Lithium isotopes track magmatic fluids in submarine hydrothermal systems | Scientific Reports

2021-11-12 09:59:56 By : Ms. Phoebe Wu

Thank you for visiting Nature. The browser version you are using has limited support for CSS. For the best experience, we recommend that you use a newer version of the browser (or turn off the compatibility mode in Internet Explorer). At the same time, to ensure continued support, we will display sites without styles and JavaScript.

Scientific Reports Volume 5, Article Number: 13812 (2015) Cite this article

The fluid inclusions and the lithium isotopic composition of gangue quartz from China's huge volcanic massive sulfide deposits provide strong evidence for the input of magmatic fluids into the Triassic submarine hydrothermal system. For fluid inclusions, the δ7Li results vary from 4.5‰ to 13.8‰, and for gangue quartz (9 gangue quartz samples containing primary fluid inclusions) from 6.7‰ to 21.0‰. These data confirm the temperature-dependent Li isotope fractionation between hydrothermal quartz and fluid (ie Δδ7Liquartz-fluid = –8.9382 × (1000/T) 22.22(R2 = 0.98; 175 °C–340 °C)), which indicates The fluid inclusion is in equilibrium with the quartz it carries, so the δ7Liquartz data can be used to determine the composition of the fluid. Therefore, we estimate that the ore-forming fluid has a δ7Li range from -0.7‰ to 18.4‰ at a temperature of 175–340 °C. This δ7Li range, together with the Li-O model, indicates that magmatic fluids play an important role in mineralization. This study shows that lithium isotopes can be effectively used to track magmatic fluids in submarine hydrothermal systems and have the potential to monitor fluid mixing and mineralization processes.

Submarine hydrothermal fluids affect the mass balance of the hydrosphere1 and form sulfide minerals on the modern seafloor2 and their ancient analogues preserved on land3. However, their source remains controversial4. A common belief is that they come from pure seawater circulating in hot rocks2,5. Another view is that there may be a mixture of magma fluids escaping from the depths of the magma4. This controversy reflects the fact that transitional isotope tracers (eg, δD, δ18O) rarely provide reliable characteristics of fluid origin. Lithium isotopes have been shown to be important geochemical tracers for fluid-related processes, especially aqueous fluids, volatiles and metals in magma-hydrothermal systems, because of the complexation behavior of lithium chlorides in aqueous fluids 6,7, Strong fluid mobility8 and large mass 9. Fractionation Li isotope fractionation can be used to track mineral precipitation and/or diffusion in hydrothermal systems. Quartz is the most common ore-forming fluid captured by gangue minerals and serves as inclusions in hydrothermal deposits. Its lithium isotope system may have the potential to limit fluid sources and quantify fluid flux from different sources.

In this study, we first reported the lithium isotopic composition of the Gacun Zn-Pb-Cu deposit, a huge volcanic massive sulfide deposit in the Yindun arc basin system 10, 11 in southwest China (Figure 1). Our data shows that magmatic fluids escaping from rhyolite melts play an important role in ore formation. In addition, we demonstrate that Li isotope measurements can be used to monitor fluid mixing related to mineralization and help locate particularly mineralized horizons.

The geological map of the Gacun deposit (modified according to References 10 and 12) shows the sample location and spatial distribution of the main ore bodies on the 4100 m exploration plane.

Due to the later deformation, the Triassic volcanic-sedimentary strata folded and steeply dipped eastward, providing an ideal cross-section showing the rhyolite pack and the occurrence of ore bodies from deep (west) to top (east). The illustration shows the structural framework of the righteous arc formed by the westward subduction of the Triassic Garze-Litang marine lithosphere. The Gacun deposit is located in the intra-arc rift zone within the Traissic arc and consists of three main ore bodies. The location of U-Pb dating samples and the boundary of the feeder area of ​​the hydrothermal system 10 are also shown. 53 Li analysis samples were collected along 8 exploration lines of ore bodies of different heights (4050 m, 4100 m, 4150 m) and displayed on the exploration plane of 4100 m. The positions of all samples are marked with solid circles (at 4100 m) and hollow circles (at 4050 m and 4160 m). Through binary mixture modeling of Li-O isotope data of 53 quartz samples, the minimum amount of seawater (Xseawater) in the ore-forming fluid of each sample was estimated (see Appendix II, Table S1 and Figure 3). All data in Table S1.

We evaluated the submarine hydrothermal system by studying the lithium isotopes on quartz and the fluid inclusions formed by the Triassic submarine hydrothermal fluid in the Gacun deposit (Figure 1). The deposit is located in a 500-meter thick volcanic pack of steeply dipping rhyolite exposed in Gacun (Figure 1). It includes three mineralized zones: (1) flaky massive sulfide zone (UMO), with associated smoke (barite, flint, jasper) covered by Triassic black phyllite, (2) discordant The low intermediate stratigraphic bound network zone (MSO) is located in the ~233 Ma rhyolite tuff unit and (3) the semi-layered lower stringer zone (LSO) part is located in the ~221 Ma rhyolite dome. The three mining areas are formed by hydrothermal fluids that pass through a subsea feeder area 10, 13 rooted in a rhyolite dome (Figure 1). The ~233 Ma rhyolite volcanic rock comes from a single, relatively shallow, differentiated felsic magma. The approximately 221 Ma rhyolite dome is the latest intrusive phase, acting as a "heat engine", driving fluid circulation through the overlying volcanic pack12.

In this study, nine rhyolite samples were collected for lithium isotope analysis (Table S1). The fresh rhyolite (n = 3) outside the Gacun deposit produced a narrow δ7Li range from 1.0‰ to 2.3‰, similar to continental crust (δ7Li = 1.2 ~ 1.7‰)15. The altered rhyolite from the feed zone (n = 6) shows a δ7Li value ( Table S1).

The 53 gangue quartz samples collected in the main ore body (Figure 1) are divided into two groups: pure quartz (n = 41) and quartz sericite (n = 12). The former occurs in all three mining areas; the latter is limited to LSO and MSO. The lithium concentration and isotopic composition of 53 quartz samples and 29 fluid samples extracted from these matrix quartz are listed in Table S1 and plotted in Figures 2 and S1. The Li concentration of pure quartz (n = 41) is 0.03–1.61 μg/g, and the δ7Li varies from 4.1‰ to 22.5‰. The extracted fluid (n = 17) produced a δ7Li value range from 3.7‰ to 15.0‰ (Figure 2a). Quartz sericite mixture has similar Li concentration (0.02–1.75 μg/g) but low δ7Li value (0.8‰ to 4.2‰), and the corresponding fluid (n = 12) has relatively high δ7Li, ranging from 1.3‰ to 10.2 ‰ (Figure 2a). Generally speaking, the quartz samples from each mining area have a variable δ7Li value, which increases laterally from the feeder area (Table S1). The extracted fluid showed that the average δ7Li value gradually increased from 5.4‰ in LSO to 9.8‰ in UMO (Table S1).

(a) The δ7Li of the bulk quartz and fluid inclusions varies with the measured homogenization temperature. (b) The relationship between the Li isotope fractionation factor (△quartz fluid) and the homogenization temperature (1000/T) of fluid inclusions in 9 pure quartz samples (including primary fluid inclusions). All data in Table S1.

All analyzed quartz samples (n = 53) measured O isotopic composition, and the δ18O values ​​produced by them were between 2.6‰ and 16.4‰ (Table S1). Based on the analysis of the homogenization temperature (Th: 117°C–368°C) of fluid inclusions in quartz (Table S1), the formula of Matsuhisa et al. was used. (1979) 16, the δ18O value of the fluid is estimated to vary from 0.9‰ to 8.7‰, showing a gradual decrease from rhyolite dome (av. 7.6‰) to LSO (av. 7.1‰) and MSO (av. 5.4‰), To UMO (av. 2.8‰).

The δ7Li values ​​of the fluid inclusions and the main quartz of 17 pure quartz samples were obtained, and the Δδ7Liquartz fluid values ​​obtained were 1.1-7.6‰ (Table S1). Extensive petrographic observations show that only 9 pure quartz samples contain primary fluid inclusions, which can be achieved through their arrangement along the quartz growth zone, regular form (smooth grains, columnar and polygonal) and variable size (usually 3- 10 μm) 10 to prove. The value of Δδ7Liquartz-fluid produced by these 9 samples varied from 4.5‰ to 13.8‰ (Table S2), and showed a negative correlation between Δδ7Liquartz-fluid and 1/T (Figure 2b), which can be described by the following linear equation: Δδ7Liquartz- fluid = −8.9382 × (1000/T) 22.22 (R2 = 0.98; 175 °C–340 °C). The Δδ7Liquartz fluid shows a strong temperature-dependent fractionation.

Recent experiments on different mineral-fluid systems have given conflicting results on the fractionation of lithium isotopes between fluids and minerals. Linton et al. (2005)17 studied the fractionation of Li isotopes in the quartz-muscovite-fluid system and found that the isotopes of the fluid are lighter than the minerals. Studies of natural samples (such as granite pegmatite) have reached similar conclusions18. In contrast, the experiment of the synthetic spodumene-hydrothermal fluid system confirmed the temperature-dependent fractionation, but it was found that the isotopes of the fluid were heavier than the coexisting spodumene. This result is consistent with some empirical studies, which show that the isotope of the fluid is 20,21,22 heavier than metamorphic basalt at low to moderate temperatures. Our results are consistent with the study of Lynton et al. (2005)17 and Teng et al. (2006) 18.

Diffusion-driven fractionation predicts that after the fluid inclusions are wrapped, there should be a difference of δ7Li between the quartz and the fluid. The measurement result (ie Δδ7Liquartz-fluid) should conform to the traditional stable isotope theory: heavier isotopes tend to enter the Millions of years of diffusion and transport enter the liquid phase. In addition, since this difference is usually uncertain, it lacks a good correlation with temperature. However, our data shows that fluid inclusions are much lighter than the quartz they host, and show a strong correlation between Δδ7Liquartz-fluid and 1/T (Figure 2b). Therefore, we believe that the result is mainly the equilibrium fractionation that occurs when quartz is formed.

The above empirical equation shows that for these 9 pure quartz samples, the fluid in equilibrium with the quartz is 2.1‰ to 7.6‰ lighter than their host quartz. For other pure quartz samples (n = 32) that contain secondary inclusions and fail the δ7Li fluid measurement, their δ7Liquartz data can indicate the variable composition of the fluid. Based on the measured Th (Table S1), using the above equation, we estimate that their δ7Lifluid varies between -0.7‰ and 18.4‰ (Table S1).

It is worth noting that the quartz sericite group (n = 12) produced different Δδ7Liquartz fluids, varying from -6.7‰ to -0.5‰ (Table S1). The'δ7Liquartz' of the quartz sericite group (n = 12) is a mixed value because it is mixed with a small amount of sericite. 'δ7Liquartz' tends to be lighter than pure quartz because the δ7Li of mica tends to be lighter23. However, during the extraction of fluid inclusions, we grind the quartz sericite sample to >200 mesh (<44 μm) and centrifuge 30 mL of percolate (distilled water and deionized water: 5 rounds of 6 mL), and finally use a 0.22 μm nylon filter Filter the pores (see appendix 1) to avoid the mixing of residual mineral powder. Therefore, the δ7Li (1.3~10.2‰) of the fluid extracted from the quartz-sericite sample can represent the δ7Li value of the Gacun ore-forming fluid.

Our above data show that in 11°–13° EPR (8.1‰ to 10.9‰)20, it is estimated that the δ7Li (-0.7‰ to 18.5‰) range of the Gacun ore-forming fluid is wider than that of the active vent fluid, 21 ° EPR (6.8‰ to 8.9‰)20, efforts in Judan de Fuca Ridge (7.2‰ to 8.9‰)21 and Guyamas (~ 10‰)22.

The source of these vent fluids is attributable to (1) gradually interacting sea water, passing down the hot rock and being heated by the magma chamber 20, 21, (2) a mixture of hot salt water and cold sea water derived from sea water 24 and (3) magma Fluid and sea water 21. Our Li isotope data can be used to test the above hypothesis. The reaction process model shows that as the interaction depth and fluid temperature increase, the δ7Li of the reaction seawater gradually decreases, and it is predicted that the reaction seawater at the lowest water/rock ratio has the lightest δ7Li value of 20. However, this model cannot explain the low δ7Li characteristics of Gacun fluid (<4‰), because experiment 25, modeling 24, and research on natural samples 26 show that seawater reacting at <450 °C rarely produces as low as <4 a large amount of light δ7Li ‰. In addition, the lightest δ7Lifluid values ​​are mostly concentrated in or near the feeder belt in Ga Village (Figure S1), where there is no organic-rich sediment infiltrated by the circulating fluid (Figure 1), and it comes from the rhyolite dome. Heat transfer may produce a high water/rock ratio of 27, which also opposes this model. The phase separation of heated seawater can form brine and vapor phases28, 29, but this process does not reduce the δ7Li value by 21. In Ga Village, the δ7Li value of the strongly altered dome rhyolite (1.2‰, GC4100-9-1, Table S1) is similar to that of fresh rhyolite (~ 1.2‰), requiring extremely low δ7Li (≤1.2 ‰) ) Thermal fluid or salt water interacts with subvolcanic rocks. The formation of this low δ7Li brine (fluid) requires an abnormally high temperature (>500°C)30 during the seawater/rock interaction, which is inconsistent with the microthermal results of fluid inclusions (Table S1). The last candidate is the injection of magma fluid escaping from felsic magma, which results in the release of floating steam-rich fluid, local heating, and supercritical phase separation 21. This situation is supported by the high temperature (> 350 °C) and high salinity (41.0 wt% NaCl) fluid inclusions carried by the phenocryst in the dome rhyolite, and it is also compatible with ore and rhyolite. The close spatial relationship is consistent13 and the relatively high gas content, high salinity and 18O enrichment of the Gacun hydrothermal system10.

So far, some experiments have been carried out to determine the Li isotope fractionation between magma fluid and melt30, 31. The theoretical model shows that the δ7Li of the residual melt and the dissolved fluid will not change significantly with the gradual dissolution of the fluid17. Based on the lithium isotope composition of the Gacun rhyolite (δ7Li: 1.0‰ to 2.3‰), we estimate that the δ7Li value of the initial fluid dissolved from the rhyolite melt varies from 1.0‰ to 3.0, which is close to the lowest measured value. The value of δ7Li LSO​​. This means that the changes in δ7Lifluid (-0.7 to 18.5‰) observed in Ga Village may be caused by other processes, such as interaction with marine sediments (δ7Li = -1.0 ∼ 24‰) 32,33 and mixing with seawater (Δ7Li = 31.5‰) 24,26. However, due to the lack of hydrothermal alteration of the wall (Triassic shale) and the absence of marine sediments in the Ga Village host inclusions (Figure 1), the first possibility was ruled out. All δ7Lifluid data can be explained by the binary mixing between magma fluid and seawater, which is supported by the Gacun Li-O isotope system (Figure 3).

The oxygen-lithium isotopic composition of the Gacun ore-forming fluid can be reproduced by mixing different amounts of seawater (δ18O = 0‰; δ7Li = 31.5‰) and magmatic fluid (δ18O = 8‰; δ7Li) = 1.5‰) with variable Limagmatic /Liseawater mass mixing ratio.

Dotted lines with open loops (10% intervals) show the binary mixing between magma fluid and seawater. Lim/Lic refers to the ratio of lithium concentration in magma fluid and seawater. See appendix 2 for modeling calculation.

Figure 3 shows that the δ18O-δ7Li data partially overlaps with the magmatic fluid, indicating the important role of magmatic fluid in mineralization. However, most of them show a systematic shift to seawater, and there is a general negative correlation between δ7Li and δ18O in each ore zone (Figure 3). UMO has higher δ7Lifluid (7.4‰ to 18.4‰) and lower δ18Ofluid (0.9‰ to 4.9‰), which means more seawater contribution, while LSO in the dome produces lower δ7Lifluid (0 to 6.5‰) ) And δ18Ofluid are relatively high (6.9‰~8.7‰), recording the footprint of magma fluid (Figure 3). The Li-O mixing model (see Appendix II) shows that all sample data can be reproduced by mixing different amounts of seawater with magma fluid. By using a variable Limagmatic/Liseawater mass mixing ratio, the best-fitting model curve connecting each mining area can be obtained (Figure 3). This two-component modeling requires UMO's Limagmatic/Liseawater mixing ratio (1.5:1 to 5:1), which is much lower than LSO (5:1 to 20:1). The highly variable ratio may be controlled by the chemical composition of the degassed magma fluid, which changes as the magma evolves. This is because the Li partition coefficient (Dfluid/melt) increases with temperature, the mole fraction of H2O, and the concentration of Cl in the fluid by 6,34. A previous study showed that quartz from LSO (av. 320 °C) contains CO2-rich multiphase inclusions, while quartz from UMO (av. 200 °C) contains Gacun10 liquid-gas two-phase aqueous inclusions body. This indicates that the ore-forming fluid in Gacun has changed from a mixture of CO2 and H2O to H2O dominating over time, which is consistent with the composition trend of continuous degassing of evolutionary magma4,35. Under similar temperature and pressure, it can be predicted that the relatively high CO2/H2O ratio in the LSO fluid may lead to a lower lithium concentration in the magma fluid, while the higher concentration of H2O and Cl in the UMO fluid may lead to a stronger lithium concentration . Li is distributed into the fluid (Figure 3). Assuming that the δ7Li of the magmatic fluid is 3‰ and the δ18O is 8‰36, the Triassic seawater is the same as modern seawater (31.5‰). It is estimated that about 80-90% and <50% of the magmatic fluids occupy the stratum as LSO and UMO (Table S1). Considering that the δ7Li (~26‰) of Mesozoic seawater may be 37,38 lighter than modern seawater, our estimation result is the largest proportion of magmatic fluid in the Gacun ore-forming hydrothermal system.

The spatial variation of the smallest percentage of seawater (Xseawater) in the ore-forming fluid system shows the outline of the Triassic submarine hydrothermal system in Gacun (Figures 1 and 4), where the convective circulation of the fluid through the ~233 Ma volcanic unit is caused by the location of ~ 221 Ma rhyolite dome drive. The initial fluid is dominated by magmatic water, which escapes from the rhyolite dome, first forms LSO, then mixes with the seawater circulating in the hot rock, and discharges upward through the sub-vertical replenishment zone (Figure 4). The mixture of seawater and magmatic fluid migrated laterally along the permeable layer in the rhyolite pack to form the MSO ore body (Figure 4). The intermittent input of seawater-based fluids into the submarine salt pool 10 in the submarine basin leads to the formation of UMO10,11.

Two-dimensional composition changes of the Triassic submarine hydrothermal system in the Gacun geological section[J].

This figure shows that the initial fluid, dominated by magma vapor escaping from the rhyolite dome, is discharged upwards through the sub-vertical feed zone, and drives the convective circulation of the fluid dominated by seawater through the volcanic unit. Based on all the point data of Xseawater for each sample in Figure 1, outline the quantitative scale (quantity) of seawater in each mineralized zone.

The δ7Li production of fluid inclusions in the Gacun deposit ranged from 4.5‰ to 13.8‰, while the δ7Li of gangue quartz (9 samples of gangue quartz containing primary fluid inclusions) ranged from 6.7‰ to 21.0‰. The estimated δ7Li values ​​of ore-forming fluids (-0.7‰ to 18.4‰) indicate the important role of magmatic fluids in ore-forming. Our data shows that the combination of δ7Li and δ18O data can monitor fluid mixing related to mineralization and help locate particularly mineralized horizons.

Atomic absorption spectrophotometer (AAS) and dual-focus multi-receiver inductively coupled plasma mass spectrometer (MC-ICP-MS) were used to measure the lithium concentration and lithium concentration of quartz and carrier fluid inclusions at the MLR Key Laboratory of Metal Generation and Mineral Evaluation. Isotope ratio, Institute of Mineral Resources, CAGS, China, follow Rudnick et al. (2004) 39 and Tian et al. (2012) 40. The detailed information of sample processing and analysis methods, as well as the results of microscopic temperature measurement and Li-O isotope analysis are given in Appendix I and Tables S1 and S2.

Fluid inclusions are extracted by crushing and leaching methods at the MLR Key Laboratory of Mineralization and Mineral Evaluation of CAGS Institute of Mineral Resources. Before extracting fluid inclusions, put about 4 g of finely selected quartz particles (60 ~ 80 mesh, 178 ~ 250 μm) on a hot plate (T <120 °C) with chloronitric acid (~3:1 of HCl-HNO3) The mixture) was heated for 3 hours. Then wash the quartz particles with distilled deionized water until the conductivity of the leachate is consistent with that of deionized water (>18.2 MΩ cm resistivity). Use an agate mortar in the ultra-clean cabinet to dry and grind clean quartz particles into fine powder (>200 mesh, <44μm). The leachate is extracted with fine powder in 30ml deionized water (6ml each time, 5 times), centrifuged, and then filtered with a 0.22μm nylon filter (to remove silica powder). Based on the traditional structural analysis of fluid inclusion populations and homogeneous temperature, primary inclusions of magma origin dominate (>90%) of the quartz samples selected for grinding and leaching analysis. The fluid inclusion extract is dried and re-dissolved in 4 M HCl to prepare for chromatographic separation.

Dissolve the dried quartz powder in the concentrated HNO3-HF mixture in the Savillex screw-cap beaker (about 1g sample in 0.5 mL HNO3 5 mL HF) on a hot plate (T <120 °C) overnight, and then replenish the dry The residue was overnight with concentrated HNO3 and dried again, then extracted in concentrated HCl until the solution was clear. The solution was then dried and re-dissolved in 4M HCl in preparation for chromatographic separation. Tian et al. describe a purification procedure using three columns. (2012) 40. Lithium isotope composition was measured at the Institute of Mineral Resources using the standard bracket method and Thermo Finnigan Neptune MC-ICP-MS instrument. The δ7Li value of each unknown is calculated relative to the average of the two bracket IRMM-016 runs. The results of IRMM-016 (δ7Li = 0.2 ± 0.4‰, 2σ, n = 6) are consistent with previously published data (for example, Rudnick et al., 200439; Halama et al., 200741, 200842). In the course of this research, two international rock standards were analyzed to evaluate the accuracy of the measurement. The δ7Li value produced by the basalt BHVO-2 standard is 4.7‰±1.0‰ (2σ, n = 53; Zhao et al., 201543) and The δ7Li value produced by the andesite AGV-2 standard is 6.1 ± 0.4‰ (2σ, n = 14; Zhao et al., 201543). The external accuracy of Li isotope analysis is based on the 2σ value of the pure lithium standard solution and rock solution repeated in three years, which is less than or equal to ±1.0‰ (Zhao et al., 201543). The accuracy of this method was established by Teng et al. (2006a)18 Based on the isotope dilution method is ±5%, and the accuracy is less than ±10% (Teng et al., 2006b)44.

The O isotope composition of quartz is analyzed in the CAGS Isotope Laboratory (Beijing). Separate individual quartz particles for oxygen isotope measurement. The analysis procedure of O isotopes follows the method of Clayton and Mayeda (1963) 45 and Hou et al. (2014) 46, respectively. At ~550 °C, quartz reacts with BrF5 to form O2, and SiF4 and BrF5 are separated from O2 by liquid nitrogen. At 700 °C, O2 reacts with krypton gas and converts to CO2, which is collected in a sample tube. A Thermo Finnigan MAT 253 mass spectrometer was used to measure the oxygen isotope ratio. The result is expressed in δ, with V-SMOW as the standard, as follows: δ18OV-SMOW = [(18O/16O)sample/(18O/18O)V-SMOW -1] × 1000. The analytical accuracy of δ18O measurement is 0.2 ‰ (1SD), based on repeated measurements of standard samples46.

How to cite this article: Yang, D. et al. Lithium isotopes track magmatic fluids in submarine hydrothermal systems. science. Rep. 5, 13812; doi: 10.1038/srep13812 (2015).

Damn Feng, KL, etc. 21°Hydrothermal chemistry of the seafloor in the East Pacific Uplift. Geochim Cosmochim Acta 49, 2197-2220 (1985).

Rhone, PA, etc. Main activities and remaining submarine hydrothermal mineralization: TAG hydrothermal field, mid-Atlantic ridge 26°N, 45°W. Economic Geology 88, 1989–2017 (1993).

Franklin, JM, Gibson, HL, Jonasson, IR & Galley, AG Massive sulfide deposits of volcanic origin. Economic Geology 100, 523–560 (2005).

Yang, K. & Scott, SD Degass volatiles and ore metal magma into a hydrothermal system in the modern seafloor of the Manus Backarc Basin in the eastern western Pacific Ocean. Economic Geology 97, 1079–1100 (2002).

Lydon, JW Deposit Model #14. Massive sulfide deposits of volcanic origin. Part 2: Genetic model. Canadian Earth Science 15, 43–65 (1988).

Webster, JD, Holloway, JR & Hervig, RL distribute philophilic trace elements between H2O and H2O CO2 fluids and topaz rhyolite melts. Economic Geology 84, 116–134 (1989).

Candela, PA & Piccoli, PM simulates the distribution of ore metals from the melt to the steam and steam/brine mixture. In Thompson, JFH, Ed., Canadian Mineralogy Association, Quebec. Granite, Fluids, and Mineral Deposits 23, 101–128 (1995).

You, CF, Castillo, PR & Gieskes, JM. The behavior of trace elements in hydrothermal experiment: the influence on the shallow fluid process in the subduction zone. Earth Planet Sci Lett 140, 41–52 (1996).

CAS ADS article Google Scholar 

Tomascak, the development of the understanding and application of PB lithium isotopes in earth and planetary sciences. In: Johnson C., Beard B. & Albarede F., eds. Geochemistry of non-traditional isotope systems. America. The Mineral Society of Amer Geochem Soc 153–195 (2004).

Hou, ZQ, etc. Genesis of the Gacun Volcanic Massive Sulfide Deposit in Sichuan, China: Evidence from Fluid Inclusions and Oxygen Isotope. Economic Geology 96, 1491-1512 (2001).

Hou, ZQ, etc. The Ga Village VHMS deposit in Southwest China: From field observations to metallogenic models[J]. Mineral Deposits 20, 44–56 (2001) (Chinese, English abstract).

Yang, D. The origin and evolution of ore-forming fluids of volcanic massive sulfide (VMS) deposits: New constraints of lithium isotopes on the genesis of the Gacun deposit in Sichuan. Doctoral Dissertation of China University of Geosciences (Beijing) 47-57 (2014).

Xu, MJ etc. Silver-rich polymetallic deposit in Ga Village, Sichuan Province, China. Chengdu University of Science and Technology Press 1-164 (Chinese and English abstract) (1993).

Hou, ZQ, etc. Etonian arc collisional orogeny in southwest China: Evidence from Triassic granites. Acta Geology 75, 484–497 (2001).

Lithium isotope systematics of Teng, FZ, Rudnick, RL, McDonough, WF and Wu, FY A-type granites and their basic enclaves: further restrictions on the lithium isotope composition of continental crust. Chemistry Geol 262, 370–379 (2009).

CAS ADS article Google Scholar 

Matsuhisa, Y. etc. Fractionation of oxygen isotopes in a quartz-albite-anorthite-water system. Geochimica et Cosmochimica Acta 43, 1131–1140 (1979).

CAS ADS article Google Scholar 

Lynton, SJ, Walker, RJ & Candela, PA Lithium isotopes in the Qz-Ms-fluid system: an experimental study. Geochemistry. Cosmic chemistry. Journal 69, 3337–3347 (2005).

CAS ADS article Google Scholar 

Teng, FZ etc. Lithium isotope systematics of granite and pegmatites in the Black Hills, South Dakota. Am Mineral 91, 1488–1498 (2006a).

CAS ADS article Google Scholar 

Wunder, B. etc. Temperature-dependent isotope fractionation of lithium between clinopyroxene and high-pressure aqueous fluid. Contributions to mineralogy and petrology 151, 112–120 (2006).

CAS ADS article Google Scholar 

Chan, LH, Edmond, JM & Thompson, GA Lithium isotopes of hot springs and metabasalts in the hydrothermal system of the mid-ocean ridge. Chinese Journal of Geophysics 98, 9653-9659 (1993).

CAS ADS article Google Scholar 

Foustoukos, DI, James, RH, Berndt, ME & Seyfried, WE Jr. Lithium isotopic system of hydrothermal vent fluids in the Main Endeavor oil field in the northern part of the Juan de Fuca ridge. Chem Geol 212, 17–26 (2004).

CAS ADS article Google Scholar 

Chen, LH et al. Lithium isotopic geochemistry of sediments and hydrothermal fluids in the Guaymas Basin, Gulf of California. Geochim Cosmochim Acta 58, 4443–4453 (1994).

CAS ADS article Google Scholar 

Wunder, B. etc. Fractionation of lithium isotopes between lithium-containing staurolite, lepidolite and aqueous fluid: an experimental study. Chem Geol 238, 277–290 (2007).

CAS ADS article Google Scholar 

Millot, R., Scaillet, B. and Sanjuan, B. Lithium isotopes in island arc geothermal systems: Guadeloupe, Martinique (French West Indies) and experimental methods. Geochim Cosmochim Acta 74, 1852–1871 (2009).

Seyfried, Jr., WE, Chen, X. & Chan, LH Trace element mobility and lithium isotope exchange during hydrothermal alteration of submarine weathered basalt: An experimental study at 350C and 500 bar. Geochim Cosmochim Acta 62, 949–960 (1998).

CAS ADS article Google Scholar 

Chan, LH and Edmond, JM Changes in the composition of lithium isotopes in the marine environment: a preliminary report. Geochim Cosmochim Acta 52, 1711-1717 (1988).

CAS ADS article Google Scholar 

Cann, JR, Strens, MR & Rice, AA A simple magma-driven heat balance model of volcanic origin and massive sulfide formation. Earth Planet Sci Lett 76, 122–136 (1986).

Butterfield, DA etc. The gradient of the composition of the hydrothermal fluid from the endeavor vent oil field: phase separation and brine loss. Chinese Journal of Geophysics 99, 9561-9583 (1994).

Little Seyfried et al. The chemistry of hydrothermal vent fluids from the Main Endeavor oil field in the northern part of Juan de Fuca Ridge: Geochemical control after the June 1999 earthquake. Chinese Journal of Geophysics 108 (B9), 2429 (2003).

Chacko, T., Cole, DR & Horita, J. Equilibrium oxygen, hydrogen, and carbon isotopic fractionation factors applicable to geological systems. Edited in JW Valley and DR Cole, stable isotope geochemistry. Review of the American Society of Mineralogy and Geochemical Mineralogy, Chantilly, Virginia. 43, 1-82 (2001).

CAS ADS article Google Scholar 

Marshall, human resources, etc. Lithium isotopic composition of orogenic eclogites and deep subducted plates. Earth Planet Sci Lett 262, 563–580 (2007).

CAS ADS article Google Scholar 

Chen, LH et al. Lithium isotopic composition of volcanic arc lava in Central America: The influence of slab-derived fluids on the modification of the subarc mantle. Chem Geol 160, 255–280 (1999).

CAS ADS article Google Scholar 

Hoefs, J. & Sywall, M. Lithium isotopic composition and global lithium isotopic balance of Quaternary and Tertiary biocarbonates. Geochim Cosmochim Acta 61, 2679-2690 (1997).

CAS ADS article Google Scholar 

London, D., Hervig, RL & Morgan, GB Melt vapor solubility and element distribution in a peraluminous granite-pegmatite system-experimental results using Macusani glass at 200 MPa. Contrib Mineral Petrol 99, 360–373 (1988).

CAS ADS article Google Scholar 

Yang, K. & Scott, SD The possible contribution of metal-rich magma fluid to the submarine hydrothermal system. Nature 383, 410–413 (1996).

Taylor, HP Jr. & Sheppard, SMF Igneous Rock: I. Isotope Fractionation and Isotope System Process. Rev. Mineralogy 16, 227–272 (1986).

Misra, S., Philip N. & Froelich, PN Seawater: Changes in silicate weathering and upwinding. Science 335, 818–822 (2012).

CAS ADS article Google Scholar 

Pogge von Strandmann, PAE, Jenkyns, HC & Woodfine, RG Lithium isotopic evidence for enhanced weathering during marine hypoxia event 2. Natural Earth Science 10.1038/NGEO1875 (2013).

Rudnick, RL, Tomascak, PB, Njoa, HB & Gardnerb, R. Extreme lithium isotope fractionation during continental weathering in the saprolite of South Carolina. Chem Geol 212, 45–57 (2004).

CAS ADS article Google Scholar 

Tian, ​​SH, Hou, ZQ and Su use MC-ICPMS to separate and accurately measure the lithium isotopes in the three standard materials. Acta Geology 86, 1297–1305 (2012).

Halama, R. etc. Lithium isotopic composition of Oldoinyo Lengai: the nature of the mantle origin and the lack of isotopic fractionation during the diagenesis of carbonate rocks. Earth Planet Sci Lett 254, 77–89 (2007).

CAS ADS article Google Scholar 

Halama, R. etc. Use carbonate rocks to track the evolution of lithium isotopes in the mantle. Earth Planet Sci Lett 265, 726–742 (2008).

CAS ADS article Google Scholar 

Zhao, Y. et al. The study of using MC-ICP-MS to measure the lithium isotope composition of common standard reference materials. Rock miner analysis. 34, 28-39 (Chinese, English abstract) (2015).

Teng, FZ etc. Diffusion in the Xishan pegmatite surrounding rock drives the fractionation of lithium isotopes. Earth Planet Sci Lett 243, 701–710 (2006b).

CAS ADS article Google Scholar 

Clayton, RN & Mayeda, TK uses bromine pentafluoride to extract oxygen from oxides and silicates for isotopic analysis. Geochim Cosmochim Acta 27, 43–45 (1963).

CAS ADS article Google Scholar 

Hou, ZQ, Liu, Y., Yang, ZM, Tian, ​​SH & Xie, YL Recycle marine sediments to form giant REE deposits related to carbonate rocks. Scientific Reports 5, 10231, 10.1038/srep10231 (2015).

CAS ADS article PubMed PubMed Central Google Scholar 

This work was funded by the IGCP/SIDA-600 project and NSFC (41103005, 41573038). Thanks to Chen Weishi for his help in measuring fluid inclusions and Wan Defang for his help in oxygen isotope analysis.

Institute of Mineral Resources, Chinese Academy of Geological Sciences, Beijing, 100037

Danyang, Zhao Yue, Hou Kejun and Tian Shihong

Institute of Geology, Chinese Academy of Geological Sciences, Beijing, 100037

Hou Zengqian, Yang Zhiming, Fu Qiang

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

ZQH designed and initiated the research and interpretation of the data and wrote the draft. DY contributed to field investigations and Li isotope analysis and data interpretation. YZ, KH, and ST contributed to isotope measurement, while ZY and QF contributed to field work and micro-calorimetry.

The author declares that there are no competing economic interests.

This work is licensed under the Creative Commons Attribution 4.0 International License Agreement. The images or other third-party materials in this article are included in the Creative Commons license of the article, unless otherwise stated in the credit line; if the material is not included under the Creative Commons license, the user will need to obtain permission from the license holder to copy The material. To view a copy of this license, please visit http://creativecommons.org/licenses/by/4.0/

Yang, D., Hou, Z., Zhao, Y. etc. Lithium isotopes track magmatic fluids in submarine hydrothermal systems. Scientific Report 5, 13812 (2015). https://doi.org/10.1038/srep13812

DOI: https://doi.org/10.1038/srep13812

Anyone you share the following link with can read this content:

Sorry, there is currently no shareable link in this article.

Provided by Springer Nature SharedIt content sharing program

By submitting a comment, you agree to abide by our terms and community guidelines. If you find content that is abusive or does not comply with our terms or guidelines, please mark it as inappropriate.

Scientific Report (Sci Rep) ISSN 2045-2322 (online)