Next Article in Journal
Relationship of the Warming of Red Sea Surface Water over 140 Years with External Heat Elements
Next Article in Special Issue
The Geochemical Features and Genesis of Ferromanganese Deposits from Caiwei Guyot, Northwestern Pacific Ocean
Previous Article in Journal
Microalgae-Based PUFAs for Food and Feed: Current Applications, Future Possibilities, and Constraints
Previous Article in Special Issue
A Novel Device for the In Situ Enrichment of Gold from Submarine Venting Fluids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strontium, Hydrogen and Oxygen Behavior in Vent Fluids and Plumes from the Kueishantao Hydrothermal Field Offshore Northeast Taiwan: Constrained by Fluid Processes

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 266237, China
3
University of Chinese Academy of Sciences, Beijing 101408, China
4
Center for Ocean Mega-Science, Chinese Academy of Sciences, Qingdao 266071, China
5
Department of Oceanography, National Sun Yat-Sen University, Kaohsiung 80424, Taiwan, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2022, 10(7), 845; https://doi.org/10.3390/jmse10070845
Submission received: 7 May 2022 / Revised: 9 June 2022 / Accepted: 10 June 2022 / Published: 21 June 2022

Abstract

:
Strontium (Sr), hydrogen (H) and oxygen (O) in vent fluids are important for understanding the water–rock interaction and hydrothermal flux in hydrothermal systems. We have analyzed the Sr, H and O isotopic compositions of seawater, vent fluid and hydrothermal plume samples in the Kueishantao hydrothermal field, as well as their calcium (Ca), total sulfur (S), Sr, arsenic (As), stibium (Sb), chlorine (Cl) and manganese (Mn) concentrations for understanding the origin and processes of fluids. The results suggest that most As, Sb and Mn are leached from andesitic rocks into the fluids, and most Ca and Cl remained in the deep reaction zone during the fluid–andesitic rock interaction. The ranges of 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values in the yellow spring, white spring and plumes are small. The 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of fluids and plumes are like those of ambient seawater, indicating that the Sr, H and O of vent fluids and hydrothermal plumes are derived primarily from seawater. This suggests that the interaction of andesite and subseafloor fluid is of short duration and results in the majority of As, Sb and Mn being released into fluids, while most Ca and Cl remained in the deep reaction zone. In addition, there was no significant variation of Sr, H and O isotopic compositions in the upwelling fluid, keeping the similar isotopic compositions of seawater. There are obvious correlations among the pH values, As and Sb concentrations, and H isotopic compositions of the vent fluids and hydrothermal plumes, implying that the As and Sb concentrations and H isotopic compositions can trace the dispersion of plumes in the ambient seawater. According to the Sr concentrations and 87Sr/86Sr values, the water/rock ratios are 3076~8124, which is consistent with the idea that the interaction between fluid and andesite at the subseafloor is of short duration. The hydrothermal flux of Sr discharged from the yellow spring into the seawater is between 2.06 × 104 and 2.26 × 104 mol/yr, and the white spring discharges 1.18 × 104~1.26 × 104 mol/yr Sr if just andesites appear in the reaction zone.

1. Introduction

Extensive strontium (Sr), hydrogen (H), and oxygen (O) isotopic and chemical exchange depending on fluid temperatures, hydrothermal alteration and the water/rock ratios can be caused by hydrothermal circulation [1,2,3,4,5]. The intensity of the water–rock interaction has been recorded by these isotopes in the hydrothermal fluids, and they are particularly useful in the research of fluid origin and hydrothermal processes in subseafloor hydrothermal systems [1,6,7]. The mobile Sr, H and O isotopes are good markers of host rock composition and subseafloor alteration processes in hydrothermal systems [2,4,8,9]. Sr, H and O in hydrothermal fluids come from different sources, relying on the hydrothermal circulation system, seawater and fluid mixing, host rocks and magma [2,4,10,11]. Their isotopic ratios are used to identify different sources (rock, seawater and sediment) [2,4,12] and mixing versus conductive cooling or heating processes and may be valuable for constraining the phase separation conditions in seafloor hydrothermal systems.
However, it has been implied that the high-temperature hydrothermal circulation is insufficient to balance the other Sr input fluxes, as an increase of ~5.4 × 10−5 Myr−1 [13] has been observed in oceanic 87Sr/86Sr over the past few million years [14,15,16]. Many explanations have been put forward to explain the missing unradiogenic 87Sr/86Sr flux, such as hydrothermal systems occurring on the ridge flank [16], low-temperature alteration processes and the precipitation of carbonate [17,18]. The reactions happening in the hydrothermal systems are of great importance and are believed to be the main source of unradiogenic 87Sr/86Sr in seawater; however, the Sr flux into the oceans from high- and particularly low-temperature hydrothermal systems remain poorly quantified [14,17].
During subcritical phase separation, H isotope fractionation has been shown to depend on both temperature and salinity [19]. Furthermore, H isotope ratios of fluids are relatively insensitive to reactions with the host rock because H is abundant in the fluid relative to the mineral content. However, the H isotope fractionation data of brine and vapor coexisting in the two-phase supercritical region of seawater is completely lacking. The H isotope fractionation related to the phase separation of NaCl-H2O fluid has only been obtained up to 350 °C [19]. Although the temperature of the vent fluid is generally close to 350 °C [20], phase separation occurs in the deep reaction zones at temperatures significantly higher than this. Many hydrothermal experiments studying the fractionation of seawater phase separation at 450 °C had obtained H isotopic fractionation factors, and these data were then used to explain the phase separation and segregation in deep-sea hydrothermal systems [2].
The first O isotope studies of black smoker hydrothermal fluids were published by the East Pacific Rise Study Group [21] and Welhan and Craig [22], which showed that the δ18O values are close to that of ambient seawater; however, they increase by 1–2‰ owing to the reaction between basalt and seawater at elevated temperatures. Furthermore, Teagle et al. [23] used Sr and O isotope systematics to confirm that anhydrite on the Trans-Atlantic Geotraverse (TAG) Mound forms through the mixing between fluids and seawater, which is conductively heated to 100–180 °C before mixing; thus, anhydrite forms by a mixture of heated seawater and vent fluid at 230–320 °C. According to Bach and Humphris [24], the 87Sr/86Sr and δ18O of the vent fluids on the oceanic ridge show a global correction with the rate of spreading (or the rate of magma supply). They believe that at low spreading rates (or low magma supply rates), there is significantly less Sr from seawater and higher δ18O from seawater in vent fluids. This is due to a greater exchange of Sr and O with the crust during hydrothermal fluid circulation, possibly because of longer reaction paths under slow-spreading ridges. Using the datasets of hydrothermal vent fluids associated with ridges, linear regression equations and R2 values show that the δ18O proposed by Bach and Humphris [24] is not statistically related to the spreading rate.
Field and experimental studies, as well as isotopic exchange computations [25,26,27,28,29,30,31], have clearly shown that both O and H isotope values increase due to water/rock interactions with the igneous crust. Due to the decrease of water/rock mass ratios caused by the evolution of hydrothermal fluids, the O and H isotope values of end-member vent fluids follow the calculated seawater/basalt reaction vector [25]. For slower spreading ridges, higher δD values are consistent with longer fluid paths and more fluid-rock interactions. However, δ18O values would be even more responsive to fluid–rock interactions and have insignificant variation with the rate of spreading. With the increase of the hydrothermal vent fluid database, the short-term time variation in each field becomes more and more important.
The chemical characteristics of deep-sea hydrothermal fluids from mid-ocean ridges and back-arc basins, including Sr, H and O isotopic compositions, have been described by [20,32,33,34,35]. However, little is known about the chemical and Sr, H and O isotopic compositions of shallow hydrothermal vents fluids. The quantification of global hydrothermal flux into the ocean is also biased due to the lack of budget caused by the interaction of seawater and andesite. To better understand the fluid–andesitic rock interaction, the fluid and plume samples in the Kueishantao shallow hydrothermal field were collected and analyzed. The main purposes of this study are (1) to investigate the geochemistry characteristics and evolution of hydrothermal fluids and their plumes, (2) to study the rock–fluid interactions recorded by fluids and plumes, and (3) to assess the hydrothermal flux of Sr in the shallow hydrothermal field.

2. Geologic Setting

The Kueishantao shallow hydrothermal field (121°55′ E, 24°50′ N) is in the southeast of Kueishantao, northeastern Taiwan, and near the southern Okinawa Trough. The hydrothermal field covers an area of about 0.5 square kilometers (Figure 1). Kueishantao’s last major eruption occurred about 7000 years ago [36], and the area around the hydrothermal vents is characterized by andesitic lava and pyroclastic flows. The andesitic magma is thought to be produced by a MORB-type magma with an assimilated 30% local continental crust [37]. The difference in Fe-Cu-Zn isotopic compositions between the Kueishantao andesites and the MORBs and the continental crust might indicate entrainment of carbonate sediment components into the andesitic magma [38].
The Kueishantao hydrothermal field has more than 30 hydrothermal vents at water depths of 10–30 m. The hydrothermal products in the form of chimneys, mounds and balls are mainly composed of natural sulfur. On 12 August 2000, a large yellow chimney approximately 6 m high was discovered at a water depth of 20 m. The geochemical study of the native sulfur chimneys and balls indicated that trace elements are derived primarily from the andesite and partly from seawater, and the interaction between the subseafloor fluid and the andesite has a short duration [39,40].
Figure 1. (a) The bathymetric map of Taiwan and the Okinawa Trough (based on [41]). (b) The tectonic map with the location of Kueishantao island (from [42]). (c) The location of the yellow spring (yellow star, 108 °C) and white spring (red star, 51 °C) in the Kueishantao hydrothermal field (based on [41]).
Figure 1. (a) The bathymetric map of Taiwan and the Okinawa Trough (based on [41]). (b) The tectonic map with the location of Kueishantao island (from [42]). (c) The location of the yellow spring (yellow star, 108 °C) and white spring (red star, 51 °C) in the Kueishantao hydrothermal field (based on [41]).
Jmse 10 00845 g001
There are two types of vents, “yellow spring” (78~116 °C) and “white spring” (30~65 °C) [42]. The yellow spring fluids are characterized by very low pH (~1.52) and a wide range of chemical compositions. The white spring fluids have a relatively low content of methane, iron and copper [42]. The temperature of the vent fluids varied with tides and reached the maximum temperature approximately 3.5 h after each high tide [42,43,44]. The Ca, Mg and K are significantly correlated with each other, indicating the same source of seawater [45]. The gas component is mainly CO2, followed by N2, CH4 and H2S [46].

3. Sampling and Methods

3.1. Sample Collection

On 31 May 2011, the samples of hydrothermal fluids and plumes were collected from the Kueishantao hydrothermal field by divers (Table 1, Figure 1). Four liter Pyrex bottles were used to sample the hydrothermal fluids at and in the vent. Two-valve polyethylene tubes were also employed to sample the fluids in order to assess the reliability of the data gained by the Pyrex bottles. The hydrothermal plume was sampled in 1 L Nalgene polypropylene bottles at approximately 2, 5 and 7 m below sea level above the yellow spring vent (water depth, 7.2 m) and approximately 5, 10, 13 and 15 m below sea level above the white spring vent (water depth, 15.1 m). Shallow seawater was sampled at a depth of 10 m near Kueiwei to eliminate the effect of hydrothermal activity. The fluid fluxes and temperatures were determined in situ. Kuo [43] described in detail the methods of fluid collection and the measurement of temperature and flux in situ.

3.2. Analytical Methods

In the lab on the shore, the pH of the liquid was analyzed by a portable pH meter with a resolution of 0.01 (JENCO 6010, San Diego, CA, USA). Before measurement, calibrate the pH meter with buffer solutions of pH 4.00 and 6.86 (i.e., 0.05 mol/L potassium hydrogen phthalate (25 °C, pH = 4.00) and 0.025 mol/L mixed phosphate (25 °C, pH = 6.86) buffers). All liquids were filtered into 1 L Nalgene polypropylene bottles, which were previously soaked in 1:1 HNO3 for 48 h, washed with distilled water, and then dried.
Calcium (Ca), total sulfur (S) and Sr were analyzed by the inductively coupled plasma optical emission spectrometer at the Shandong Institute of Geophysical and Geochemical Exploration, and the precision was <±5%. Arsenic (As) and stibium (Sb) were determined using atomic fluorescence spectrophotometer at the Qingdao Institute of Marine Geology, China Geological Survey Bureau. Manganese (Mn) was analyzed by inductively coupled plasma mass spectrometer, and chlorine (Cl) was determined using ion chromatography at the Institute of Oceanology, Chinese Academy of Sciences. The detailed chemical separation and measurement procedures of As, Sb, Mn and Cl were described in previous studies [47]. Seawater standard NASS-5 was used. The analytical precision of As, Sb and Mn was better than 5%, and that of Cl was ±1%.
The 87Sr/86Sr was measured using thermal ionization mass spectrometry (TIMS; Triton) at the State Key Laboratory of Geological Processes and Mineral Resources, China University of Geosciences. The procedures of chemical separation and determination were described in detail in previous studies [48,49]. For Sr, the procedural blanks were less than 50 pg. The determination of the NBS987 Sr isotopic standard gained over one year yielded an average value of 87Sr/86Sr = 0.710266 ± 0.000009 (2σ, n = 38). Rock standard AGV-2 was used for 87Sr/86Sr. The analytical precision was <0.006%.
H and O isotopic ratios were measured by the stable isotope ratio mass spectrometer (MAT-253, Thermo Fisher, USA) at the Institute of Mineral Resources, Chinese Academy of Geological Sciences. Fluid samples were prepared for H isotope analyses using the Zn-reduction method and for O isotope analyses using the CO2 equilibration technique. Values of H and O isotopes are calibrated relative to the V-SMOW standard. The analytical precision of the δD and δ18O are better than 1.0‰ and 0.1‰, respectively.

4. Results

We report the element (Ca, total S, Sr, As, Sb, Cl and Mn) concentrations and 87Sr/86Sr, δD and δ18Ocompositions of fluid and plume samples collected from the Kueishantao hydrothermal field. At both yellow and white spring sites, the hydrothermal fluids are released directly from the andesite host rock due to an absence of sediment cover. Figure 1 displays the sample locations and element concentrations, and the 87Sr/86Sr, δD and δ18O values are listed in Table 2.
The Ca, total S, Sr As, Sb, Cl and Mn concentrations in the yellow spring varied from 381.3 mg/L, 892 mg/L, 6.37 mg/L, 5.59 µg/L, 0.29 µg/L, 20,814 mg/L and 10.8 µg/L at 0 mbsl (meters below sea level) to 371.2 mg/L, 810 mg/L, 5.87 mg/L, 22.4 µg/L, 0.77 µg/L, 18,052 mg/L and 4.55 µg/L at 7.2 mbsl (Figure 2), respectively. The Ca, total S, Sr As, Sb, Cl and Mn concentrations in the white spring varied from 381.3 mg/L, 832 mg/L, 6.35 mg/L, 5.65 µg/L, 0.27 µg/L, 19,068 mg/L and 6.24 µg/L at 0 mbsl to 369.4 mg/L, 963 mg/L, 6.12 mg/L, 15.8 µg/L, 0.36 µg/L, 16,951 mg/L and 23.3 µg/L at 15.1 mbsl (Figure 2), respectively.
Compared with the Ca, Sr As and Cl concentrations in the yellow spring hydrothermal plumes (381.3 to 385.0 mg/L, 6.34 to 6.37 mg/L, 3.81 to 6.01 µg/L and 20,422 to 20,814 mg/L), those in the white spring hydrothermal plumes have a slightly wider range of variation (376.2 to 383.0 mg/L, 6.27 to 6.35 mg/L, 4.08 to 14.5 µg/L and 18,925 to 20,454 mg/L) than in (Figure 3). Most of the total S and Sr concentrations for both hydrothermal plumes are slightly lower than that of shallow seawater (835 and 6.62 mg/L) (Figure 2). The Ca, Sr and Cl concentrations of most hydrothermal fluids are clearly lower than that of shallow seawater and consistent with those (Cl 15,811–19,888 mg/L) of [39,42].
The δDV-SMOW and δ18OV-SMOW values in the yellow spring range from 1‰ and 0.1‰ in fluid to 5‰ and 0.2‰ in the plume, and the 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values in the white spring range from 0.709154, −4‰, and 0.2‰ in fluid to 0.709202, 2‰ and 0.2‰ in the plume (Table 2). The 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of the two hydrothermal plumes were similar (Table 2), and near the 0.709177, 2‰ and 0.1‰ values of ambient seawater (Figure 3). The 87Sr/86Sr values of the white spring hydrothermal fluids are distinctly lower than those of shallow seawater and the white spring hydrothermal plume (Figure 3).
The hydrothermal fluids have a pH range of 2.29 to 5.11, and those of the plumes are 5.51~6.15. The pH values of the hydrothermal fluids and plumes are distinctly lower than that of shallow seawater (8.02), and from the hydrothermal fluids to the hydrothermal plumes, the pH value increases gradually as the water depth decreases (Figure 2). The As, Sb and Mn concentrations and δDV-SMOW values of the hydrothermal fluids and plumes display a good positive and negative correlation with pH values (Figure 4).
From vent fluids to hydrothermal plumes, the pH values, As and Sb concentrations and H isotope compositions have a good correlation, and the concentrations of As and Sb decrease with the increase of pH value (Figure 4), which is like the boron (B) concentrations and isotopic compositions [42], indicating that the As and Sb concentrations and H isotopic compositions can be used to trace the variation of chemical characteristics during the hydrothermal plumes that disperse in the seawater.
Moreover, from the surface to the water depth of 5 m, the Cl and total S concentrations and δ18OV-SMOW values in the hydrothermal plumes are similar (Figure 2), which is due to the mixing of the hydrothermal plumes with shallow seawater (Figure 4B,C).

5. Discussion

5.1. Sources of Strontium, Hydrogen and Oxygen

Because Sr, H and O isotopes in seawater and andesite are different, the Sr, H and O in hydrothermal fluids might be derived from two components: seawater and andesite. In the Kueishantao hydrothermal field, the 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values in ambient seawater are 0.709177, 2‰ and 0.1‰ with a Sr concentration of 6.62 mg/L, and an average Sr content in the andesite is 206 ppm (Table 2). The andesite in the Kueishantao hydrothermal field have a large range of 87Sr/86Sr (from 0.70577 to 0.70688) and δ18OSMOW (from 7.1‰ to 7.9‰) values, with Sr concentrations of 221–317 ppm [37].
However, the hydrothermal plumes have similar values of 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW to those of the shallow seawater (Table 2) (Figure 3), and the 87Sr/86Sr ratios of the hydrothermal fluids and plumes in the Kueishantao hydrothermal field were significantly higher than those (from 0.70382 to 0.7056) of hydrothermal fluids from the Eastern Lau spreading center [35], suggesting that the hydrothermal plumes and fluids are derived primarily from seawater.
A simple two-end-member mixing model using an equation:
Mmix = X × Mseawater + (1 − X) × Mandesite
where X is the amount of the seawater component; M is the 87Sr/86Sr and δ18O; Mmix, Mseawater (87Sr/86Srseawater 0.709177, δ18Oseawater 0.1‰) and Mandesite (87Sr/86Sr andesite from 0.70577 to 0.70688, δ18Oandesite from 7.1‰ to 7.9‰) are the Sr and O isotopic compositions of hydrothermal plumes and fluids, seawater and andesite, respectively. The model shows that approximately 99% and 97% of the Sr and O are derived from seawater (Table 3).

5.2. Seawater-Rock Interaction

Sr, H and O behave conservatively, as their behavior during hydrothermal alteration demonstrates [6]. The geochemistry of Sr, H and O in submarine hydrothermal systems has been used to constrain subseafloor fluid-rock interaction processes [23,30]. When high-temperature fluids reacted with fresh basalt in mid-ocean ridges [50,51], Sr is believed to be extracted from basalt quantitatively during hydrothermal reactions without subsequent re-equilibration with secondary mineral phases [6,23,30], while O and H may be primarily derived from seawater. The relative concentrations of Sr, As, Sb and Mn in the hydrothermal fluids could reflect those in the volcanic rocks where they were leached [52,53].
Past studies of seafloor hydrothermal fields imply that hydrothermal fluids often have similar 87Sr/86Sr, δD and δ18O values to those of the nearby rocks (87Sr/86Sr) and seawater (δD and δ18O), indicating the leaching of Sr from the host rocks or mixing of H and O from seawater [6,23,30]. Therefore, the Sr, H and O isotopic compositions of hydrothermal fluids can be used to clarify different source rocks.
In both springs, the 87Sr/86Sr values of the hydrothermal fluids are obviously higher than that in hydrothermal fluids at the mid-ocean ridge (the East Pacific Rise near 9°N 0.70381–0.70790, [7]) and back-arc basin (Lau basin 0.70382–0.7056, [20]) (Figure 3), and the 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of Kueishantao vent fluids and hydrothermal plumes are like those of the ambient seawater, indicating a short-term interaction between subseafloor fluids and andesite.
Most As, Sb and Mn concentrations in hydrothermal fluids are much higher than those in sea level hydrothermal plumes, which have lower temperatures. However, the high As, Sb and Mn concentrations cannot be produced solely by local seawater contributions. The apparently higher concentrations of As, Sb and Mn in the hydrothermal fluids may show an extra source from the andesite basement. The seawater-andesite interaction will result in higher exchangeable As, Sb and Mn concentrations. The high As, Sb and Mn concentrations of most hydrothermal fluids (above seawater, Table 2) partly reflect the constraints by the andesite, which reacts with source fluids. These As-, Sb- and Mn-enriched fluids in both springs have low pH values and can be interpreted by the interaction between seawater and andesite, which explains the difference between Kueishantao and mid-ocean ridge hydrothermal systems. The 87Sr/86Sr values in the Kueishantao hydrothermal fluids are obviously higher than those in the hydrothermal fluids on the Valu Fa ridge in the Lau basin (Figure 3), indicating a short interaction between the andesite basement and seawater.
In the Kueishantao hydrothermal field, the concentrations of As, Sb and Mn in the hydrothermal plume and fluid are enriched relative to those in the seawater. This is different from the basalt-hosted hydrothermal systems in which As, Sb and Mn losses are caused by the entrainment of seawater during the formation of altered minerals at low temperatures, and subsequently, the mobilization of As, Sb and Mn in the high-temperature reaction zone [52,53], usually resulting in the slight enrichment of As, Sb and Mn in high-temperature hydrothermal fluids. The concentrations of As, Sb and Mn in the Kueishantao hydrothermal system are like those in the back-arc basins. At these two locations, the weak water–rock interaction results in the smaller removal of As, Sb and Mn in the hydrothermal fluid circulation. However, the unique enrichment of As, Sb and Mn in the hydrothermal fluids is a feature of hydrothermal systems with andesite basements because trace As, Sb and Mn have been leached out by seawater.
We calculated the water/rock ratios of the fluid-rock reaction as follows:
W/R = ((87Sr/86Srh87Sr/86Srandesite) × Srandesite)/((87Sr/86Srsw87Sr/86Srh) × Srsw)
where 87Sr/86Srh, 87Sr/86Srandesite and 87Sr/86Srsw (0.709177) are the isotopic ratios in hydrothermal fluids, andesite and seawater, respectively, and Srandesite (206 ppm) and Srsw (6.62 mg/L) are the Sr concentrations of andesite and seawater, respectively. The water/rock ratios for 87Sr/86Sr andesite = 0.70577 and 0.70688 are listed in Table 3, ranging from 4578 to 8124 and 3076 to 5467, respectively. The significant difference between these results and those (approximately 1.6) from fluids from the DSDP Hole 504B, Costa Rica Rift reported by Kawahata et al. [6] is ascribed to the shorter fluid pathways and seawater–rock interaction. However, the Sr/Ca ratio of KHF hydrothermal fluids and plumes (0.0165–0.0171, n = 15) is like that (0.0178) of ambient seawater and significantly higher than that (0.0042–0.0068, n = 35) of KHF andesites, which indicates the precipitation of anhydrite is negligible in the subseafloor recharge zone. This suggests that the fluids in the Kueishantao hydrothermal field have weaker water and rock interactions and shorter reaction paths, and the isotopic fractionation seems to decrease accompanying the increasing water/rock ratio, resulting in the 87Sr/86Sr, δD and δ18O values of upwelling fluids in the Kueishantao hydrothermal field like those of ambient seawater.

5.3. From Vent Fluid to Plume

The Ca and Cl concentrations of vent fluids are less than those of hydrothermal plumes and ambient seawater (Table 2), and the Ca and Cl concentrations of hydrothermal plumes are higher than those of shallow seawater. This suggests that the mixing of hydrothermal fluids and seawater leads to the enrichment of Ca and Cl in the hydrothermal plumes, and most Ca and Cl are retained in the deep reaction zone during the fluid-andesite interaction. Thus, the resulting Ca and Cl concentrations of upwelling and vent fluids are less than that of shallow seawater.
The As, Sb and Mn concentrations of most vent fluids are higher than those of hydrothermal plumes and shallow seawater (Table 2), and the As, Sb and Mn concentrations of most vent fluids are higher than those of hydrothermal plumes. This suggests that As, Sb and Mn in the hydrothermal fluids are leached from andesite during the fluid–rock interaction, and the mixing of the hydrothermal fluid and shallow seawater leads to the leaching of the As, Sb and Mn in the fluids into hydrothermal plumes.
In a plot of pH versus As, Sb and δDV-SMOW, the linear correlation (Figure 4) indicates that the hydrothermal plumes are a two-component mixture of seawater and fluid. Therefore, the As, Sb and δDV-SMOW values of the Kueishantao hydrothermal fluids and plumes can be interpreted as the result of mixing seawater As, Sb and δDV-SMOW with fluid As and Sb extracted from the andesitic rocks.

5.4. Strontium Flux

The hydrothermal circulation under the sea floor leads to extensive chemical and isotopic exchange due to changes in fluid temperature. For Sr, the reaction direction changes from uptake by the rock at low temperature to release from the rock at high temperature [51]. However, the Sr and Cl are clearly like those of the hydrothermal fluids in the North Cleft segment on the Juan de Fuca Ridge [54] and those vented from the Plume vent at the South Cleft [55] (Figure 2). The Sr concentrations and 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of the white spring (51 °C) are like those of the yellow spring (108 °C), indicating that the exchange magnitude of Sr, H and O between the andesite and seawater is consistent within these two types of springs.
For calculating the Sr flux, it was presumed that the yellow and white springs have stable flow rates and that andesite is the only host rock in the reaction zone. Using measured Sr concentrations in hydrothermal fluids, the Sr flux calculated for the white spring vent into the ocean is between 2.06 × 104 and 2.26 × 104 mol/yr, and the Sr flux for the yellow spring vent is between 1.18 × 104 and 1.26 × 104 mol/yr (Table 3). Therefore, the Sr flux from the yellow spring is slightly higher than that from the white spring. Assuming that more than 30 vents occur in the Kueishantao hydrothermal field [42], the hydrothermal Sr flux is about 6 × 105 mol/yr with Fsrsw (fraction of Sr from seawater) values from 0.990 to 0.996, much higher than the Fsrsw values of 0.083—0.318 at hydrothermal vent fields on the Lau basin, Mariana Trough, Mid-Atlantic Ridge, Juan de Fuca Ridge and East Pacific Rise [24]. The difference also shows that there is a short lifetime of subseafloor fluid–rock interaction and/or short reaction paths at the Kueishantao hydrothermal field.

6. Conclusions

The element concentrations and 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of the Kueishantao hydrothermal fluids and plumes have smaller ranges of variation than those of fluids released from back-arc basins and mid-ocean ridges. The 87Sr/86Sr values and Sr concentrations of fluids range from 0.709154 to 0.709164 and 5.87 to 6.56 mg/L, respectively. For the plumes, the 87Sr/86Sr values range from 0.709147 to 0.709202, and the Sr concentrations range from 6.27 to 6.37 mg/L. From hydrothermal fluid to hydrothermal plume, the As, Sb and Mn concentrations decrease with the reduction of water depth, whereas Cl and Ca concentrations tend to increase. The As, Sb and Mn concentrations of the hydrothermal fluids and plumes of both yellow and white springs are higher than those of shallow seawater (As, Sb and Mn are 1.15, 0.20 and 0.91 mg/L), and the 87Sr/86Sr values of most vent fluids of the white spring are obviously less than that of shallow seawater. The variations of As, Sb and Mn concentration and δDV-SMOW values of the hydrothermal plumes can be interpreted by the mixing of vent fluids with seawater. Using As, Sb and H isotopes of hydrothermal plumes can describe their diffusive processes.
Sr, H and O in Kueishantao vent fluids are primarily derived from the ambient seawater. The As, Sb, Mn, Ca and Cl concentrations of hydrothermal fluid are influenced slightly by weak interactions between seawater and rock in the deep reaction zones, which is characterized by leaching of As, Sb and Mn from andesite. The hydrothermal flux of Sr in the Kueishantao hydrothermal field is between 1.18 × 104 and 2.26 × 104 mol/yr. Compared with the fluxes in other hydrothermal fields, the low flux might be caused by lower formation temperatures, higher seawater/andesite ratios (3076–8124), the shorter duration of seawater–andesite interaction and/or the shorter fluid paths.
However, most As, Sb, and Mn in the Kueishantao hydrothermal fluids are leached from andesitic rocks into the fluids. Most Ca and Cl remained in the reaction zone of the subseafloor hydrothermal field during seawater and rock interactions. At present, the sampling numbers of hydrothermal fluids, plumes and andesite are not enough. In the future, it will be possible to understand the material and heat flux of hydrothermal fluids and plumes formed by subseafloor magma chambers to seawater and their effect on seawater and ecologic environments by carrying out long-term continual observations.

Author Contributions

Conceptualization, data curation, formal analysis, funding acquisition, writing and editing, Z.Z.; methodology, formal analysis, data curation, writing—review and editing, X.W.; methodology, writing—review and editing, X.Y. and S.C.; data curation, methodology, validation, H.Q.; writing—review and editing, C.-T.A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the NSFC Major Research Plan on West-Pacific Earth System Multi-spheric Interactions (grant number 91958213), the Strategic Priority Research Program of the Chinese Academy of Sciences (grant number XDB42020402), the Special Fund for the Taishan Scholar Program of Shandong Province (grant number ts201511061), the National Key Basic Research Program of China (grant number 2013CB429700), and the Science and Technology Development Program of Shandong Province (grant number 2013GRC31502).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data that support the findings of this study are given in the main text.

Acknowledgments

We thank Bing-Jye Wang and Sea-watch Company for sampling the hydrothermal fluids and plumes.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shanks, W.C., III; Seyfried, W.E., Jr. Stable isotope studies of vent fluids and chimney minerals, southern Juan de Fuca Ridge: Sodium metasomatism and seawater sulfate reduction. J. Geophys. Res. 1987, 92, 11387–11399. [Google Scholar] [CrossRef]
  2. Berndt, M.E.; Seal, R.R.; Shanks, W.C.; Seyfried, W.E., Jr. Hydrogen isotope systematics of phase separation in submarine hydrothermal systems: Experimental calibration and theoretical models. Geochim. Cosmochim. Acta 1996, 60, 1595–1604. [Google Scholar] [CrossRef]
  3. Liakhovitch, V.; Quick, J.E.; Gregory, R.T. Hydrogen and Oxygen Isotope Constraints on Hydrothermal Alteration of the Trinity Peridotite, Klamath Mountains, California. Int. Geol. Rev. 2005, 47, 203–214. [Google Scholar] [CrossRef]
  4. Gena, K.R.; Chiba, H.; Mizuta, T.; Matsubaya, O. Hydrogen, oxygen and sulfur isotope studies of seafloor hydrothermal system at the Desmos caldera, Manus back-arc basin, Papua New Guinea: An analogue of terrestrial acid hot crater-lake. Resour. Geol. 2006, 56, 183–190. [Google Scholar] [CrossRef]
  5. Voigt, M.; Pearce, C.R.; Baldermann, A.; Oelkers, E.H. Stable and radiogenic strontium isotope fractionation during hydrothermal seawater-basalt interaction. Geochim. Cosmochim. Acta 2018, 240, 131–151. [Google Scholar] [CrossRef]
  6. Kawahata, H.; Kusakabe, M.; Kikuchi, Y. Strontium, oxygen, and hydrogen isotope geochemistry of hydrothermally altered and weathered rocks in DSDP Hole 504B, Costa Rica Rift. Earth Planet. Sci. Lett. 1987, 85, 343–355. [Google Scholar] [CrossRef]
  7. Ravizza, G.; Blusztajn, J.; Von Damm, K.L.; Bray, A.M.; Bach, W.; Hart, S.R. Sr isotope variations in vent fluids from 9°46′–9°54′ N East Pacific Rise: Evidence of a non-zero-Mg fluid component. Geochim. Cosmochim. Acta 2001, 65, 729–739. [Google Scholar] [CrossRef]
  8. Dekov, V.M.; Cuadros, J.; Shanks, W.C.; Koski, R.A. Deposition of talc-kerolite-smectite-smectite at seafloor hydrothermal vent fields: Evidence from mineralogical, geochemical and oxygen isotope studies. Chem. Geol. 2008, 247, 171–194. [Google Scholar] [CrossRef]
  9. Wilckens, F.K.; Reeves, E.P.; Bach, W.; Seewald, J.S.; Kasemann, S.A. Application of B, Mg, Li and Sr isotopes in acid-sulfate vent fluids and volcanic rocks as tracers for fluid-rock interaction in back-arc hydrothermal systems. Geochem. Geophys. Geosyst. 2019, 20, 5849–5866. [Google Scholar] [CrossRef] [Green Version]
  10. Palmer, M.R. Controls over the chloride concentration of submarine hydrothermal vent fluids: Evidence from Sr/Ca and 87Sr/86Sr ratios. Earth Planet. Sci. Lett. 1992, 109, 37–46. [Google Scholar] [CrossRef]
  11. Webber, A.P.; Roberts, S.; Burgess, R.; Boyce, A.J. Fluid mixing and thermal regimes beneath the PACMANUS hydrothermal field, Papua New Guinea: Helium and oxygen isotope data. Earth Planet. Sci. Lett. 2011, 304, 93–102. [Google Scholar] [CrossRef]
  12. Jean-Baptise, P.; Dapoigny, A.; Stievenard, M.; Charlou, J.L.; Fouquet, Y.; Donval, J.P.; Auzende, J.M. Helium and oxygen isotope analyses of hydrothermal fluids from the East Pacific Rise between 17° S and 19° S. Geo-Mar. Lett. 1997, 17, 213–219. [Google Scholar] [CrossRef]
  13. Hodell, D.A.; Mead, G.A.; Mueller, P.A. Variation in the strontium isotopic composition of seawater (8 Ma to present): Implications for chemical weathering rates and dissolved fluxes to the oceans. Chem. Geol. Isot. Geosci. Sect. 1990, 80, 291–307. [Google Scholar] [CrossRef]
  14. Davis, A.C.; Bickle, M.J.; Teagle, D.A.H. Imbalance in the oceanic strontium budget. Earth Planet. Sci. Lett. 2003, 211, 173–187. [Google Scholar] [CrossRef]
  15. Palmer, M.R.; Edmond, J.M. The strontium isotope budget of the modern ocean. Earth Planet. Sci. Lett. 1989, 92, 11–26. [Google Scholar] [CrossRef]
  16. Butterfield, D.A.; Nelson, B.K.; Wheat, C.G.; Mottl, M.J.; Roe, K.K. Evidence for basaltic Sr in midocean ridge-flank hydrothermal systems and implications for the global oceanic Sr isotope balance. Geochim. Cosmochim. Acta 2001, 65, 4141–4153. [Google Scholar] [CrossRef]
  17. Coogan, L.A.; Dosso, S.E. Alteration of ocean crust provides a strong temperature dependent feedback on the geological carbon cycle and is a primary driver of the Sr isotopic composition of seawater. Earth Planet. Sci. Lett. 2015, 415, 38–46. [Google Scholar] [CrossRef] [Green Version]
  18. Antonelli, M.A.; Pester, N.J.; Brown, S.T.; DePaolo, D.J. Effect of paleoseawater composition on hydrothermal exchange in midocean ridges. Proc. Natl. Acad. Sci. USA 2017, 114, 12413–12418. [Google Scholar] [CrossRef] [Green Version]
  19. Horita, J.; Cole, D.R.; Weslowski, D.J. The activity composition relationship of oxygen and hydrogen isotopes in aqueous salt solutions: III. Vapor-liquid water equilibration of NaCl solutions to 350 °C. Geochim. Cosmochim. Acta 1995, 6, 1139–1152. [Google Scholar] [CrossRef]
  20. Campbell, A.C.; Palmer, M.R.; Klinkhammer, G.P.; Bowers, T.S.; Edmond, J.M.; Lawrence, J.R.; Casey, J.F.; Thompson, G.; Humphris, S.; Rona, P.; et al. Chemistry of hot springs on the Mid-Atlantic Ridge. Nature 1988, 335, 514–519. [Google Scholar] [CrossRef]
  21. East-Pacific-Rise-Study-Group. Crustal processes of the mid-ocean ridge. Science 1981, 213, 31–40. [Google Scholar] [CrossRef] [PubMed]
  22. Welhan, J.A.; Craig, H. Methane, hydrogen and helium in hydrothermal fluids at 21° N on the East Pacific Rise. In Hydrothermal Processes at Seafloor Spreading Centers; NATO conference series: IV, Marine Sciences; Rona, P.A., Bostrom, K., Laubier, L., Smith, K.L., Jr., Eds.; Plenum Press: New York, NY, USA, 1983; pp. 391–409. [Google Scholar]
  23. Teagle, D.A.H.; Alt, J.C.; Chiba, H.; Humphris, S.E.; Halliday, A.N. Strontium and oxygen isotopic constraints on fluid mixing, alteration and mineralization in the TAG hydrothermal deposit. Chem. Geol. 1998, 149, 1–24. [Google Scholar] [CrossRef]
  24. Bach, W.; Humphris, S.E. Relationship between the Sr and O isotope compositions of hydrothermal fluids and the spreading and magma-supply rates at oceanic spreading centers. Geology 1999, 27, 1067–1070. [Google Scholar] [CrossRef]
  25. Shanks, W.C., III; Bohlke, J.K.; Seal, R.R., II. Stable isotopes in Mid-Ocean Ridge Hydrothermal Systems: Interactions between fluids, minerals and organisms. In Physical, Chemical, Biological, and Geological Interactions within Hydrothermal Systems; Geophysical Monograph Series; Humphris, R.A., Zierenberg, R.A., Mullineaux, L.S., Thomson, R.E., Eds.; American Geophysical Union: Washington, DC, USA, 1995; Volume 91, pp. 194–221. [Google Scholar]
  26. Muehlenbachs, K.; Clayton, R.N. Oxygen isotope geochemistry of submarine greenstones. Can. J. Earth Sci. 1972, 9, 471–478. [Google Scholar] [CrossRef]
  27. Stakes, D.S.; O’Neil, J.R. Mineralogy and stable isotope geochemistry of hydrothermally altered oceanic rocks. Earth Planet. Sci. Lett. 1982, 57, 285–304. [Google Scholar] [CrossRef]
  28. Bowers, T.S.; Taylor, H.P., Jr. An Integrated chemical and stable isotope model of the origin of mid-ocean ridge hot spring systems. J. Geophys. Res. 1985, 90, 12583–12606. [Google Scholar] [CrossRef] [Green Version]
  29. Cole, D.R.; Mottl, M.J.; Ohmoto, H. Isotopic exchange in mineral-fluid systems: II. Oxygen and hydrogen isotopic investigation of the experimental basalt-seawater system. Geochim. Cosmochim. Acta 1987, 51, 1523–1538. [Google Scholar] [CrossRef]
  30. Bowers, T.S. Stable isotope signatures of water-rock interaction in mid-ocean ridge hydrothermal systems: Sulfur, oxygen and hydrogen. J. Geophys. Res. 1989, 94, 5775–5786. [Google Scholar] [CrossRef]
  31. Böhlke, J.K.; Shanks, W.C., III. Stable isotope study of hydrothermal vents at Escanaba Trough, NE Pacific: Observed and calculated effects of sediment-seawater interaction. In Geologic, Hydrothermal, and Biologic Studies at Escanaba Trough, Gorda Ridge, Offshore Northern California, Gorda Ridge; Morton, J.L., Zierenberg, R.A., Reiss, C.A., Eds.; U.S. Geological Survey Bull: Reston, VA, USA, 2022; Volume 1994, pp. 223–239. [Google Scholar]
  32. Fouquet, Y.; Von Stackelberg, U.; Charlou, J.L.; Donval, J.P.; Erzinger, J.; Foucher, J.P.; Herzig, P.; Mühe, R.; Soakai, S.; Wiedicke, M.; et al. Hydrothermal activity and metallogenesis in the Lau back-arc basin. Nature 1991, 349, 778–781. [Google Scholar] [CrossRef]
  33. James, R.H.; Elderfield, H.; Palmer, M.R. The chemistry of hydrothermal fluids from the Broken Spur site, 29°N Mid-Atlantic Ridge. Geochim. Cosmochim. Acta 1995, 59, 651–659. [Google Scholar] [CrossRef]
  34. Schmidt, K.; Garbe-Schönberg, D.; Koschinsky, A.; Strauss, H.; Jost, C.L.; Klevenz, V.; Königer, P. Fluid elemental and stable isotope composition of the Nibelungen hydrothermal field (8°18′ S, Mid-Atlantic Ridge): Constraints on fluid-rock interaction in heterogeneous lithosphere. Chem. Geol. 2011, 280, 1–18. [Google Scholar] [CrossRef]
  35. Mottl, M.J.; Seewald, J.S.; Wheat, C.G.; Tivey, M.K.; Michael, P.J.; Proskurowski, G.; McCollom, T.M.; Reeves, E.; Sharkey, J.; You, C.-F.; et al. Chemistry of hot springs along the Eastern Lau Spreading Center. Geochim. Cosmochim. Acta 2011, 75, 1013–1038. [Google Scholar] [CrossRef]
  36. Chen, Y.-G.; Wu, W.-S.; Chen, C.-H.; Liu, T.-K. A date for volcanic eruption inferred from a siltstone xenolith. Quat. Sci. Rev. 2001, 20, 869–873. [Google Scholar] [CrossRef]
  37. Chen, C.-H.; Lee, T.; Shieh, Y.-N.; Chen, C.-H.; Hsu, W.-Y. Magmatism at the onset of back-arc basin spreading in the Okinawa Trough. J. Volcanol. Geotherm. Res. 1995, 69, 313–322. [Google Scholar]
  38. Zeng, Z.; Li, X.; Chen, S.; Zhang, Y.; Chen, Z.; Chen, C.-T.A. Iron-Copper-Zinc Isotopic Compositions of Andesites from the Kueishantao Hydrothermal Field off Northeastern Taiwan. Sustainability 2022, 14, 359. [Google Scholar] [CrossRef]
  39. Zeng, Z.G.; Liu, C.H.; Chen, C.-T.A.; Yin, X.B.; Chen, D.G.; Wang, X.Y.; Wang, X.M.; Zhang, G.L. Origin of a native sulfur chimney in the Kueishantao hydrothermal field, offshore northeast Taiwan. Sci. China (Ser. D Earth Sci.) 2007, 50, 1746–1753. [Google Scholar] [CrossRef]
  40. Zeng, Z.G.; Chen, C.-T.A.; Yin, X.B.; Zhang, X.Y.; Wang, X.Y.; Zhang, G.L.; Wang, X.M.; Chen, D.G. Origin of native sulfur ball from the Kueishantao hydrothermal field offshore northeast Taiwan: Evidence from trace and rare earth element composition. J. Asian Earth Sci. 2011, 40, 661–671. [Google Scholar] [CrossRef]
  41. Zeng, Z.G.; Wang, X.Y.; Chen, C.-T.A.; Yin, X.B.; Chen, S.; Ma, Y.Q.; Xiao, Y.K. Boron isotope compositions of fluids and plumes from the Kueishantao hydrothermal field off northeastern Taiwan: Implications for fluid origin and hydrothermal processes. Mar. Chem. 2013, 157, 59–66. [Google Scholar] [CrossRef]
  42. Chen, C.T.A.; Zeng, Z.G.; Kuo, F.W.; Yang, T.F.; Wang, B.J.; Tu, Y.Y. Tide-influenced acidic hydrothermal system offshore NE Taiwan. Chem. Geol. 2005, 224, 69–81. [Google Scholar] [CrossRef]
  43. Kuo, F.W. Preliminary Investigation of Shallow Hydrothermal Vents on Kueishantao Islet of Northeastern Taiwan. Master’s Thesis, Institute of Marine Geology and Chemistry, National Sun Yat-Sen University, Kaohsiung, Taiwan, China, 2001; p. 81. (In Chinese). [Google Scholar]
  44. Chen, C.T.A.; Wang, B.J.; Huang, J.F.; Lou, J.Y.; Kuo, F.W.; Tu, Y.Y.; Tsai, H.S. Investigation into extremely acidic hydrothermal fluids off Kueishantao islet, Taiwan. Acta Oceanol. Sin. 2005, 24, 125–133. [Google Scholar]
  45. Chen, X.-G.; Lyu, S.-S.; Garbe-Schönberg, D.; Lebrato, M.; Li, X.; Zhang, H.-Y.; Zhang, P.-P.; Chen, C.-T.A.; Ye, Y. Heavy metals from Kueishantao shallow-sea hydrothermal vents, offshore northeast Taiwan. J. Mar. Syst. 2018, 180, 211–219. [Google Scholar] [CrossRef]
  46. Chen, X.G.; Zhang, H.Y.; Li, X.H.; Chen, C.-T.A.; Yang, T.F.; Ye, Y. The chemical and isotopic compositions of gas discharge from shallow-water hydrothermal vents at Kueishantao, offshore northeast Taiwan. Geochem. J. 2016, 50, 341–355. [Google Scholar] [CrossRef]
  47. Zeng, Z.; Wang, X.; Qi, H.; Zhu, B. Arsenic and Antimony in Hydrothermal Plumes from the Eastern Manus Basin, Papua New Guinea. Geofluids 2018, 2018, 6079586. [Google Scholar] [CrossRef] [Green Version]
  48. Chen, F.; Li, X.-H.; Wang, X.-L.; Li, Q.-L.; Siebel, W. Zircon age and Nd-Hf isotopic composition of the Yunnan Tethyan belt, southwestern China. Int. J. Earth Sci. 2007, 96, 1179–1194. [Google Scholar] [CrossRef]
  49. Chen, F.; Zhu, X.-Y.; Wang, W.; Wang, F.; Pham, T.-H.; Siebel, W. Single-grain detrital muscovite Rb-Sr isotopic composition as an indicator of provenance for the Carboniferous sedimentary rocks in northern Dabie, China. Geochem. J. 2009, 43, 257–273. [Google Scholar] [CrossRef] [Green Version]
  50. Von Damm, K.L.; Edmond, J.M.; Grant, B.; Measures, C.I.; Walden, B.; Weiss, R.F. Chemistry of submarine hydrothermal solutions at 21° N, East Pacific Rise. Geochim. Cosmochim. Acta 1985, 49, 2197–2220. [Google Scholar] [CrossRef]
  51. Spivack, A.J.; Edmond, J.M. Boron isotope exchange between seawater and the oceanic crust. Geochim. Cosmochim. Acta 1987, 51, 1033–1043. [Google Scholar] [CrossRef]
  52. Alt, J.C.; Honnorez, J.; Laverne, C.; Emmermann, R. Hydrothermal alteration of 1 km section through the upper oceanic crust, deep sea drilling project Hole 504B: Mineralogy, chemistry, and evolution of seawater-basalt interaction. J. Geophys. Res. 1986, 91, 10309–10335. [Google Scholar] [CrossRef]
  53. Patten, C.G.C.; Pitcairn, I.K.; Teagle, D.A.H.; Harris, M. Mobility of Au and related elements during the hydrothermal alteration of the oceanic crust: Implications for the sources of metals in VMS deposits. Miner. Depos. 2015, 51, 179–200. [Google Scholar] [CrossRef]
  54. Butterfield, D.A.; Massoth, G.J. Geochemistry of north Cleft segment vent fluids: Temporal changes in chlorinity and their possible relation to recent volcanism. J. Geophys. Res. 1994, 99, 4951–4968. [Google Scholar] [CrossRef]
  55. Von Damm, K.L. Seafloor hydrothermal activity: Black smoker chemistry and chimneys. Annu. Rev. Earth Planet. Sci. Lett. 1990, 18, 173–204. [Google Scholar] [CrossRef]
Figure 2. Changes in pH (a), Ca (b), total S (c), Sr (d), As (e), Sb (f), Cl (g) and Mn (h) concentrations from hydrothermal fluid to plume at the yellow and white spring sites.
Figure 2. Changes in pH (a), Ca (b), total S (c), Sr (d), As (e), Sb (f), Cl (g) and Mn (h) concentrations from hydrothermal fluid to plume at the yellow and white spring sites.
Jmse 10 00845 g002
Figure 3. Variation of 87Sr/86Sr (a), δD (b) and δ18O (c) values from hydrothermal fluid to plume at the yellow and white spring sites.
Figure 3. Variation of 87Sr/86Sr (a), δD (b) and δ18O (c) values from hydrothermal fluid to plume at the yellow and white spring sites.
Jmse 10 00845 g003
Figure 4. pH vs. As concentration (a), pH vs. Sb concentration (b) and pH vs. δD (c) values for fluids and plumes in the yellow and white springs.
Figure 4. pH vs. As concentration (a), pH vs. Sb concentration (b) and pH vs. δD (c) values for fluids and plumes in the yellow and white springs.
Jmse 10 00845 g004
Table 1. Location of the yellow and white springs in the Kueishantao hydrothermal field as well as the fluid temperature and flux [45].
Table 1. Location of the yellow and white springs in the Kueishantao hydrothermal field as well as the fluid temperature and flux [45].
Spring TypeLatitude (°N)Longitude (°E)Depth (m)Fluid Temperature (°C)Fluid Flux (m3/h)
Yellow spring24.8349121.961947.210835.1
White spring24.83412121.9619615.15119.3
Table 2. pH, Ca, total S, Sr, As, Sb, Cl and Mn concentrations and 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of fluid and plume samples collected from the Kueishantao hydrothermal field.
Table 2. pH, Ca, total S, Sr, As, Sb, Cl and Mn concentrations and 87Sr/86Sr, δDV-SMOW and δ18OV-SMOW values of fluid and plume samples collected from the Kueishantao hydrothermal field.
SampleDepth (m)pHCa (mg/L)Total S (mg/L)Sr (mg/L)As (µg/L)Sb (µg/L)Cl (mg/L)Mn (µg/L)87Sr/86Sr2σ(±)δDV-SMOW (‰)δ18OV-SMOW (‰)
Ambient
seawater
108.02372.68356.621.150.2019,3520.910.7091770.00000820.1
YSP, 0 m06.15381.38296.375.590.2920,81410.80.7091470.00000820.1
YSP, −2 m26.12385.08406.363.810.2320,4223.930.7091560.00000730.2
YSP, −5 m55.60382.28336.346.010.2820,8103.450.7091680.00000650.2
YSF, out, bottle7.22.81371.28105.8722.40.7718,0524.55--10.1
YSF, in, bottle>7.22.29378.98126.1246.41.2218,2543.58--10.1
YSF, out, tube7.2-345.08386.32-0.0317,38018.5----
YSF, in, tube>7.2-359.110696.430.29-18,32314.2----
WSP, 0 m06.14381.38326.355.650.2719,0686.240.7092020.00000600.1
WSP, −2 m26.12383.08336.354.080.2918,9253.480.7091860.00000500.1
WSP, −5 m55.91379.38236.306.460.3220,2054.340.7091710.00000620.1
WSP, −10 m105.51376.28236.2714.50.4820,45410.40.7091780.00000500.2
WSF, out, bottle15.15.11380.58266.4210.50.3616,5297.480.7091580.000004−40.2
WSF, in, bottle15.14.67369.49636.1215.80.3616,95123.3--−20.2
WSF, out, tube15.15.10383.28026.5627.60.4717,81114.40.7091540.00000710.2
WSF, in, tube15.14.67372.58056.2321.10.3917,30913.20.7091640.00000610.3
SampleRange of Sr (ppm) Average of Sr (ppm)
Kueishantao
andesite
175–264 206 (n = 41)
“-” no detect; YSP represents yellow spring plume; YSF represents yellow spring fluid; WSP represents white spring plume; WSF represents white spring fluid.
Table 3. Calculated water/rock ratios, Sr and O source contributions and Sr fluxes in the Kueishantao hydrothermal field.
Table 3. Calculated water/rock ratios, Sr and O source contributions and Sr fluxes in the Kueishantao hydrothermal field.
Sping FluidWater/Rock RatioStrontium Source Contributions (%)Strontium Flux (mole/yr)Oxygen Source Contributions (%)
87Sr/86Sr andesite = 0.7057787Sr/86Sr andesite = 0.70688Andesite 1Seawater 2Andesite 3Seawater 2Andesite 4Seawater 5Andesite 6Seawater 5
YSF, out, bottle------2.06 × 1041.4398.571.2898.72
YSF, in, bottle------2.15 × 1041.4398.571.2898.72
YSF, out, tube------2.22 × 1041.4398.571.2898.72
YSF, in, tube------2.26 × 1042.8697.142.5697.44
WSF, out, bottle554837300.5699.410.8399.171.24 × 1040.001000.00100
WSF, in, bottle------1.18 × 104----
WSF, out, tube457830760.6899.321.0099.001.26 × 104----
WSF, in, tube812454670.3899.620.5799.431.20 × 1040.001000.00100
“-” no data. Sr of spring fluid from andesite and seawater. Sr source contribution (X) can be calculated by using the simple two endmember mixing method: X = ((87Sr/86Srspring fluid − 87Sr/86Srseawater)/(87Sr/86Srandesite87Sr/86Srseawater) × 100); YSF represents yellow spring fluid; WSF represents white spring fluid; 1 represents 87Sr/86Srandesite = 0.70577 from [37]; 2 represents 87Sr/86Srseawater = 0.709177; 3 represents 87Sr/86Srandesite = 0.70688 from [37]. O source contribution (Y) can be calculated by using simple two endmember mixing method: Y = ((δ18Ospring fluid − δ18Oseawater)/(δ18O andesite − δ18O seawater) × 100); 4 represents δ18Oandesite = 7.1‰ from [40]; 5 represents δ18Oseawater = 0.1; 6 means δ18Oandesite = 7.9‰ from [37].
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.; Wang, X.; Yin, X.; Chen, S.; Qi, H.; Chen, C.-T.A. Strontium, Hydrogen and Oxygen Behavior in Vent Fluids and Plumes from the Kueishantao Hydrothermal Field Offshore Northeast Taiwan: Constrained by Fluid Processes. J. Mar. Sci. Eng. 2022, 10, 845. https://doi.org/10.3390/jmse10070845

AMA Style

Zeng Z, Wang X, Yin X, Chen S, Qi H, Chen C-TA. Strontium, Hydrogen and Oxygen Behavior in Vent Fluids and Plumes from the Kueishantao Hydrothermal Field Offshore Northeast Taiwan: Constrained by Fluid Processes. Journal of Marine Science and Engineering. 2022; 10(7):845. https://doi.org/10.3390/jmse10070845

Chicago/Turabian Style

Zeng, Zhigang, Xiaoyuan Wang, Xuebo Yin, Shuai Chen, Haiyan Qi, and Chen-Tung Arthur Chen. 2022. "Strontium, Hydrogen and Oxygen Behavior in Vent Fluids and Plumes from the Kueishantao Hydrothermal Field Offshore Northeast Taiwan: Constrained by Fluid Processes" Journal of Marine Science and Engineering 10, no. 7: 845. https://doi.org/10.3390/jmse10070845

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