Next Article in Journal
Assessing Risks from Groundwater Exploitation and Utilization: Case Study of the Shanghai Megacity, China
Previous Article in Journal
Trade-Offs between Human and Environment: Challenges for Regional Water Management under Changing Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geochemical and Isotopic Compositions and Geothermometry of Thermal Waters in the Magumsan Area, South Korea

1
Department of Construction and Disaster Prevention Engineering, Daejeon University, Daejeon 34520, Korea
2
Groundwater Research Center, Korea Institute of Geoscience and Mineral Resources, Daejeon 34132, Korea
3
EGI Consulting Co., Incheon 22698, Korea
4
Division of Polar Earth System Science, Korea Polar Research Institute, Incheon 21990, Korea
5
Department of Science Education, Ewha Womans University, Seoul 100-744, Korea
*
Authors to whom correspondence should be addressed.
Water 2019, 11(9), 1774; https://doi.org/10.3390/w11091774
Submission received: 16 July 2019 / Revised: 16 August 2019 / Accepted: 19 August 2019 / Published: 26 August 2019
(This article belongs to the Section Hydrology)

Abstract

:
The Magumsan thermal waters of the southeastern Korean Peninsula are pumped out of six deep wells (average depth, 300 m) at temperatures of 30.8–49 °C. The thermal waters are chemically classified into two groups: NaHCO3 type (<31 °C) and NaCl (HCO3, SO4) type (>40 °C), both of which have chemical compositions that are distinct from local groundwater (Ca–HCO3 type). δ18O and δD values suggest that the thermal waters originate from meteoric water and they are isotopically fractionated by silicate hydration or H2S exchange. δ34S values (+7.0 to +15%) of dissolved sulfate in the thermal waters reflect enrichment in 34S through kinetically controlled oxidation of magmatic pyrite in the thermal aquifer and mixing with paleo-seawater. On the 3He/4He vs. 4He/20Ne diagram, the thermal waters plot along a single air mixing line of dominant crustal He, which indicates that the heat source for the thermal waters is non-volcanogenic thermal energy that is generated from the decay of radioactive elements in crustal rocks. Chalcedony geothermometry and thermodynamic equilibrium calculations using the PHREEQC program indicate a reservoir temperature for the immature thermal waters of 54–86 °C and 55–83 °C, respectively.

1. Introduction

Hot-spring thermal water has been developed at about 400 sites in South Korea for spa and medical purposes, with −68% of the springs presenting temperatures of 25–30 °C.
Thermal water in the Magumsan area was first developed as a commercial hot spa in 1926 from a deep well in andesitic rocks of the Cretaceous Kyeongsang basin. Nowadays, the thermal waters in this area come from six deep wells (average depth −300 m), which are distributed in an area of 200 m diameter. The Magumsan thermal water is reported to have a chemical composition, including Na+, Cl, and SO42− as the major ions, with a maximum temperature of 55 °C [1]. Although Kim and Nakai analyzed the δ34S of dissolved sulfate in one sample of the Magumsan thermal waters [2], they suggested that sulfate S originated from mixed sources of seawater and magmatic pyrite sulfate. Park [3] explained that the chemical composition of the Magumsan thermal water with high Cl has been controlled by the flushing effect of saline thermal water by fresh water, owing to the seawater regression.
However, until the present day, helium and sulfur isotopic, and geothermometric methods have not been applied in elucidating the sulfur origin, heat source, geothermal reservoir temperature, or mixing ratio of thermal and cold water.
Geothermal systems in non-volcanic stable continental regions are unlikely to have magmatic heat sources. Rather, their heat source is mostly of crustal origin and is associated with the decay of radioactive elements (e.g., 238U, 232Th, 40K) [4]. The circulation depth of meteoric water and the presence of cap rock preventing heat loss from the system controls maximum reservoir temperatures of such geothermal systems. Another heat source is that developed along interfaces between lithospheric sectors, where friction and viscous shearing forces along boundaries are converted into heat [5].
Helium isotopic compositions of thermal water can provide critical constraints for heat sources in geothermal systems, owing to (1) the distinctive He isotopic ratios in mantle and crustal reservoirs, (2) contemporaneous release of heat and He from a magmatic mass or from decay of radioactive elements in the crust, and (3) the upward transport of mantle and crustal He accompanying transport of heat within the crust [4]. Jeong et al. [6,7,8] and Park et al. [9] have conducted helium isotopic applications for thermal water in South Korea to reveal heat sources by the determination of mixing ratios among three end-members (i.e., atmosphere, crust, and mantle) via 3He/4He–4He/20Ne ratio plots. They suggested that the heat source of thermal alkaline waters in South Korea is mainly constrained by an energy source that is produced from decay of radioactive elements in granitic rocks, which is based on dominant contribution of crustal helium (4He) with different ratios of 3He contribution from the deep source, such as magma or mantle according to thermal sites.
Cation and silica geothermometers, the Na–K–Mg ternary diagram, the silica–enthalpy model, and the thermodynamic equilibrium method have widely been used as tools in estimating the reservoir temperatures of geothermal systems, based on chemical analyses of the thermal waters [10,11,12,13]. After a cross validation of the results from each methods in this research, suitable methods were adapted to estimate the reservoir temperature of Magumsan thermal waters, because of the different range of their applicable temperature and given chemical data condition.
The aims of the present study were to determine the chemical composition, recharge origin, and sulfur origin (as sulfate) of the Magumsan thermal waters, in order to quantitatively estimate the ratios of He of different origins in thermal waters, including crust, mantle, and atmosphere, and to identify the major subsurface heat source of the geothermal reservoir. The geothermal reservoir temperature and mixing ratio of thermal water and shallow cold water were estimated while using geothermometers, the silica–enthalpy model, and the thermodynamic equilibrium method, which have their intrinsic applicable temperature range.

2. Study Area

The Korean Peninsula is situated on the southeastern margin of the Eurasian Plate, in a transitional setting between island-arc and back-arc terranes that formed through the subduction of the Philippine Sea Plate beneath the Eurasian Plate (Figure 1a). The geology comprises five main geological provinces: the Gyeonggi and Youngnam massifs, the Ogcheon Belt, Jurassic–Cretaceous granitoids, and the Kyungsang basin (Figure 1b). Precambrian basement rocks, which are exposed in the Gyeonggi and Youngnam massifs, comprise 2.7–1.1 Ga high-grade gneisses and schists. These massifs are separated by the Ogcheon Belt, which is a northeast-trending fold-and-thrust belt that comprises low- to medium-grade metasedimentary rocks [14]. Most of the thermal waters in South Korea are related to the fault zone in the Jurassic and Cretaceous granitoids, whereas thermal water in the Gyeongsang basin, which mainly comprises andesitic and sedimentary rocks, has been locally developed in areas with much higher than average surface heat flow of 66.6 mW m−2 [15]. The Magumsan area, in the southeastern Korean Peninsula, is surrounded to the east by Mt. Cheonma (365 m) and to the west by Mt. Magum (276 m). The average annual temperatures recorded at the nearest meteorological observatory over the last decade are 14.0–15.7 °C, with an annual average precipitation of 1021–1866 mm (overall average 1531 mm year−1). About half of the annual precipitation occurs during the rainy season from July–September.
The Magumsan thermal water arises from andesitic rocks in the Cretaceous Gyeongsang basin, which is a non-marine environment. The geology of the Magumsan area mainly comprises Chusan andesite, which is intruded by granodiorite and biotite granite in the northwest of the study area (Figure 1c). The Chusan andesitic rocks include andesite, trachyte–andesite, and brecciated andesite that occur variously as lava flows, intruded rock, and basement rock. The andesite is highly altered and silicified by hydrothermal reactions with intruding granitic rocks. A drill core in the thermal area has revealed silicified chert as a secondary mineral that results from hydrothermal alteration in the host andesite and it occurs as a thick layer at 170–250 m depth (Figure 2). Calcite veins were likely to have been intercalated before or after the epithermal stage. Pyrite occurs as a secondary mineral in the core and it is more abundant in the deeper parts of the core. Pb and Zn were also found in the core at 125 m depth. An epithermal ore deposit bearing Cu, Pb, and Zn as major ores is located 700 m northwest of the thermal area [1].
Granodiorite in the study area predominantly consists of the oligoclase–labradorite, amphibole, quartz, and biotite. The Chindong Formation, which mainly comprises dark-gray shale and mud rock, is distributed in the northeastern part of the Magumsan thermal area and it is locally intercalated with greenish-gray and dark-brown shales altered to chert and hornfels. The shale and mud rock are composed principally of quartz, feldspar, and fine biotite. Some feldspars are altered to chlorite as a result of hydrothermal alteration by intruding granite [1]. Faults in the study area trend northwest and they dip steeply across the thermal area, and they likely play a role as circulating conduits for thermal water [1,16].

3. Analytical Methods

3.1. Sample Collection and in Situ Measurements

Magumsan thermal waters are pumped out of wells of 300 m average depth. Six thermal water (>30 °C) and five groundwater samples were collected in the Magumsan area from the sites that are shown in Figure 3.
Measurements of pH, oxidation–reduction potential (ORP), dissolved-oxygen content (DO), electrical conductivity (EC), and temperature of all water samples were made in situ while using portable instruments (Orion three star and five star, Thermo Scientific, Beverly, MA, USA) 10–20 min after purging to achieve stable values. Alkalinity, being expressed as HCO3 concentration, was quantified on-site by titration with 0.05 M HCl. All of the water samples were filtered through a 0.45 μm membrane filter and then stored in 60 mL polythene bottles. Samples for cation analysis were acidified to pH <2 with ultrapure HNO3.

3.2. Chemical and Stable-Isotope Analyses

The concentrations of major cations and minor elements were determined by atomic absorption spectrometry (AAS; Unicam model 989, Unicam Sstemas Analiticos, Lda, Lisbon, Portugal), inductively coupled plasma–atomic emission spectrometry (ICP–AES; Shimadzu model ICPS-1000 III, Kyoto, Japan), and inductively coupled plasma–mass spectrometry (ICP–MS; Fison model PQ III, Thermo Scientific, Massachusetts, MA, USA) at the Korea Basic Science Institute (KBSI) of Daejeon, South Korea. The anion concentrations were determined by ion chromatography (Dionex DX-120 resin, Dionex Corporation, California, CA, USA). The reliability of chemical analysis was estimated by calculating the charge imbalance between major cations and anions, with results of +4.09% to −5.74% for all of the samples.
δ18O, δ2H, and δ34S values of water samples were determined by stable isotope ratio–mass spectrometry (SIR–MS; Isotoprime model, GV Instruments, Manchester, UK). All of the isotopic analyses were carried out at the KBSI. O and H isotopic ratios are expressed relative to Vienna Standard Mean Ocean Water (V-SMOW), with analytical precisions that were within ±0.1% and ±1.0%, respectively. The determination of sulfate δ34S values involved scavenging by precipitation of BaSO4 after the addition of BaCl2 powder to the thermal waters. Sulfur isotopic ratios are expressed relative to Canon Diablo Troilite (CDT), with analytical precisions within ±0.2%.

3.3. Noble-Gas Analysis

The water samples for noble-gas analysis were collected in 50 cm3 vacuum-tight glass bottles with high-vacuum stopcocks on both sides. Analyses were performed at the University of Tokyo, Japan. Noble gases that were dissolved in the water samples were extracted by an all-metal Toepler pump system, which enabled analyses to be performed under ultra-low blank conditions. The extracted gases were compressed into small volumes that were connected to the noble-gas purification line of the MS system. The noble gases were purified by two Ti–Zr getters and then separated into three fractions, He, Ne, and Ar, by charcoal traps and a cryogenically cooled sintered-stainless-steel trap. He, Ne, and Ar isotopic ratios and absolute abundances were determined while using a modified noble-gas mass spectrometer (VG5400, MS-III, Micromass Communication, North Carolina, NC, USA) at the Laboratory for Earthquake Chemistry, University of Tokyo, Japan. The sensitivities and mass discrimination correction factors were determined by analysis of known amounts of atmosphere using the same procedure as that applied to samples (details are described by [17]). Here, only He and Ne isotopic ratios were considered due to the atmospheric origin of Ar.

4. Results and Discussion

4.1. Geochemical Compositions

Table 1 provides the in situ measurement data and chemical compositions of thermal waters and groundwaters analyzed in this study.
The temperatures of thermal waters and groundwaters were 30.8–49.8 °C and 17.9–23.5 °C, respectively. pH values of thermal waters were 7.42–7.80, which is slightly higher than those of the groundwaters (5.86–7.35). ORP values of thermal waters indicate reducing conditions with values of −71 to −108 mV (cf. groundwater: +150 to +402 mV). Average EC values of thermal waters were higher than those of groundwaters.
Major-ion contents (meq L−1) of thermal waters (MH) and groundwaters (MG) are compared in a Schöeller diagram (Figure 4). Na+, K+, SO42−, and Cl generally have higher concentration ranges in thermal waters than in groundwater, whereas Mg and Mn contents display more striking differences: the Mg content of the thermal waters is very low (<1 mg L−1) when compared with that of groundwaters (up to 12.9 mg L−1), and the Mn content of groundwater (<5 µg L−1) is much lower than that of the thermal waters (up to 18.4 mg L−1). The low Mg content of thermal waters can be explained by the absorption of Mg in high-temperature geothermal systems. As the geothermal fluids flow from a high temperature to a low temperature environment, it appears to adsorb large amounts of Mg into contacting host rocks [13,18]. The occurrence of Mn in thermal water might be due to the dissolution of secondary Mn-bearing minerals that are associated with epithermal activity.
Figure 5 indicates the influence of temperature on pH, EC, and major-ion concentrations in thermal waters and groundwaters, in which the correlation coefficients (R) between variables are given. All of the parameters, except HCO3 concentration in thermal waters, are positively correlated with temperature with R2 values of 0.59–0.89, indicating that temperature plays an important role in the promotion of water–rock interactions. The inverse trend of HCO3 in thermal waters may be explained by increasing temperature causing a decrease in the dissolution rate of CO2. In Figure 5, Na++K+, SO42−, Cl, and SiO2 concentrations in thermal water as a function of temperature can be separated into two types: (1) A type of a higher-ion concentration at higher temperatures and (2) B type of a lower-ion concentration at lower temperatures, with these types corresponding to chemical types of Piper diagram.
Major cations and anions (expressed as meq percentiles) are shown in a trilinear plot to delineate the geochemical evolution and chemical types of thermal waters and groundwaters (Piper diagram; Figure 6). Chemical compositions of the groundwaters (MG) are of the Ca–HCO3 type, whereas the thermal waters (MH) can be chemically divided into a Na–HCO3 type at lower temperatures (type I) and an Na–Cl (HCO3, SO42−) type at higher temperatures (type II). It is generally considered that, in the geochemical evolution of thermal waters and deep groundwater in granite and gneiss areas of South Korea, the waters are of the Ca–HCO3 type initially, progressing through the Ca(Na)–HCO3 type to the alkaline Na–HCO3 type in the final stage [13,19]. The two different types of thermal water appear to have evolved in different geological environments.

4.2. O, H, and S Isotopic Compositions

Table 1 shows the oxygen (δ18O) and hydrogen (δ2H) isotopic compositions of thermal waters and local groundwater. The δ18O and δ2H values of thermal water are in the ranges of −8.6% to −7.9%, and −55% to −51%, respectively. The δ18O and δ2H values of the local groundwater are in the ranges of −8.4% to −7.6% and −58% to −54%, respectively.
In a δ2H–δ18O diagram (Figure 7), δ18O and δ2H values plot close to the global meteoric water line (GMWL) [20] and local meteoric water line (LMWL), with thermal waters above and groundwaters below or on the GMWL/LMWL, with the following local isotopic relationship applying: δ2D = 7.78δ18O + 5.3 [21]. The isotopic fractionation of thermal waters after meteoric water recharge results in D enrichment, with such an enrichment in other thermal waters of South Korea having been previously reported [2]. It may be explained that the clustering of thermal waters above the LMWL that indicates deuterium enrichment is a result of H2S exchange [22,23]. H2S isotopic exchange reaction, which may lead to a significant deviation from the initial stable isotope content of thermal water, is the following: 2H1HS(g) + 1H1HO(l) = 1H1HS(g) + 2H1HO(l) [23].
The oxygen shift due to oxygen isotopic exchange in water–rock interactions, as observed in high-temperature thermal waters of Japan and other countries [24], is not usually observed in the Magumsan thermal waters. Although isotopic fractionation in thermal waters occurs in a narrow range (within 1% for δ18O), the δ18O and δD values of thermal waters are negatively correlated with temperature (R2 = 0.79–0.82) (Figure 8), with higher-temperature thermal waters having more isotopically depleted values. This may reflects the altitude effect, with the more depleted water being recharged at higher altitudes and migrating deeper into the heat source, and also reflecting another effect, such as silicate hydration. During the hydration of silicate minerals in thermal water contact with rocks, the fractured rocks would preferentially use heavy isotopes, which cause the fractionation of thermal water enriched in light isotopes [25].
The sulfur isotopic ratios (δ34S) were determined to elucidate the origin of sulfur dissolved as sulfate in Magumsan thermal waters. The δ34S values of dissolved sulfate ranged from +7% to +15% (Table 1), with the thermal waters being classified into two types A type of a lower level of +7.0% to +7.2%, and B type of a higher level of +9.9% to +15.0% (Figure 9). This classification corresponds to the chemical types of thermal waters in the trilinear plot of Figure 6, with the former type representing the lower-temperature Na–HCO3 type, and the latter type the higher-temperature Na–Cl (HCO3, SO4) type.
The δ34S values of sulfide minerals in ore deposits of South Korea are known to have ranges of −2.1% to 9.2% [25,26,27,28], whereas the δ34S values of atmospheric and seawater sulfur have ranges of +6% to +7% and values of +20%, respectively [29]. Kim and Nakai [2] reported δ34S values of dissolved sulfate in 17 hot-spring waters in South Korea in the range of +5.5% to +29.3%, and classified the thermal waters into three groups: (1) those including sulfate of seawater origin (δ34S > +19%), (2) those with sulfate formed by oxidation of igneous sulfide minerals (δ34S +5% to +10%), and (3) those with a mixture of these sources (δ34S +10% to 19%). Although only one Magumsan thermal water sample (at 40.7 °C) was analyzed by Kim and Nakai [2], those authors suggested that sulfate S originated from mixed sources of seawater and magmatic pyrite sulfate.
Pyrite is commonly found in andesitic rocks and drill cores, and it is generally considered that the oxidation of pyrite is a source of sulfate in low-δ34S (<7.2%) thermal waters. Here, we consider that there are two possible sources of higher δ34S (>9.9%) values: (1) the derivation of sulfate from the kinetic oxidation of magmatic pyrites in the thermal aquifer system, with kinetically controlled chemical oxidation of reduced sulfides (and sulfur) forming SO42− that was enriched in 34S [30,31]; and, (2) paleo-seawater mixing, as indicated by the positive correlation (R2 = 0.66–0.95) between temperature, Cl, SO42−, and δ34S in thermal waters (Figure 9). The δ34S values are classified into two distinct groups (higher and lower) in terms of temperature and SO42− and Cl concentrations in thermal waters.

4.3. Helium Analysis

The isotopes of noble gases are useful for studying heat sources [32,33]. In thermal systems, the atmospheric 3He/4He ratio (Ra = 1.384 × 10−6) is generally used as a reference value [34]. He isotopic ratios are generally interpreted in terms of the mixing of a MORB-type upper-mantle source end-member (average 3He/4He = 12 × 10−6, or 8.5 ± 1 Ra [35]) and a crustal source end-member (3He/4He 0.01 × 10−6, or 0.01–0.1 Ra [36]), where 3He is mantle derived and 4He is predominantly of crustal origin and associated with radioactive decay of U–Th series elements.
Here, the He and Ne isotopic ratios were studied in the Magumsan thermal waters, with Table 2 presenting concentrations and isotopic ratios of noble gases. 3He/4He and 4He/20Ne ratios of thermal waters were in the ranges 0.335–0.442 (× 10−6) and 2.08–21.7, respectively, which indicated the predominance of crustal He in thermal waters.
In a 3He/4He–4He/20Ne diagram (Figure 10), the 4He/20Ne ratio can be considered as an indicator of atmospheric He contamination in the thermal waters, with He ratios for atmospheric, crustal, and upper-mantle end-members being shown for comparison. The air–mantle mixing line represents a mixture of isotopic ratios for atmospheric He and Ne (3He/4He = 1.384 × 10−6; 3He/20Ne = 0.317 [37]) and those of MORB-type mantle origins (3He/4He = 12 × 10−6, or 8.5 ± 1 Ra; 4He/20Ne > 10,000 [35]). The air–crust mixing line represents a mixture of atmospheric and crustal He and Ne (3He/4He = 5 × 10−9; 4He/20Ne > 10,000). Although the crustal 3He/4He ratios range from 4 × 10−9 to 1 × 10−8, depending on the chemical composition of the crust [36], a crustal 3He/4He ratio of 5 × 10−9 remains valid for the entire range of crustal values when being expressed in terms of orders of magnitude.
The mixing line that passes through the data points (Figure 10) was calculated by assuming two-component mixing between atmospheric He and Ne and a source with unknown 3He/4He ratio and 4He/20Ne = 10,000 [38]. Park et al. (2016) provided details of the calculation [9].
Notably, He data for thermal waters plot on a single air-mixing line between mantle and crustal He components (Figure 10) with 3He/4He = 0.34 × 10−6, which is a similar ratio to that observed in other thermal waters in South Korea not enriched in mantle 3He [6,7,8,9]. The distribution of He isotopic ratios on a single mixing line also indicates that He originates from a common source, with the range of 4He/20Ne ratios representing the degree of atmospheric He contamination through circulating meteoric thermal water that is saturated with dissolved noble gases.
The measured He concentrations in thermal waters represent a mixture of three source components (atmosphere, crust, and mantle), with mixing ratios, as given for 3He and 4He in Table 3. These ratios were calculated while using the method of Kotarba and Nagao (2008) [39]. In thermal waters, the crust-derived 4He contributed 83%–96% and mantle-derived 4He 1.6%–3.4% of the total 4He, with the atmospheric contribution being 1.5%–15.3%. The mantle- and crust-derived 3He contributions were in the ranges 47%–92.2% and 2.0%–2.8%, respectively. Sample MH-3 has a high 3He air-contamination ratio of 51% (15.3% for 4He).
Figure 11 shows the relationships between the 4He concentrations and 4He/20Ne ratios and temperature, where both are positively correlated with temperature (R2 = 0.72 and 0.57, respectively). These relationships indicate that the heat source for Magumsan thermal waters is strongly associated with the radioactive decay of U and Th in the crustal rocks.

4.4. Geothermometry and a Mixing Model for the Thermal Reservoir

Cation and silica geothermometers are often used to estimate the temperatures of thermal waters. A requirement for the successful application of cation geothermometers is the attainment of water–rock chemical equilibrium in the geothermal reservoir [13]. The mixing of thermal and cold waters may cause errors in chemical geothermometry, but silica geothermometry can provide realistic estimates of reservoir temperature, even over the ranges of 120–180 °C [40]. Chalcedony geothermometry can also be utilized at temperatures of <190 °C [13,41,42].
Table 4 presents Magumsan reservoir temperatures that are estimated using cation, quartz, and chalcedony geothermometers. Cation geothermometry results show little agreement, with uncertainties mainly resulting from the mixing of thermal and shallow cold water, and a lack of equilibrium [43]. Each geothermometer has its own limitations [44,45], and water–rock equilibrium in geothermal reservoirs is a prerequisite for accuracy. Even so, it is necessary to select an appropriate method that is based on the chemical characteristics of the thermal water as determined by a Na–K–Mg ternary diagram (Figure 12) [10].
The Na–K–Mg ternary diagram has been previously used to estimate the reservoir temperatures and to gauge whether equilibrium has been attained between thermal water and mineral assemblages comprising albite, K-feldspar, muscovite, and clinochlore [11]. The diagram classifies water into three groups (zones): (I) fully equilibrated water (mature), (II) partially equilibrated water (mixed water), and (III) immature waters (Figure 12). For waters within zone III, where water–rock equilibrium does not exist, cation geothermometry would not provide reliable results [46]. The thermal waters and groundwaters in the study area plot along (or around) the boundary between zones II and III, and in zone III (Figure 12). The thermal waters are most likely on the boundary between partial and no equilibrium due to their ascent from a deep aquifer and mixing with cold shallow waters at certain ratios. This mixing causes changes in both temperature and chemical composition, and full water–rock equilibrium may not be attained [13,18].
Silica (quartz and chalcedony) and cation (Na–K–Ca and Na–K) geothermometers were used to estimate the reservoir temperatures of the Magumsan geothermal system while using the chemical analyses of the thermal waters using the equations that are listed in Table 4.
Reservoir temperatures of thermal waters that are calculated by quartz and chalcedony geothermometers are in the ranges 85–113 °C and 54–86 °C, respectively. Thermal/cold water mixing ratios that are calculated by these geothermometers are 68%–85% and 54%–74%, respectively (Table 5). Therefore, the silica geothermometer is considered to be the most suitable for application in the Magumsan area, and the silica–enthalpy model [47] was applied. In this method, enthalpy is used as a coordinate rather than temperature, because the combined heat contents of two waters at different temperatures are conserved when the waters are mixed but their combined temperatures are not [12].
In a silica-enthalpy diagram (Figure 13), the mean temperature of cold shallow water was assumed to equal the annual mean temperature (14.5 °C) in the study area, and the SiO2 concentrations of thermal waters and shallow cold water were assumed to be 45–69 mg L−1 and 17.6 mg L−1, respectively, based on the data for surface water of Choi and Kim [1]. With reference to the steam table [52], the enthalpies of the thermal waters and cold shallow water were calculated to be 129–208 KJ kg−1, and 65 KJ kg−1, respectively. The intersection point of the mixing line with the solubility of chalcedony indicates an enthalpy of 555−707 KJ kg−1 and a maximum silica content of 63 mg L−1 in thermal water. The estimated reservoir temperature was in the range of 132–167 °C. Table 6 gives the reservoir enthalpies, silica concentrations, and mixing ratios that were calculated while using the silica–enthalpy method.
Regarding quartz solubility, it was assumed that the reservoir temperature is in the range 171–206 °C, which is higher than the range (54–113 °C) that was calculated by chalcedony and quartz geothermometry. The overestimated temperature probably resulted from the occurrence of conductive cooling [12]. The mixing ratio of thermal and cold water can be estimated while using a graphical method with the silica–enthalpy model [51]. The mixing ratio of cold/thermal water was estimated at 76%–86% for chalcedony and 81%–90% for quartz, with these ratios being higher than those that were given by chalcedony geothermometry, which was considered to be a reliable method.
The thermodynamic equilibrium geothermometer that is based on the calculation of saturation index (SI) values in water–mineral interaction systems can also be used to assess thermal reservoir temperatures. Here, SI values for carbonate (calcite and dolomite), silicate minerals (albite, anorthite, K-mica, chlorite, illite, and kaolinite), and silica minerals (quartz and chalcedony), which are potential reactive primary minerals and secondary minerals, were used to predict the tendency for precipitation or the dissolution of these minerals, with SI = 0 indicating thermodynamic equilibrium, SI > 0 indicating oversaturation (precipitation conditions), and SI < 0 indicating undersaturation (dissolution conditions).
The chemical data of Table 1 were used to calculate SI values of thermal waters while using the software program PHREEQC [53,54] under the assumption that no conductive cooling occurs. The equilibrium state between the minerals and water is temperature dependent. Saturation indices between chemical compositions of thermal waters and potential reactive minerals were calculated as a function of temperature, with the results being shown in Figure 14. The results indicate that the thermal waters are oversaturated with respect to kaolinite, K-mica, calcite, chalcedony, and quartz, and they are undersaturated with respect to albite, anorthite, illite, chlorite, and dolomite at their discharge temperatures. The composition of thermal water is more likely to reflect the state of water–rock equilibrium at deeper reservoir temperatures, rather than at the discharge temperature. This value is most likely the reservoir temperature when the equilibrium lines of a series of potential reactive minerals converge at a certain temperature [55].
SI values of kaolinite, K-mica, dolomite, chalcedony, and quartz for the thermal waters converge on and lie close to SI = 0 in the temperature range of 55–83 °C. This may be the temperature range in which the maximum number of mineral phases is in equilibrium with the thermal waters, and can thus be interpreted as corresponding to the reservoir temperature, which is consistent with chalcedony geothermometry.

5. Conclusions

Magumsan thermal waters from six deep wells at temperatures of 30.8–49.8 °C can be chemically classified into two groups: lower-temperature (<31 °C) Na–HCO3 type and higher-temperature (>40 °C) Na–Cl (HCO3, SO42−) type, which are both quite different from local groundwater of Ca–HCO3 type. It is inferred that major ions, such as Na+, Cl, and SO42−, and SiO2 in thermal waters are derived from interactions of water with andesitic host rock, pyrite, and chert associated with epithermal activity, and mixing with paleo-seawater in the deep aquifer, which may be clarified by age dating, such as 36Cl isotope analysis as further research.
δ18O and δD values of the thermal waters are clustered above the LMWL, resulting in silicate oxidation and H2S exchange. Isotopic depletion with increasing temperature suggests that isotopically lighter meteoric water that was recharged at higher altitudes migrates into a deeper and higher-temperature thermal reservoir. The δ34S values of dissolved sulfate in thermal waters can be classified into two groups: low-δ34S (7.0%–7.2%) and high-δ34S (9.9%–15%) waters, with the former corresponding to the lower-temperature Na–HCO3 type water and the latter to the higher-temperature Na–Cl (HCO3, SO4) type. The sulfate in higher-δ34S (>9.9%) waters has two possible sources: kinetic-controlled oxidation of magmatic pyrite in the thermal aquifer causing enrichment in 34S, and paleo-seawater in the deep aquifer mixing with thermal water of meteoric origin.
In a ternary end-member mixing diagram, the 3He/4He ratios in thermal waters plot along a single air-mixing line between mantle and crustal origins with a dominant contribution (80.6%–95.7%). A positive correlation between 4He contents and 4He/20Ne ratios and temperatures of thermal water, together with the predominant crustal origin of He, strongly implies that the heat source of thermal waters in the Magumsan area is thermal energy generated by the decay of radioactive elements in crustal rocks.
Magumsan thermal waters plot on the boundary between partial equilibrium and immaturity in a Na–K–Mg ternary diagram, and cation geothermometers and the silica–enthalpy model overestimate the reservoir temperature. A reliable temperature estimate of 55–83 °C was obtained with the thermodynamic equilibrium method, which was consistent with a range of 54–86 °C estimated by chalcedony geothermometry. The mixing ratio of cold shallow water with thermal water is estimated to be 54%–74%.

Author Contributions

C.-H.J., B.-D.L., S.-W.A., J.-H.Y., Y.-C.L., Y.-J.L., and H.-W.J. conceived and designed the research project. C.-H.J., K.N., S.-W.A., and K.-H.K. performed the fieldwork and collected the water samples. K.N. and K.-H.K. conducted noble-gas analyses. All authors read and approved the manuscript.

Funding

This research was financially supported by a Basic Research Project (19-3411) of the Korea Institute of Geoscience and Mineral Resources (KIGAM), funded by the Ministry of Science and ICT of Korea, and by the Daejeon University research fund (20173610).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choi, C.I.; Kim, H.C. Geological Survey of Korea Bulletin No. 11: The Hot Spring of Makeumsan Area; Korea Institute of Geoscience and Mineral Resources: Daejeon, Korea, 1969; pp. 207–226. [Google Scholar]
  2. Kim, K.H.; Nakai, N. A study on hydrogen, oxygen and sulfur isotopic ratios of the hot spring waters in South Korea. Geochemistry 1981, 15, 6–16. [Google Scholar]
  3. Park, S.S. Hydrogeochemical Studies on the Origin and Geochemical Environments of Thermal Groundwaters in Bugok and Magumsan Areas, Southern Korea. Ph.D. Thesis, Korea University, Seoul, Korea, 2005. [Google Scholar]
  4. Karakuş, H. Helium and carbon isotope composition of gas discharges in the Simav Geothermal Field, Turkey: Implications for the heat source. Geochemistry 2013, 57, 213–223. [Google Scholar] [CrossRef]
  5. Wang, X.; Zhao, L.; Liu, X.; Lili, A.; Zhang, Y. Temperature effect on the transport of nitrate and ammonium ions on a loose-pore geothermal reservoir. J. Geochem. Exp. 2013, 124, 59–66. [Google Scholar] [CrossRef]
  6. Jeong, C.H.; Hur, H.S.; Nagao, K.; Kim, K.H. Hydrochemical and isotopic characteristics, and origin of noble gas for low-temperature hot spring waters in the Honam area. Econ. Environ. Geol. 2007, 40, 635–649. [Google Scholar]
  7. Jeong, C.H.; Nagao, K.; Kim, K.H.; Choi, H.K.; Sumino, H.; Park, J.S.; Park, C.H.; Lee, J.I.; Hur, S.D. Hydrochemistry and noble gas origin of various hot spring waters from the eastern area in South Korea. J. Soil Groundw. Environ. 2008, 13, 1–12. [Google Scholar]
  8. Jeong, C.H.; Koh, Y.K.; Shin, S.H.; Nagao, K.; Kim, K.H.; Kim, G.Y. Hydrochemistry and noble gas origin of hot spring waters of Icheon and Pocheon area in Korea. J. Eng. Geol. 2009, 19, 529–541. [Google Scholar]
  9. Park, J.; Jeong, C.H.; Nagao, K.; Yang, J.H.; Sumino, H.; Kim, K.H.; Kim, M.S.; Lee, J.I.; Park, C.H.; Koh, Y.K.; et al. Hydrochemistry and noble gas geochemistry of geothermal waters in Chungcheong Province, South Korea. Geochem. J. 2016, 50, 89–103. [Google Scholar] [CrossRef]
  10. Shestakova, A.; Gusva, N.; Kopylova, Y.; Khvaschevskaya, A.; Polya, D.A.; Tokarev, I. Geothermometry and isotope geochemistry of CO2-rich thermal waters in Choygan, East Tuva, Russia. Water 2018, 10, 729. [Google Scholar] [CrossRef]
  11. Giggenbach, W.F.; Werner, F. Geothermal solute equilibria. Deviation of Na-K-Mg-Ca geoindicators. Geochim. Cosmochim. 1988, 52, 2749–2765. [Google Scholar] [CrossRef]
  12. Ozler, H. Hydrogeology and geochemistry in the Curuksu (Denizli) hydrothermal field, western Turkey. Environ. Geol. 2000, 39, 1169–1180. [Google Scholar]
  13. Yuce, G.; Taşkıran, L. Isotope and chemical composition of thermal fluids at Tekman geothermal area (Eastern Turkey). Geochem. J. 2013, 47, 423–435. [Google Scholar] [CrossRef]
  14. Chough, S.K.; Kwon, S.T.; Ree, J.H.; Choi, D.K. Tectonic and sedimentary evolution of the Korean peninsula: A review and new view. Earth Sci. Rev. 2000, 52, 175–235. [Google Scholar] [CrossRef]
  15. Kim, H.C.; Lee, Y. Heat flow in the Republic of Korea. J. Geophys. Res. 2007, 112, B05413. [Google Scholar] [CrossRef]
  16. Kim, N.J.; Lee, H.K. Explanatory text of the geological map of Yongsan sheet (1:50,000); Korea Institute of Geoscience and Mineral Reosources: Daejeon, Korea, 1964; pp. 1–31. [Google Scholar]
  17. Aka, F.T.; Kusakabe, M.; Nagao, K.; Tanyileke, G. Noble gas isotopic compositions and water/gas chemistry of soda springs from the islands of Bioko, Sao Tome and Annobon, along with Cameroon Volcanic Line, West Africa. Appl. Geochem. 2001, 16, 323–338. [Google Scholar] [CrossRef]
  18. Ongay, E. Hydrogeological and Hydrochemical Investigation of Balıkesir-Pamukçu Geothermal Area and its Vicinity. Master’s Thesis, Dokuz Eylul University, İzmir, Turkey, 2004. [Google Scholar]
  19. Jeong, C.H.; Kim, C.S.; Kim, T.K.; Kim, S.J. Reaction path modeling on geochemical evolution of groundwater and formation of secondary minerals in water-gneiss reaction system. J. Miner. Soc. Korea 2007, 10, 33–44. [Google Scholar]
  20. Craig, H. Isotopic variations in meteoric water. Science 1961, 133, 1702–1703. [Google Scholar] [CrossRef]
  21. Koh, Y.K.; Bae, D.S.; Kim, C.S.; Kim, K.S.; Chung, H.J.; Kim, S.Y. Consideration of the groundwater recharge based on environmental isotopic characteristics of the small basin in the Yeosu Area. J. Soil Groundw. Environ. 2001, 6, 93–106. [Google Scholar]
  22. Geyh, M. Groundwater Saturated and Unsaturated Zone. In Environmental Isotopes in the Hydrological Cycle Principles and Applications, Technical Document in Hydrology; UNESCO: Paris, France, 2000; Volume 4, pp. 1–196. [Google Scholar]
  23. D’amore, F.; Panichi, C. Geochemistry in geothermal exploration. Energy Res. 1985, 9, 277–298. [Google Scholar] [CrossRef]
  24. Matsubaya, O.; Sakai, H.; Kusachi, I.; Setake, H. Hydrogen and oxygen isotopic ratios and major element chemistry of Japanese thermal water system. Geochem. J. 1973, 7, 123–151. [Google Scholar] [CrossRef]
  25. Allegre, C.J. Isotope Geology; Cambridge University Press: Cambridge, UK, 2008; 534p. [Google Scholar]
  26. Yoo, B.C. Genesis of the Ogcheon gold-silver deposit in Republic of Korea: Ore minerals, fluidinclusion and stable isotope studies. Econ. Environ. Geol. 2013, 46, 153–163. [Google Scholar] [CrossRef]
  27. Yu, H.M.; Shin, D.B. Mineralization and genetic environments of the central and main orebodies in the Manjang deposit, Geosan. J. Miner. Soc. Korea 2018, 31, 87–101. [Google Scholar] [CrossRef]
  28. Yoo, B.C.; You, B.W. Geopung copper deposit in Ogcheon, Chungcheongbuk-do: Mineralogy, fluid inclusion and stable isotope studies. Econ. Environ. Geol. 2001, 44, 193–201. [Google Scholar] [CrossRef]
  29. Koh, Y.K.; Yun, S.T.; Kim, C.S.; Bae, D.S.; Park, S.S. Geothermal evolution and deep environment of the Geothermal waters in the Bugok area: Reconsideration on the origin of sulfate-type geothermal water. Econ. Environ. Geol. 2001, 34, 329–343. [Google Scholar]
  30. Harrison, A.G.; Thode, H.G. The kinetic isotope effect in the chemical reduction of sulphate. Trans. Faraday Soc. 1957, 53, 1648–1651. [Google Scholar] [CrossRef]
  31. Ohmoto, H.; Rye, R.O. Isotopes of Sulfur and Carbon. In Geochemistry of Hydrothermal Ore Deposits, 2nd ed.; Barnes, H.L., Ed.; Join Wiley and Sons, Inc.: New York, NY, USA, 1979; pp. 509–567. [Google Scholar]
  32. Torgersen, T.; Jenkins, W.J. Helium isotopes in geothermal system: Iceland, the geysers, raft river and steamboat springs. Geochim. Cosmochim. Acta 1982, 46, 739–748. [Google Scholar] [CrossRef]
  33. Yokoyama, T.; Nakai, S.I.; Wakita, H. Helium and carbon isotopic compositions of hot spring gases in the Tibetan Plateau. J. Volcanol. Geotherm. Res. 1999, 88, 99–107. [Google Scholar] [CrossRef]
  34. Hulston, J.R.; Lupton, J.E. Helium isotope studies of geothermal fluids in the Taupo Volcanic Zone, New Zealand. J. Volcanol. Gemtherm. Res. 1996, 74, 297–321. [Google Scholar] [CrossRef]
  35. Graham, D.W. Noble gas isotope geochemistry of mid-ocean ridge and ocean island basalts: Characterization of mantle source reservoirs. Rev. Miner. Geochem. 2002, 47, 247–317. [Google Scholar] [CrossRef]
  36. Ballentine, C.J.; Burnard, P.G. Production, release and transport of noble gases in continental crust. Rev. Miner. Geochem. 2002, 47, 481–538. [Google Scholar] [CrossRef]
  37. Ozima, M.; Podosek, F.A. Noble Gases in the Earth. In Noble Gas Geochemistry, 2nd ed.; Cambridge University Press: Cambridge, UK, 2002; pp. 217–249. [Google Scholar]
  38. Sano, Y.; Wakita, H. Geographical distribution of 3He/4He ratios in Japan: Implications for arc tectonics and incipient magmatism. J. Geophys. Res. Solid Earth 1985, 90, 8729–8741. [Google Scholar] [CrossRef]
  39. Kotaba, M.J.; Nagao, K. Composition and origin of natural gases accumulated in the Polish and Ukrainian parts of the Carpathian region: Gaseous hydrocarbons, noble gases, carbon dioxide and nitrogen. Chem. Geol. 2008, 255, 426–438. [Google Scholar] [CrossRef]
  40. Arnórsson, S.; Gunnlaugsson, E. New gas geothermometers for geothermal exploration-calibration and application. Geochim. Cosmochim. 1985, 49, 1307–1325. [Google Scholar] [CrossRef]
  41. Henley, R.W.; Truesdell, A.H.; Barton, P.B., Jr.; Whitney, J.A. Fluid-Mineral Equilibria in Hydrothermal Systems; Society of Economic Geologist, Inc.: Littleton, CO, USA, 1984; pp. 1–267. [Google Scholar]
  42. Fournier, R.O. Chemical geothermometers and mixing models for geothermal systems. J. Geotherm. 1977, 5, 41–50. [Google Scholar] [CrossRef]
  43. Gu, X.; Zhang, Q.; Cui, Y.; Shao, J.; Xiao, Y.; Zhang, P.; Liu, J. Hydrogeochemistry and Genesis Analysis of Thermal and Mineral Springs in Arxan, Northeastern China. Water 2017, 9, 61. [Google Scholar] [CrossRef]
  44. Dulanya, Z.; Morales-Simfors, N.; Sivertun, Å. Comparative study of the silica and cation geothermometry of the Malawi hot springs: Potential alternative energy source. J. Afr. Earth Sci. 2010, 57, 321–327. [Google Scholar] [CrossRef]
  45. Pope, L.A.; Hajash, A.; Popp, R.K. An experimental investigation of the quartz, Na-K, Na-K-Ca geothermometers and the effects of fluid composition. J. Volcanol. Geotherm. Res. 1987, 31, 151–161. [Google Scholar] [CrossRef]
  46. Tarcan, G. Geothermal Water Chemistry, Centre of Geothermal Energy. Res. Appl. (Jenarum) Summer Sch. Publ. 2002, 87–113. [Google Scholar]
  47. Trusdell, A.H.; Fournier, R.O. Procedure for estimating the temperature of a hot-water component in a mixed water by using a plot of dissolved silica versus enthalpy. J. Res. USGS 1977, 5, 49–52. [Google Scholar]
  48. Fournier, R.O.; Truesdell, A.H. An empirical N-K-Ca chemical geothermometer for natural waters. Geochem. Cosmochim. Acta 1973, 37, 1255–1275. [Google Scholar] [CrossRef]
  49. Truesdell, A.H. Summary of Section III: Geochemical Techniques in Exploration; Lawrence Berkeley Labortory: Berkeley, CA, USA, 1976; pp. 54–80. [Google Scholar]
  50. Fournier, R.O. A revised equation for the Na/K geothermometer. Geotherm. Resour. Counc. Trans. 1979, 3, 221–224. [Google Scholar]
  51. Arnórsson, S. Chemical equilibria in icelandic geothermal systems: Implications of chemical geothermometry investigations. Geothermics 1983, 12, 119–128. [Google Scholar] [CrossRef]
  52. Heat Exchanger Design Handbook. Available online: http://hedhme.com/content_map/?link_id=22140&article_id=526 (accessed on 19 June 2019).
  53. PHREEQC. (Computer File) Water Chemistry, Computer Programs Geochemistry; Version 3.0; US Geological Survey: Denver, CO, USA, 1999.
  54. Parkhurst, D.L. User’s Guide to PHREEQC–A Computer Program for Speciation, Reaction-Path, Advective-Transport, and Inverse Geochemical Calculations; US Geological Survey: Denver, CO, USA, 1995.
  55. Tole, P.M.; Ármannsson, H.; He, P.Z.; Arnórsson, S. Fluid/mineral equilibrium calculations for geothermal chemical geothermometry. Geothermics 1993, 22, 17–37. [Google Scholar] [CrossRef]
Figure 1. Location and geological maps. (a) Tectonic framework of northeast Asia; (b) major geological units of South Korea; (c) geology of the Magumsan area, including a cross-section (A–A′) [16].
Figure 1. Location and geological maps. (a) Tectonic framework of northeast Asia; (b) major geological units of South Korea; (c) geology of the Magumsan area, including a cross-section (A–A′) [16].
Water 11 01774 g001
Figure 2. Lithological log and temperature gradient in a drilled core from the Magumsan thermal area [1].
Figure 2. Lithological log and temperature gradient in a drilled core from the Magumsan thermal area [1].
Water 11 01774 g002
Figure 3. Map showing sample locations and temperatures of thermal water (MH) and groundwater (MG) in the Magumsan area.
Figure 3. Map showing sample locations and temperatures of thermal water (MH) and groundwater (MG) in the Magumsan area.
Water 11 01774 g003
Figure 4. Schöeller diagram comparing concentration ranges of major ions in thermal water (MH) and groundwater (MG). Filled and open symbols indicate maximum and minimum values, respectively.
Figure 4. Schöeller diagram comparing concentration ranges of major ions in thermal water (MH) and groundwater (MG). Filled and open symbols indicate maximum and minimum values, respectively.
Water 11 01774 g004
Figure 5. Relationships between chemical parameters and the temperatures of thermal waters (filled circles) and groundwater (open triangles).
Figure 5. Relationships between chemical parameters and the temperatures of thermal waters (filled circles) and groundwater (open triangles).
Water 11 01774 g005
Figure 6. Piper diagram of the major chemical compositions of thermal waters (MH: filled triangles and circles) and groundwater (MG: blue pluses) expressed in milli-equivalent percentages. See Section 4.1 for details.
Figure 6. Piper diagram of the major chemical compositions of thermal waters (MH: filled triangles and circles) and groundwater (MG: blue pluses) expressed in milli-equivalent percentages. See Section 4.1 for details.
Water 11 01774 g006
Figure 7. δ18O–δD relationship for thermal waters and groundwaters in the study area. Numbers indicate the sample number of thermal waters (MH) and groundwaters (MG). (GMWL: Global Meteoric Water Line; LMWL: Local Meteoric Water Line).
Figure 7. δ18O–δD relationship for thermal waters and groundwaters in the study area. Numbers indicate the sample number of thermal waters (MH) and groundwaters (MG). (GMWL: Global Meteoric Water Line; LMWL: Local Meteoric Water Line).
Water 11 01774 g007
Figure 8. Relationships between temperature and δ18O (a) and δD (b) of the thermal waters (MH) and groundwaters (MG).
Figure 8. Relationships between temperature and δ18O (a) and δD (b) of the thermal waters (MH) and groundwaters (MG).
Water 11 01774 g008
Figure 9. Relationships between chemical composition, temperature, and δ34S of the thermal water.
Figure 9. Relationships between chemical composition, temperature, and δ34S of the thermal water.
Water 11 01774 g009
Figure 10. Three-end-member mixing diagram for He isotopic ratios in Magumsan thermal waters. (MH-3: 15.3% air contaminated sample; other samples: 1.5%–4.0% air contaminated samples).
Figure 10. Three-end-member mixing diagram for He isotopic ratios in Magumsan thermal waters. (MH-3: 15.3% air contaminated sample; other samples: 1.5%–4.0% air contaminated samples).
Water 11 01774 g010
Figure 11. Correlations between 4He concentration (a) and 4He/20Ne isotopic ratio (b) and temperature of thermal waters.
Figure 11. Correlations between 4He concentration (a) and 4He/20Ne isotopic ratio (b) and temperature of thermal waters.
Water 11 01774 g011
Figure 12. Ternary Na–K–Mg plot of thermal waters (filled circles with number) and groundwaters (open circles), showing water–rock equilibrium temperatures [11].
Figure 12. Ternary Na–K–Mg plot of thermal waters (filled circles with number) and groundwaters (open circles), showing water–rock equilibrium temperatures [11].
Water 11 01774 g012
Figure 13. Enthalpy–silica mixing diagram of thermal waters (filled circles) and cold shallow water (open circle) in the Magumsan area.
Figure 13. Enthalpy–silica mixing diagram of thermal waters (filled circles) and cold shallow water (open circle) in the Magumsan area.
Water 11 01774 g013
Figure 14. Changes in saturation indices of thermal waters with respect to potential reactive minerals, as a function of temperature.
Figure 14. Changes in saturation indices of thermal waters with respect to potential reactive minerals, as a function of temperature.
Water 11 01774 g014
Table 1. Geochemical and isotopic compositions of thermal water and groundwater samples that were collected in the Magumsan area.
Table 1. Geochemical and isotopic compositions of thermal water and groundwater samples that were collected in the Magumsan area.
(Unit: mg L−1)
Sample
ID
Temp
(°C)
pHORP
(mV)
EC
(μS cm−1)
DONa+K+Ca2+Mg2+Sr2+Fe
MH-142.27.78−10412723.401888.9644.20.670.830.008
MH-245.07.52−7110441.301547.8439.00.910.650.003
MH-331.37.48−885511.6092.12.1620.90.550.320.01
MH-449.87.8−8214896.4020713.347.10.521.070.013
MH-539.47.53−7510380.501636.6933.90.560.600.004
MH-630.87.42−1087400.101072.3034.60.510.680.006
MG-120.87.35+1902224.0014.10.6026.64.130.130.002
MG-220.97.09+2281884.9010.10.6124.63.660.900.117
MG-323.57.33+4024364.8047.28.5727.212.90.260.002
MG-417.95.86+1504801.9039.55.0938.611.00.340.002
MG-521.07.20+1616643.1056.40.9385.13.781.850.002
Sample
ID
MnSiO2HCO3F2212ClSO42−E (%)δ18O
(%)
δ2H
(%)
δ34S
(%)
MH-10.6155.82191.9217757.63.82−8.4−5513.0
MH-210.254.12140.1713149.63.94−8.4−5410.0
MH-34.0135.12300.3824.019.52.76−7.9−517.0
MH-48.8562.51860.4824061.32.70−8.6−5515.0
MH-52.4950.52300.2614364.0−1.26−8.2−539.9
MH-618.434.43070.3635.524.7−0.65−8.2−527.2
MG-10.00233.2134<0.057.056.88−5.81−8.4−57
MG-20.00233.2118<0.053.134.89−5.18−8.3−57
MG-30.00222.9177<0.0521.637.24.08−8.1−58
MG-40.00218.0126<0.0550.621.8−3.58−7.6−55
MG-50.00526.9328<0.0528.767.8−4.54−7.8−54
Notes: E: charge balance error.
Table 2. Helium and neon isotopic data for Magumsan thermal waters.
Table 2. Helium and neon isotopic data for Magumsan thermal waters.
Sample ID4He (×10−7 cm3 g−1)20Ne (×10−9 cm3 g−1)3He/4He (×10−6)4He/20Ne
MH-10.3202710.33511.8
MH-20.6083440.44217.6
MH-30.1024910.4202.08
MH-50.4602120.34021.7
MH-60.2383010.3667.89
Table 3. Mixing ratios of three sources of 3He and 4He in Magumsan thermal waters.
Table 3. Mixing ratios of three sources of 3He and 4He in Magumsan thermal waters.
Sample No.4He3He
Air (%)Mantle (%)Crust (%)Air (%)Mantle (%)Crust (%)
MH-12.72.494.911.385.92.8
MH-21.83.494.85.792.22.1
MH-315.31.683.051.047.02.0
MH-51.52.696.06.091.12.8
MH-64.02.593.515.482.12.6
Table 4. Mathematical equations to estimate reservoir temperatures.
Table 4. Mathematical equations to estimate reservoir temperatures.
GeothermometersEquationsSource
Qtz a (°C)   T =   1309 5.19 l o g SiO 2 273.15     [42]
Qtz b (°C)   T =   1522 5.75 l o g SiO 2 273.15     [42]
Chal a (°C)   T =   1309 4.69 l o g SiO 2 273.15     [42]
Chal b (°C)   T =   1309 5.09 l o g SiO 2 273.15     [42]
Na–K–Ca (°C)   T =   1647 ( l o g Na K + β l o g Ca 0.5 Na + 2.06 ) + 2.47 273.15 ,     β = 4 3   f o r   T < 100     [48]
Na–K1 (°C)   T = 856 0.857 + l o g Na K 273.15     [49]
Na–K2 (°C)   T = 1217 1.483 + l o g Na K 273.15     [50]
Na–K3 (°C)   T = 933 0.933 + l o g Na K 273.15     [51]
K–Mg (°C)   T = 4410 14.0 l o g Na K 273.15     [11]
Note: a no steam loss; b maximum steam loss (at 100 °C); Na–K1, Na–K2 and Na–K3 are the equations suggested by Truesdell, Fournier and Arnórsson, respectively; T is the model reservoir temperature in °C; S is the silica concentration of the thermal water in mg/L; and Na, K, Ca and Mg concentrations are expressed in mol/L.
Table 5. Estimated reservoir temperatures and mixing ratios of thermal waters using quartz, chalcedony, and cation geothermometers.
Table 5. Estimated reservoir temperatures and mixing ratios of thermal waters using quartz, chalcedony, and cation geothermometers.
Sample ID.Estimated Reservoir Temperature and Mixing Ratio
Outflow (°C)Qtz a (°C)Qtz b (°C)Mix q (%)Chal a (°C)Chal b (°C)Mix ch (%)Na–K–Ca (°C)Na–K1 (°C)Na–K2 (°C)Na–K3 (°C)K–Mg (°C)
MH-142.21071070.7577.580.40.6463.382.1127.8102.348.1
MH-245.01061060.7275.979.00.5859.986.4131.6106.642.4
MH-331.386.088.80.8555.060.30.7331.941.990.960.923.3
MH-449.81131120.6883.485.70.5575.2102.4145.7122.959.1
MH-539.4102.31030.7772.475.90.6642.140.189.259.032.7
MH-630.885.188.10.8554.159.50.7426.537.586.856.425.1
Note: Qtz, Quartz; Chal, Chalcedony; a no steam loss; b maximum steam loss (at 100 °C); Mix q and Mix ch are the mixing ratio about quartz and chalcedony, respectively; Na–K1, Na–K2 and Na–K3 are the equations suggested by Truesdell, Fournier and Arnórsson, respectively.
Table 6. Estimated reservoir temperatures of thermal waters and their mixing ratio with cold water based on the mixing model of enthalpy and silica.
Table 6. Estimated reservoir temperatures of thermal waters and their mixing ratio with cold water based on the mixing model of enthalpy and silica.
Sample ID.Enthalpy (KJ kg−1)Mixing Ratio (%)SiO2 (mg L−1)Mixing Ratio (%)
OutflowReservoirCold WaterOutflowReservoirCold Water
MH-117788062.50.8655.827926.80.89
MH-218879562.50.8354.121326.80.85
MH-313372662.50.8935.113126.80.92
MH-420883062.50.8162.524126.80.83
MH-516585562.50.8750.525426.80.90
MH-612972362.50.9034.412526.80.92
Note: Enthalpy and SiO2 of shallow cold water are assumed to be 14.5 °C (60.8 KJ kg−1) and 26.8 mg L−1, respectively.

Share and Cite

MDPI and ACS Style

Jeong, C.-H.; Lee, B.-D.; Yang, J.-H.; Nagao, K.; Kim, K.-H.; Ahn, S.-W.; Lee, Y.-C.; Lee, Y.-J.; Jang, H.-W. Geochemical and Isotopic Compositions and Geothermometry of Thermal Waters in the Magumsan Area, South Korea. Water 2019, 11, 1774. https://doi.org/10.3390/w11091774

AMA Style

Jeong C-H, Lee B-D, Yang J-H, Nagao K, Kim K-H, Ahn S-W, Lee Y-C, Lee Y-J, Jang H-W. Geochemical and Isotopic Compositions and Geothermometry of Thermal Waters in the Magumsan Area, South Korea. Water. 2019; 11(9):1774. https://doi.org/10.3390/w11091774

Chicago/Turabian Style

Jeong, Chan-Ho, Byeong-Dae Lee, Jae-Ha Yang, Keisuke Nagao, Kyu-Han Kim, Sang-Won Ahn, Yong-Cheon Lee, Yu-Jin Lee, and Hyeon-Woo Jang. 2019. "Geochemical and Isotopic Compositions and Geothermometry of Thermal Waters in the Magumsan Area, South Korea" Water 11, no. 9: 1774. https://doi.org/10.3390/w11091774

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