68, no. 10, pp. 2295–2311, 2004. [51] J. Sun, X. Zhu, Y. Chen, and N. Fang, “Iron isotopic constraints on the genesis of Bayan Obo ore deposit, Inner Mongolia, China,” Precambrian Research, vol. 235, pp. 88–106, 2013. [52] G. Markl, F. von Blanckenburg, and T. Wagner, “Iron isotope fractionation during hydrothermal ore deposition and alteration,” Geochimica et Cosmochimica Acta, vol. 70, no. 12, pp. 3011–3030, 2006. [53] Z. Gao, X. Zhu, J. Sun et al., “Spatial evolution of Zn-Fe-Pb isotopes of sphalerite within a single ore body: A case study from the Dongshengmiao ore deposit, Inner Mongolia, China,” Mineralium Deposita, pp. 1–11, 2017. [54] H. Chen, P. S. Savage, F. Z. Teng, R. T. Helz, and F. Moynier, “Zinc isotope fractionation during magmatic differentiation and the isotopic composition of the bulk Earth,” Earth and Planetary Science Letters, vol. 369, pp. 34–42, 2013. [55] C. N. Maréchal, N. Emmanuel, D. Chantal, and A. Francis, “Abundance of zinc isotopes as a marine biogeochemical tracer,” Geochemistry, Geophysics, Geosystems, vol. 1, no. 5, 2000. [56] D. J. Weiss, N. Rausch, T. F. D. Mason et al., “Atmospheric deposition and isotope biogeochemistry of zinc in ombrotrophic peat,” Geochimica et Cosmochimica Acta, vol. 71, no. 14, pp. 3498–3517, 2007. [57] Y. Bentahila, D. Ben Othman, and J.-M. Luck, “Strontium, lead and zinc isotopes in marine cores as tracers of sedimentary provenance: A case study around Taiwan orogen,” Chemical Geology, vol. 248, no. 1-2, pp. 62–82, 2008. [58] J.-P. Toutain, J. Sonke, M. Munoz et al., “Evidence for Zn isotopic fractionation at Merapi volcano,” Chemical Geology, vol. 253, no. 1-2, pp. 74–82, 2008. [59] J. Viers, P. Oliva, A. Nonell et al., “Evidence of Zn isotopic fractionation in a soil-plant system of a pristine tropical watershed (Nsimi, Cameroon),” Chemical Geology, vol. 239, no. 1-2, pp. 124–137, 2007. [60] G. F. Herzog, F. Moynier, F. Albarède, and A. A. Berezhnoy, “Isotopic and elemental abundances of copper and zinc in lunar samples, Zagami, Pele’s hairs, and a terrestrial basalt,” Geochimica et Cosmochimica Acta, vol. 73, no. 19, pp. 5884– 5904, 2009. [61] S. Pichat, C. Douchet, and F. Albarède, “Zinc isotope variations in deep-sea carbonates from the eastern equatorial Pacific over the last 175 ka,” Earth and Planetary Science Letters, vol. 210, no. 1-2, pp. 167–178, 2003. [62] S. G. John, O. J. Rouxel, P. R. Craddock, A. M. Engwall, and E. A. Boyle, “Zinc stable isotopes in seafloor hydrothermal vent fluids 21 [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] and chimneys,” Earth and Planetary Science Letters, vol. 269, no. 1-2, pp. 17–28, 2008. S. G. John and T. M. Conway, “A role for scavenging in the marine biogeochemical cycling of zinc and zinc isotopes,” Earth and Planetary Science Letters, vol. 394, pp. 159–167, 2014. J.-X. Zhou, Z.-L. Huang, M.-F. Zhou, X.-K. Zhu, and P. Muchez, “Zinc, sulfur and lead isotopic variations in carbonate-hosted Pb-Zn sulfide deposits, southwest China,” Ore Geology Reviews, vol. 58, no. C, pp. 41–54, 2014. F. Albarède, “The stable isotope geochemistry of copper and zinc,” Reviews in Mineralogy and Geochemistry, vol. 55, pp. 409– 427, 2004. G. Liu and G. Einsele, “Sedimentary history of the Tethyan basin in the Tibetan Himalayas,” Geologische Rundschau, vol. 82, no. 1, pp. 32–61, 1994. P. Lefort, A. Yin, and T. M. Harrison, Evolution of the Himalaya. The tectonics of Asia, Cambridge University Press, New York, NY, USA, 1996. E. Garzanti, “Stratigraphy and sedimentary history of the Nepal Tethys Himalaya passive margin,” Journal of Asian Earth Sciences, vol. 17, no. 5-6, pp. 805–827, 1999. R. Armijo, P. Tapponnier, J. L. Mercier, and T. L. Han, “Quaternary extension in southern Tibet: field observations and tectonic implications,” Journal of Geophysical Research: Solid Earth, vol. 91, no. B14, pp. 13803–13872, 1986. P. England and G. Houseman, “Extension during continental convergence, with application to the Tibetan Plateau,” Journal of Geophysical Research: Solid Earth, vol. 94, no. B12, pp. 17561– 17579, 1989. A. Yin, “Mode of Cenozoic east-west extension in Tibet suggesting common origin of rifts in Asia during the Indo-Asian collision,” Journal of Geophysical Research: Solid Earth, vol. 105, no. B9, pp. 21745–21759, 2000. Z. Li, Z. Hou, F. Nie, and X. Meng, “Characteristic and distribution of the partial melting layers in the upper crust: Evidence from active hydrothermal fluid in south Tibet,” Acta Geologica Sinica, vol. 79, no. 1, pp. 69–77, 2005. M. Coleman and K. Hodges, “Evidence for Tibetan plateau uplift before 14 Myr ago from a new minimum age for east-west extension,” Nature, vol. 374, no. 6517, pp. 49–52, 1995. P. Kapp, M. Taylor, D. Stockli, and L. Ding, “Development of active low-angle normal fault systems during orogenic collapse: Insight from Tibet,” Geology, vol. 36, no. 1, pp. 7–10, 2008. A. Yin, P. A. Kapp, M. A. Murphy et al., “Significant late Neogene east-west extension in northern Tibet,” Geology, vol. 27, no. 9, pp. 787–790, 1999. H. Williams, S. Turner, S. Kelley, and N. Harris, “Age and composition of dikes in Southern Tibet: New constraints on the timing of east-west extension and its relationship to postcollisional volcanism,” Geology, vol. 29, no. 4, pp. 339–342, 2001. S. H. Jiang, F. J. Nie, P. Hu, Y. Liu, and X. R. Lai, “Geochemical characteristics of the mafic dyke swarms in South Tibet,” Acta Geological Sinica (English Edition), vol. 81, no. 1, pp. 60–71, 2007. J. S. Tong, J. Liu, H. M. Zhong, J. Xia, R. K. Lu, and Y. H. Li, “Zircon U-Pb dating and geochemistry of mafic dike swarms in the Lhozag area, southern Tibet, China, and their tectonic implications,” Geological Bulletin of China, vol. 26, no. 12, pp. 1654–1664, 2007. Q. Bian and L. Ding, “Discovery of the Zhegucuo Au (As)bearing fine granular quartz diorite in the eastern Tethyan 22 [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] Geofluids Himalayan Belt and its significance,” Acta Petrologica Sinica, vol. 22, no. 4, pp. 977–988, 2006. G. T. Pan, L. Q. Wang, and D. C. Zhu, “Thoughts on some important scientific problems in regional geological survey of the Qinghai-Tibet plateau,” Regional Geology of China, vol. 23, pp. 12–19, 2004. H. Zhong, J. Xia, J. Tong, R. Lu, Y. Li, and S. Xu, “New results and major progress in regional geological survey of the Lhozag County Sheet,” Regional Geology of China, vol. 23, pp. 451–457, 2004. D.-C. Zhu, S.-L. Chung, X.-X. Mo et al., “The 132 Ma Comei-Bunbury large igneous province: Remnants identified in present-day southeastern Tibet and southwestern Australia,” Geology, vol. 37, no. 7, pp. 583–586, 2009. B. B. Qiu, D. C. Zhu, Z. D. Zhao, and L. Q. Wang, “The westward extension of Comei fragmented large igneous province in southern Tibet and its implications.,” Acta Petrologica Sinica, vol. 26, no. 7, pp. 2207–2216, 2010. F. J. Nie, P. Hu, S. H. Jiang, Z. Q. Li, and Y. Liu, “Type and temporal-spatial distribution of gold and antimony deposits (prospects) in Southern Tibet, China,” Acta Geological SinicaEnglish Edition, vol. 79, no. 3, pp. 374–384, 2005. L. Zeng, L.-E. Gao, C. Dong, and S. Tang, “High-pressure melting of metapelite and the formation of Ca-rich granitic melts in the Namche Barwa Massif, southern Tibet,” Gondwana Research, vol. 21, no. 1, pp. 138–151, 2012. M. P. Searle, R. R. Parrish, K. V. Hodges, A. Hurford, M. W. Ayres, and M. J. Whitehouse, “Shisha Pangma leucogranite, south Tibetan Himalaya: Field relations, geochemistry, age, origin, and emplacement,” The Journal of Geology, vol. 105, no. 3, pp. 295–317, 1997. T. M. Harrison, P. Copeland, W. S. F. Kidd, and O. M. Lovera, “Activation of the Nyainqentanghla Shear Zone: Implications for uplift of the southern Tibetan Plateau,” Tectonics, vol. 14, no. 3, pp. 658–676, 1995. Z. Hou, S. Tian, Z. Yuan et al., “The Himalayan collision zone carbonatites in western Sichuan, SW China: Petrogenesis, mantle source and tectonic implication,” Earth and Planetary Science Letters, vol. 244, no. 1-2, pp. 234–250, 2006. Y. Y. Zheng, X. Sun, S. B. Gao et al., Target optimization and demonstrate research of Au-Sb Polymetallic Metallogenic Belt in Southern Tibet, Project Report. China University of Geoscience, Wuhan, China, 2016. B. Lin, J. X. Tang, W. B. Zheng et al., “Petrochemical features, zircon U-Pb dating and Hf isotopic composition of the rhyolite in Zhaxikang deposit, Southern Xizang (Tibet),” Geological Review, vol. 60, no. 1, pp. 178–189, 2014. V. N. Korolyuk, Y. G. Lavrent’ev, L. V. Usova, and E. N. Nigmatulina, “Some features of X-ray registration using a JXA-8100 electron-probe microanalyzer,” Journal of Analytical Chemistry, vol. 65, no. 3, pp. 249–254, 2010. W. Tian, B. Chen, T. R. Ireland, D. H. Green, K. Suzuki, and Z. Chu, “Petrology and geochemistry of dunites, chromitites and mineral inclusions from the Gaositai Alaskan-type complex, North China Craton: Implications for mantle source characteristics,” Lithos, vol. 127, no. 1-2, pp. 165–175, 2011. T. Yesavage, G. E. Stinchcomb, M. S. Fantle, P. B. Sak, A. Kasznel, and S. L. Brantley, “Investigation of a diabase-derived regolith profile from Pennsylvania: Mineralogy, chemistry and Fe isotope fractionation,” Geoderma, vol. 273, pp. 83–97, 2016. C. Archer and D. Vance, “Mass discrimination correction in multiple-collector plasma source mass spectrometry: an [95] [96] [97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] example using Cu and Zn isotopes,” Journal of Analytical Atomic Spectrometry, vol. 19, no. 5, pp. 656–665, 2004. J. B. Chapman, T. F. D. Mason, D. J. Weiss, B. J. Coles, and J. J. Wilkinson, “An adapted column chemistry procedure for separation of Fe, Cu and Zn from geological matrices, and natural Zn isotopic variations in geological standard reference materials BCR-027, BCR-030 and NOD-P-1,” Abstracts with Programs-Geological Society of America, vol. 36, p. 448, 2004. J. B. Chapman, T. F. D. Mason, D. J. Weiss, B. J. Coles, and J. J. Wilkinson, “Chemical separation and isotopic variations of Cu and Zn from five geological reference materials,” Geostandards and Geoanalytical Research, vol. 30, no. 1, pp. 5–16, 2006. K. Peel, D. Weiss, J. Chapman, T. Arnold, and B. Coles, “A simple combined sample-standard bracketing and interelement correction procedure for accurate mass bias correction and precise Zn and Cu isotope ratio measurements,” Journal of Analytical Atomic Spectrometry, vol. 23, no. 1, pp. 103–110, 2008. C. Archer, M. B. Andersen, C. Cloquet et al., “Inter-calibration of a proposed new primary reference standard AA-ETH Zn for zinc isotopic analysis,” Journal of Analytical Atomic Spectrometry, vol. 32, no. 2, pp. 415–419, 2017. C. Archer, D. Vance, and I. Butler, “Abiotic Zn isotope fractionations associated with ZnS precipitation,” 14th Annual VM Goldschmidt Conference. Geochimica et Cosmochimica Acta, vol. 68, no. 11, pp. A325–A325, 2004. D. Gagnevin, A. J. Boyce, C. D. Barrie, J. F. Menuge, and R. J. Blakeman, “Zn, Fe and S isotope fractionation in a large hydrothermal system,” Geochimica et Cosmochimica Acta, vol. 88, pp. 183–198, 2012. S. Graham, N. Pearson, S. Jackson, W. Griffin, and S. Y. O’Reilly, “Tracing Cu and Fe from source to porphyry: In situ determination of Cu and Fe isotope ratios in sulfides from the Grasberg Cu-Au deposit,” Chemical Geology, vol. 207, no. 3-4, pp. 147–169, 2004. Z. Sharp, Principles of Stable Isotope Geochemistry, Pearson education, Upper Saddle River, NJ, USA, 2007. J. Yao, R. Mathur, W. Sun et al., “Fractionation of Cu and Mo isotopes caused by vapor-liquid partitioning, evidence from the Dahutang W-Cu-Mo ore field,” Geochemistry, Geophysics, Geosystems, vol. 17, no. 5, pp. 1725–1739, 2016. M. Yu, Characteristics of ore geology and ore-forming fluid in the Zhaxikang Sb-Pb-Zn-Ag deposit, Southern Tibet, China [Msc. thesis], China University of Geoscience, Beijing, China, 2015. W. Dansgaard, “Stable isotopes in precipitation,” Tellus, vol. 16, no. 4, pp. 436–468, 1964. V. B. Polyakov and D. M. Soultanov, “New data on equilibrium iron isotope fractionation among sulfides: Constraints on mechanisms of sulfide formation in hydrothermal and igneous systems,” Geochimica et Cosmochimica Acta, vol. 75, no. 7, pp. 1957–1974, 2011. T. Fujii, F. Moynier, M.-L. Pons, and F. Albarède, “The origin of Zn isotope fractionation in sulfides,” Geochimica et Cosmochimica Acta, vol. 75, no. 23, pp. 7632–7643, 2011. S. H. Little, D. Vance, C. Walker-Brown, and W. M. Landing, “The oceanic mass balance of copper and zinc isotopes, investigated by analysis of their inputs, and outputs to ferromanganese oxide sediments,” Geochimica et Cosmochimica Acta, vol. 125, pp. 673–693, 2014. Y. Zhao, D. Vance, W. Abouchami, and H. J. W. de Baar, “Biogeochemical cycling of zinc and its isotopes in the Southern Ocean,” Geochimica et Cosmochimica Acta, vol. 125, pp. 653– 672, 2014. Geofluids [110] J.-B. Chen, J. Gaillardet, C. Dessert et al., “Zn isotope compositions of the thermal spring waters of la soufrière volcano, guadeloupe island,” Geochimica et Cosmochimica Acta, vol. 127, pp. 67–82, 2014. [111] H. C. Urey, “The thermodynamic properties of isotopic substances,” Journal of the Chemical Society (Resumed), pp. 562–581, 1947. [112] J. S. Tong, H. M. Zhong, J. Xia, R. K. Lu, and S. Y. Yang, “Geochemical features and tectonic setting of peraluminous granite in the Lhozag area, southern Tibet,” Regional Geology of China, vol. 5, pp. 308–318, 2003 (Chinese). [113] J. Y. Zhang, Q. A. Liao, D. W. Li, X. H. Zhang, and Y. M. Yuan, “Laguigangri leucogranites and its relation with laguigangri metamorphic core complex in Sajia, south Tibet,” Earth Science (Journal of China University of Geosciences), vol. 28, no. 6, pp. 695–700, 2003 (Chinese). 23 International Journal of Chemistry Geophysics Scientifica Hindawi www.hindawi.com Volume 2018 Hindawi www.hindawi.com Volume 2018 The Scientific World Journal Hindawi Publishing Corporation http://www.hindawi.com www.hindawi.com 2013 Volume 2018 Advances in Public Health Hindawi www.hindawi.com Volume 2018 Journal of Environmental and Public Health Advances in Meteorology Hindawi www.hindawi.com Hindawi www.hindawi.com Volume 2018 Volume 2018 Submit your manuscripts at www.hindawi.com Applied & Environmental Soil Science Hindawi www.hindawi.com Journal of Geological Research Volume 2018 Hindawi www.hindawi.com International Journal of International Journal of Ecology Agronomy International Journal of Forestry Research Hindawi www.hindawi.com Volume 2018 Hindawi www.hindawi.com Volume 2018 International Journal of Biodiversity Hindawi www.hindawi.com Hindawi www.hindawi.com International Journal of Advances in Microbiology Agriculture Hindawi www.hindawi.com Volume 2018 International Journal of Archaea Volume 2018 Volume 2018 Analytical Chemistry Volume 2018 Hindawi www.hindawi.com Volume 2018 Hindawi www.hindawi.com Journal of Journal of Chemistry Marine Biology Hindawi www.hindawi.com Volume 2018 Hindawi www.hindawi.com Volume 2018 Volume 2018