Skip to main content

Origins and pathways of deeply derived carbon and fluids observed in hot spring waters from non-active volcanic fields, western Kumamoto, Japan

Abstract

Natural springs containing volcanic and magmatic components occur in association with these activities. However, features of deeply originated fluids and solutes were less documented from fields, where active volcanic and magmatic activities are not distributed. To characterize the presence of deep components and identify their major pathways 28 groundwater samples (~ 1230 m deep) were collected from hot spring sites located at western coast of Kumamoto, southwestern Japan, where the typical subduction related magmatisms are absent. The samples were measured for dissolved ion concentrations and stable isotope ratios (δ2HH2O, δ18OH2O, δ13CDIC and δ34SSO4) that were compared with data of 33 water samples from vicinity surface systems. The groundwaters were classified into three types based on major hydrochemistry: high Cl− fluid, low concentration fluid, and high HCO3− fluid. Our data set suggests that the high Cl− fluid was formed by saline water mixing with aquifer waters of meteoric origin and subsequently evolved by reverse cation exchange. The low concentration fluid is identical to regional aquifer water of meteoric origin that was subjected to cation exchange. The high HCO3− fluid showed the highest HCO3− concentrations (~ 3,888 mg/l) with the highest δ13CDIC (up to − 1.9‰). Based on our carbon mixing model and observed δ2HH2O and δ18OH2O shifts, it is suggested that dissolved carbon of mantle origin and small fraction of fluids generated in deep crust were transported towards surface through structural weakness under open tectonic setting. These deeply derived components were then mixed with waters in the surface systems and diluted. Their impacts on surface hydrological systems were limited in space except few locations, where deeply connected pathways are anticipated along active structural deformations.

Introduction

Artesian springs with high temperature and/or elevated concentrations occur in major active volcanic and seismotectonic regions globally. These highly concentrated dissolved solutes are facilitated with incorporation of volcanic and magmatic gases and fluids (Aiuppa et al. 2009, 2017), and simultaneous high temperature weathering or alterations that can occur in crust deeper than active hydrological systems in surface (Hedenquist and Lowenstern 1994). Discharge of these fluids is commonly observed in vicinity of active volcano-hydrothermal fields. For example, relevant number of studies tried to depict origins and pathways of deeply sourced components to quantify the impact of volcanic and magmatic activities on lateral material transportations from land to ocean (Chiodini et al. 2000; Dessert et al. 2009; Gaillardet et al. 2011; Rive et al. 2013; Hosono et al. 2018; Perez-Fodich and Derry 2019). Similar high concentration hot spring waters are found in regions, where active volcanic and magmatic activities are not existing (Tomaru et al. 2007; Kusuda et al. 2014; Togo et al. 2014; Morikawa et al. 2016; Kusuhara et al. 2020 and references therein). However, origin, distribution, pathway, and thus their impact of hydrochemical features on surface systems are less well known for these waters.

The major sources invoked to explain for the origin of deeply originated fluids include magmatic fluids, slab fluids and deep crustal fluids (Hedenquist and Lowenstern 1994; Ague 2014; Morikawa et al. 2016). The magmatic fluids present in a crust beneath active volcanoes that are located typically along the volcanic front with some possible exceptions on back arc. Dehydration of subducting oceanic plates releases waters and fluids from slab sediments and rocks (here we call slab fluids) into mantle wedges. Some many hot springs in southwest Japan are thought to be originated directly of these fluids on fore arc system (Tomaru et al. 2007; Kusuda et al. 2014; Togo et al. 2014; Yoshida et al. 2015; Morikawa et al. 2016; Kusuhara et al. 2020). Alternatively, fluids generated in lower crust (here we call deep crustal fluids) can also contribute to surface systems along major pathways especially at regions, where substantial heat source exists beneath lower crust, such as upwelling asthenospheric mantle (e.g., Zhao et al. 2018); however, this phenomenon has been less advocated from nature of lived hot spring discharges due to rare opportunity to test this hypothesis.

Previous studies (Nakajima and Hasegawa 2007; Hirose et al. 2008; Zhao et al. 2021) showed the depth contours of subducting slabs under Japanese archipelago demonstrating that western Kyushu is located on unique setting, where possible subduction plate is in mantle deeper than 200 km and typical subduction related magmatisms are absent. Some hot springs occur along the coast of western Kumamoto in western Kyushu without linkage of active volcanisms in surface (Fig. 1a). Notably, these historically recognized hot springs and other recently developed hot spring facilities are located along and on extension of the major faults (Fig. 1) that possibly allow deeply originated gases and fluids in the crust passing toward surface. There have been no reports comprehensively investigating the origin and formation of these hot springs; however, we thought that precise geochemical characterization on these hot springs may be able to put more constraint on a presence, properties, and discharging processes of these components.

Fig. 1
figure 1

Map showing the study area and its geological background. a Tectonic structure of the Kyushu Islands with location of active volcanoes (red triangle) and some important localities explained in the main text. The Beppu–Shimabara and Kagoshima grabens are after Matsumoto (1979) and Hayasaka (1987). Yellow hatch indicates an area, where subsurface high conductivity blocks distribute in and beneath lower crust which were depicted from recent electrical resistivity survey (Hata et al. 2017). b Sampling locations shown on topographical map with locality name and major fault systems. c A–A’ cross section in panel a showing the locations of low frequency micro earthquakes during 2002–2016 (purple circles) and low seismic velocity zones (red color) in the mantle wedge and crust (Zhao et al. 2018). d Simplified geological map modified after Moreno et al. (2016). Geographical distribution of three types of fluids, High-Cl−, Low-concentration and High-HCO3− fluids, are also shown

Stable isotopes are useful for identifying the origin of waters and dissolved ions, and for interpreting the biogeochemical processes in aquifers (Kendall 1998; Cook and Herczeg 2000). Stable isotope ratios of oxygen and hydrogen in water molecule (δ2H and δ18O) have been used to examine the contribution of fluids that had experienced high temperature water–rock interaction resulting in δ18O isotopic shift towards enriched compositions leaving δ2H unchanged (Giggenbach 1992). Stable carbon isotope ratio of dissolved inorganic carbon (δ13C) is a fundamental tracer for identifying a contribution of magmatic or mantle derived carbon in fluids (e.g., Chiodini et al. 2000; Yamada et al. 2011; Rive et al. 2013). Sulfur isotope ratio of dissolved sulfate (δ34S) is also applicable for evaluating volcanic influence on spring waters and for interpreting redox processes in aquifer environments (e.g., Kusakabe et al. 2000; Hosono et al. 2014a, 2018). This study integrates the results from these isotopic analyses and characterized geochemical features of hot spring waters with comparison of those from other surface waters for trying to reveal the upwelling signatures of deeply originated components in crust and to assess their impact on surface hydrological systems.

Outline of the study area

The study area is located on the western coast of the Kumamoto Prefecture in Kyushu Islands and consists of four local domains, Uto, Amakusa, Yatsushiro and Ashikita districts (Fig. 1b). Uto and Amakusa are located on western peninsula and small islands, while Yatsushiro and Ashikita are on western coast of the main island Kyushu, respectively. Coastal flat plain in Yatsushiro is the reclaimed land on tidal flat. The Kyushu has humid subtropical climate and average annual temperature and precipitation of Amakusa and Yatsushiro are 16.4 and 16.8 °C and 2,076 and 1,979 mm during 1981–2010, respectively (http://www.jma.go.jp/jma/menu/report.html). Around 40% of total precipitation occurs during summer rainy season in June and July.

In the Kyushu Islands the active volcanoes are generally situated in two major graben structures, Beppu–Shimabara and Kagoshima grabens (Fig. 1a) (Nakada et al. 2016) that are facilitated by a unique open tectonic setting partly associated with wet upwelling flow in the mantle wedge (Zhao et al. 2021). The study area is located outside of these graben structures, lying between two major active fault systems, Hinagu and Futagawa strike-slip faults (Fig. 1b), for example, which activities generated the 2016 Mw 7.0 Kumamoto destructive earthquakes (e.g., Hosono et al. 2019, 2020). As stated earlier, neither active volcanic activities nor typical subduction related magmatisms are recognized within and beneath the study area. For instance, low frequency micro earthquakes generally occur beneath active volcanoes such as Aso, Kuju and Tsurumi somehow accompanied with low seismic velocity anomalies in the crust, while there is no significant evidence of major magmas or melts anticipated in the crust beneath the study area (Fig. 1c). In contrast, some recent tomography images (Zhao et al. 2018, 2021) depicted the presence of upwelling asthenospheric mantle in the corner of mantle wedge beneath crust under the study area.

Paleozoic to Mesozoic low-temperature and high-pressure metamorphic rocks and metasedimentary rocks of accretionary prism are widely distributed in the southern part of the Usuki–Yatsushiro Tectonic Line (Fig. 1d). In the northern part of this tectonic line Cretaceous Himenoura Group, alternative sedimentary layers of sandstone, mudstone, shale and conglomerate, comprises of the basement rocks of the study area. Cretaceous metamorphic rocks are also outcropped in Amakusa but their distributions are quite limited. These geological basements are overline by Paleogene sedimentary formations which are further intruded by Miocene dikes and overline by Miocene to Pleistocene volcanic rocks mainly of andesitic to dacitic compositions. Most of these young volcanic rocks are found at Uto peninsula (1.7 to 4.2 Ma; Yokose et al. 2009) and Ashikita district (Fig. 1d). Some limestone formations are interlayered but their distribution is quite limited that are not seen near majority hot spring sites.

Samples and methods

In total 28 groundwater samples were collected from borehole of the hot spring facilities except one sample that was collected from artesian spring during June 2019 and May 2020 (Fig. 1b, Additional file 1: Table S1). The depth of borehole is ranging from 70 to 1,230 m and are generally deeper than 800 m and temperature of the waters were ranging from 19.8 to 45.2 °C (Additional file 1: Table S1). Shallow boreholes (70–100 m) are of Yunoko and Shimoda hot springs, which hot springs with Oyano hot spring were first discovered as artesian flowing from the ground (Fig. 1a) and then later explored subsurface by drilling. Other facilities explored groundwater with elevated temperature though originally there were no hot spring discharge symptoms on a surface land. In addition, 19 stream waters, 10 mountain spring waters, and 4 shallow well (~ 45 m in depth) waters were collected within the same area during July 2019 and April 2020 for isotopic and geochemical comparisons (Fig. 1b), which samples are representative for waters in meteologically driven surface hydrological systems (we termed ‘surface water’ for all these waters). No precipitation occurred during 3 days before the sampling.

Groundwaters from hot spring sites were collected directly from the pipes of the pumping or artesian flows. Stream and artesian mountain spring waters were collected on sites, while shallow well waters were obtained using electric pumping or Baylor tube sampler. Field parameters such as temperature, electrical conductivity, pH, oxidation reduction potential, and dissolved oxygen were monitored using field meters (HORIBA D-54) and water samples were collected after these values became constant. The HCO3− concentration was determined by in situ alkalinity measurement following titrating with sulfuric acid. Water samples were stored in 100, 500 and 2,000 ml polyethylene bottles for major ion concentrations (Na+, K+, Ca2+, Mg2+, Cl−, SO42−, and NO3−), δ13C and δ34S analyses, while 20 ml glass vials were used for collecting water samples for δ2H and δ18O analyses, respectively. The samples were filtered through ADVANTEC 0.2 µm cellulose–acetate filters onsite before storing for 100 ml bottles and immediately after taking back to laboratory for 500 and 2,000 ml bottles, respectively. The filtered water samples in latter two bottles were adjusted to pH 12 to 12.5 with NaOH and to pH 2.5 to 3 with 6 M HCl, respectively, then dissolved inorganic carbon and SO42− were collected as BaCO3 and BaSO4 by adding 10% BaCl2 for δ34S and δ13C analyses.

Major ion concentrations were measured by ion chromatography (Compact IC 761, Metrohm, Switzerland). Both δ13C of BaCO3 and δ34S of BaSO4 were determined using a continuous-flow gas ratio mass spectrometer (Delta V Advantage, Thermo Fisher Scientific, USA) interfaced with an elemental analyzer (Thermo Fisher Flash 2000, USA) (Hosono et al. 2014a, 2015). International standards (NBS123 and OGS-1) and several carbonate and sulfate materials were used for quality check of the measurement data. The precision of both δ13C and δ34S analyses were generally better than ± 0.2‰. The δ2H and δ18O of water molecular were determined by a gas-ratio mass spectrometer (Delta V Advantage, Thermo Fisher Scientific, USA) coupled with an automatic water–gas equilibrization devise (Nakano Denshi Co., Ltd., Japan). Based on replicate measurements of standards and samples, the analytical precisions for δ2H and δ18O were better than ± 0.5‰ and ± 0.05‰, respectively.

Results and discussion

Hydrochemical classification

According to trilinear diagram hydrochemical features of groundwater samples from hot spring sites clearly differ from those for surface waters (Fig. 2a). The surface water samples are broadly plotted on compositional field between Ca–HCO3− type with pristine signature and modern sea water, suggesting presence of sea water mixing into fresh surface waters with lower dissolved solutes (electric conductivity of around 50 μS/cm increased near 1,000 μS/cm, Additional file 1: Table S1). On the other hand, the groundwater samples from hot spring sites are exclusively plotted along two different trends with increased concentrations (electric conductivity of ~ 42,500 μS/cm), high Cl− with variable Ca2+ ratio and high Na+ with variable HCO3− ratio (Fig. 2b). Thus, these groundwaters changed their hydrochemical features owing to several factors including mixing of waters with other origins and geochemical evolutions, such as cation exchange and redox processes.

Fig. 2
figure 2

Hydrochemical classification based on trilinear diagram. a Comparison between surface waters and groundwaters from hot spring sites. b Classification of three different types of groundwaters from hot spring sites. c Difference in Cl− concentrations among three types of fluids. d Diversity in HCO3− concentrations among three types of fluids

The groundwater samples categorized by the former trend are relatively more enriched in Cl− and Na+ and thus we classified this type of waters as High-Cl− fluid (Fig. 2b, c, Additional file 1: Table S1). The groundwater samples categorized by the latter trend are subdivided into two groups, one characterized with greater HCO3− concentrations (~ 3,888 mg/l) and the other with least dissolved concentrations but with relatively higher SO42− ratio (Fig. 2d). We classified these two types of groundwaters and named High-HCO3− fluid and Low-concentration fluid, respectively (Fig. 2b, d, Additional file 1: Table S1). Geographical distribution of three types of fluids, High-Cl−, Low-concentration and High-HCO3− fluids, are shown in Fig. 1d. This hydrochemical classification and terminology are used in the following discussion.

Origin of water

Figure 3 shows δ2H versus δ18O diagrams plotting all collected samples and data of some hypothesized source waters for comparison. This comparison allows us to identify the possible origins of water samples. All surface water samples were closely plotted on local meteoric water line except one stream water sample that is slightly deviated towards composition of modern sea water (Fig. 3a). This suggests that all these waters are mainly of meteoric origin except one sample that was clearly affected by a mixing of sea water. We admit that some of these waters are affected by sea water component in terms of major ion chemistry. However, this effect was small and not obviously reflected on δ2H and δ18O shifts. Thus their compositions were regard as representative feature of waters in surface hydrological systems over the study area.

Fig. 3
figure 3

δ2H versus δ18O diagrams. a–d Isotopic signatures for the surface waters, High-Cl−, Low-concentration, and High-HCO3− fluids, respectively. Isotopic compositional field for hypothesized magmatic water is after Giggenbach (1992). Local meteoric water line in Kumamoto (ca. 20 km northeast from Uto district) is after Okumura et al. (2018)

In Fig. 3b, samples of High-Cl− fluid are broadly plotted along an array between compositional field of surface waters and modern sea water. In combination with Cl− concentration (Additional file 2: Fig. S1), our simple mixing model suggests that this fluid was originated of the surface fresh water admixed with modern sea water with variable mixing proportions ranging from few to nearly 100% due to sea water intrusion and/or contribution of saline pore waters in sediments. This interpretation agrees well with the fact that all High-Cl− fluids are located just on the coasts or on the reclaimed land in Yatsushiro (Fig. 1d). In turn, for Low-concentration fluid all sample plots were fallen on the compositional field of the surface water (Fig. 3c), indicating their intrinsic meteoric origin analogous to surface circulated waters.

In contrast to High-Cl− and Low-concentration fluids, some samples of High-HCO3− fluid are plotted on right to the local meteoric water line towards hypothesized magmatic water proposed by Giggenbach (1992) (Fig. 3d), although this isotopic shift seems not to be much drastic compared to ones reported from some volcanic fields and other hot spring waters of subduction slab origin (e.g., Kita et al. 2009; Kusuda et al. 2014; Morikawa et al. 2016). Nevertheless, observed isotopic shift possibly indicates incorporation of deep fluids that have experienced high temperature water–rock interaction in deep crust as explained more later.

Geochemical evolution

Cation exchange

High-Cl− fluid showing variable Ca2+ ratio (Fig. 2b) possibly indicates the occurrence of other processes than source inputs, i.e., geochemical evolution. The chloro-alkaline indices (CAI) (Schoeller 1977) expressed by the following equations in milliequivalent concentrations can be used to test the occurrence of cation exchange reactions that control cation concentrations in aquifers.

$${\text{CAI}} - 1 = \frac{{{\text{Cl}}^{ - } - \left( {{\text{Na}}^{ + } + {\text{K}}^{ + } } \right)}}{{{\text{Cl}}^{ - } }}$$
(1)
$${\text{CAI}} - 2 = \frac{{{\text{Cl}}^{ - } - \left( {{\text{Na}}^{ + } + {\text{K}}^{ + } } \right)}}{{{\text{SO}}_{4}^{2 - } + {\text{HCO}}_{3}^{ - } + {\text{CO}}_{3}^{2 - } + {\text{NO}}_{3}^{ - } }}$$
(2)

Here, the cation exchange and reverse cation exchange are expected to occur when both CAI-1 and CAI-2 show negative and positive values, respectively. The parameter CO32− can be neglected when pH of solutions is lower than 10.3. The pH measured for our samples were all below 10.3 (Additional file 1: Table S1). Figure 4 displays both CAI-1 and CAI-2 yielded positive values and the values are tended to increase with increase in Ca2+ ratio for the High-Cl− fluid. We can thus conclude that hydrochemical features of High-Cl− fluid was formed by the sea component mixing and subsequent progressive reverse cation exchange reaction in aquifer. Exchange of calcium ion in exchangeable site in aquifer’s solid phase was released by Na+ in water, assuming to easily be enhanced by additives of saline water with high Na+ concentration. Moderately, but similar variation in Ca2+ ratio was observed for Low-concentration and High-HCO3− fluids (Fig. 2). However, both fluids had negative CAI-1 and CAI-2 values which tended to decrease with decrease in Ca2+ ratios (Fig. 4). Thus, these fluids had evolved along normal cation exchange reaction in contrast to High-Cl− fluid.

Fig. 4
figure 4

Biplot diagram between Ca2+ ratio and chloro-alkaline indices. a, b Plot diagrams compare for CAI-1 and CAI-2, respectively, three types of groundwaters from hot spring sites, High-Cl−, Low-concentration, and High-HCO3− fluids. Ca2+ ratio denotes proportion of calcium ion among four major cations (Na+, K+, Ca2+ and Mg2+) in milliequivalent concentrations

Redox process (sulfur cycle)

The redox process is another important geochemical evolutional process that must be considered. This is true, because some groundwaters from hot spring sites showed low dissolved oxygen (< 2 mg/l) and negative oxidation reduction potential (mV) although surface waters generally showed aerobic regime (Additional file 1: Table S1), implying the occurrence of some anaerobic microbial reductions facilitated in some aquifers. In general, Cl− concentrations are not affected by redox changes and NO3− was not highly concentrated (mostly < 2.0 mg/l except three samples with > 10 mg/l due to local contamination) in the water samples investigated (Additional file 1: Table S1). Thus Cl− and NO3− are not of major interests for the subject to discuss about the issue. Here, additional source characteristics and redox processes will be discussed for SO42− with conjunction of δ34S (Fig. 5).

Fig. 5
figure 5

SO42− concentration versus δ34S diagrams. a, b Sample plots for three types of groundwaters from hot spring sites, High-Cl−, Low-concentration, and High-HCO3− fluids, with comparison of those for surface waters. Compositions of modern sea water are after Rees et al. (1978)

The δ34S and SO42− variations for surface waters were considered as juvenile indices that are defined as − 7‰ < δ34S < 15‰ and 2 mg/l < SO42− < 30 mg/l, respectively (Fig. 5a), taking their overall aerobic (Additional file 1: Table S1) and the least contaminated hydrochemical features. In Fig. 5b, some High-Cl− fluids showed elevated SO42− (~ 1,700 mg/l) towards compositions of modern sea water (δ34S: ca. 20.3‰, SO42−: ca. 2649 mg/l; Rees et al. 1978), suggesting significant saline water contribution as explained previously. However, other samples showed lower concentrations with elevated δ34S up to 28‰, indicating an occurrence of microbial reduction of concentrated sulfate of marine origin (Strebel et al. 1990; Tuttle et al. 2009). Some samples were not measured for δ34S because of their low SO42− concentrations probably due to progressive sulfate reduction reaction. Such phenomenon is commonly seen for coastal aquifers either at near study fields in Kumamoto (Hosono et al. 2014a) or some other countries (Hosono et al. 2010, 2011, 2014b).

The Low-concentration fluid displayed compositions almost identical to those for surface water except two samples (HS19007 and HS20004) with moderately elevated SO42− (~ 280 mg/l) and δ34S (~ 17.4‰) and one sample (HS19002) with slightly elevated SO42− (57 mg/l) and quite depleted δ34S (− 13.7‰) (Fig. 5b). The observed signature for former two samples can be explained by mixing of sulfate of marine origin similar to High-Cl− fluid but that intensity was too small to clearly be detected on stable isotope ratios of water molecular (Fig. 3c). The signature of later sample can be explained by the oxidation of sulfide with lower δ34S (Taylor et al. 1984; van Everdingen and Krouse 1985). In summary, overall hydrochemical formation process of Low-concentration fluid resembles to that for surface waters except cation exchange reaction.

The High-HCO3− fluid has the least SO42− concentrations, which are lower than 16 mg/l, and for majority samples, ranging between detection limit and 5 mg/l (Additional file 1: Table S1). Two samples had similar δ34S compositions to those for surface waters but other four samples showed most enriched δ34S signature (~ 45‰) among all samples with depleted SO42− concentrations (Fig. 5b). No striking signatures of contribution from marine sulfate are observed, corresponding to results of δ2H and δ18O analyses. Overall, hydrochemistry of the High-HCO3− fluid was changed their features with similar manner as found for Low-concentration fluid, except least contribution from marine sulfate and tremendous HCO3− enrichment (Fig. 2d).

Upwelling signatures of deeply derived carbon and fluids

Significant contribution of sulfur from some heat sources was not detected from any water samples collected from our study area (Fig. 5), which δ34S values resemble a feature of arc volcanic magmas typically in a range between 0 and 10‰ (Ueda and Sakai 1984). On the other hand, elevated HCO3− in High-HCO3− fluid may be originated of CO2 gas released from some heat sources that was dissolved in waters and preferentially transported towards surface leaving SO2 and other gasses (e.g., Ohsawa et al. 2002; Gultekin et al. 2011). The use of δ13C tracer in dissolved inorganic carbon can provide clue to test this hypothesis (Chiodini et al. 2000; Yamada et al. 2011; Rive et al. 2013).

On the δ13C versus HCO3− concentration diagram (Fig. 6a), all surface water samples fall on δ13C compositions between − 31.3 and − 11.8‰ with the lowest HCO3− concentrations generally less than ca. 100 mg/l (Additional file 1: Table S1), except one sample (GW20002) with relatively elevated δ13C (− 8.9‰) and HCO3− concentration (295 mg/l). These geochemical features are of the typical ecosystem-mediated carbon globally (Clark and Fritz 1997) and aquifer systems near the study areas (Ohsawa et al. 2002; Hosono et al. 2014a) and thus are reasonably assigned as a representative composition of surface water systems before alteration by additional source inputs. On the same diagram, almost all groundwater samples from hot spring sites are plotted along the exponential curved trends based on two end-component mixing models between hypothesized soil CO2 and deep originated CO2, suggesting progressive contribution of CO2 from deep heat source into shallower hydrological systems.

Fig. 6
figure 6

Identification for deeply originated components. a HCO3− concentration versus δ13C diagram for all collected water samples. Simple two components mixing curved lines with relative contribution rates from hypothesized mantle sourced carbon in 10% interval are shown in the graph. Mixing of two sets of end-components, soil carbon (HCO3− and δ13C: 20 mg/l and − 30‰, 160 mg/l and − 30‰, respectively) and mantle sourced carbon (HCO3− and δ13C: 4,000 mg/l and − 4.5‰, 4,000 mg/l and − 1.5‰, respectively) were hypothesized taking compositional variations observed from this study, with a same method applied elsewhere (e.g., Chiodini et al. 2000; Yamada et al. 2011; Rive et al. 2013). b Geographical distribution of relative contribution from mantle sourced carbon in groundwater samples deduced from mixing model shown in panel a. c δ2H versus δ18O plot diagram for High-HCO3− fluid shown in different colors with variable δ18O. d Geographical distribution of δ18O values. The symbol color is same as in panel c

A contribution of deep originated CO2 dominates in High-HCO3− fluid compared to other fluids (High-Cl− and Low-concentration fluids) which contain only ~ 5% of carbon from that source (Fig. 6a). Therefore, significant contribution of CO2 from some heat source is responsible for the formation of major hydrochemical features of High-HCO3− fluid. Figure 6b displays geographical distribution of relative contribution of carbon from the heat source. It is notable from this map that groundwaters with the greater contributions are sporadically distributed along the fault systems, and interestingly, these anomalies are characteristically located near the historical hot spring sites, i.e., Yunoko, Shimoda and Oyano, regardless of their host geological formations (Fig. 1). It seems plausible, therefore, that the surface geology is not necessarily associated with the occasion of observed deeply sourced carbon, instead, the presence of deeper heat source must be more critically important. The groundwaters with greater δ2H and δ18O compositional shifts tend to be seen on locations, where deeply originated CO2 occurs (Fig. 6c, d), supporting above scenario.

If there are still lived magmas or melts beneath surface, the observed isotopic features of High-HCO3− fluid could be explained by involvement of these magmatic gases and fluids; however, the presence of these heat sources has not been reported for particular fields based on scientific evidences. Alternatively, based on well-known geotectonic setting for the study area, upwelling asthenospheric mantle (Zhao et al. 2018, 2021) is the most plausible origin of deeply sourced CO2 that was released into metasomatized lower crust triggered by crustal assimilation with this huge heat source. Similar phenomenon was reported from the Hitoyoshi Basin in non-volcanic field at western Kumamoto (Fig. 1a) based on the same isotopic systematics (Sakai et al. 2011), with some additional noble gas isotopes, e.g., 3He/4He (Ohsawa et al. 2011).

Hata et al. (2017, 2020) provided key information for putting more constraint on the realization of the deep crustal fluids. Their recent electrical resistivity image finely identified significant high conductivity blocks exist in western Kyushu under shallow depth in the upper mantle and lower crust (~ 50 km), and argued the presence of deep hydrothermal fluids particularly under Hishikari (hydrothermal gold–silver ore deposits) (Fig. 1a), Hitoyoshi, Ashikita and Amakusa, where deep crustal fluid upwelling was anticipated. Thus, we propose fluids generated in a lower crust possibly contributes for the formation of our hypothesized deep fluids, although a possibility of partial involvement of recycled subduction slab fluids can’t be neglected as their ultimate origin (Nakamura et al. 2019; Sano and Marty 1994). However, observed degree of isotopic shift on δ2H and δ18O compositions (Fig. 6c) was not much dramatic as seen for deeply originated CO2 (Fig. 6a) probably due to both dispersion of CO2 gas and mitigation by the mixing with water of meteoric origin, while the deep crustal fluids ascent through long distance in the crust.

Discharging slab originated fluids ubiquitous in southwest Japan is thought to be transported through major tectonic structures, such as Median Tectonic Line, forearc thrust faults in the accretionary prisms and branching faults, driven by geo- and heat pressures under compressional tectonic setting (Kusuda et al. 2014; Togo et al. 2014; Yoshida et al. 2015; Morikawa et al. 2016). Deeply originated gases and fluids proposed in this study could also ascend through structural breaks in the Hinagu and Futagawa strike-slip faulting systems (Fig. 1) and possibly in relation to pressure release under open tectonic setting, with a similar law demonstrated from previous works in Kyushu (Hosono and Nakano, 2004; Hosono et al. 2018) and from other open tectonic regions (e.g., Cerón et al. 1998). It’s interesting to note that stronger signature of upwelling of the deeply sourced components (Fig. 6b, d) characteristically occurs near the locations, where more than two different faults are crossing (Fig. 1b). This may be due to generation of lower pressure around cross point of the different stresses in the crust, where pressurized deep fluids can more easily release towards air pressure near surface (Cappa et al. 2009; Takashima and Kano 2005).

The results of this study showed δ13C was the most sensible tracers among all measured stable isotope ratios for fingerprinting deeply originated components in groundwaters. It is also useful for identifying the locations of major structural pathways and could be applicable for further investigating deeply connecting fault systems and seismic responses (Barbieri et al. 2020; Barberio et al. 2021; Luccio et al. 2018). Giving that the mantle sourced carbon mediated groundwaters (High-HCO3− fluid) shared nearly half of total analyzed groundwaters (46%), this type of waters must to be extended wide in space bellow few hundred meters deeper in crust. Despite their prosperity in deep aquifers their discharges to near surface systems look somehow restrictive at, only where major fluid pathways exist, and thus, their overall impacts on surface hydrological systems seem to be small on this particular study setting. In our study area, only minor contamination with mantle originated carbon, i.e., ca. 5%, was observed locally in shallow well water (GW20002, 20 m in depth) near Yunoko hot spring (Fig. 6a). These aspects can be taken when we consider carbon cycle on earth surface systems in other fields.

Conclusions

The hydrochemical features of surface and subsurface waters were characterized to understand the origin, distribution and pathway of deeply originated components in hot springs in western Kumamoto using multiple stable isotope ratios (δ2H, δ18O, δ13C and δ34S) for an area, where subduction-related active volcanic and magmatic activities are not anticipated. Our results implied that dissolved carbon and waters assumed to be originated from upper mantle and lower crust, respectively, could be transported to shallower aquifers along the major pathways. This study illustrates a possible genesis of some carbon rich hot springs appearing on ‘back-arc-side’ non-volcanic fields. Additional analysis on noble gas isotopes may provide firmer evidence. The aspects from our results are useful for interpreting data from other fields and for more global generalization, and beneficial to users for hot spring resources.

Availability of data and materials

All data used in this study are listed in figures and Supplementary Information.

Abbreviations

MTL:

Median Tectonic Line

TL:

Tectonic line

DIC:

Dissolved inorganic carbon

References

  • Ague JJ (2014) Fluid Flow in the Deep Crust. In: Rudnick RL (ed) The Crust. In: Holland HD, Turekian KK (eds) Treatise on Geochemistry, vol 4. Elsevier Science, Second Edition, pp 203–247

    Chapter  Google Scholar 

  • Aiuppa A, Baker D, Webster J (2009) Halogens in volcanic systems. Chem Geol 263:1–18

    Article  Google Scholar 

  • Aiuppa A, Fischer TP, Plank T, Robidoux P, Di Napoli R (2017) Along-arc, inter-arc and arc-to-arc variations in volcanic gas CO2/ST ratios reveal dual source of carbon in arc volcanism. Earth Sci Rev 168:24–47

    Article  Google Scholar 

  • Barberio MD, Gori F, Barbieri M, Boschetti T, Caracausi A, Cardello GL, Petitta M (2021) Understanding the origin and mixing of deep fluids in shallow aquifers and possible implications for crustal deformation studies: San Vittorino Plain. Central Apennines Appl Sci 11:1353. https://doi.org/10.3390/app11041353

    Article  Google Scholar 

  • Barbieri M, Boschetti T, Barberio MD, Billi A, Franchini S, Iacumin P, Selmo E, Petitta M (2020) Tracing deep fluid source contribution to groundwater in an active seismic area (central Italy): A combined geothermometric and isotopic (δ13C) perspective. J Hydrol 582:124495. https://doi.org/10.1016/j.jhydrol.2019.124495

    Article  Google Scholar 

  • Cappa F, Rutqvist J, Yamamoto K (2009) Modeling crustal deformation and rupture processes related to upwelling of deep CO2-rich fluids during the 1965–1967 Matsushiro earthquake swarm in Japan. J Geophys Res 114:B10304. https://doi.org/10.1029/2009JB006398

    Article  Google Scholar 

  • Cerón JC, Pulido-Bosch A, Sanz de Galdeano C (1998) Isotopic identification of CO2 from a deep origin in thermomineral waters of southeastern Spain. Chem Geol 149:251–258

    Article  Google Scholar 

  • Chiodini G, Frondini F, Cardellini C, Parello F, Peruzzi L (2000) Rate of diffuse carbon dioxide Earth degassing estimated from carbon balance of regional aquifers: the case of central Apennine. Italy J Geophys Res Solid Earth 105(B4):8423–8434

    Article  Google Scholar 

  • Clark ID, Fritz P (1997) Environmental isotopes in hydrogeology. Lewis Publishers, New York

    Google Scholar 

  • Cook P, Herczeg AL (2000) Environmental Tracers in Subsurface Hydrology. Kluwer Academic Publishers, USA

    Book  Google Scholar 

  • Dessert C, Gaillardet J, Dupre B, Schott J, Pokrovsky OS (2009) Fluxes of high-versus low-temperature water–rock interactions in aerial volcanic areas: example from the Kamchatka Peninsula, Russia. Geochim Cosmochim Acta 73:148–169

    Article  Google Scholar 

  • Gaillardet J, Rad S, Rive K, Louvat P, Gorge C, Allègre CJ, Lajeunesse E (2011) Orography-driven chemical denudation in the Lesser Antilles: evidence for a new feed-back mechanism stabilizing atmospheric CO2. Am J Sci 311:851–894

    Article  Google Scholar 

  • Giggenbach WF (1992) Isotopic shifts in waters from geothermal and volcanic systems along convergent plate boundaries and their origin. Earth Planet Sci Lett 113:495–510

    Article  Google Scholar 

  • Gultekin F, Hatipoglu E, Ersoy AF (2011) Hydrogeochemistry, environmental isotopes and the origin of the Hamamayagi-Ladik thermal spring (Samsun, Turkey). Env Earth Sci 62:1351–1360

    Article  Google Scholar 

  • Hata M, Uyeshima M, Handa S, Shimoizumi M, Tanaka Y, Hashimoto T, Kagiyama T, Utada H, Munekane H, Ichiki M, Fuji-ta K (2017) 3-D electrical resistivity structure based on geomagnetic transfer functions exploring the features of arc magmatism beneath Kyushu, Southwest Japan Arc. J Geophys Res Solid Earth 122:172–190. https://doi.org/10.1002/2016JB013179

    Article  Google Scholar 

  • Hata M, Munekane H, Utada H, Kagiyama T (2020) Three-dimensional electrical resistivity structure beneath a volcanically and seismically active island, Kyushu, Southwest Japan Arc. J Geophys Res Solid Earth. https://doi.org/10.1029/2019JB017485

    Article  Google Scholar 

  • Hayasaka S (1987) Geologic structure of Kagoshima Bay, South Kyushu, Japan. Assoc Geol Collab Japan 33:225–233 (in Japanese with English abstract)

    Google Scholar 

  • Hedenquist J, Lowenstern J (1994) The role of magmas in the formation of hydrothermal ore deposits. Nature 370:519–527. https://doi.org/10.1038/370519a0

    Article  Google Scholar 

  • Hirose F, Nakajima J, Hasegawa A (2008) Three-dimensional seismic velocity structure and configuration of the Philippine Sea slab in southwestern Japan estimated by double-difference tomography. J Geophys Res Solid Earth. https://doi.org/10.1029/2007JB005274

    Article  Google Scholar 

  • Hosono T, Nakano T (2004) Pb-Sr isotopic evidence for contribution of deep crustal fluid of the Hishikari epithermal gold deposit, southwestern Japan. Earth Planet Sci Lett 222:61–69

    Article  Google Scholar 

  • Hosono T, Siringan F, Yamanaka T, Umezawa Y, Onodera S, Nakano T, Taniguchi M (2010) Application of multi-isotope ratios to study the source and quality of urban groundwater in Metro Manila, Philippines. Appl Geochem 25:900–909

    Article  Google Scholar 

  • Hosono T, Delinom R, Nakano T, Kagabu M, Shimada J (2011) Evolution model of δ34S and δ18O in dissolved sulfate in volcanic fan aquifers from recharge to coastal zone and through the Jakarta urban area, Indonesia. Sci Total Env 409:2541–2554

    Article  Google Scholar 

  • Hosono T, Lorphensriand O, Onodera S, Okawa H, Nakano T, Yamanaka T, Tsujimura M, Taniguchi M (2014a) Different isotopic evolutionary trends of δ34S and δ18O compositions of dissolved sulfate in an anaerobic deltaic aquifer system. Appl Geochem 46:30–42

    Article  Google Scholar 

  • Hosono T, Tokunaga T, Tsushima A, Shimada J (2014b) Combined use of δ13C, δ15N, and δ34S tracers to study anaerobic bacterial processes in groundwater flow systems. Water Res 54:284–296

    Article  Google Scholar 

  • Hosono T, Alvarez K, Lin I-T, Shimada J (2015) Nitrogen, carbon, and sulfur isotopic change during heterotrophic (Pseudomonas aerofaciens) and autotrophic (Thiobacillus denitrificans) denitrification reactions. J Contam Hydrol 183:72–81

    Article  Google Scholar 

  • Hosono T, Hartmann J, Louvat P, Amann T, Washington KE, West AJ, Okamura K, Böttcher ME, Gaillardet J (2018) Earthquake-induced structural deformations enhance long-term solute fluxes from active volcanic systems. Sci Rep 8:14809. https://doi.org/10.1038/s41598-018-32735-1

    Article  Google Scholar 

  • Hosono T, Yamada C, Shibata T, Tawara Y, Wang CY, Manga M, Rahman ATMS, Shimada J (2019) Coseismic groundwater drawdown along crustal ruptures during the 2016 Mw 7.0 Kumamoto earthquake. Water Resour Res 55:5891–5903

    Article  Google Scholar 

  • Hosono T, Yamada C, Manga M, Wang CY, Tanimizu M (2020) Stable isotopes show that earthquakes enhance permeability and release water from mountains. Nat Commun 11:2776. https://doi.org/10.1038/s41467-020-16604-y

    Article  Google Scholar 

  • Kendall C (1998) Tracing nitrogen source and cycling in catchments. In: Kendall C, McDonnell JJ (eds) Isotope tracers in catchment hydrology. Elsevier Science B.V, The Netherland, pp 519–576

    Chapter  Google Scholar 

  • Kita I, Kai T, Ito R, Ishida M, Ueda A (2009) Magmatic fluid input to the Kuju-Iwoyama hydrothermal system prior to the 1995 eruption of the Kuju volcano (Kyushu, Japan). Geothermics 38:294–302

    Article  Google Scholar 

  • Kusakabe M, Komoda Y, Takano B, Abiko T (2000) Sulfur isotopic effects in the disproportionation reaction of sulfur dioxide in hydrothermal fluids: implications for the δ34S variations of dissolved bisulfate and elemental sulfur from active crater lakes. J Volcanol Geotherm Res 97:287–307

    Article  Google Scholar 

  • Kusuda C, Iwamori H, Nakamura H, Kazahaya K, Morikawa N (2014) Arima hot spring waters as a deep-seated brine from subducting slab. Earth Planets Space 66:119. http://www.earth-planets-space.com/content/66/1/119

  • Kusuhara F, Kazahaya K, Morikawa N, Yasuhara M, Tanaka H, Takahashi M, Tosaki Y (2020) Original composition and formation process of slab-derived deep brine from Kashio mineral spring in central Japan. Earth Planets Space 72:107. https://doi.org/10.1186/s40623-020-01225-y

    Article  Google Scholar 

  • Luccio FD, Chiodini G, Caliro S, Cardellini C, Convertito V, Pino NA, Tolomei C, Ventura G (2018) Seismic signature of active intrusions in mountain chains. Sci Adv 4:e1701825. https://doi.org/10.1126/sciadv.1701825

    Article  Google Scholar 

  • Matsumoto Y (1979) Some problems on volcanic activities and depression structures in Kyushu, Japan. Memoir Geol Soc Japan 16:127–139 (in Japanese with English abstract)

    Google Scholar 

  • Moreno T, Wallis S, Kojima T, Gibbons W (2016) The geology of Japan. Geological Society, London

    Book  Google Scholar 

  • Morikawa N, Kazahaya K, Takahashi M, Inamura A, Takahashi HA, Yasuhara M, Ohwada M, Sato T, Nakama A, Handa H, Sumino H, Nagao K (2016) Widespread distribution of ascending fluids transporting mantle helium in the fore-arc region and their upwelling processes: Noble gas and major element composition of deep groundwater in the Kii Peninsula, southwest Japan. Geochim Cosmochim Acta 182:173–196

    Article  Google Scholar 

  • Nakada S, Yakamoto T, Maeno F (2016) Miocene-Holocene volcanism. In: Moreno T, Wallis S, Kojima T, Gibbons W (eds) The geology of Japan. Geological Society, London, pp 237–308

    Google Scholar 

  • Nakajima J, Hasegawa A (2007) Subduction of the Philippine Sea plate beneath southwestern Japan: slab geometry and its relationship to arc magmatism. J Geophys Res Solid Earth. https://doi.org/10.1029/2006JB004770

    Article  Google Scholar 

  • Nakamura H, Iwamori H, Nakagawa M, Shibata T, Kimura JI, Miyazaki T, Chang Q, Vaglarov BS, Takahashi T, Hirahara Y (2019) Geochemical mapping of slab-derived fluid and source mantle along Japan arcs. Gondwana Res 70:36–49

    Article  Google Scholar 

  • Ohsawa S, Kazahaya K, Yasuhara M, Kono T, Kitaoka K, Yusa Y, Yamaguchi K (2002) Escape of volcanic gas into shallow groundwater systems at Unzen Volcano (Japan): evidence from chemical and stable carbon isotope compositions of dissolved inorganic carbon. Limnology 3:169–173

    Article  Google Scholar 

  • Ohsawa S, Sakai T, Yamada M, Mishima T, Yoshikawa S, Kagiyama T (2011) Dissolved inorganic carbon extremely rich in mantle component of hot spring waters from the Hitoyoshi Basin located in a non-volcanic region of central Kyushu, Japan. J Hot Spring Sci 60:410–417

    Google Scholar 

  • Okumura A, Hosono T, Boateng D, Shimada J (2018) Evaluations of the downward velocity of soil water movement in the unsaturated zone in a groundwater recharge area using δ18O tracer: the Kumamoto region, southern Japan. Geol Croat 71:2. https://doi.org/10.4154/gc.2018.09

    Article  Google Scholar 

  • Perez-Fodich A, Derry LA (2019) Organic acids and high soil CO2 drive intense chemical weathering of Hawaiian basalts: Insights from reactive transport models. Geochim Cosmochim Acta 249:173–198

    Article  Google Scholar 

  • Rees CE, Jenkins WJ, Monster J (1978) The sulfur isotopic composition of ocean water sulphate. Geochim Cosmochim Acta 42:377–381

    Article  Google Scholar 

  • Rive K, Gaillardet J, Agrinier P, Rad S (2013) Carbon isotopes in the rivers from the Lesser Antilles: origin of the carbonic acid consumed by weathering reactions in the Lesser Antilles. Earth Surf Proc Land 38:1020–1035

    Article  Google Scholar 

  • Sakai T, Ohsawa S, Yamada M, Mishima T, Yoshikawa S, Kagiyama T, Oue K (2011) Origin of dissolved inorganic carbon of hot spring waters discharged from the non-volcanic region of central Kyusyu, Japan. J Hot Spring Sci 60:418–443

    Google Scholar 

  • Sano Y, Marty B (1994) Origin of carbon in fumarolic gas from island arcs. Chem Geol 119:265–274

    Article  Google Scholar 

  • Schoeller H (1977) Geochemistry of groundwater. An international guide for research and practice. UNESCO 15:1–18

    Google Scholar 

  • Strebel O, Böttcher J, Fritz P (1990) Use of isotope fractionation of sulfate-sulfur and sulfate-oxygen to assess bacterial desulfurication in a sandy aquifer. J Hydrol 121:155–172

    Article  Google Scholar 

  • Takashima C, Kano A (2005) Depositional processes of travertine developed at Shionoha hot spring, Nara Prefecture, Japan. J Geol Soc Japan 111:751–764

    Article  Google Scholar 

  • Taylor BE, Wheeler MC, Nordstrom DK (1984) Stable isotope geochemistry of acid mine drainage: experimental oxidation of pyrite. Geochim Cosmochim Acta 48:2669–2678

    Article  Google Scholar 

  • Togo Y, Kazahaya K, Tosaki Y, Morikawa N, Matsuzaki H, Takahashi M, Sato T (2014) Groundwater, possibly originated from subducted sediments, in Joban and Hamadori areas, southern Tohoku, Japan. Earth Planets Space 66:131. http://www.earth-planets-space.com/content/66/1/131

  • Tomaru H, Ohsawa S, Amita K, Lu Z, Fehn U (2007) Influence of subduction zone settings on the origin of forearc fluids: Halogen concentrations and 129I/I ratios in waters from Kyushu, Japan. Appl Geochem 22:676–691

    Article  Google Scholar 

  • Tuttle MLW, Breit GN, Cozzarelli IM (2009) Processes affecting δ34S and δ18O values of dissolved sulfate in alluvium along the Canadian River, central Oklahoma, USA. Chem Geol 265:455–467

    Article  Google Scholar 

  • Ueda A, Sakai H (1984) Sulfur isotope study of Quaternary volcanic-rocks from the Japanese Islands Arc. Geochim Cosmochim Acta 48:1837–1848

    Article  Google Scholar 

  • van Everdingen RO, Krouse HR (1985) Isotope composition of sulphates generated by bacterial and abiological oxidation. Nature 315:395–396

    Article  Google Scholar 

  • Yamada M, Ohsawa S, Kazahaya K, Yasuhara M, Takahashi H, Amita K, Mawatari H, Yoshikawa S (2011) Mixing of magmatic CO2 into volcano groundwater flow at Aso volcano assessed combining carbon and water stable isotopes. J Geochem Explor 108:81–87

    Article  Google Scholar 

  • Yokose H, Yanashima T, Kikuchi W, Sugiyama N, Shinohara A, Takeuchi T, Nagao K, Kodama K (2009) Episodic magmatism since 5Ma in the western part of Beppu-Shimabara graben, Kyusyu, Japan. J Mineral Petrol Econ Geol 94:338–348 (in Japanese with English abstract)

    Article  Google Scholar 

  • Yoshida K, Hirajima T, Ohsawa S, Kobayashi T, Mishima T, Sengen Y (2015) Geochemical features and relative B-Li–Cl compositions of deep-origin fluids trapped in high-pressure metamorphic rocks. Lithos 226:50–64

    Article  Google Scholar 

  • Zhao D, Yamashita K, Toyokuni G (2018) Tomography of the 2016 Kumamoto earthquake area and the Beppu-Shimabara graben. Sci Rep 8:15488. https://doi.org/10.1038/s41598-018-33805-0

    Article  Google Scholar 

  • Zhao D, Wang J, Huang Z, Liu X (2021) Seismic structure and subduction dynamics of the western Japan arc. Tectonophysics 802:228743. https://doi.org/10.1016/j.tecto.2021.228743

    Article  Google Scholar 

Download references

Acknowledgements

We thank the hot spring facilities to kindly let us collect water samples for chemical and isotopic analyses. We are grateful for team members of the Kumamoto University for their cordial help during field survey and laboratory analysis.

Funding

This study was funded by the JSPS Grant-in-Aid for Scientific Research B (17H01861) and Fostering Joint International Research A (19KK0291).

Author information

Authors and Affiliations

Authors

Contributions

TH conceived the idea and wrote the manuscript. CY conducted sampling survey and analyzed all data sets. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Takahiro Hosono.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Table S1.

Location, property, and measurement and analytical results for all water samples collected in this study from western coast of Kumamoto in western Kyushu, Japan.

Additional file 2: Fig. S1.

Cl− concentration versus δ2H diagram for High-Cl− fluid. Simple two components (surface water and modern sea water) mixing lines with relative contribution rates from modern sea water in 10% interval are shown. End-components for surface water with variable compositions (Cl− and δ2H: 13.8 mg/l and − 2.6‰ ~ 17.2 mg/l and − 37.3‰) were considered.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hosono, T., Yamanaka, C. Origins and pathways of deeply derived carbon and fluids observed in hot spring waters from non-active volcanic fields, western Kumamoto, Japan. Earth Planets Space 73, 155 (2021). https://doi.org/10.1186/s40623-021-01478-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40623-021-01478-1

Keywords