Next Article in Journal
A Study on the Control Solution of Ship’s Central Fresh Water-Cooling System for Efficient Energy Control Based on Merchant Training Ship
Next Article in Special Issue
Volatile Characteristics of Apatite in Dacite from the Eastern Manus Basin and Their Geological Implications
Previous Article in Journal
Effect of Lateral Stress on the Mechanical Properties of Rock Fracture and Its Implication on the Stability of Underground Oil Storage Caverns
Previous Article in Special Issue
Detrital Zircon Provenance in the Sediments in the Southern Okinawa Trough
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical and Isotopic Composition of Sulfide Minerals from the Noho Hydrothermal Field in the Okinawa Trough

1
Seafloor Hydrothermal Activity Laboratory, CAS Key Laboratory of Marine Geology and Environment, Institute of Oceanology, Chinese Academy of Sciences, Qingdao 266071, China
2
Laboratory for Marine Mineral Resources, Qingdao National Laboratory for Marine Science and Technology, Qingdao 266071, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
4
Center for Ocean Mega-Science, Chinese Academy of Sciences, 7 Nanhai Road, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2022, 10(5), 678; https://doi.org/10.3390/jmse10050678
Submission received: 1 April 2022 / Revised: 7 May 2022 / Accepted: 9 May 2022 / Published: 16 May 2022

Abstract

:
Studies of the element contents and isotopic characteristics of sulfide minerals from seafloor hydrothermal sulfide deposits are a significant method of investigating seawater-fluid mixing and fluid-rock and/or sediment interactions in hydrothermal systems. The seafloor hydrothermal sulfide ores from the Noho hydrothermal field (NHF) in the Okinawa Trough (OT) consist of pyrrhotite, isocubanite, sphalerite, galena, and amorphous silica. The Rh, Ag, Sb, and Tl contents mostly increase in galena as the fluid temperature decreases in the late ore-forming stage. In the sulfide minerals, the rare earth elements are mainly derived from the hydrothermal fluids, while the volcanic rocks and/or sediments are the sources of the sulfur and lead in the sulfide minerals. After the precipitation of galena, the redox state becomes oxidizing, and the pH value of the fluid increases, which is accompanied by the formation of amorphous silica. Finally, neither pyrite nor marcasite has been observed in association with pyrrhotite in the NHF sulfides, likely indicating that the amount of sulfur was limited in this hydrothermal system, and most of the residual Fe was incorporated into the sphalerite. This suggests that the later pyrite and/or marcasite precipitation in the seafloor hydrothermal sulfide deposit is controlled by the sulfur content of the fluid. Furthermore, it is possible to use hydrothermal sulfides and their inclusions to trace subseafloor fluid circulation processes.

1. Introduction

The majority of seafloor hydrothermal sulfide deposits contain a mixture of sulfide minerals (e.g., pyrrhotite, pyrite, marcasite, chalcopyrite, isocubanite, sphalerite, wurtzite, and galena), secondary Cu sulfides (bornite, covellite, digenite, and chalcocite), sulfosalts (tennantite-tetrahedrite among many others) [1], sulfates (anhydrite and barite), and amorphous silica in different abundancies. For example, pyrrhotite, which contains Cu, Co, Mn, Ni, Sn, Pb, Cr, Ti, Ag, and Se [2,3,4], is commonly found as a major phase in many seafloor hydrothermal sulfide deposits (e.g., [5,6,7,8,9,10]). Its in-situ chemical composition and genesis have been investigated in several hydrothermal fields, including the Bent Hill field on the northern Juan de Fuca Ridge (JdFR) [11], the Broken Spur vent field on the Mid-Atlantic Ridge (MAR) [12], the Lucky Strike vent field on the MAR [13], and the Ashadze hydrothermal field on the MAR [14]. However, the trace element contents of pyrrhotite are variable [1]. It is known that the pyrrhotite from the southern JdFR has higher silver (Ag) contents (0.04–0.07 wt.%; [15]) than that from the Endeavour Segment of the JdFR (<0.04 wt.%; [16]) and the Rainbow hydrothermal field on the MAR (0.02 wt.%; [17]). In addition, the Cu content of pyrrhotite from the Endeavour Segment of the JdFR (0.065 wt.%; [16]) is higher than that from the southern JdFR (0.01–0.06 wt.%; [15]).
Sphalerite, also one of the most common minerals in seafloor hydrothermal sulfides and chimneys (e.g., [18,19,20,21]), contains Mn, Fe, Co, Ni, Cu, Ag, Sn, Pb, Hg, Cd, Ge, Ga, In, and Sb [22,23,24]. Its Cd content can be very high (200–11,900 ppm) [1]. Sphalerite associated with arc-related geologic settings generally contains higher levels of As (1000–10,000 s ppm), Pb (1000–10,000 s ppm), and Sb (100–1000 s ppm) than that in mid-ocean ridge (MOR) hydrothermal systems (10–100 s ppm) [1]. The FeS content of sphalerite can be useful in determining its formation temperature and the sulfur fugacity of the fluid in the chimneys in seafloor hydrothermal fields [20].
The presence of significant amounts of galena is characteristic of arc and back-arc basin (BAB) hydrothermal sulfide deposits [10,20,25,26,27]. Galena contains Cu, Ag, Sb, Bi, Cd, As, In, Mn, Tl, and Zn [28]. Particularly high Cd (800–1100 ppm) and Zn (2000–3400 ppm) contents have been measured in the galena from the Palinuro Volcanic Complex, Hook Ridge, and the Endeavour Segment of the JdFR [16,29]. Furthermore, the highest Sb (2160 ppm) and Ag (9400 ppm) contents in galena have been observed in arc-related seafloor hydrothermal fields [1].
However, the in-situ element contents and their distributions in sulfide minerals from the Noho hydrothermal field (NHF) in the Okinawa Trough (OT) are poorly documented. To reveal the formation processes and metal sources of the sulfide minerals in the seafloor hydrothermal systems, we studied and characterized the mineralogical, chemical, and isotopic compositions of the sulfide minerals from the seafloor hydrothermal sulfide deposits in the NHF.

2. Geological Setting

The OT is an active BAB located behind the Ryukyu trench–arc system, extending approximately 1200 km from Taiwan island to Kyushu Island in the western Pacific (Figure 1). It was formed by the active subduction of the Philippine Sea Plate under the Eurasian Plate in the Early Pleistocene (2–1.5 Ma; [30,31,32]). However, the OT is divided into the southern (SOT), middle (MOT), and northern (NOT) segments by the Kerama and Tokara faults [33]. The latest phase of rifting started <100 ka ago according to the ages of the zircon in the OT volcanic rocks, but the exact ages of the earlier rifting phases may be considerably older (108 Ma to 2.7 Ga) [34]. Furthermore, the SOT is characterized by a sedimentary cover that is much thinner (1–3 km thick) than the very thick cover in the NOT (~8 km) [31,35,36]. The OT volcanic rocks vary from tholeiitic basalt to calc-alkaline andesite to rhyolite [37], including basaltic andesite and dacite in the MOT [38].
However, the presence of high 3He contents in the seawater [40] and a high heat flow [41] suggest that there are numerous active seafloor hydrothermal fields in the MOT [42]. Indeed, several seafloor hydrothermal fields have been discovered in the MOT [43,44,45,46,47]. For example, the NHF (27°31.1′ N, 126°59.0′ E, 1581 m) was surveyed in 2016 during seafloor hydrothermal activity off the Iheya Ridge in the MOT using a television (TV)-log grabber (Figure 1) during the HOBAB4 cruise. The survey was conducted near the Noho site (27°31.3′ N, 126°59.1′ E, 1595 m), approximately 3 km SE of the Clam hydrothermal field in the Sakai field discovered by the Japan Oil, Gas, and Metals National Corporation (JOGMEC) ([48]; JOGMEC news release on 4 December 2014; NT15-13 cruise report available at http://www.godac.jamstec.go.jp/darwin/cruise/Natsushima/NT15-13/). The exact position of the Noho (JOGMEC) site had not yet been published at that time [42]. The NHF likely corresponded to the acoustic water column anomaly site [42]. It is mainly associated with basaltic and trachyandesitic rocks (H4-TVG5-2; Figure 1) which are located on the outside slope of the Iheya Ridge and are covered by several meters of pumice-rich sediments. The site includes active hydrothermal vents, chimneys, hydrothermal sulfide deposits, and chemosynthetic communities. However, there is a coupling relationship between the tectonic setting, magmatism, mineralization, fluid-rock interactions, and sedimentary processes in the NHF hydrothermal system and the response, adaptation, record, and activity of organisms. It is interesting to investigate the synergetic metallogenic mechanism of the magma, fluid, rocks, sediments, seawater, and organisms in the NHF hydrothermal field. In the future, it will be possible to use hydrothermal sulfide and its inclusions to trace the deep circulation processes of the fluids in the NHF. Furthermore, we can combine geophysical prospecting, geochemical exploration, and biological exploration techniques with drilling techniques to explore the seafloor polymetallic sulfides and hydrothermal vents, and we can use the four systems of exploration techniques to understand the resource potential of the seafloor polymetallic sulfide deposits in the NHF.

3. Sampling and Methods

Seafloor hydrothermal sulfide samples H4-TVG5-1-1 and H4-TVG5-3-1 were collected from the NHF using a TV grabber (Figure 1; Table S1). To determine the morphology, structure, and chemical compositions of the minerals, reflected and transmitted light microscopy, scanning electron microscopy (SEM), and energy dispersive spectrometry (EDS) were used to analyze thin sections of selected samples. Furthermore, the in-situ major and trace element contents of the samples were analyzed, as well as their S and Pb isotopic compositions.

3.1. SEM Analysis

Pyrrhotite, isocubanite, sphalerite, galena, and amorphous silica were studied using the TESCAN VEGA 3 LMH SEM with an Oxford INCA X-Max EDS at the Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China. The standard analytical operating conditions included an accelerating voltage of 20 kV, a beam intensity of 15 nA, and a working distance of ~15 mm. The following standards were used for the EDS measurements and calibrations: (1) sulfides–pyrrhotite (Fe, S), galena (Pb, S), sphalerite (Zn, S), chalcopyrite (Cu, Fe, S), greenote (Cd), stibnite (Sb), and proustite (Ag, As); (2) silicates–olivine (Ni) and basalt glass (Ti, Si); (3) carbonates–calcite (O); (4) metals–cobalt (Co); (5) oxides–Cr-spinel (Mn); and (6) system standard library (Au, Se). The EDS detection limits for Au, S, Pb, Ag, Cd, Sb, O, Se, As, Zn, Cu, Ni, Co, Fe, Mn, and Ti were 0.1 wt.% (Table S2).

3.2. Major Element Analysis

The major element compositions of the pyrrhotite, sphalerite, and galena were analyzed using the JXA-8230 electron microprobe analyzer (EMPA) at the State Key Laboratory of Continental Dynamics, Northwest University, Xi’an, China. The instrument was operated with a 15 kV acceleration voltage, 10 nA beam current, and 2 μm beam diameter. The following reference materials were used for the wavelength-dispersive spectrometry measurements and calibrations: (1) metals–Co, Au, Se, Ni, Mn, and Ti and (2) natural sulfides–FeS2, CuFeS2, ZnS, PbS, Sb2S3, Ag3AsS3, and CdS. The counting time for the peak and background of Au, S, Pb, Ag, Cd, Sb, Se, As, Zn, Cu, Ni, Co, Fe, Mn, and Ti are 20–60 s and 15 s, respectively. The detection limits of these elements were 0.02–0.11 wt.% (Table S3).

3.3. Trace Element Analysis

The trace element contents of the pyrrhotite, sphalerite, and galena were determined using the laser ablation inductively coupled plasma mass spectrometer (LA-ICP-MS) at the State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Wuhan. A GeoLas Pro 193 nm ArF excimer laser was used to analyze the samples. The laser energy was 80 mJ, the frequency was 6 Hz, and the diameter ablation spot was 44 μm. The ion-signal intensities were acquired using an Agilent 7500 x ICP-MS instrument. Helium was used as the carrier gas which was mixed with argon via a T-connector before entering the ICP-MS. Each measurement incorporated approximately 30 s of background acquisition (gas blank) followed by 50 s of data acquisition from the sample. The ICPMSDataCal software was used to quantitatively calibrate the trace element contents [49]. Fe, Cu, Zn, and Pb were selected as the internal standards for the data reduction. Reference glass NIST 610 was used as the external standard, which was analyzed every 10 spots to monitor the instrument drift. The accuracy was determined with respect to reference glass NIST 610 and was assessed to be better than 10% (1σ).

3.4. In-Situ S Isotope Analysis

The in-situ S isotope compositions of the sphalerite, pyrrhotite, and galena were determined using double side-polished slices (DSPS) and the 193 nm femtosecond laser-ablation multi-collector ICP-MS (fs-LA-MC-ICP-MS) at the State Key Laboratory of Continental Dynamics (SKLCD), Northwest University, Xi’an, China. The equipment consisted of a 193 nm NWRfemto (RESOlution M-50, ASI) laser ablation system coupled with a Nu Plasma II MC-ICP-MS (NU Instruments, Ltd., Wrexham, UK). The ablated materials were transported into the plasma using He as the carrier gas. In a cyclone coaxial mixer, Ar gas was mixed with the carrier gas before being transported into the ICP’s torch. The energy fluence of the laser was approximately 3.5–4 J/cm2. The beam diameter was 35 μm with a laser repletion rate of 3–4 Hz for a single spot analysis. Standard PSPT-3 (sphalerite; δ34SV-CDT = 26.5 ± 0.2‰) was used as the certified reference material. The obtained δ34SV-CDT of Standard PSPT-3 in this study is 26.36 ± 0.29‰ (2SD; n = 21; Table S4), which is consistent with the recommended values (δ34SV-CDT = 26.5 ± 0.2‰; [50]). The detailed analysis parameters are described in Bao et al. [50], Chen et al. [51], and Yuan et al. [52].

3.5. In-Situ Pb Isotope Analysis

The in-situ Pb isotope compositions of the galena were determined using DSPSs and the fs-LA-MC-ICP-MS at the SKLCD. To remove potential contamination, anhydrous ethanol was used to carefully clean the surfaces of the DSPSs prior to the laser ablation analysis. The analytical spots were carefully selected to prevent the possible influence of inclusions and impurities. High temperature-activated carbon was used to filter the Hg contained in the carrier gas, which lowered both the Hg background and the detection limit. Internal Tl isotope reference NIST SRM997 was used in conjunction with external reference NIST SRM 610 to correct for fractionation and mass discrimination effects. The Tl solution was introduced through the CETAC Aridus II desolvation nebulizer system. The exponential law correction method for Tl normalization with an optimally adjusted Tl ratio was used to obtain Pb isotope data with high precision and accuracy [53]. The measured isotope ratios were in agreement with the reference and the published values within 2σ measurement uncertainties [50,52,53,54]. Standard CBI-3 (Natural galena) was used as certified reference material. The measured Pb isotope ratios of standard CBI-3 in this study (208Pb/206Pb = 2.1246 ± 0.0010, 207Pb/206Pb = 0.8696 ± 0.00006, 206Pb/204Pb = 17.963 ± 0.0017, 207Pb/204Pb = 15.621 ± 0.0025, and 208Pb/204Pb = 38.165 ± 0.0048; SD; N = 5; Table S4) were consistent with the recommended values (208Pb/206Pb = 2.1247 ± 0.0000, 207Pb/206Pb = 0.8694 ± 0.00001, 206Pb/204Pb = 17.964 ± 0.0014, 207Pb/204Pb = 15.622 ± 0.0013, and 208Pb/204Pb = 38.168 ± 0.0028; [52]).

4. Results

4.1. Mineralogy of Hydrothermal Sulfide Samples

Seafloor hydrothermal sulfide ore samples H4-TVG5-1-1 (~8 cm in size) and H4-TVG5-3-1 (~24 cm in size) are black to dark grey in color and contained pores up to 0.2 mm (H4-TVG5-1-1) and 1 mm (H4-TVG5-3-1) in diameter. They are composed of sulfide minerals (Figure 2), i.e., pyrrhotite, galena, sphalerite, and isocubanite, as well as amorphous silica. The coarser-grained aggregates of pyrrhotite (up to 100 μm), sphalerite (up to 50 μm), galena (up to 30 μm), and isocubanite (up to 20 μm) comprise >70% of the sulfide samples by volume in both samples (Figure 2 and Figure 3).

4.1.1. Pyrrhotite

Pyrrhotite is the most abundant mineral. It occurs as randomly oriented laths that range in size from a few to hundreds of micrometers. The laths form boxwork textures (Figure 2) shaped similar to blades or hexagonal plates (Figure 2 and Figure 4a). Some of the pyrrhotite boxwork laths contain interstitial sphalerite, minor isocubanite, and late galena (Figure 2). The whitish rim around pyrrhotite is likely a thin oxidized pyrrhotite-rim, similar to the rims commonly observed in pyrrhotite from other hydrothermal fields (Figure 2f).

4.1.2. Sphalerite

The sphalerite from the NHF is red and Fe-rich (Table S3). It is intimately connected with the pyrrhotite and galena (Figure 2 and Figure 4), intergrown with the pyrrhotite and isocubanite (Figure 2h), overgrown by amorphous silica, and contains local inclusions of galena. The sphalerite commonly occurs as hydrothermal aggregates that are intimately intergrown within the interstices between the pyrrhotite laths where it forms rounded or elongated anhedral grains up to several tens of micrometers across (Figure 2f).

4.1.3. Galena

Galena is the major sulfide mineral in both samples (Figure 2 and Figure 4). It is disseminated throughout the samples, appearing as subhedral or anhedral crystals (Figure 2), and ranges in size from a few to tens of micrometers (Figure 2 and Figure 4). Most of the drop-like, oval, or elongated inclusions of galena crystals up to 25 μm in size are found in the sphalerite (Figure 2d). This indicates that there is a positive correlation between the Zn and Pb in the sulfide minerals [10]. The galena is often surrounded by amorphous silica (Figure 2e). It is also found as intergrowths of minute (5–40 μm) euhedral to subhedral crystals in the pores within the pyrrhotite–sphalerite aggregates (Figure 2h).

4.1.4. Isocubanite and Amorphous Silica

The isocubanite coexists with the sphalerite and galena in hydrothermal sulfide samples H4-TVG5-1-1 and H4-TVG5-3-1 (Figure 2). When present, amorphous silica occurs as a thin late-stage rim around the sulfide minerals and may overgrow the sphalerite, pyrrhotite, and galena (Figure 2).

4.2. Major and Trace Element Contents of Sulfide Minerals

4.2.1. Major Element Contents of Sulfide Minerals

The major element concentrations of the sulfide minerals are presented in Tables S2 and S3. According to Belzile et al. [55] and Fallon et al. [1], pyrrhotite can be monoclinic (less Fe-deficient: 45.9–46.8 mol% Fe) or hexagonal (more Fe-deficient: 47.2–48.0 mol% Fe) (Table S2). Most sphalerite has low Cu contents (0.10–0.46 wt.%) and very high Fe (12.46–19.53 wt.%) and Mn (0.10–6.27 wt.%) contents (Figure 4; Table S2). The FeS contents of the sphalerite slabs range from 21.95 to 34.78 mol% (Table S2). In contrast, the Fe contents (1.09 to 2.61 wt.%) of the FeS-bearing sphalerite from the sulfide chimneys in the North Konll, Iheya Ridge, are reportedly very low (~4 mol%; Ueno et al. [20]). The major elements in the galena are Pb (83.68–87.17 wt.%) and S (13.31–13.69 wt.%) (Tables S2 and S3), and those in the isocubanite are Zn (0.80–2.12 wt.%) and Pb (0.38–2.35 wt.%) (Table S2), while most of the amorphous silica contains significant amounts of Fe (0.46–9.64 wt.%) and Zn (0.13–4.86 wt.%) (Table S2).

4.2.2. Trace Element Contents of Sulfide Minerals

The trace element concentrations of the sulfide minerals are presented in Table S5. The pyrrhotite is relatively enriched in Ge (28.0–97.3 ppm) and Mo (1.21–3.40 ppm). The Ge content (28.0 to 97.3 ppm) of the pyrrhotite samples (H4-TVG5-1-1) from the NHF are noticeably higher than those of the galena (0.51–1.09 ppm) and sphalerite (3.63–4.84 ppm) crystals (Table S5; Figure 5).
The sphalerite contains Cd (318–388 ppm), Sn (145–182 ppm), Au (0.67–0.82 ppm), Th (4.85–5.09 ppm), Mo (0.08–0.34 ppm), Tl (0.11–0.67 ppm), and Bi (0.03–0.15 ppm). The Th contents of the sphalerite samples from the NHF are usually higher than those of the pyrrhotite and galena from sample H4-TVG5-1-1 (Table S5; Figure 5).
The Mo (0.07–2.44 ppm), Cd (3.11–70.1 ppm), V (0.41–22.3 ppm), Cr (2.82–369 ppm), Ni (0.45–6.48 ppm), and Cu (12.9–4420 ppm) contents of the galena from the NHF are much more variable than those of the pyrrhotite and sphalerite, covering the largest range (Tables S2, S3 and S5; Figure 5). The As (1.74–2233 ppm), Se (17.0–109 ppm), Fe (3128–101,010 ppm), and Zn (829–127,104 ppm) contents of the galena samples from the NHF also span large ranges (Tables S2, S3 and S5). Se is concentrated in the galena (17.0–109 ppm) and pyrrhotite (13.2–52.7 ppm) (Table S5), whereas Tl was detected in the galena (0.49–4.09 ppm), pyrrhotite (0.12–0.64 ppm), and sphalerite (0.11–0.67 ppm) (Table S5). The galena is more enriched in Ag (693–1940 ppm) and Sb (1608–5129 ppm) than the pyrrhotite and sphalerite (Table S5; Figure 5). In particular, the Ag contents (693–1940 ppm) of most of the galena samples are significantly higher than those of the pyrrhotite (0.49–9.28 ppm) and sphalerite (42.4–53.9 ppm) samples (Figure 5). There is a positive correlation between the Ag and Bi contents of the galena. Most of the pyrrhotite (Th 0.08–2.38 ppm) and sphalerite (Th 4.85–5.09 ppm) contain more Th than the galena (Th 0.04–2.37 ppm), while the galena (7.52–16.8 ppm) has significantly higher Rh contents than the pyrrhotite (0.08–0.14 ppm) and sphalerite (0.04–0.05 ppm) (Table S5; Figure 5). Furthermore, a significant positive correlation was observed between the Au and Mo contents of the galena from the NHF.

4.3. Rare Earth Element Compositions of Sulfide Minerals

The rare earth element (REE) contents were normalized to Chondrite values (subscript CN) [56]. The Eu and Ce anomalies were assessed using the following equations: (Eu/Eu*)CN = (2 × EuCN)/(SmCN + GdCN) and (Ce/Ce*)CN = (2 × CeCN)/(LaCN + PrCN). The REE data for the Yonaguni Knoll IV vent fluids, Iheya Ridge volcanic rocks, and MOT sediments used for comparison are from Hongo et al. [57], Li et al. [39], and Xu et al. [58], respectively.
The total REE contents (∑REEs) of the pyrrhotite, sphalerite, and galena from the NHF are highly variable (2.01–36.7 ppm) (Table S5). Most of the ∑REE contents of the pyrrhotite (2.32–36.5 ppm) and galena (2.01–36.7 ppm) are higher than those of the sphalerite (10.2–13.2 ppm) (Table S5). Among the studied samples, the pyrrhotite from the NHF had the highest ∑REE contents (36.5 ppm, H4-TVG5-1-1, point 2–3) (Table S5).
The Chondrite-normalized REE distribution patterns of the pyrrhotite, sphalerite, and galena from the NHF are presented in Figure 6. The REE patterns of most of the sulfide minerals exhibit light rare earth element (LREE) enrichment (LREE/HREE = 8.09–32.0), variable LaCN/LuCN ratios (0.98 and 50.0), and Eu ((Eu/Eu*)CN = 0.47–4.13) and Ce ((Ce/Ce*)CN = 1.70–3.85) anomalies (Figure 6a–c). Compared with the Eu anomalies of the pyrrhotite ((Eu/Eu*)CN = 1.44–1.68) and sphalerite ((Eu/Eu*)CN = 3.61–4.13) from the NHF, the galena has smaller Eu anomalies ((Eu/Eu*)CN = 0.47–0.83). Moreover, most HREE contents are around the detection limit, which has been marked in dash lines (Figure 6) and were only used to qualitatively understand their low contents.

4.4. In-Situ S and Pb Isotopic Compositions of the Sulfide Minerals

Five sulfur isotope analyses were conducted on the sphalerite, pyrrhotite, and galena grains from sample H4-TVG5-1-1. The δ34S values obtained for the sulfide minerals are consistent and range from 3.58 to 5.69‰. One sphalerite sample from hydrothermal sulfide sample H4-TVG5-1-1 yielded the highest δ34S values (5.69‰). The galena and pyrrhotite are characterized by lower sulfur isotope values (<4.16‰). The lowest δ34S value (3.58‰) was for a galena sample (Table S6).
The galena from sample H4-TVG5-1-1 from the NHF has uniform Pb isotopic compositions, with a narrow range of Pb isotopic values (Table S6; Figure 7). The 207Pb/204Pb and 208Pb/204Pb ratios of the galena from the NHF are within or close to the range of values for the Iheya Ridge volcanic rocks (basalt and trachyandesite) and the JADE sediments in the MOT (Figure 7).

5. Discussion

5.1. Seawater–Fluid Mixing and Element Enrichments

The NHF sulfide minerals often contain well-developed pyrrhotite blades, xenomorphic isocubanite, interstitial sphalerite, and sharp-edged galena crystals, which are the result of rapid cooling during seawater–fluid mixing (Figure 2) [10]. Mo-enriched sulfide is characteristic of high-temperature paragenetic mineral associations [61]. The Mo contents (1.21–3.40 ppm) of the pyrrhotite samples from the NHF are mainly higher than those of the sphalerite (0.08–0.34 ppm) and galena (0.07–2.44 ppm) in sample H4-TVG5-1-1 (Table S5; Figure 5). This suggests that the Mo-enrichment of the pyrrhotite may have occurred at high temperatures.
The sphalerite grains from the NHF are Fe-rich (Tables S2 and S3); their FeS mol% contents are significantly higher than those from the PACMANUS hydrothermal field (0.00–2.22) in the eastern Manus Basin, SW Pacific [62]. These values are identical to previously reported values [63] for Vienna Woods (3.35–10.65) and sediment-poor JADE (0.00–1.93) samples and are consistent with those (23.48–50.81) reported for the sediment-rich Hakurei samples from the OT [62]. Such FeS contents indicate low-sulfidation conditions [64]. The sphalerite from the NHF is enriched in Th compared to the pyrrhotite and galena (Table S5; Figure 5). This suggests that the abundance of sphalerite controls the Th enrichment in seafloor hydrothermal sulfides. The Th/U ratios (1.86–4.76) of the pyrrhotite from the NHF are consistent with those of the Iheya Ridge basalt (2.64; Li et al. [39]), trachyandesite (4.13; Li et al. [39]), and the MOT metalliferous sediments (0.18–6.11; Yang et al. [65]), so pyrrhotite serves as the source of the Th and U in the subseafloor hydrothermal fluids. Furthermore, the Ge contents (0.51–97.3 ppm) of the sulfide minerals from the NHF in the OT are higher than those (Ge 0.0055 ppm) of seawater [59]. The positive correlation between the Ge and Pd contents indicates an affinity between these elements and the pyrrhotite (Figure 8a).
The galena from the NHF is commonly associated with sphalerite and was observed to form partial rims around the sphalerite crystals (Figure 2d). The galena inclusions in the sphalerite indicate that the galena post-dates the sphalerite (Keith et al., 2016), and it may have formed through the following reaction: ZnS (s) + Pb2+ (aq) → PbS (s) + Zn2+ (aq). This would result in relatively Zn-rich galena (up to 1.83 wt.%) (Tables S2 and S3) and needs to be confirmed in a future study. However, this indicates a varying fluid chemistry and a change in the solubility as the temperature decreases or increased dilution during seawater–fluid mixing [66]. In the latter case, pyrrhotite would form later than sphalerite/galena, since the fluid is in the interior of both samples H4-TVG5-1-1 and H4-TVG5-3-1.
The Ag contents (0.49–9.28 ppm) of the pyrrhotite from sample H4-TVG5-1-1 (Table S5) are lower than those from the Rainbow hydrothermal vent field on the MAR (200 ppm), the southern JdFR (400–700 ppm), and the Endeavour Segment of the JdFR (<400 ppm) [15,16,17]. In addition, the Ag contents exhibit a positive correlation with the higher Bi contents (Figure 8c).
The Rh (7.52–16.8 ppm), Bi (0.14–1.38 ppm), Ag (693–1940 ppm), Tl (0.49–4.09 ppm), and Sb (1608–5129 ppm) contents of the galena from the NHF are variable and are noticeably higher than the Rh contents of the sulfides from the PACMANUS hydrothermal field (0.0004–0.3556 ppm) in the eastern Manus Basin, southwestern (SW) Pacific [67], the Turtle Pits, and the Logatchev hydrothermal field on the MAR (0.0012–0.149 ppm) [68]. The values of the galena from the NHF suggest that the chemical compositions of the fluids changed with time during the galena deposition. These values are also significantly higher than those of the sphalerite and pyrrhotite from the NHF (Table S5; Figure 5), the vent fluid (Bi = 0.000647 ppm) in the OT [69], and seawater (Bi = 0.00003 ppm, Ag = 0.002 ppm, Tl = 0.013 ppm, Sb = 0.2 ppm) [59]. Furthermore, the enrichment of Rh (7.52–16.8 ppm), Bi (0.14–1.38 ppm), Ag (693–1940 ppm), Tl (0.49–4.09 ppm), and Sb (1608–5129 ppm) in the galena may be the result of the galena precipitating later from late-stage fluids and rapid cooling due to seawater–fluid mixing (Table S5; Figure 5). However, Rh is more compatible in a monosulfide-solid solution (MSS) than in an intermediate solid solution (ISS) and has been found to be more soluble than Pt and Pd [70]. Due to these properties, Rh may crystallize as chalcogen (Se, Te, As, Sb, Bi)-rich discrete precipitate (e.g., RhSeS or RhSbS) before or synchronously with other chalcogen-bearing phases that are associated with late-stage low-sulfur precious metal haloes around hydrothermal sulfides. Similar associations have been documented at various locations around the Sudbury metallogenic deposit, Ontario [71,72]. Consequently, the Ag content is positively correlated with the Sb content for the galena from the NHF and in neighboring sphalerite (Figure 8d), which may also indicate the presence of Sb with the Ag incorporated into the galena from the sphalerite during mineralization. The Rh (7.52–16.8 ppm), Bi (0.14–1.38 ppm), Ag (693–1490 ppm), Tl (0.49–4.09 ppm), and Sb (1608–5129 ppm) contents of the galena are higher than those of the original hydrothermal sphalerite and the earlier pyrrhotite in the NHF (Table S5; Figure 5), which indicates that the incorporation of these trace elements into the galena, pyrrhotite, and sphalerite is controlled by their partition coefficients. This implies that the back-arc galena-rich hydrothermal sulfide deposit may be an Rh-Ag-Sb enriched deposit.
Compared with the PACMANUS hydrothermal mineralization (average Se content of pyrite is 5.97–7.39 ppm; [1]), most of the Noho sulfide minerals are enriched in Se (13.2–109 ppm) (Table S5), which was concentrated in the galena and pyrrhotite via seawater–fluid mixing (Table S5). This is similar to the seafloor hydrothermal mineralization in the Kulo Lasi volcano, Futuna, and SW Pacific. This implies that the Noho basalt-, trachyandesite-, and sediment-hosted hydrothermal mineralization may be enriched in Se [10] based on the fact that it was found in large quantities in the pyrrhotite and based on the reducing conditions during the fluid-rock and/or sediment interaction.
In addition, the Pb contents of the pyrrhotite (up to 1.33 wt.%) and sphalerite (up to 2.93 wt.%) are comparable to other BAB mineralization in the PACMANUS field (0.176 wt.% to 2.35 wt.% Pb in sphalerite) in the SW Pacific [73,74], but they are higher than those in the Trans-Atlantic Geotraverse (TAG) hydrothermal field (<0.1 to 0.16 wt.% Pb in sphalerite) on the MAR [1,75]. However, some of the sphalerite grains exhibit noticeably high Pb contents (up to 2.93 wt.%) (Table S2) which may indicate spectral contamination from nearby galena grains, or true chemical contamination, caused by the diffusion of the chemical components across grain boundaries. The seafloor pyrrhotite (Fe = 58.5–62.0 wt.%; Pb = 0.02–1.33 wt.%; S = 37.8–40.9 wt.%; Zn = 632–5300 ppm; Cu = 130–529 ppm) and galena (Fe = 0.10–1.32 wt.%; Pb = 81.1–88.8 wt.%; S = 12.7–14.5 wt.%; Zn = 829–127,104 ppm; Cu = 12.9–4420 ppm) from the NHF contain different proportions of major (Fe, Pb, and S) and minor (Cu and Zn) elements which may also indicate probable spectral contamination from the adjacent sphalerite and subsurface Cu-sulfide inclusions (isocubanite, up to 30.83 wt.% Cu) (Table S2).

5.2. Variable REE Contents and Sources of REEs

The ∑REEs contents of the pyrrhotite analyzed in this study vary significantly (2.01–36.7 ppm), and they do not exhibit systematic variations with the Fe (58.5–59.8 wt.%) and S (39.5–40.3 wt.%) contents. The degree of LREE and heavy rare earth element (HREE) fractionation in the NHF pyrrhotite (LREE/HREE) is highly variable, up to 32.0 (Table S5). The ∑REE contents and ranges of the pyrrhotite (Table S5) exceed those of the sphalerite. The substitutions of REEs into the pyrrhotite are similar to those involving galena and sphalerite [76,77], indicating the significant influence of the REEs’ larger ionic radii.
The REE patterns of the pyrrhotite and sphalerite from the NHF contain positive Eu anomalies (1.44–4.13) (Table S5; Figure 6), similar to the values for the vent fluids in the OT [57,78] and almost all hydrothermal fields globally. These similarities indicate that the pyrrhotite and sphalerite inherited the positive Eu anomalies of the vent fluids [77,79,80,81]. The stability of the soluble Eu2+ species increases in contact with Cl complexation, low- to high-temperature acidic fluids, and under reducing conditions [77,82,83,84,85,86]. Therefore, the positive Eu anomalies of the pyrrhotite and sphalerite with high Eu content (0.04 to 0.27 ppm) may have been induced by Cl-complexation under high-temperature, low-pH, and strongly reducing fluid conditions [77,81].
The REE patterns of the galena from the NHF are characterized by negative Eu anomalies (0.47–0.83) (Table S5; Figure 6c), which are indicative of relatively low-temperature seawater [87]. A large proportion of the Eu in the low-temperature fluid is trivalent because divalent Eu is stable at temperatures of greater than ~250 °C [83]. The decreasing formation temperature of galena strongly correlates with the decreasing Eu2+/Eu3+ ratios of the vent fluids [77]. The accumulation of Eu2+ in galena formed at medium to low temperatures is also reduced. The stability of the soluble Eu2+ species has been reported to decrease in association with SO42− complexation, high- to low-temperature acidic fluids, and under reducing conditions [83]. Therefore, the negative Eu anomaly of the galena were most likely induced by lower Eu contents, medium (300–200 °C) to low (<200 °C) temperatures (Figure 3), and/or mixing between fluids and seawater and the galena chemistry. Consequently, the Eu content of the fluids may have influenced the Eu anomaly in the precipitated galena and its depositional conditions and processes. The negative Eu anomalies are related to the lower Eu contents of the galena (0.03–0.14 ppm) (Table S5), which have been interpreted to have formed at medium to low temperatures and from less concentrated Eu-bearing- fluids [80,81,88].
The REEs in the sulfide minerals from the NHF may indicate the sources and evolution of the seafloor hydrothermal fluids. Previous studies of seafloor hydrothermal sulfides from the Rainbow, Broken Spur, and TAG hydrothermal fields on the MAR have revealed that the REEs were mainly derived from the fluids [79,89] and were incorporated into the sulfide minerals during seawater–fluid mixing (e.g., Bau and Dulski [90] and Zeng et al. [81]). The REE patterns of the sulfide minerals from the NHF are comparable to those of the Yonaguni Knoll IV vent fluids (Figure 6d) [57]. Thus, the REEs in the sulfide minerals from the NHF were most likely derived from fluids which leached the REEs from the local sub-seafloor volcanic rocks and/or sediments and deposited them in the Noho hydrothermal sulfide deposits (Figure 6d) [58,81,91]. Admittedly, more detailed works focusing on how REE enters sulfides need to be conducted in the future.

5.3. Sources of Sulfur in Sulfide Minerals

The sulfur isotope compositions of the sulfides and their possible geological reservoirs are important indicators that constrain the sources of the reduced sulfur [92,93]. The NHF hydrothermal sulfide deposit is hosted in volcanic rocks and sediments. Igneous rock-derived sulfur, including that in the MORBs, mantle peridotites, and calc-alkaline volcanic rocks (e.g., andesites or rhyolites; Zeng et al. [21]) show δ34S values very close to 0‰ (+0.1 ± 0.5‰) [94,95,96,97]. However, the δ34S values of the Quaternary volcanic rocks from the Japanese Island Arc (+4.4 ± 2.1‰; Ueda and Sakai [98]) and the volcanic rocks from the northern Mariana Trough (+2.0 to +20.7‰; Woodhead et al. [99]) are significantly higher. Although volcanic sulfur is rapidly leached by high-temperature (300–500 °C) fluids [100], no significant isotopic fractionation occurs during the leaching, transport, and reprecipitation of sulfide minerals [101]. The seawater sulfur in the NHF was most likely derived from seawater and/or fluids that leached sulfur from the sediments. Corresponding to volcanic sulfur, seawater sulfate has a δ34S value of +21‰ [102] and H2S-containing seafloor hydrothermal fluids related to sulfide and sulfate mineralization usually have δ34S values of 1.5‰ to 7‰ (e.g., Michard et al. [103]; Butterfield et al. [104]; Shanks et al. [96]; and Shanks [105]). These values are between the δ34S values of volcanic rocks (+0.1 ± 0.5‰; [94,95,96,97], the Japanese Island Arc volcanic rocks (+4.4 ± 2.1‰; Ueda and Sakai [98]), the northern Mariana Trough volcanic rocks (+2.0 to +20.7‰; Woodhead et al. [99]), and sediments (δ34S values similar to those of seawater sulfate, which can be affected by microorganisms >+21‰) and seawater (+21‰). Therefore, the sulfur isotopic compositions of sulfide minerals can serve as evidence of the sources of the sulfur [21,106].
The δ34S values of the SO4 and H2S from the OT vent fluids range from 20.6‰ to 25.7‰ (average of 22.4‰, n = 9) and –0.2‰ to 12‰ (average of 5.9‰, n = 26), respectively [69,78,107,108,109,110,111]. The δ34S values of the sulfide mineral samples from the NHF are all within a narrow range (3.58–5.69‰; Table S6) which overlaps with the overall range for sulfide minerals from global seafloor hydrothermal fields (0.0–9.6‰; n = 1901; Zeng et al. [21]). This indicates a heavy sulfur contribution from reduced seawater sulfate, falling within the range between the seawater δ34S values for SO4 (20.99‰; [102]) and vent fluid H2S (−0.2 to 12‰; [69,78,107,108,109,110,111]).
The sulfur isotope compositions of the sulfide minerals from the NHF may have been influenced by the mixing of seawater and/or sediment SO4 with volcanic rock-derived sulfur during the sulfide formation according to the following reaction:
M2+ (aq) + SO42− (aq) + 4H2S (aq) → MSx (s) + yS (s) + 4H2O (aq)
where M2+ is Fe2+, Pb2+, or Zn2+; and MS is pyrite (x = 2, y = 3), galena (x = 1, y = 4), or sphalerite (x = 1, y = 4).
According to a simple two end-member mixing model,
34Smix = X × δ34Sseawater and/or sediment SO4 + (1 − X) × δ34Svolcanic rock
where X is the amounts of seawater and/or the sediment SO4 component, and δ34Smix, δ34Sseawater and/or sediment SO4 (21‰), and δ34Svolcanic rock (0.1‰, volcanic rock-derived sulfur) are the sulfur isotope compositions of sulfide minerals or vent fluids, seawater and/or sediment SO4 and volcanic rock-derived sulfur, respectively, the sulfur isotope composition of sulfide minerals can be obtained via mixing sulfur from seawater and/or sediment SO4 (17–19%, n = 5) with sulfur from volcanic rocks (81–83%, n = 5). This mixing likely occurred in the downwelling limb of the hydrothermal convection cell in the NHF. This further suggests that the proportions of the sulfur in NHF sulfide minerals contributed by seawater and/or sediment SO4 were less than that contributed by the volcanic rock-derived sulfur. The sulfide minerals in sample H4-TVG5-1-1 precipitated in the NHF primarily where the hydrothermal fluids are dominant along the tube interiors. In addition, disulfides and sulfates tend to be more abundant where mixing with seawater promotes oxidation of S2− in the distal portions of the sample. The δ34S values of the sphalerite from the NHF are higher than those of the galena and pyrrhotite from the NHF (Table S6), and they are significantly lower than those of the sphalerite from the JADE hydrothermal field which has been reported to be +5 to +6‰ [20] and +6.6 to +7.1‰ [112]. These values have been significantly influenced by the reduction in seawater and/or sediment sulfate under highly oxidizing fluid conditions during sphalerite formation [113,114].
In addition, the decrease in the δ34S values from the sphalerite to the galena [115] suggests that the sphalerite-galena pair (point 8-9-3 for sphalerite and point 8-9-5 for galena) in the NHF represent sulfur isotopic equilibrium. Based on the known temperature-dependent fractionation factors [93,116], the formation temperature of the coexisting sphalerite-galena pair was estimated to be 314 °C and 318 °C using the equation for the sulfur isotope geothermometer (Table S6).

5.4. Fluid-Rock Interactions and Source of Lead in Galena

The majority of the Pb isotope ratios of the galena plot were in or near the fields of the basalts and andesites from the NHF and the JADE sediments from the MOT (Figure 7). This suggests that the MOT volcanic rocks and/or sediments are the main source of the Pb in the seafloor vent fluids (e.g., Vidal and Clauer [117]; Chen [118]; Hegner and Tatsumoto [119]; Hinkley and Tatsumoto [120]; Fouquet and Marcoux [121]; Halbach et al. [38]; Charlou et al. [122]; Yao et al. [123]; and Zeng et al. [21]). The range of the Pb isotope compositions of the galena (Table S6) is small compared to those of the volcanic rocks and sediments in the OT. The Pb isotope ratios of the galena from the NHF are the same as those of the volcanic rocks and sediments in the OT, which indicates that the heterogeneous Pb isotopes (e.g., Allègre et al. [124] and Hamelin et al. [125]) were extracted from the volcanic rocks and/or sediments via fluid-rock and/or sediment interactions and were homogenized during fluid circulation, supporting the conclusions of previous studies [21,121,126]. This indicates that the Pb isotope compositions of the galena can be used to infer the Pb isotope compositions of the local volcanic rocks, sediments, and fluids. The Pb isotope ratios of the galena from the NHF are close to the average Pb isotope ratios of the volcanic rocks and sediments in the MOT (Figure 7b), suggesting that the volcanic rock and/or sediments were likely the sources of the Pb in the galena from the NHF. In addition, these ratios plot in or near the field of the west Philippine Sea sediments and altered oceanic crust basalts (Figure 7b) [60], implying that sedimentary and subducted oceanic crust components were a possible source of the Pb in the galena from the MOT.

5.5. Hydrothermal Temperature and Redox Conditions during Sulfide Formation

The equilibrium sulfur isotope temperature (314 °C and 318 °C) of the sphalerite-galena pair (Table S6) suggests that the dominance of sphalerite is a characteristic of the Noho hydrothermal sulfides with formation temperatures of >300 °C. Based on the observed microtextural relationships of the minerals and the interpreted paragenetic sequences (Figure 3), the sulfide mineral crystallization sequence was as follows. First, the pyrrhotite formed, followed by isocubanite and sphalerite. In the final stage, galena, and then amorphous silica is precipitated. The amorphous silica is concluded to have formed during the late stage since it precipitates at low temperatures (i.e., from ~50 °C to <200 °C; Figure 3) [10]. The overall mineral crystallization sequence from pyrrhotite to isocubanite to sphalerite to galena and finally to amorphous silica is indicative of hot, reduced hydrothermal fluids mixing with cooler, more oxygenated seawater [18]. This also emphasizes the importance of the interactions between the mineralizing fluid and the coexisting seawater during the formation of this hydrothermal sulfide deposit in the NHF. However, replacement of fine-grained sulfides with coarse-crystalline sulfides indicates a reaction with hotter and, at least in part, more reduced ascending fluids from which pyrrhotite and sphalerite may also have precipitate [127,128]. In addition, the decrease in the Fe/S ratio of the pyrrhotite from 0.96 to 0.92 (Table S2) indicates an increase in the level of pyrrhotite oxidation (fO2 increase or fS2 decrease) [1].
Interestingly, no pyrite or marcasite was observed in samples H4-TVG5-1-1 or H4-TVG5-3-1. The lack of these minerals sheds important light on the formation processes of the seafloor hydrothermal sulfides in the NHF. For instance, it is known that pyrite is much less soluble than pyrrhotite (approximately 6–10 orders of magnitude less soluble at 200–250 °C; [129]). Just after the pyrrhotite precipitated, the residual Fe precipitated to form Fe-rich sphalerite. This did not leave enough sulfur to form abundant pyrite and/or marcasite. Thus, the lack of pyrite and/or marcasite when pyrrhotite is present merely indicates low sulfur activity. However, according to the mineral paragenesis, the pyrrhotite, isocubanite, and sphalerite were formed during the early ore-forming stage, while the sphalerite, galena, and amorphous silica were formed during the late stage. The pyrrhotite, which represents the earliest mineralization stage, is occasionally overgrown by sphalerite and late-stage galena (Figure 2 and Figure 3).
Even so, the accurate sulfide mineralizations model requires drilling samples and data. Therefore, our TV-grabber collected samples are focused on exploring the element sources (e.g., REE, S, and Pb) and physicochemical conditions (e.g., temperature, fO2, and fS2) of the sulfide mineral formation.

6. Conclusions

The mineral assemblage of the seafloor hydrothermal sulfide ores from the NHF is dominated by pyrrhotite, sphalerite, galena, minor isocubanite, and amorphous silica. The early sulfides (pyrrhotite, isocubanite, and sphalerite) are partly or completely surrounded by the later minerals (sphalerite, galena, and amorphous silica). The late overgrowth of the pyrrhotite by sphalerite and the sphalerite by galena indicates an unstable hydrothermal system.
Most of the measured Rh, Bi, Ag, Tl, Se, and Sb contents of the pyrrhotite and sphalerite are significantly lower than those of the galena, indicating enrichment of these trace elements in the galena during the seawater–fluid mixing. This suggests that the galena-rich hydrothermal sulfide deposit in this BAB may be Rh-Ag-Sb-enriched.
The REEs in the sulfide minerals from the NHF were likely sourced from the seafloor hydrothermal fluids. The REE contents and patterns of the sulfide minerals are related to the mineral chemistry, but they were also influenced by the physiochemical compositions, REE contents and patterns of the fluids, the degree of mixing between the fluids and seawater, and interactions with sub-seafloor rocks and/or sediments. The pyrrhotite and the sphalerite precipitated at higher temperatures in association with acidic or reducing fluids have positive Eu anomalies. Thus, the Eu anomaly values may indicate the Eu content and temperature of the source fluids.
The sphalerite has higher δ34S values than the pyrrhotite or galena. The higher δ34S values of the sulfide minerals from the NHF indicate a heavy sulfur contribution from reduced seawater and/or sediment sulfate, which may indicate that the sulfur in the sulfides was mainly derived from the volcanic rocks.
The Pb isotope compositions of the galena are similar to those of the associated volcanic rocks and/or sediments in the NHF and the JADE sediments in the MOT, suggesting that the galena inherited the Pb isotope compositions of the host rocks in the Noho sub-seafloor hydrothermal systems during the galena formation via seawater-fluid mixing and fluid-rock and/or sediment interactions. The Pb isotope compositions of the galena from the NHF are very homogenous and have a narrow range, plotting in or near the field of the large Pb isotope dataset for the MOT volcanic rocks, western Philippine Sea sediments, and altered oceanic crust basalts. This implies that the isotopic composition of the Pb in the galena is important for understanding the influence of plate subduction on back-arc hydrothermal and magmatic systems.
The formation temperature of the coexisting sphalerite and galena in the NHF was estimated to be 314 °C and 318 °C. The higher sulfur fugacity of the fluid resulted in lower formation temperatures and greater Fe contents in the sphalerite. Altogether, these facts suggest that the sphalerite was deposited under reducing fluid conditions. After the pyrrhotite precipitated, most of the remaining Fe was used for the formation of Fe-rich sphalerite precipitates. These processes consumed all of the available sulfur and prevented the formation of pyrite and/or marcasite.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jmse10050678/s1, Table S1: Sampling data for this study, Table S2: Oxford INCA X-Max energy dispersive spectrometry (EDS) analyses of the sulfide minerals and amorphous silica in samples H4-TVG5-1-1 and H4-TVG5-3-1 from the NHF (wt.%), Table S3: Electron microprobe analyses of the sulfide minerals in sample H4-TVG5-1-1 from the NHF (wt.%) and their atoms per formula unit for elements, Table S4: Detective limit of the trace elements analyzed via LA-ICPMS, and standards of the S-Pb isotopes analyzed via LA-ICPMS. Table S5: Trace element contents of the sulfide minerals in sample H4-TVG5-1-1 from the NHF determined via LA-ICP-MS (in ppm), Table S6: Sulfur and lead isotopic compositions of the sulfide minerals in sample H4-TVG5-1-1 from the NHF determined via LA-MC-ICP-MS and the formation temperature of the coexisting sphalerite and galena.

Author Contributions

Z.Z.—Conceptualization, Data curation, formal analysis, funding acquisition, writing, and editing. Z.C.—data curation, formal analysis, validation. H.Q.—Data curation, methodology, validation. B.Z.—Data curation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the NSFC Major Research Plan for West-Pacific Earth System Multispheric Interactions [project number 91958213], the Strategic Priority Research Program of the Chinese Academy of Sciences [grant number XDB42020402], the National Programme on Global Change and Air-Sea Interaction [grant number GASI-GEOGE-02], the International Partnership Program of the Chinese Academy of Sciences [grant number 133137KYSB20170003], the Special Fund for the Taishan Scholar Program of Shandong Province [grant number ts201511061], and the National Key Basic Research Program of China [grant number 2013CB429700].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data can be obtained from the corresponding author and are available online at Supplementary Materials.

Acknowledgments

We would like to thank the crews of the R/V Kexue during the HOBAB 4 cruise for their help with the sample collection. We are most grateful for the detailed and constructive comments and suggestions provided by two anonymous reviewers, which greatly improved an earlier version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fallon, E.K.; Petersen, S.; Brooker, R.A.; Scott, T.B. Oxidative dissolution of hydrothermal mixed-sulphide ore: An assessment of current knowledge in relation to seafloor massive sulphide mining. Ore Geol. Rev. 2017, 86, 309–337. [Google Scholar] [CrossRef] [Green Version]
  2. Hawley, J.E.; Nichol, I. Trace elements in pyrite, pyrrhotite and chalcopyrite of different ores. Econ. Geol. 1961, 56, 467–487. [Google Scholar] [CrossRef]
  3. Arnold, R. Range in composition and structure of 82 natural terrestrial pyrrhotites. Can. Mineral. 1967, 9, 31–50. [Google Scholar]
  4. Cabri, L.J.; Campbell, J.L.; Laflamme, J.G.; Leigh, R.G.; Maxwell, J.A.; Scott, J.D. Proton-microprobe analysis of trace elements in sulfides from some massive-sulfide deposits. Can. Mineral. 1985, 23, 133–148. [Google Scholar]
  5. Törmänen, T.O.; Koski, R.A. Gold enrichment and the Bi-Au association in pyrrhotite-rich massive sulfide deposits, Escanaba Trough, Southern Gorda Ridge. Econ. Geol. 2005, 100, 1135–1150. [Google Scholar] [CrossRef]
  6. Berkenbosch, H.A.; De Ronde, C.E.J.; Gemmell, J.B.; McNeill, A.W.; Goemann, K. Mineralogy and formation of black smoker chimneys from Brothers submarine volcano, Kermadec arc. Econ. Geol. 2012, 107, 1613–1633. [Google Scholar] [CrossRef]
  7. Tivey, M.; Becker, E.; Beinart, R.; Fisher, C.; Girguis, P.; Langmuir, C.; Michael, P.; Reysenbach, A.-L. Reysenbach. Links from mantle to microbe at the Lau Integrated Study Site: Insights from a back-arc spreading center. Oceanography 2012, 25, 62–77. [Google Scholar] [CrossRef] [Green Version]
  8. Melekestseva, I.Y.; Tret’Yakov, G.A.; Nimis, P.; Yuminov, A.M.; Maslennikov, V.V.; Maslennikova, S.P.; Kotlyarov, V.A.; Beltenev, V.E.; Danyushevsky, L.; Large, R. Barite-rich massive sulfides from the Semenov-1 hydrothermal field (Mid-Atlantic Ridge, 13 30.87′ N): Evidence for phase separation and magmatic input. Mar. Geol. 2014, 349, 37–54. [Google Scholar] [CrossRef]
  9. Evans, G.N.; Tivey, M.K.; Seewald, J.S.; Wheat, C. Wheat. Influences of the Tonga subduction zone on seafloor massive sulfide deposits along the eastern Lau spreading center and Valu Fa Ridge. Geochim. Cosmochim. Acta 2017, 215, 214–246. [Google Scholar] [CrossRef]
  10. Fouquet, Y.; Pelleter, E.; Konn, C.; Chazot, G.; Dupré, S.; Alix, A.-S.; Chéron, S.; Donval, J.-P.; Guyader, V.; Etoubleau, J. Volcanic and hydrothermal processes in submarine calderas: The Kulo Lasi example (SW Pacific). Ore Geol. Rev. 2018, 99, 314–343. [Google Scholar] [CrossRef] [Green Version]
  11. Goodfellow, W.D.; Franklin, J.M. Geology, mineralogy, and chemistry of sediment-hosted clastic massive sulfides in shallow cores, Middle Valley, Northern Juan de Fuca Ridge. Econ. Geol. 1993, 88, 2037–2068. [Google Scholar] [CrossRef]
  12. Bogdanov, Y.A.; Lein, A.Y.; Sagalevich, A.M.; Ul’yanov, A.A.; Dorofeev, S.; Ul’yanova, N.V. Hydrothermal sulfide deposits of the Lucky Strike vent field, Mid-Atlantic Ridge. Geochem. Int. 2006, 44, 403–418. [Google Scholar] [CrossRef]
  13. Bogdanov, Y.A.; Lein, A.Y.; Maslennikov, V.V.; Li, S.; Ul’Yanov, A.A. Mineralogical-geochemical features of sulfide ores from the Broken Spur hydrothermal vent field. Oceanology 2008, 48, 679–700. [Google Scholar] [CrossRef]
  14. Mozgova, N.N.; Trubkin, N.V.; Borodaev, Y.S.; Cherkashev, G.A.; Stepanova, T.V.; Semkova, T.A.; Uspenskaya, T.Y. Mineralogy of massive sulfides from the Ashadze hydrothermal field, 13 N, Mid-Atlantic Ridge. Can. Mineral. 2008, 46, 545–567. [Google Scholar] [CrossRef]
  15. Paradis, S.; Jonasson, I.R.; le Cheminant, G.M.; Watkinsons, D.H. Two zinc-rich chimneys from the Plume Site, southern Juan de Fuca Ridge. Can. Mineral. 1988, 26, 637–654. [Google Scholar]
  16. Tivey, M.K.; Stakes, D.S.; Cook, T.L.; Hannington, M.D.; Petersen, S. A model for growth of steep-sided vent structures on the Endeavour Segment of the Juan de Fuca Ridge: Results of a petrologic and geochemical study. J. Geophys. Res. Solid Earth 1999, 104, 22859–22883. [Google Scholar] [CrossRef]
  17. Marques, A.F.A.; Barriga, F.; Chavagnac, V.; Fouquet, Y. Mineralogy, geochemistry, and Nd isotope composition of the Rainbow hydrothermal field, Mid-Atlantic Ridge. Miner. Depos. 2006, 41, 52–67. [Google Scholar] [CrossRef] [Green Version]
  18. Hannington, M.D.; Jonasson, I.R.; Herzig, P.M.; Petersen, S. Physical and chemical processes of seafloor mineralization at mid-ocean ridges. In Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions; Washington DC American Geophysical Union Geophysical Monograph Series; American Geophysical Union: Washington, DC, USA, 1995; Volume 91, pp. 115–157. [Google Scholar]
  19. Monecke, T.; Petersen, S.; Hannington, M.D.; Grant, H.; Samson, I.M. The minor element endowment of modern sea-floor massive sulfides and comparison with deposits hosted in ancient volcanic successions. Rev. Econ. Geol. 2016, 18, 245–306. [Google Scholar]
  20. Ueno, H. Ore and gangue minerals of sulfide chimneys from the North Knoll, Iheya Ridge, Okinawa Trough, Japan. JAMSTEC Deep-Sea Res. 2003, 22, 49–62. [Google Scholar]
  21. Zeng, Z.; Ma, Y.; Chen, S.; Selby, D.; Wang, X.; Yin, X. Sulfur and lead isotopic compositions of massive sulfides from deep-sea hydrothermal systems: Implications for ore genesis and fluid circulation. Ore Geol. Rev. 2017, 87, 155–171. [Google Scholar] [CrossRef] [Green Version]
  22. Tong, X.; Song, S.X.; He, J. Research on Mineral Processing of Marmatite Ore. Metal Mine 2006, 6, 8–12. [Google Scholar]
  23. Cook, N.J.; Ciobanu, C.L.; Pring, A.; Skinner, W.; Shimizu, M.; Danyushevsky, L.; Saini-Eidukat, B.; Melcher, F. Trace and minor elements in sphalerite: A LA-ICPMS study. Geochim. Cosmochim. Acta 2009, 73, 4761–4791. [Google Scholar] [CrossRef]
  24. Chen, Y.; Chen, J.; Guo, J. A DFT study on the effect of lattice impurities on the electronic structures and floatability of sphalerite. Miner. Eng. 2010, 23, 1120–1130. [Google Scholar] [CrossRef]
  25. Waida, C.; Ueno, H. Ore and gangue minerals of seafloor hydrothermal deposits in the Mariana trough. JAMSTEC Rep. Res. Dev. 2005, 1, 1–12. [Google Scholar]
  26. Suzuki, R.; Ishibashi, J.-I.; Nakaseama, M.; Konno, U.; Tsunogai, U.; Gena, K.; Chiba, H. Diverse range of mineralization induced by phase separation of hydrothermal fluid: Case study of the Yonaguni Knoll IV Hydrothermal Field in the Okinawa Trough Back-Arc Basin. Resour. Geol. 2008, 58, 267–288. [Google Scholar] [CrossRef]
  27. Petersen, S.; Monecke, T.; Westhues, A.; Hannington, M.D.; Gemmell, J.B.; Sharpe, R.; Peters, M.; Strauss, H.; Lackschewitz, K.; Augustin, N.; et al. Drilling shallow-water massive sulfides at the Palinuro volcanic complex, Aeolian island arc, Italy. Econ. Geol. 2014, 109, 2129–2158. [Google Scholar] [CrossRef]
  28. Chen, J.; Wang, L.; Chen, Y.; Guo, J. A DFT study of the effect of natural impurities on the electronic structure of galena. Int. J. Miner. Process. 2011, 98, 132–136. [Google Scholar] [CrossRef]
  29. Petersen, S.; Herzig, P.M.; Schwarz-Schampera, U.; Hannington, M.D.; Jonasson, I.R. Hydrothermal precipitates associated with bimodal volcanism in the Central Bransfield Strait, Antarctica. Miner. Depos. 2004, 39, 358–379. [Google Scholar] [CrossRef]
  30. Lee, C.-S.; Shor, G.G.; Bibee, L.; Lu, R.S.; Hilde, T.W. Okinawa Trough: Origin of a back-arc basin. Mar. Geol. 1980, 35, 219–241. [Google Scholar] [CrossRef]
  31. Sibuet, J.-C.; Deffontaines, B.; Hsu, S.-K.; Thareau, N.; Le Formal, J.-P.; Liu, C.-S. Okinawa trough backarc basin: Early tectonic and magmatic evolution. J. Geophys. Res. Solid Earth 1998, 103, 30245–30267. [Google Scholar] [CrossRef] [Green Version]
  32. Kimura, M. Genesis and formation of the Okinawa Trough, Japan. Mem. Geol. Soc. 1990, 34, 77–88. [Google Scholar]
  33. Shinjo, R.; Kato, Y. Geochemical constraints on the origin of bimodal magmatism at the Okinawa Trough, an incipient back-arc basin. Lithos 2000, 54, 117–137. [Google Scholar] [CrossRef]
  34. Zeng, Z.-G.; Chen, Z.-X.; Zhang, Y.-X. Zhang. Zircon record of an Archaean crustal fragment and supercontinent amalgamation in quaternary back-arc volcanic rocks. Sci. Rep. 2021, 11, 12367. [Google Scholar] [CrossRef] [PubMed]
  35. Sibuet, J.-C.; Letouzey, J.; Barbier, F.; Charvet, J.; Foucher, J.-P.; Hilde, T.W.C.; Kimura, M.; Chiao, L.-Y.; Marsset, B.; Muller, C.; et al. Back arc extension in the Okinawa Trough. J. Geophys. Res. Solid Earth 1987, 92, 14041–14063. [Google Scholar] [CrossRef] [Green Version]
  36. Gena, K.; Chiba, H.; Kase, K.; Nakashima, K.; Ishiyama, D. The Tiger Sulfide Chimney, Yonaguni Knoll IV Hydrothermal Field, Southern Okinawa Trough, Japan: The First Reported Occurrence of Pt-Cu-Fe-Bearing Bismuthinite and Sn-Bearing Chalcopyrite in an Active Seafloor Hydrothermal System. Resour. Geol. 2013, 63, 360–370. [Google Scholar] [CrossRef] [Green Version]
  37. Shinjo, R.; Chung, S.-L.; Kato, Y.; Kimura, M. Geochemical and Sr-Nd isotopic characteristics of volcanic rocks from the Okinawa Trough and Ryukyu Arc: Implications for the evolution of a young, intracontinental back arc basin. J. Geophys. Res. Solid Earth 1999, 104, 10591–10608. [Google Scholar] [CrossRef]
  38. Halbach, P.; Hansmann, W.; Köppel, V.; Pracejus, B. Whole-rock and sulfide lead-isotope data from the hydrothermal JADE field in the Okinawa back-arc trough. Miner. Depos. 1997, 32, 70–78. [Google Scholar] [CrossRef]
  39. Li, X.; Zeng, Z.; Chen, S.; Ma, Y.; Yang, H.; Zhang, Y.; Chen, Z. Geochemical and Sr-Nd-Pb isotopic compositions of volcanic rocks from the Iheya Ridge, the middle Okinawa Trough: Implications for petrogenesis and a mantle source. Acta Oceanolog. Sin. 2018, 37, 73–88. [Google Scholar] [CrossRef]
  40. Ishibashi, J.-I.; Gamo, T.; Sakai, H.; Nojiri, Y.; Igarashi, G.; Shitashima, K.; Tsubota, H. Geochemical evidence for hydrothermal activity in the Okinawa Trough. Geochem. J. 1988, 22, 107–114. [Google Scholar] [CrossRef] [Green Version]
  41. Yamano, M.; Uyeda, S.; Foucher, J.-P.; Sibuet, J.-C. Heat flow anomaly in the middle Okinawa Trough. Tectonophysics 1989, 159, 307–318. [Google Scholar] [CrossRef]
  42. Nakamura, K.; Kawagucci, S.; Kitada, K.; Kumagai, H.; Takai, K.; Okino, K. Water column imaging with multibeam echo-sounding in the mid-Okinawa Trough: Implications for distribution of deep-sea hydrothermal vent sites and the cause of acoustic water column anomaly. Geochem. J. 2015, 49, 579–596. [Google Scholar] [CrossRef] [Green Version]
  43. Halbach, P.; Nakamura, K.-I.; Wahsner, M.; Lange, J.; Sakai, H.; Käselitz, L.; Hansen, R.-D.; Yamano, M.; Post, J.; Prause, B.; et al. Prause. Probable modern analogue of Kuroko-type massive sulphide deposits in the Okinawa Trough back-arc basin. Nature 1989, 338, 496–499. [Google Scholar] [CrossRef]
  44. Sakai, H.; Gamo, T.; Kim, E.-S.; Tsutsumi, M.; Tanaka, T.; Ishibashi, J.; Wakita, H.; Yamano, M.; Oomori, T. Venting of carbon dioxide-rich fluid and hydrate formation in mid-Okinawa trough backarc basin. Science 1990, 248, 1093–1096. [Google Scholar] [CrossRef] [PubMed]
  45. Gamo, T. Growth mechanism of the hydrothermal mounds at the CLAM Site, Mid Okinawa Trough, inferred from their morphological, mineralogical and chemical characteristics. JAMSTEC Deep Sea Res. 1991, 7, 163–184. [Google Scholar]
  46. Kawagucci, S.; Toki, T.; Ishibashi, J.; Takai, K.; Ito, M.; Oomori, T.; Gamo, T. Isotopic variation of molecular hydrogen in 20–375 C hydrothermal fluids as detected by a new analytical method. J. Geophys. Res. Biogeosci. 2010, 115, G03021. [Google Scholar] [CrossRef] [Green Version]
  47. Ishibashi, J.-I.; Ikegami, F.; Tsuji, T.; Urabe, T. Hydrothermal activity in the Okinawa Trough back-arc basin: Geological background and hydrothermal mineralization. In Subseafloor Biosphere Linked to Hydrothermal Systems: TAIGA Concept; Ishibashi, J.-I., Okino, K., Sunamura, M., Eds.; Springer: Tokyo, Japan, 2015; pp. 337–359. [Google Scholar]
  48. Chiba, M.; Koizumi, A.; Oshika, J.; Tanahashi, M.; Ueda, S.; Ishikawa, N.; Okamoto, N.; Ishizuka, O.; Shimoda, G. Discovery of Sulfide Mounds in the Okinawa Trough; Abstracts with Programs; The Society of Resource: Tokyo, Japan, 2014; Volume 64, p. 24. [Google Scholar]
  49. Liu, Y.; Hu, Z.; Gao, S.; Günther, D.; Xu, J.; Gao, C.; Chen, H. In situ analysis of major and trace elements of anhydrous minerals by LA-ICP-MS without applying an internal standard. Chem. Geol. 2008, 257, 34–43. [Google Scholar] [CrossRef]
  50. Bao, Z.; Chen, L.; Zong, C.; Yuan, H.; Chen, K.; Dai, M. Development of pressed sulfide powder tablets for in situ sulfur and lead isotope measurement using LA-MC-ICP-MS. Int. J. Mass Spectrom. 2017, 421, 255–262. [Google Scholar] [CrossRef]
  51. Chen, L.; Chen, K.; Bao, Z.; Liang, P.; Sun, T.; Yuan, H. Preparation of standards for in situ sulfur isotope measurement in sulfides using femtosecond laser ablation MC-ICP-MS. J. Anal. At. Spectrom. 2017, 32, 107–116. [Google Scholar] [CrossRef]
  52. Yuan, H.; Liu, X.; Chen, L.; Bao, Z.; Chen, K.; Zong, C.; Li, X.-C.; Qiu, J.W. Simultaneous measurement of sulfur and lead isotopes in sulfides using nanosecond laser ablation coupled with two multi-collector inductively coupled plasma mass spectrometers. J. Asian Earth Sci. 2018, 154, 386–396. [Google Scholar] [CrossRef]
  53. Chen, K.; Yuan, H.; Bao, Z.; Zong, C.; Dai, M. Precise and accurate in situ determination of lead isotope ratios in NIST, USGS, MPI-DING and CGSG glass reference materials using femtosecond laser ablation MC-ICP-MS. Geostand. Geoanal. Res. 2014, 38, 5–21. [Google Scholar]
  54. Yuan, H.; Yin, C.; Liu, X.; Chen, K.; Bao, Z.; Zong, C.; Dai, M.; Lai, S.; Wang, R.; Jiang, S. High precision in-situ Pb isotopic analysis of sulfide minerals by femtosecond laser ablation multi-collector inductively coupled plasma mass spectrometry. Sci. China Earth Sci. 2015, 58, 1713–1721. [Google Scholar] [CrossRef]
  55. Belzile, N.; Chen, Y.-W.; Cai, M.; Li, Y. A review on pyrrhotite oxidation. J. Geochem. Explor. 2004, 84, 65–76. [Google Scholar] [CrossRef]
  56. Evensen, N.; Hamilton, P.; O’Nions, R. Rare-earth abundances in chondritic meteorites. Geochim. Cosmochim. Acta 1978, 42, 1199–1212. [Google Scholar] [CrossRef]
  57. Hongo, Y.; Obata, H.; Gamo, T.; Nakaseama, M.; Ishibashi, J.; Konno, U.; Saegusa, S.; Ohkubo, S.; Tsunogai, U. Rare Earth Elements in the hydrothermal system at Okinawa Trough back-arc basin. Geochem. J. 2007, 41, 1–15. [Google Scholar] [CrossRef] [Green Version]
  58. Xu, Z.; Lim, D.; Li, T.; Kim, S.; Jung, H.; Wan, S.; Chang, F.; Cai, M. REEs and Sr-Nd isotope variations in a 20 ky-sediment core from the middle Okinawa Trough, East China Sea: An in-depth provenance analysis of siliciclastic components. Mar. Geol. 2019, 415, 105970. [Google Scholar] [CrossRef]
  59. Steele, J.H.; Thorpe, S.A.; Turekian, K.K. Marine Chemistry and Geochemistry: A Derivative of Encyclopedia of Ocean Sciences, 2nd ed.; Elsevier Science: London, UK, 2010; pp. 601–602. [Google Scholar]
  60. Shu, Y.; Nielsen, S.G.; Zeng, Z.; Shinjo, R.; Blusztajn, J.; Wang, X.; Chen, S. Tracing subducted sediment inputs to the Ryukyu arc-Okinawa Trough system: Evidence from thallium isotopes. Geochim. Cosmochim. Acta 2017, 217, 462–491. [Google Scholar] [CrossRef]
  61. Hannington, M.; Herzig, P.; Scott, S.; Thompson, G.; Rona, P. Comparative mineralogy and geochemistry of gold-bearing sulfide deposits on the mid-ocean ridges. Mar. Geol. 1991, 101, 217–248. [Google Scholar] [CrossRef]
  62. Kawasumi, S.; Chiba, H. Redox state of seafloor hydrothermal fluids and its effect on sulfide mineralization. Chem. Geol. 2017, 451, 25–37. [Google Scholar] [CrossRef]
  63. Moss, R.; Scott, S.D. Geochemistry and mineralogy of gold-rich hydrothermal precipitates from the eastern Manus Basin, Papua New Guinea. Can. Mineral. 2001, 39, 957–978. [Google Scholar] [CrossRef]
  64. Scott, S.D.; Kissin, S.A. Sphalerite composition in the Zn-Fe-S system below 300 degrees C. Econ. Geol. 1973, 68, 475–479. [Google Scholar] [CrossRef]
  65. Yang, B.; Liu, J.; Shi, X.; Zhang, H.; Wang, X.; Wu, Y.; Fang, X. Mineralogy and sulfur isotope characteristics of metalliferous sediments from the Tangyin hydrothermal field in the southern Okinawa Trough. Ore Geol. Rev. 2020, 120, 103464. [Google Scholar] [CrossRef]
  66. Keith, M.; Häckel, F.; Haase, K.M.; Schwarz-Schampera, U.; Klemd, R. Trace element systematics of pyrite from submarine hydrothermal vents. Ore Geol. Rev. 2016, 72, 728–745. [Google Scholar] [CrossRef]
  67. Pašava, J.; Vymazalová, A.; Petersen, S.; Herzig, P. PGE distribution in massive sulfides from the PACMANUS hydrothermal field, eastern Manus basin, Papua New Guinea: Implications for PGE enrichment in some ancient volcanogenic massive sulfide deposits. Miner. Depos. 2004, 39, 784–792. [Google Scholar] [CrossRef]
  68. Pašava, J.; Vymazalová, A.; Petersen, S. PGE fractionation in seafloor hydrothermal systems: Examples from mafic-and ultramafic-hosted hydrothermal fields at the slow-spreading Mid-Atlantic Ridge. Miner. Depos. 2007, 42, 423–431. [Google Scholar] [CrossRef]
  69. Gamo, T. Wide variation of chemical characteristics of submarine hydrothermal fluids due to secondary modification processes after high temperature water-rock interaction: A review. In Biogeochemical Processes and Ocean Flux in the Western Pacific; Sakai, H., Nozaki, Y., Eds.; Terra Scientific Publishing Company (TERRAPUB): Tokyo, Japan, 1995; pp. 425–451. [Google Scholar]
  70. Chen, L.-M.; Song, X.-Y.; Danyushevsky, L.; Wang, Y.-S.; Tian, Y.-L.; Xiao, J.-F. A laser ablation ICP-MS study of platinum-group and chalcophile elements in base metal sulfide minerals of the Jinchuan Ni–Cu sulfide deposit, NW China. Ore Geol. Rev. 2015, 65, 955–967. [Google Scholar] [CrossRef]
  71. Farrow, C.E.; Watkinson, D.H. Diversity of precious-metal mineralization in footwall Cu-Ni-PGE deposits, Sudbury, Ontario; implications for hydrothermal models of formation. Can. Mineral. 1997, 35, 817–839. [Google Scholar]
  72. Liu, Y.; Brenan, J. Partitioning of platinum-group elements (PGE) and chalcogens (Se, Te, As, Sb, Bi) between monosulfide-solid solution (MSS), intermediate solid solution (ISS) and sulfide liquid at controlled fO2fS2 conditions. Geochim. Cosmochim. Acta 2015, 159, 139–161. [Google Scholar] [CrossRef]
  73. Binns, R.A.; Scott, S.D. Actively forming polymetallic sulfide deposits associated with felsic volcanic rocks in the eastern Manus back-arc basin, Papua New Guinea. Econ. Geol. 1993, 88, 2226–2236. [Google Scholar] [CrossRef]
  74. Wohlgemuth-Ueberwasser, C.C.; Viljoen, F.; Petersen, S.; Vorster, C. Distribution and solubility limits of trace elements in hydrothermal black smoker sulfides: An in-situ LA-ICP-MS study. Geochim. Cosmochim. Acta 2015, 159, 16–41. [Google Scholar] [CrossRef]
  75. Tivey, M.; Humphris, S.E.; Thompson, G.; Hannington, M.D.; Rona, P.A. Deducing patterns of fluid flow and mixing within the TAG active hydrothermal mound using mineralogical and geochemical data. J. Geophys. Res. Solid Earth 1995, 100, 12527–12555. [Google Scholar] [CrossRef]
  76. Alt, J.C. The chemistry and sulfur isotope composition of massive sulfide and associated deposits on Green Seamount, eastern Pacific. Econ. Geol. 1988, 83, 1026–1033. [Google Scholar] [CrossRef]
  77. Mills, R.; Elderfield, H. Rare earth element geochemistry of hydrothermal deposits from the active TAG Mound, 26 N Mid-Atlantic Ridge. Geochim. Cosmochim. Acta 1995, 59, 3511–3524. [Google Scholar] [CrossRef]
  78. Ishibashi, J.-I.; Urabe, T. Hydrothermal activity related to arc-backarc magmatism in the western Pacific. In Backarc Basins; Springer: New York, NY, USA, 1995; pp. 451–495. [Google Scholar]
  79. Barrett, T.J.; Jarvis, I.; Jarvis, K.E. Rare earth element geochemistry of massive sulfides-sulfates and gossans on the Southern Explorer Ridge. Geology 1990, 18, 583–586. [Google Scholar] [CrossRef]
  80. Gillis, K.M.; Smith, A.D.; Ludden, J.N. Trace element and Sr isotopic contents of hydrothermal clays and sulfides from the Snakepit hydrothermal field: ODP site 649. Proc. ODP Sci. Res. 1990, 106, 315–319. [Google Scholar]
  81. Zeng, Z.; Ma, Y.; Yin, X.; Selby, D.; Kong, F.; Chen, S. Factors affecting the rare earth element compositions in massive sulfides from deep-sea hydrothermal systems. Geochem. Geophys. Geosyst. 2015, 16, 2679–2693. [Google Scholar] [CrossRef] [Green Version]
  82. Schade, J.; Cornell, D.H.; Theart, H.F.J. Rare earth element and isotopic evidence for the genesis of the Prieska massive sulfide deposit, South Africa. Econ. Geol. 1989, 84, 49–63. [Google Scholar] [CrossRef]
  83. Sverjensky, D.A. Europium redox equilibria in aqueous solution. Earth Planet. Sci. Lett. 1984, 67, 70–78. [Google Scholar] [CrossRef]
  84. Wood, S.A.; Williams-Jones, A.E. The aqueous geochemistry of the rare-earth elements and yttrium 4. Monazite solubility and REE mobility in exhalative massive sulfide-depositing environments. Chem. Geol. 1994, 115, 47–60. [Google Scholar] [CrossRef]
  85. Haas, J.R.; Shock, E.L.; Sassani, D.C. Rare earth elements in hydrothermal systems: Estimates of standard partial molal thermodynamic properties of aqueous complexes of the rare earth elements at high pressures and temperatures. Geochim. Cosmochim. Acta 1995, 59, 4329–4350. [Google Scholar] [CrossRef]
  86. Allen, D.E.; Seyfried, W., Jr. REE controls in ultramafic hosted MOR hydrothermal systems: An experimental study at elevated temperature and pressure. Geochim. Cosmochim. Acta 2005, 69, 675–683. [Google Scholar] [CrossRef]
  87. De Baar, H.J.; Brewer, P.G.; Bacon, M.P. Anomalies in rare earth distributions in seawater: Gd and Tb. Geochim. Cosmochim. Acta 1985, 49, 1961–1969. [Google Scholar] [CrossRef]
  88. Michard, A.; Albarede, F. The REE content of some hydrothermal fluids. Chem. Geol. 1986, 55, 51–60. [Google Scholar] [CrossRef]
  89. Rimskaya-Korsakova, M.N.; Dubinin, A.V. Rare earth elements in sulfides of submarine hydrothermal vents of the Atlantic Ocean. Dokl. Earth Sci. 2003, 389, 432–436. [Google Scholar]
  90. Bau, M.; Dulski, P. Comparing yttrium and rare earths in hydrothermal fluids from the Mid-Atlantic Ridge: Implications for Y and REE behaviour during near-vent mixing and for the Y/Ho ratio of Proterozoic seawater. Chem. Geol. 1999, 155, 77–90. [Google Scholar] [CrossRef]
  91. Langmuir, C.; Humphris, S.; Fornari, D.; Van Dover, C.; Von Damm, K.; Tivey, M.; Colodner, D.; Charlou, J.-L.; Desonie, D.; Wilson, C.; et al. Hydrothermal vents near a mantle hot spot: The Lucky Strike vent field at 37 N on the Mid-Atlantic Ridge. Earth Planet. Sci. Lett. 1997, 148, 69–91. [Google Scholar] [CrossRef]
  92. Rye, R.O.; Ohmoto, H. Sulfur and carbon isotopes and ore genesis: A review. Econ. Geol. 1974, 69, 826–842. [Google Scholar] [CrossRef]
  93. Ohomoto, H.; Rye, R.O. Isotopes of sulfur and carbon. In Geochemistry of Hydrothermal Ore Deposits, 2nd ed.; Barnes, H.L., Ed.; J Wiley and Sons: New York, NY, USA, 1979; pp. 509–567. [Google Scholar]
  94. Sakai, H.; Marais, D.; Ueda, A.; Moore, J. Concentrations and isotope ratios of carbon, nitrogen and sulfur in ocean-floor basalts. Geochim. Cosmochim. Acta 1984, 48, 2433–2441. [Google Scholar] [CrossRef]
  95. Alt, J.C.; Anderson, T.F.; Bonnell, L. The geochemistry of sulfur in a 1.3 km section of hydrothermally altered oceanic crust, DSDP Hole 504B. Geochim. Cosmochim. Acta 1989, 53, 1011–1023. [Google Scholar] [CrossRef]
  96. Shanks, W.C.; Böhlke, J.K.; Seal, R.R. Stable isotopes in mid-ocean ridge hydrothermal systems: Interactions between fluids, minerals, and organisms. Geophys. Monogr. Am. Geophys. Union 1995, 91, 194. [Google Scholar]
  97. Alt, J.C.; Shanks, W.C., III. Serpentinization of abyssal peridotites from the MARK area, Mid-Atlantic Ridge: Sulfur geochemistry and reaction modeling. Geochim. Cosmochim. Acta 2003, 67, 641–653. [Google Scholar] [CrossRef]
  98. Ueda, A.; Sakai, H. Sulfur isotope study of Quaternary volcanic rocks from the Japanese Islands Arc. Geochim. Cosmochim. Acta 1984, 48, 1837–1848. [Google Scholar] [CrossRef]
  99. Woodhead, J.D.; Harmon, R.S.; Fraser, D. O, S, Sr, and Pb isotope variations in volcanic rocks from the Northern Mariana Islands: Implications for crustal recycling in intra-oceanic arcs. Earth Planet. Sci. Lett. 1987, 83, 39–52. [Google Scholar] [CrossRef]
  100. Mottl, M.J.; Holland, H.D.; Corr, R.F. Chemical exchange during hydrothermal alteration of basalt by seawater—II. Experimental results for Fe, Mn, and sulfur species. Geochim. Cosmochim. Acta 1979, 43, 869–884. [Google Scholar] [CrossRef]
  101. Shanks, W.; Jeffrey, N. Sulfur isotope studies of hydrothermal anhydrite and pyrite, Deep Sea Drilling Project leg 64, Guaymas Basin, Gulf of California. Init. Rep. Deep Sea Drill. Proj. 1982, 64, 1137–1142. [Google Scholar]
  102. Rees, C.; Jenkins, W.; Monster, J. The sulphur isotopic composition of ocean water sulphate. Geochim. Cosmochim. Acta 1978, 42, 377–381. [Google Scholar] [CrossRef]
  103. Michard, G.; Albarede, F.; Michard, A.; Minster, J.-F.; Charlou, J.-L.; Tan, N. Chemistry of solutions from the 13 N East Pacific Rise hydrothermal site. Earth Planet. Sci. Lett. 1984, 67, 297–307. [Google Scholar] [CrossRef]
  104. Butterfield, D.; Massoth, G.J.; McDuff, R.E.; Lupton, J.E.; Lilley, M.D. Geochemistry of hydrothermal fluids from Axial Seamount hydrothermal emissions study vent field, Juan de Fuca Ridge: Subseafloor boiling and subsequent fluid-rock interaction. J. Geophys. Res. Solid Earth 1990, 95, 12895–12921. [Google Scholar] [CrossRef]
  105. Shanks, W.C., III. Stable isotopes in seafloor hydrothermal systems: Vent fluids, hydrothermal deposits, hydrothermal alteration, and microbial processes. Rev. Mineral. Geochem. 2001, 43, 469–525. [Google Scholar] [CrossRef]
  106. Fouquet, Y.; Knott, R.; Cambon, P.; Fallick, A.; Rickard, D.; Desbruyères, D. Formation of large sulfide mineral deposits along fast spreading ridges. Example from off-axial deposits at 12 43′ N on the East Pacific Rise. Earth Planet. Sci. Lett. 1996, 144, 147–162. [Google Scholar] [CrossRef]
  107. Gamo, T.; Sakai, H.; Kim, E.-S.; Shitashima, K.; Ishibashi, J.-I. High alkalinity due to sulfate reduction in the CLAM hydrothermal field, Okinawa Trough. Earth Planet. Sci. Lett. 1991, 107, 328–338. [Google Scholar] [CrossRef]
  108. Nakagawa, S.; Takai, K.; Inagaki, F.; Chiba, H.; Ishibashi, J.-I.; Kataoka, S.; Hirayama, H.; Nunoura, T.; Horikoshi, K.; Sako, Y. Variability in microbial community and venting chemistry in a sediment-hosted backarc hydrothermal system: Impacts of subseafloor phase-separation. FEMS Microbiol. Ecol. 2005, 54, 141–155. [Google Scholar] [CrossRef] [PubMed]
  109. Kawagucci, S.; Chiba, H.; Ishibashi, J.-I.; Yamanaka, T.; Toki, T.; Muramatsu, Y.; Ueno, Y.; Makabe, A.; Inoue, K.; Yoshida, N.; et al. Hydrothermal fluid geochemistry at the Iheya North field in the mid-Okinawa Trough: Implication for origin of methane in subseafloor fluid circulation systems. Geochem. J. 2011, 45, 109–124. [Google Scholar] [CrossRef] [Green Version]
  110. Kawagucci, S. Fluid geochemistry of high-temperature hydrothermal fields in the Okinawa Trough. In Subseafloor Biosphere Linked to Hydrothermal Systems: TAIGA Concept; Ishibashi, J.-I., Okino, K., Sunamura, M., Eds.; Springer: Tokyo, Japan, 2015; pp. 387–403. [Google Scholar]
  111. Takai, K.; Nakagawa, S.; Nunoura, T. Comparative investigation of microbial communities associated with hydrothermal activities in the Okinawa trough. In Subseafloor Biosphere Linked to Hydrothermal Systems: TAIGA Concept; Ishibashi, J.-I., Okino, K., Sunamura, M., Eds.; Springer: Tokyo, Japan, 2015; pp. 421–435. [Google Scholar]
  112. Lüders, V.; Pracejus, B.; Halbach, P. Fluid inclusion and sulfur isotope studies in probable modern analogue Kuroko-type ores from the JADE hydrothermal field (Central Okinawa Trough, Japan). Chem. Geol. 2001, 173, 45–58. [Google Scholar] [CrossRef]
  113. Alt, J.C.; Shanks, W.C.; Bach, W.; Paulick, H.; Garrido, C.J.; Beaudoin, G. Hydrothermal alteration and microbial sulfate reduction in peridotite and gabbro exposed by detachment faulting at the Mid-Atlantic Ridge, 15 20′ N (ODP Leg 209): A sulfur and oxygen isotope study. Geochem. Geophys. Geosyst. 2007, 8, Q08002. [Google Scholar] [CrossRef]
  114. Delacour, A.; Früh-Green, G.L.; Bernasconi, S.M.; Kelley, D.S. Sulfur in peridotites and gabbros at Lost City (30 N, MAR): Implications for hydrothermal alteration and microbial activity during serpentinization. Geochim. Cosmochim. Acta 2008, 72, 5090–5110. [Google Scholar] [CrossRef]
  115. Sakai, H. Isotopic properties of sulfur compounds in hydrothermal processes. Geochem. J. 1968, 2, 29–49. [Google Scholar] [CrossRef]
  116. Li, Y.; Liu, J. Calculation of sulfur isotope fractionation in sulfides. Geochim. Cosmochim. Acta 2006, 70, 1789–1795. [Google Scholar] [CrossRef]
  117. Vidal, P.; Clauer, N. Pb and Sr isotopic systematics of some basalts and sulfides from the East Pacific Rise at 21 N (project RITA). Earth Planet. Sci. Lett. 1981, 55, 237–246. [Google Scholar] [CrossRef]
  118. Chen, J. U, Th, and Pb isotopes in hot springs on the Juan de Fuca Ridge. J. Geophys. Res. Solid Earth 1987, 92, 11411–11415. [Google Scholar] [CrossRef]
  119. Hegner, E.; Tatsumoto, M. Pb, Sr, and Nd isotopes in basalts and sulfides from the Juan de Fuca Ridge. J. Geophys. Res. Solid Earth 1987, 92, 11380–11386. [Google Scholar] [CrossRef]
  120. Hinkley, T.K.; Tatsumoto, M. Metals and isotopes in Juan de Fuca Ridge hydrothermal fluids and their associated solid materials. J. Geophys. Res. Solid Earth 1987, 92, 11400–11410. [Google Scholar] [CrossRef]
  121. Fouquet, Y.; Marcoux, E. Lead isotope systematics in Pacific hydrothermal sulfide deposits. J. Geophys. Res. Solid Earth 1995, 100, 6025–6040. [Google Scholar] [CrossRef]
  122. Charlou, J.; Donval, J.; Fouquet, Y.; Jean-Baptiste, P.; Holm, N. Geochemistry of high H2 and CH4 vent fluids issuing from ultramafic rocks at the Rainbow hydrothermal field (36 14′ N, MAR). Chem. Geol. 2002, 191, 345–359. [Google Scholar] [CrossRef]
  123. Yao, H.-Q.; Zhou, H.-Y.; Peng, X.-T.; Bao, S.-X.; Wu, Z.-J.; Li, J.-T.; Sun, Z.-L.; Chen, Z.-Q.; Li, J.-W.; Chen, G.-Q. Metal sources of black smoker chimneys, Endeavour Segment, Juan de Fuca Ridge: Pb isotope constraints. Appl. Geochem. 2009, 24, 1971–1977. [Google Scholar] [CrossRef]
  124. Allègre, C.J.; Hamelin, B.; Dupré, B. Statistical analysis of isotopic ratios in MORB: The mantle blob cluster model and the convective regime of the mantle. Earth Planet. Sci. Lett. 1984, 71, 71–84. [Google Scholar] [CrossRef]
  125. Hamelin, B.; Dupré, B.; Allègre, C.J. Lead-strontium isotopic variations along the East Pacific Rise and the Mid-Atlantic Ridge: A comparative study. Earth Planet. Sci. Lett. 1984, 67, 340–350. [Google Scholar] [CrossRef]
  126. Andrieu, A.; Honnorez, J.; Lancelot, J. Lead isotope compositions of the TAG mineralization, Mid-Atlantic Ridge, 26 08′ N. In Ocean Drilling Program; Scientific Results; Elsevier Science: Amsterdam, The Netherlands, 1998; pp. 101–109. [Google Scholar]
  127. Thompson, G.; Humphris, S.E.; Schroeder, B.; Sulanowska, M.; Rona, P.A. Active vents and massive sulfides at 26 degrees N (TAG) and 23 degrees N (Snakepit) on the Mid-Atlantic Ridge. Can. Mineral. 1988, 26, 697–711. [Google Scholar]
  128. Davis, A.S.; Clague, D.A.; Zierenberg, R.; Wheat, C.; Cousens, B.L. Sulfide formation related to changes in the hydrothermal system on Loihi Seamount, Hawai’i, following the seismic event in 1996. Can. Mineral. 2003, 41, 457–472. [Google Scholar] [CrossRef]
  129. Murowchick, J.B. Marcasite inversion and the petrographic determination of pyrite ancestry. Econ. Geol. 1992, 87, 1141–1152. [Google Scholar] [CrossRef]
Figure 1. Locations of seafloor hydrothermal sulfide samples H4-TVG5-1-1 and H4-TVG5-3-1 in the NHF and trachyandesite samples H4-TVG5-2 (Li et al. [39]) near the Iheya Ridge in the OT.
Figure 1. Locations of seafloor hydrothermal sulfide samples H4-TVG5-1-1 and H4-TVG5-3-1 in the NHF and trachyandesite samples H4-TVG5-2 (Li et al. [39]) near the Iheya Ridge in the OT.
Jmse 10 00678 g001
Figure 2. (a) Brown H4-TVG5-1-1 sample with a fluid channel and pores; (b) grey-green H4-TVG5-3-1 sample with reddish-brown Fe-Mn oxides; (c) bladed pyrrhotite (Pyh) surrounded by galena (Gn); (d) pyrrhotite and black-grey sphalerite with a lattice texture and later coarse and fine galena filling the cavities; (e) fine isocubanite (Icb) inclusion in the galena and amorphous silica (Sil); (f) pyrrhotite (Pyh) aggregates surrounded by sphalerite (Sp) and galena (Gn). Oxidized pyrrhotite appears as a thin whitish rim around the pyrrhotite (Pyh) lath; (g) coarse bladed pyrrhotite and fine isocubanite (Icb) surrounded by black-grey sphalerite; and (h) pyrrhotite partially surrounded by galena and black-grey sphalerite with fine isocubanite (Icb).
Figure 2. (a) Brown H4-TVG5-1-1 sample with a fluid channel and pores; (b) grey-green H4-TVG5-3-1 sample with reddish-brown Fe-Mn oxides; (c) bladed pyrrhotite (Pyh) surrounded by galena (Gn); (d) pyrrhotite and black-grey sphalerite with a lattice texture and later coarse and fine galena filling the cavities; (e) fine isocubanite (Icb) inclusion in the galena and amorphous silica (Sil); (f) pyrrhotite (Pyh) aggregates surrounded by sphalerite (Sp) and galena (Gn). Oxidized pyrrhotite appears as a thin whitish rim around the pyrrhotite (Pyh) lath; (g) coarse bladed pyrrhotite and fine isocubanite (Icb) surrounded by black-grey sphalerite; and (h) pyrrhotite partially surrounded by galena and black-grey sphalerite with fine isocubanite (Icb).
Jmse 10 00678 g002
Figure 3. Sequence of sulfide mineralization in the seafloor hydrothermal sulfide deposits of the NHF in the MOT. The thicknesses of the horizontal bars indicate the relative abundances of the minerals. Pyh means pyrrhotite.
Figure 3. Sequence of sulfide mineralization in the seafloor hydrothermal sulfide deposits of the NHF in the MOT. The thicknesses of the horizontal bars indicate the relative abundances of the minerals. Pyh means pyrrhotite.
Jmse 10 00678 g003
Figure 4. SEM BSE images of selected pyrrhotite (Pyh), sphalerite (Sp), and galena (Gn) grains from the seafloor hydrothermal sulfide deposits in the NHF: (a) SEM and EDS line analysis of pyrrhotite, sphalerite, and galena in sample H4-TVG5-3-1. The pyrrhotite was oxidized with a high O content. The Mn and S contents of the sphalerite vary; (b) SEM and EDS line analysis of the sphalerite and galena in sample H4-TVG5-3-1.
Figure 4. SEM BSE images of selected pyrrhotite (Pyh), sphalerite (Sp), and galena (Gn) grains from the seafloor hydrothermal sulfide deposits in the NHF: (a) SEM and EDS line analysis of pyrrhotite, sphalerite, and galena in sample H4-TVG5-3-1. The pyrrhotite was oxidized with a high O content. The Mn and S contents of the sphalerite vary; (b) SEM and EDS line analysis of the sphalerite and galena in sample H4-TVG5-3-1.
Jmse 10 00678 g004
Figure 5. Average element contents of sulfide minerals from the NHF hydrothermal sulfide samples. The element contents of the pyrrhotite, sphalerite, and galena were determined via LA-ICP-MS.
Figure 5. Average element contents of sulfide minerals from the NHF hydrothermal sulfide samples. The element contents of the pyrrhotite, sphalerite, and galena were determined via LA-ICP-MS.
Jmse 10 00678 g005
Figure 6. REE patterns of the sulfide minerals in the hydrothermal sulfide samples including pyrrhotite (a), sphalerite (b), and galena (c) from the NHF in the MOT compared to those of volcanic rocks, vent fluids, sediments, and seawater (d). The normalization data are from Evensen et al. [56]. The data for the Iheya Ridge volcanic rocks, Yonaguni Knoll IV vent fluids, MOT sediments, and seawater data are from Li et al. [39], Hongo et al. [57], Xu et al. [58], and Steele et al. [59]. The red dash lines are the detection limit of REEs.
Figure 6. REE patterns of the sulfide minerals in the hydrothermal sulfide samples including pyrrhotite (a), sphalerite (b), and galena (c) from the NHF in the MOT compared to those of volcanic rocks, vent fluids, sediments, and seawater (d). The normalization data are from Evensen et al. [56]. The data for the Iheya Ridge volcanic rocks, Yonaguni Knoll IV vent fluids, MOT sediments, and seawater data are from Li et al. [39], Hongo et al. [57], Xu et al. [58], and Steele et al. [59]. The red dash lines are the detection limit of REEs.
Jmse 10 00678 g006
Figure 7. Plot of (a) 208Pb/204Pb vs. 206Pb/204Pb, and (b) 208Pb/206Pb vs. 207Pb/206Pb for the galena from the NHF. The data for the NOT, MOT, and SOT volcanic rocks are from Halbach et al. [38], Shu et al. [60], and Li et al. [39]. The data for the sediments from DSDP Sites 294, 295, 296, 442B, 443, and 444 in the western Philippine sea, the JADE sediments in the MOT, and the altered oceanic crust basalt from DSDP Sites 294 and 442B are from Halbach et al. [38] and Shu et al. [60]. As we have compared our in situ Pb isotope data to the end members, to avoid being hard to understand, we do not add the previous whole-rock Pb isotopes of sulfides to the same figure.
Figure 7. Plot of (a) 208Pb/204Pb vs. 206Pb/204Pb, and (b) 208Pb/206Pb vs. 207Pb/206Pb for the galena from the NHF. The data for the NOT, MOT, and SOT volcanic rocks are from Halbach et al. [38], Shu et al. [60], and Li et al. [39]. The data for the sediments from DSDP Sites 294, 295, 296, 442B, 443, and 444 in the western Philippine sea, the JADE sediments in the MOT, and the altered oceanic crust basalt from DSDP Sites 294 and 442B are from Halbach et al. [38] and Shu et al. [60]. As we have compared our in situ Pb isotope data to the end members, to avoid being hard to understand, we do not add the previous whole-rock Pb isotopes of sulfides to the same figure.
Jmse 10 00678 g007
Figure 8. (a) Plot of Pd vs. Ge for pyrrhotite; (b) Plot of Mo vs. Au for galena; (c) Plot of Bi vs. Ag for pyrrhotite; and (d) Plot of Ag vs. Sb for galena and sphalerite from the NHF in the MOT.
Figure 8. (a) Plot of Pd vs. Ge for pyrrhotite; (b) Plot of Mo vs. Au for galena; (c) Plot of Bi vs. Ag for pyrrhotite; and (d) Plot of Ag vs. Sb for galena and sphalerite from the NHF in the MOT.
Jmse 10 00678 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zeng, Z.; Chen, Z.; Qi, H.; Zhu, B. Chemical and Isotopic Composition of Sulfide Minerals from the Noho Hydrothermal Field in the Okinawa Trough. J. Mar. Sci. Eng. 2022, 10, 678. https://doi.org/10.3390/jmse10050678

AMA Style

Zeng Z, Chen Z, Qi H, Zhu B. Chemical and Isotopic Composition of Sulfide Minerals from the Noho Hydrothermal Field in the Okinawa Trough. Journal of Marine Science and Engineering. 2022; 10(5):678. https://doi.org/10.3390/jmse10050678

Chicago/Turabian Style

Zeng, Zhigang, Zuxing Chen, Haiyan Qi, and Bowen Zhu. 2022. "Chemical and Isotopic Composition of Sulfide Minerals from the Noho Hydrothermal Field in the Okinawa Trough" Journal of Marine Science and Engineering 10, no. 5: 678. https://doi.org/10.3390/jmse10050678

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

Article Metrics

Back to TopTop