Next Article in Journal
Multi-Biomarker Responses of Asian Clam Corbicula fluminea (Bivalvia, Corbiculidea) to Cadmium and Microplastics Pollutants
Previous Article in Journal
Urbanization under a Changing Climate–Impacts on Hydrology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Using Isotopic and Hydrochemical Indicators to Identify Sources of Sulfate in Karst Groundwater of the Niangziguan Spring Field, China

1
College of Water Resources Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2
Key Laboratory of Karst Dynamics, Ministry of Land and Resources, Institute of Karst Geology, Chinese Academy of Geological Sciences, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
Water 2021, 13(3), 390; https://doi.org/10.3390/w13030390
Submission received: 16 December 2020 / Revised: 27 January 2021 / Accepted: 29 January 2021 / Published: 3 February 2021
(This article belongs to the Section Hydrology)

Abstract

:
Karst groundwater in the Niangziguan spring fields is the main source to supply domestic and industrial water demands in Yangquan City, China. However, the safety of water supply in this region has recently suffered from deteriorating quality levels. Therefore, identifying pollution sources and causes is crucial for maintaining a reliable water supply. In this study, a systematic sample collection for the karst groundwater in the Niangziguan spring fields was implemented to identify hydrochemical characteristics of the karst groundwater through comprehensive analyses of hydrochemistry (piper diagram, and ion ratios,) and stable isotopes (S and H-O). The results show that the karst groundwater in the Niangziguan spring fields was categorized as SO4·HCO3-Ca·Mg, HCO3·SO4-Ca·Mg, and SO4-Ca types. K+, Cl-, and Na+ are mainly sourced from urban sewage and coal mine drainage. In addition, SO42− was mainly supplied by the dissolution of gypsum and the oxidation of FeS2 in coal-bearing strata. It is noteworthy that, based on H-O and S isotopes, 75% of the karst groundwater was contaminated by acidic water in coal mines at different degrees. In the groundwater of the Niangziguan spring field, the proportions of SO42− derived from FeS2 oxidation were 60.6% (N50, Chengxi spring), 30.3% (N51, Wulong spring), and 26.0% (N52, Four springs mixed with water). Acid mine drainage directly recharges and pollutes karst groundwater through faults or abandoned boreholes, or discharges to rivers, and indirectly pollutes karst groundwater through river infiltration in carbonate exposed areas. The main source of rapid increase of sulfate in karst groundwater is acid water from abandoned coal mines.

1. Introduction

Karst regions occupy approximately 25 percent of the land surface of the Earth [1,2,3]. Karst groundwater is an important water resource that is used as a drinking and industrial water source by a significant proportion of the world’s population [4,5]. However, karst aquifers are much more vulnerable to pollution than other aquifers, and the restoration of contaminated groundwater is difficult [6,7,8]. Karst groundwater is a “hidden resource” in northern China and the prevention and monitoring of karst groundwater pollution are more difficult than that of surface waters due to its inaccessibility. Among numerous contaminants that contribute to groundwater pollution, Fe [9,10,11,12], nitrate [13,14,15,16,17], and sulfate [18,19,20] are cited most frequently. Fe in karst groundwater is mainly affected by Fe-rich acid mine drainage (AMD). Nitrate in karst groundwater includes synthetic and organic fertilizers, as well as municipal and industrial sewage. Sulfate minerals in karst groundwater originate from natural sources like the dissolution of sulfate, atmospheric deposition, and AMD caused by human activity.
Niangziguan spring is located in Yangquan City, Shanxi Province. The average annual flow of spring water is 9.8 m3·s−1 (1956–2019). Spring water in the Niangziguan spring fields is the main source to supply domestic and industrial water demands in Yangquan City, China. Since 2010, the sulfate radical of karst spring water has been more than 250 mg·L−1. Karst well water and springs are the primary method used for the discharge of groundwater from karst aquifers [21,22]. Therefore, the monitoring of karst well water and springs can provide more insights into the hydrogeological and hydrogeochemical processes that occur in underground environments.
The hydrogeochemical characteristics are crucial to reflect the circulation and hydrodynamic conditions of regional groundwater, and to understand the water quality and distribution characteristics. To date, methods commonly used in the karst water hydrogeochemistry are the hydrochemical type, saturation index, and ion ratio methods [23,24,25]. Advanced techniques have been also used such as isotopic analysis, statistical methods, and hydrogeochemical simulations. The interactions between precipitation and groundwater can be investigated using stable isotopes of hydrogen and oxygen [26,27,28,29]. The application of isotope tracers (δ34S-SO42, δ18O-SO42) in identifying sources and fate of SO4 in groundwater has been successfully achieved for several decades [30,31,32,33,34,35]. The geochemical evolution of karst water and its associations with urbanization impacts have not been fully addressed in the karst regions throughout the world [36,37,38,39,40,41,42]. This study characterizes spatial-temporal variations in hydrochemistry in the Niangziguan spring region, a typical karst area of fast urbanization in Shanxi, using hydrochemical reconnaissance and monitoring of S, D, and O isotopes [43,44,45,46,47,48,49]. The aim of our study is to evaluate the impacts of rapid urbanization on karst water quality. Identifying pollution sources and causes is crucial for reliable water supply.

2. Materials and Methods

2.1. Study Area

The Niangziguan spring field is located in the eastern part of Shanxi Province adjacent to Hebei Province in China, which includes: (1) the urban areas, suburbs, Pingding County, and Yu County of Yangquan City; (2) Xiyang, Heshun, Zuoquan, Shouyang, and Yuci of Jinzhong City; (3) Taiyuan City and Yangqu County. These areas are shown in Figure 1. The distribution area of coal measures strata in the spring area is 4728 km2, accounting for 63.5% of the whole spring area. Therefore, the industrial structure dominated by coal has become a feature of this area. Taking a comprehensive view of the main districts and counties in the spring area (the districts and counties of Yangquan City and Shouyang, Xiyang, Heshun, and Zuoquan of Jinzhong City), the output value of coal mining industry generally accounts for more than 50% of the total output value. The total area of the district is 7436 km2 ranging from 36°55′ N to 38°15′ N and 112°20′ E to 113°55′ E. The average annual water evaporation in the study area was 1202 mm while the average annual precipitation of the whole region was 542.4 mm during the period from 1955 to 2019 at Yangquan Station within the China Meteorological Data Network. While the average annual temperature is 10.9 °C, the average temperature in January is −4.6 °C and the extreme minimum temperature is −28 °C.

2.2. Materials and Methods

Fifty-two karst well water or springs in the Niangziguan spring field were sampled during 23–24 June 2014 and analyzed for constant indicators (water temperature, pH, conductivity, Ca2+, Mg2+, K+, Na+, Cl, SO42−, HCO3, NO3, CO32−, etc.) and trace indicators (Sr2+, F). With 18 groups of acidic drainage and surface water samples collected from coal mines in the spring field in April 2013, June 2014, October 2015, and April 2019, the constant indicators (water temperature, pH, conductivity, Ca2+, Mg2+, K+, Na+, Cl, SO42−, HCO3, CO32−, etc.) and isotope (δ34S) were measured. The temperature, pH, conductivity (EC), and ORP were directly measured by a WTW 3440 (Xylem Analytics, Munich, Germany) water quality parameter tester in the field. Ca2+ and HCO3 were titrated in the field by the Merck test box (MColortest™ 1.11110.0001, Germany), with a measurement accuracy of 1 mg·L−1 and 0.1 mmol·L−1, respectively. The concentrations of Mg2+, K+, Na+, and Sr2+ were measured with ICP-AES (ICAP 7600, Thermo Fisher, Waltham, MA USA). SO42−, HCO3, NO3, CO32−, and F- were analyzed with ion chromatography (Dionex ICS-1100, Thermo Fisher, USA). The detection limits were 0.02 mg·L−1, 0.06 mg·L−1, 0.008 mg·L−1, 0.03 mg·L−1, 0.01 mg·L−1, 0.01 mg·L−1, 0.01 mg·L−1, 0.01 mg·L−1, and 0.02 mg·L−1 for Mg2+, K+, Na+, Sr2+, SO42−, HCO3, NO3, CO32−, and F, respectively. δDH2O and δ18OH2O isotopes were determined using a stable isotope ratio mass spectrometer (IRMS, MAT253, Thermo Fisher, USA). Detection of major ions, trace elements, δDH2O, and δ18OH2O were performed at the Karst Geological Resources and Environment Supervision and Monitoring Centre of the Ministry of Land and Resources. Stable sulfur of dissolved sulfates was measured at the China University of Geosciences (Wuhan) using a stable IRMS (Delta V advantage, Thermo Fisher Scientific, USA). The stable Hydrogen and Oxygen data of water are expressed in delta (δ) as parts per thousand (‰) relative to the Vienna Standard Mean Ocean Water (V-SMOW). Stable sulfur data of dissolved sulfates are expressed in delta (δ) as parts per thousand (‰) relative to the Vienna Canyon Diablo Troilite (V-CDT).

3. Results and Discussion

3.1. Hydrogeochemistry

The chemical composition of karst water is controlled by factors such as geological conditions, hydrodynamic conditions, and human engineering activities in the region, and it can better preserve such influence information [50,51,52,53,54]. Therefore, hydrogeologists often use hydrogeochemical methods to study complex karst water systems [55,56,57,58].

3.1.1. Hydrogeochemistry of the Karst Groundwater

Based on a total of 52 sets of analytical results for the karst groundwater, it was found that the pH values of karst water range from 6.96 to 8.00, with an average of 7.60. The average concentrations of Ca2+, Mg2+, K++Na+, Cl, and SO42− are 143.5 mg·L−1, 36.2 mg·L−1, 45.6 mg·L−1, and 323.9 mg·L−1, respectively. The ranges of each indicator are shown in Table 1 and Figure 2. Figure 3 shows a piper diagram of ionic concentration for the collected samples. The karst water (spring) generally shows cationic high concentrations of Ca2+ and Mg2+ and low concentrations of K++Na+, and anionic high concentration of SO42−, low concentrations of Cl-, and high concentration of HCO3 (Table 1).
The results show that the karst groundwater in Niangziguan Spring area can be divided into three types: SO4·HCO3-Ca·Mg, HCO3·SO4-Ca·Mg, and SO4-Ca types. The sulfate content in 51.9% of the karst groundwater samples exceeds the standard (Standards for drinking water quality, GB 5749-2006 in China).

3.1.2. Hydrogeochemistry of Acidic Mine Drainage (AMD) and Surface Water (River)

Figure 4 shows pH values ranging from 2.75 to 7.48 (an average of 5.41). The concentrations of Ca2+, Mg2+, K++Na+, Cl, and SO42− were 246.4 to 676.1 mg·L−1 (an average of 505.6 mg·L−1), 69.1–1119.8 mg·L−1 (an average of 391.5 mg·L−1), 23.8–686.8 mg·L−1 (an average of 167.2 mg·L−1), 3.1–144.0 mg·L−1 (average of 45.4 mg·L−1), and 1138–16218 mg·L−1 (an average of 5119.2 mg·L−1), respectively. As shown in Figure 5, the concentrations of Cl and SO42− ranged from 14.2 to 522.9 mg·L−1 (an average of 136.1 mg·L−1) and 183.2 to 4042 mg·L−1 (an average of 1240.7 mg·L−1), respectively. These results are closely related to the Shukarev classification types for the coal mine acidic water and surface water. In other words, the coal mine acidic water is usually grouped by SO4-Ca·Mg and SO4-Mg·Ca types while the surface water is mainly classified by SO4-Ca·Na and SO4-Ca types. Therefore, the coal mine acidic water generally exhibited high concentrations of Ca2+, Mg2+, and SO42− while the surface water showed a higher concentration of Cl than that of the acidic water from coal mines (Figure 6).

3.2. Stable Isotopes

Table 1 and Figure 7 show a statistical summary of stable isotopes in the karst groundwater samples in the study area, the δ34S-SO42− values ranged from −6.11‰ to 27.68‰, with an average of 13.93‰. The δD values ranged from −74‰ to −49.8‰, with an average of −68.62‰. The δ18O values were from −10.1‰ to −5.8‰, with an average of −9.25‰.
Table 2 and Figure 7 show the chemical properties and composition of AMD and river samples. The results showed that the δ34S-SO42− values for three surface water samples varied from −5.40‰ to 4.79‰, with an average of 1.11‰ while seven samples from the acidic water at coal mines showed a range of δ34S-SO42 from −6.6‰ to 0.3‰, with an average of −3.3‰. The δD values ranged from −66‰ to −57‰, with an average of −61.7‰. Further, the δ18O values were from −8.9‰ to −7.4‰, with an average of −8.2‰.

4. Discussion

4.1. Sources of K+, Na+, and Cl

The main natural sources of K+, Na+, and Cl are atmospheric precipitation, human activity emissions, and dissolution of salt rocks. Figure 8 shows the concentrations of Na+, K+, and Cl in precipitation in the spring field were in ranges of 1.4 to 23.4 mg·L−1, 0.25 to 9.2 mg·L−1, and 1.6 to 24.7 mg·L−1, respectively. The concentrations of Na+, K+, and Cl in the surface water of the spring field showed ranges of 40.2 to 523.4 mg·L−1, 3.3 to 18.0 mg·L−1, and 3.1 to 144.0 mg·L−1, respectively. In contrast, the concentrations of Na+, K+, and Cl for the acidic water at coal mines were 23.7 to 343.0 mg·L−1, 0.1 to 3.8 mg·L−1, and 3.1 to 144.0 mg·L−1, respectively. This study also measured the concentrations of Na+, K+, and Cl for the karst groundwater, which showed ranges of 6.26 to 193.6 mg·L−1, 0.8 to 6.6 mg·L−1, and 6.0 to 164.9 mg·L−1, respectively. As shown in Figure 9, Cl has a positive correlation with Na+ and K+ in karst groundwater. Na+, K+, and Cl mainly originated from urban domestic sewage or coal mine drainage [59,60,61].

4.2. Sources of Ca2+, Mg2+ and HCO3

This study showed that the karst groundwater has the main ions of Ca2+, Mg2+, and HCO3, which are mainly derived from dissolution of carbonate rocks and gypsum as well as dedolomitization. There are large differences in the mechanisms of groundwater runoff.

4.2.1. Dissolution of Carbonate Rocks and Gypsum

The dissolution of calcite and dolomite in Ordovician strata is the main source of Ca2+, Mg2+, and HCO3. The dissolution equations are shown in Equation (1):
CaSO4 + MgCa(CO3)2 → 2 CaCO3↓+ Mg2+ + SO42−.
The dissolution of gypsum may cause an increase of Ca2+ and SO42− in the karst groundwater. Figure 10a shows a strong positive correlation between Ca2+ and Mg2+ in the karst groundwater. Figure 10b also shows a strong positive correlation between Ca2+ and SO42− in the karst groundwater, indicating that the dissolution of gypsum plays an important role in the concentration of Ca2+ and SO42−.

4.2.2. Dedolomitization

As the solubility of CaSO4 is greater than that of CaCO3 and CaMg(CO3)2 due to the coion effect, the precipitation of calcium carbonate (CaCO3) will occur in order to maintain the equilibrium of Ca2+ concentration. When the precipitation of CaCO3 takes away part of the HCO3 from the solution, dolomite will be dissoluted to maintain the mass balance of HCO3 concentration, which is called dedolomitization [62,63].
We assumed that the dedolomitization is real. Dedolomitization as described in Equation (1) leads to an increase of concentration in SO42− and Mg2+ in the karst groundwater and precipitation of calcite; subsequently, the mass concentrations of HCO3 and Ca2+ remain unchanged. Figure 11 shows that SO42− and Mg2+ have a strong correlation (R2 = 0.87), demonstrating that dedolomitization is the main source of Mg2+.

4.3. Water-Rock Interaction

The chemical composition of the karst groundwater in the Niangziguan spring field is mainly dependent on the dissolution of calcite, dolomite, and gypsum. Figure 12a shows the relationship between SO42++Cl and HCO3.The points below the 1:1 line represent that the dissolution of evaporite is accompanied. The results imply that the karst groundwater in the Niangziguan spring field is mainly affected by the dissolution of carbonate rocks, along with the dissolution of evaporites (e.g., gypsum, Cl in AMD or rivers, etc.). Figure 12b also shows the relationship between SO42++HCO3 and Ca2++Mg2+. Most analytical points are lying on the straight line or below the 1:1 line, indicating that karst groundwater is affected by acid drainage and surface water. The relationship between Na++K+-Cl and Ca2++Mg2+-SO42+-CO3 is shown in Figure 12c. Most of the analytical points from the runoff area lie near the cation exchange line, which means that the hydrochemistry of groundwater is dependent on not only the dissolution of calcite, dolomite, and gypsum but also on cation exchange. Figure 12d shows the relationship between Ca2+/Na+ and HCO3/Na+ to evaluate the effects of materials from different sources (evaporite, silicate, precipitation, and carbonate rocks) in the processes of the groundwater circulation. The result demonstrates that the main factors influencing the hydrochemical features of the karst groundwater in the Niangziguan spring field are precipitation and dissolution of carbonate rocks.
The TDS values, a comprehensive indicator of water quality, obtained from 52 water samples in the karst groundwater within the study area ranged from 280.3 to 1859 mg·L−1, an average of 775.7 mg·L−1. The TDS values in the karst groundwater are larger in order of the recharge area, runoff area, discharge area, and the detention area. Figure 13 show the relationship between TDS and SO42−. It is prominent that TDS has a strong linear correlation with SO42− contents (R2 = 0.922). SO42− in regional karst groundwater is mainly derived from the dissolution of gypsum in the Ordovician carbonate rocks while SO42− in acidic water is from coal mines and meteoric precipitation. The average TDS values in the Niangziguan spring field is 688.6 mg·L−1, lower than the average values analyzed in this study, because the TDS values in the karst water are affected by the runoff condition, the dissolution of gypsum, and the pollution caused by acidic water from coal mines.

4.4. Supply Source of Karst Groundwater

The δ34S-SO42− isotope has been widely used to trace the source of sulfate in water. In addition, H-O isotopes are ideal tracers to detect the origin and evolution of groundwater. Thus, these stable isotopes in groundwater were used to identify the recharge sources of groundwater in the study area.

4.4.1. Recharge Sources

The δD values for the karst groundwater samples in the study area ranged from −74‰ to −49.8‰, an average of −68.62‰ while the δ18O values ranged from −10.1‰ to −5.8‰, an average of −9.25‰. Figure 14 show the relationship between δD and δ18O. The result shows that the intercept and slope of the δD18O relation line are both lower than the global precipitation line (δD = 8δ18O + 10) due to an inland arid climate in the study area. The results also show that meteoric precipitation is the main recharge source of karst water in the Niangziguan spring field. The H-O isotopic composition of surface water was similar to that of the karst groundwater in the runoff area, implying that the karst groundwater is mostly recharged from the seepage of surface water.

4.4.2. Sources of SO42−

While the surface water showed SO42− concentrations ranging from 183.2 to 4042 mg·L−1, an average of 1089.6 mg·L−1, SO42− concentrations for the acidic water in coal mines ranged from 1138 to 16218 mg·L−1, an average of 5079 mg·L−1. In contrast, the karst groundwater showed SO42− concentrations ranging from 21.7 to 1043 mg·L−1, an average of 323.9 mg·L−1 (Figure 15). Precipitation and dissolution of gypsum, oxidation of FeS2 in coal-bearing strata are the sources of SO42− in the karst groundwater. However, SO42− in precipitation (23.95 mg·L−1 on average) has little influence on the karst groundwater. Therefore, it was assumed that most SO42− is sourced from both oxidation of FeS2 and dissolution of gypsum. The δ34S values of SO42− formed dissolution of gypsum and formed oxidation of FeS2 are 26.2‰ (the mean value of gypsum samples) and −3.3‰ (the mean value of AMD in Niangziguan spring catchment), respectively (Figure 16). A 34S isotopic analysis was conducted to determine the source of SO42− by calculating the proportion of SO42− sourced from FeS2 oxidation in the karst groundwater by mass balance. As shown in Table 3, the proportions for the main discharge points (N50, N51, and N52) were 60.6%, 30.3%, and 26.0%, respectively.

5. Conclusions

The karst groundwater in the Niangziguan spring field was mainly classified into SO4·HCO3-Ca·Mg, HCO3·SO4-Ca·Mg, and SO4-Ca types. The SO42− exceeding ratio of karst groundwater samples in this study was 51.9%.
K+, Cl, and Na+ are mainly sourced from urban sewage and coal mine drainage. Dedolomitization gradually led to an increase of Mg2+, especially in the processes of the groundwater runoff. While the main sources of SO42− supply were the dissolution of gypsum and the oxidation of FeS2 in coal-bearing strata, the hydrochemical characteristics of the study area were mainly affected by dedolomitization, the dissolution of dolomite, salt, rock and gypsum.
The isotopic analyses found that acidic water in coal mines contributed to the pollution of the karst groundwater at different degrees. In the groundwater of the Niangziguan spring field, the proportions of SO42− derived from FeS2 oxidation were 60.6% (N50), 30.3% (N51), and 26.0% (N52), respectively.
The acid water of coal mine directly recharges and pollutes karst groundwater through faults or abandoned boreholes, or discharges to rivers and indirectly pollutes karst groundwater through river infiltration in carbonate exposed areas. The main cause for the rapid growth of sulfate in karst groundwater is acid water from abandoned coal mines.

Author Contributions

Data curation, C.T.; Formal analysis, C.T.; Funding acquisition, Y.L.; Investigation, C.T. and Y.L.; Methodology, C.T.; Project administration, Y.L.; Software, C.T.; Supervision, H.J.; Writing—original draft, C.T.; Writing—review & editing, C.T. and H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research were funded by The National Natural Science Foundation of China, grant number 41672253, The China Geological Survey Project of the Ministry of Natural Resources, and grant number DD20190334 and Basic scientific research project of Chinese academy of geological sciences and grant number 2020010.

Institutional Review Board Statement

“Not applicable” for studies not involving humans or animals.

Informed Consent Statement

“Not applicable” for studies not involving humans.

Data Availability Statement

All data, models, and code generated or used during the study appear in the submitted article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ford, D.C.; Williams, P. Karst Hydrogeology and Geomorphology; Wiley: Hoboken, NJ, USA, 2007; ISBN 9780470849965. [Google Scholar]
  2. Nerantzaki, S.D.; Nikolaidis, N.P. The response of three Mediterranean karst springs to drought and the impact of climate change. J. Hydrol. 2020, 125296. [Google Scholar] [CrossRef]
  3. Hasan, O.; Miko, S.; Ilijani, N.; Brunović, D.; Dedić, Ž.; Miko, M.Š.; Peh, Z. Discrimination of topsoil environments in a karst landscape: An outcome of a geochemical mapping campaign. Geochem. Trans. 2020, 21, 1. [Google Scholar] [CrossRef] [PubMed]
  4. Frank, S.; Goeppert, N.; Goldscheider, N. Fluorescence-based multi-parameter approach to characterize dynamics of organic carbon, faecal bacteria and particles at alpine karst springs. Sci. Total Environ. 2018, 1446–1459. [Google Scholar] [CrossRef] [PubMed]
  5. Liang, Y.; Gao, X.; Zhao, C.; Tang, C.; Shen, H.; Wang, Z.; Wang, Y. Review: Characterization, evolution, and environmental issues of karst water systems in Northern China. Hydrogeol. J. 2018, 26, 1371–1385. [Google Scholar] [CrossRef]
  6. Yang, P.; Li, Y.; Groves, C.; Hong, A. Coupled hydrogeochemical evaluation of a vulnerable karst aquifer impacted by septic effluent in a protected natural area. Sci. Total Environ. 2019, 658, 1475–1484. [Google Scholar] [CrossRef]
  7. Ren, K.; Pan, X.; Zeng, J.; Yuan, D. Contaminant sources and processes affecting spring water quality in a typical karst basin (Hongjiadu Basin, SW China): Insights provided by hydrochemical and isotopic data. Environ. Sci. Pollut. Res. 2019, 26, 31354–31367. [Google Scholar] [CrossRef]
  8. Wu, Y.; Luo, Z.; Luo, W.; Ma, T.; Wang, Y. Multiple isotope geochemistry and hydrochemical monitoring of karst water in a rapidly urbanized region. J. Contam. Hydrol. 2018, 218, 44–58. [Google Scholar] [CrossRef]
  9. Cao, X.; Zhou, S.; Xie, F.; Rong, R.; Wu, P. The distribution of rare earth elements and sources in Maoshitou reservoir affected by acid mine drainage, Southwest China. J. Geochem. Explor. 2019, 202, 92–99. [Google Scholar] [CrossRef]
  10. Ren, K.; Zeng, J.; Liang, J.; Yuan, D.; Jiao, Y.; Peng, C.; Pan, X. Impacts of acid mine drainage on karst aquifers: Evidence from hydrogeochemistry, stable sulfur and oxygen isotopes. Sci. Total Environ. 2020, 10, 143223. [Google Scholar] [CrossRef]
  11. Vesper, D.J.; White, W.B. Metal transport to karst springs during storm flow: An example from Fort Campbell, Kentucky/Tennessee, USA. J. Hydrol. 2003, 276, 20–36. [Google Scholar] [CrossRef]
  12. Zhang, X.; Guo, J.; Hu, Q.; Gao, X.; Li, C.; Luo, M.; Wang, Y. Effects of Fe-rich acid mine drainage on percolation features and pore structure in carbonate rocks. J. Hydrol. 2020, 591, 125571. [Google Scholar] [CrossRef]
  13. Boumaiza, L.; Chesnaux, R.; Drias, T.; Walter, J.; Huneau, F.; Garel, E.; Knoeller, K.; Stumpp, C. Identifying groundwater degradation sources in a Mediterranean coastal area experiencing significant multi-origin stresses. Sci. Total Environ. 2020, 746, 141203. [Google Scholar] [CrossRef] [PubMed]
  14. Kaown, D.; Koh, D.C.; Mayer, B.; Lee, K.-K. Identification of nitrate and sulfate sources in groundwater using dual stable isotope approaches for an agricultural area with different land use (Chuncheon, mid-eastern Korea). Agric. Ecosyst. Environ. 2009, 132, 223–231. [Google Scholar] [CrossRef]
  15. Lang, Y.C.; Liu, C.Q.; Li, S.L.; Zhao, Z.Q.; Zhou, Z.-H. Tracing natural and anthropogenic sources of dissolved sulfate in a karst region by using major ion chemistry and stable sulfur isotopes. Appl. Geochem. 2011, 26, S202–S205. [Google Scholar] [CrossRef]
  16. Obeidat, M.; Awawdeh, M.; Matiatos, I.; Al-Ajlouni, A.; Al-Mughaid, H. Identification and apportionment of nitrate sources in the phreatic aquifers in Northern Jordan using a dual isotope method (δ15N and δ18O of NO3-). Groundw. Sustain. Dev. 2020. [Google Scholar] [CrossRef]
  17. Yakovlev, V.; Vystavna, Y.; Diadin, D.; Vergeles, Y. Nitrates in springs and rivers of East Ukraine: Distribution, contamination and fluxes. Appl. Geochem. 2015, 53, 71–78. [Google Scholar] [CrossRef]
  18. Jakóbczyk-Karpierz, S.; Sitek, S.; Jakobsen, R.; Kowalczyk, A. Geochemical and isotopic study to determine sources and processes affecting nitrate and sulphate in groundwater influenced by intensive human activity—Carbonate aquifer Gliwice (southern Poland). Appl. Geochem. 2017, 76, 168–181. [Google Scholar] [CrossRef]
  19. Cao, X.; Wu, P.; Zhou, S.; Sun, J.; Han, Z. Tracing the origin and geochemical processes of dissolved sulphate in a karst-dominated wetland catchment using stable isotope indicators. J. Hydrol. 2018, 562, 210–222. [Google Scholar] [CrossRef]
  20. Zhu, H.; Wu, L.; Xin, C.; Yu, S.; Guo, Y.; Wang, J. Impact of anthropogenic sulfate deposition via precipitation on carbonate weathering in a typical industrial city in a karst basin of southwest China: A case study in Liuzhou. Appl. Geochem. 2019, 110, 104417. [Google Scholar] [CrossRef]
  21. Martin, J.B.; Kurz, M.J.; Khadka, M.B. Climate control of decadal-scale increases in apparent ages of eogenetic karst spring water. J. Hydrol. 2016, 540, 988–1001. [Google Scholar] [CrossRef] [Green Version]
  22. Stange, C.; Tiehm, A. Occurrence of antibiotic resistance genes and microbial source tracking markers in the water of a karst spring in Germany. Sci. Total Environ. 2020, 742, 140529. [Google Scholar] [CrossRef] [PubMed]
  23. Jeelani, G.; Shah, R.A.; Deshpande, R.D.; Fryar, A.E.; Perrin, J.; Mukherjee, A. Distinguishing and estimating recharge to karst springs in snow and glacier dominated mountainous basins of the western Himalaya, India. J. Hydrol. 2017, 550, 239–252. [Google Scholar] [CrossRef]
  24. Karmann, I.; Cruz, F.W.; Viana, O.; Burns, S.J. Climate influence on geochemistry parameters of waters from Santana–Pérolas cave system, Brazil. Chem. Geol. 2007, 244, 232–247. [Google Scholar] [CrossRef]
  25. Qin, R.; Wu, Y.; Xu, Z.; Xie, D.; Zhang, C. Assessing the impact of natural and anthropogenic activities on groundwater quality in coastal alluvial aquifers of the lower Liaohe River Plain, NE China. Appl. Geochem. 2013, 31, 142–158. [Google Scholar] [CrossRef]
  26. Capaccioni, B.; Vaselli, O.; Tassi, F.; Santo, A.P.; Huertas, A.D. Hydrogeochemistry of the thermal waters from the Sciacca Geothermal Field (Sicily, southern Italy). J. Hydrol. 2011, 396, 292–301. [Google Scholar] [CrossRef]
  27. Kohfahl, C.; Sprenger, C.; Herrera, J.B.; Meyer, H.; Chacón, F.F.; Pekdeger, A. Recharge sources and hydrogeochemical evolution of groundwater in semiarid and karstic environments: A field study in the Granada Basin (Southern Spain). Appl. Geochem. 2008, 23, 846–862. [Google Scholar] [CrossRef] [Green Version]
  28. Pasvanoğlu, S. Hydrogeochemistry of thermal and mineralized waters in the Diyadin (Ağri) area, Eastern Turkey. Appl. Geochem. 2013, 38, 70–81. [Google Scholar] [CrossRef]
  29. Shi, M.; Wang, S.; Argiriou, A.A.; Zhang, M.; Guo, R.; Jiao, R.; Kong, J.; Zhang, Y.; Qiu, X.; Zhou, S. Stable Isotope Composition in Surface Water in the Upper Yellow River in Northwest China. Water 2019, 11, 967. [Google Scholar] [CrossRef] [Green Version]
  30. Ekholm, P.; Lehtoranta, J.; Taka, M.; Sallantaus, T.; Riihimäki, J. Diffuse sources dominate the sulfate load into Finnish surface waters. Sci. Total Environ. 2020, 748, 141297. [Google Scholar] [CrossRef]
  31. Saldanha, M.M.; Araújo, I.C.S.; Triguineli, M.V.; Vaz, D.P.; Ferreira, F.N.A.; Albergaria, J.D.S.; Fontes, D.O.; Lara, L.J.C. Relative bioavailability of manganese in relation to proteinate and sulfate sources for broiler chickens from one to 20 d of age. Poult. Sci. 2020, 99, 5647–5652. [Google Scholar] [CrossRef]
  32. Torres-Martínez, J.A.; Mora, A.; Knappett, P.S.K.; Ornelas-Soto, N.; Mahlknecht, J. Tracking nitrate and sulfate sources in groundwater of an urbanized valley using a multi-tracer approach combined with a Bayesian isotope mixing model. Water Res. 2020, 182, 115962. [Google Scholar] [CrossRef] [PubMed]
  33. Zak, D.; Hupfer, M.; Cabezas, A.; Jurasinski, G.; Audet, J.; Kleeberg, A.; McInnes, R.; Kristiansen, S.M.; Petersen, R.J.; Liu, H.; et al. Sulphate in freshwater ecosystems: A review of sources, biogeochemical cycles, ecotoxicological effects and bioremediation. Earth-Sci. Rev. 2020. [Google Scholar] [CrossRef]
  34. Zhang, Q.; Wang, H.; Lu, C. Tracing sulfate origin and transformation in an area with multiple sources of pollution in northern China by using environmental isotopes and Bayesian isotope mixing model. Environ. Pollut. 2020, 265, 115105. [Google Scholar] [CrossRef] [PubMed]
  35. Zheng, L.; Chen, X.; Dong, X.; Wei, X.; Jiang, C.; Tang, Q. Using δ34S–SO4 and δ18O–SO4 to trace the sources of sulfate in different types of surface water from the Linhuan coal-mining subsidence area of Huaibei, China. Ecotoxicol. Environ. Saf. 2019, 181, 231–240. [Google Scholar] [CrossRef]
  36. Chen, X.; Sun, Y.; Huang, R. Role of hydro-geochemical functions on karst critical zone hydrology for sustainability of water resources and ecology in Southwest China. Acta Geochim. 2017, 36, 494–497. [Google Scholar] [CrossRef]
  37. He, S.; Zhu, L.; Yang, R.; Shen, Z.; Yu, X. The geochemical characteristics of aqueous rare-earth elements in shallow karst groundwater in Guiyang City, China. Chin. J. Geochem. 2011, 30, 114–124. [Google Scholar] [CrossRef]
  38. Holland, M.; Witthüser, K.T. Geochemical characterization of karst groundwater in the cradle of humankind world heritage site. S. Afr. Environ. Geol. 2008, 57, 513–524. [Google Scholar] [CrossRef] [Green Version]
  39. Jing, M.; Chen, W.; Zheng, T.; Liao, Y.; Ellis Burnet, J.; Xu, M.; Yang, C.; Shen, L.; Liang, M. Water geochemical characteristic variations in and around a karst-dominated natural reserve area, southwestern China. Environ. Earth Sci. 2011, 64, 1051–1058. [Google Scholar] [CrossRef]
  40. Rosales Lagarde, L.; Boston, P.J.; Campbell, A.R.; Hose, L.D.; Axen, G.; Stafford, K.W. Hydrogeology of northern Sierra de Chiapas, Mexico: A conceptual model based on a geochemical characterization of sulfide-rich karst brackish springs. Hydrogeol. J. 2014, 22, 1447–1467. [Google Scholar] [CrossRef]
  41. Tuyukina, T.Y. Geochemical studies of northern taiga (gypsum) karst ecosystems and their high vulnerability to natural and anthropogenic hazards. Environ. Geol. 2008, 58, 269–274. [Google Scholar] [CrossRef]
  42. Yang, M.; Lu, Y.; Zhang, F.E.; Zhang, S.; Yin, M.; Wu, G. Identification of geochemical processes by hydrogeochemical analysis in karst aquifers of a semi-arid region, Northern China. Carbonates Evaporites 2017, 34, 297–313. [Google Scholar] [CrossRef]
  43. Brkić, Ž.; Kuhta, M.; Hunjak, T. Groundwater flow mechanism in the well-developed karst aquifer system in the western Croatia: Insights from spring discharge and water isotopes. Catena 2018, 161, 14–26. [Google Scholar] [CrossRef]
  44. Alves, M.; Galvão, P.; Aranha, P. Karst Hydrogeological Controls and Anthropic Effects in an Urban Lake. J. Hydrol. 2020. [Google Scholar] [CrossRef]
  45. Ayadi, Y.; Mokadem, N.; Besser, H.; Khelifi, F.; Harabi, S.; Hamad, A.; Boyce, A.; Laouar, R.; Hamed, Y. Hydrochemistry and stable isotopes (δ18O and δ2H) tools applied to the study of karst aquifers in southern mediterranean basin (Teboursouk area, NW Tunisia). J. Afr. Earth Sci. 2018, 137, 208–217. [Google Scholar] [CrossRef] [Green Version]
  46. Gil-Márquez, J.M.; Andreo, B.; Mudarra, M. Combining hydrodynamics, hydrochemistry, and environmental isotopes to understand the hydrogeological functioning of evaporite-karst springs. An example from southern Spain. J. Hydrol. 2019, 576, 299–314. [Google Scholar] [CrossRef]
  47. Herms, I.; Jódar, J.; Soler, A.; Vadillo, I.; Lambán, L.J.; Martos-Rosillo, S.; Núñez, J.A.; Arnó, G.; Jorge, J. Contribution of isotopic research techniques to characterize high-mountain-Mediterranean karst aquifers: The Port del Comte (Eastern Pyrenees) aquifer. Sci. Total Environ. 2019, 656, 209–230. [Google Scholar] [CrossRef]
  48. Husic, A.; Fox, J.; Adams, E.; Pollock, E.; Ford, W.; Agouridis, C.; Backus, J. Quantification of nitrate fate in a karst conduit using stable isotopes and numerical modeling. Water Res. 2020, 170, 115348. [Google Scholar] [CrossRef]
  49. Ji, H.; Chang, C.; Beckford, H.O.; Song, C.; Blake, R.E. New perspectives on lateritic weathering process over karst area—Geochemistry and Si-Li isotopic evidence. Catena 2020. [Google Scholar] [CrossRef]
  50. Ma, R.; Wang, Y.; Sun, Z.; Zheng, C.; Ma, T.; Prommer, H. Geochemical evolution of groundwater in carbonate aquifers in Taiyuan, northern China. Appl. Geochem. 2011, 26, 884–897. [Google Scholar] [CrossRef]
  51. Hong Quang, N.; Tuan, V.A.; Thi Thu Hang, L.; Manh Hung, N.; Thi The, D.; Thi Dieu, D.; Duc Anh, N.; Hackney, C.R. Hydrological/Hydraulic Modeling-Based Thresholding of Multi SAR Remote Sensing Data for Flood Monitoring in Regions of the Vietnamese Lower Mekong River Basin. Water 2019, 12, 71. [Google Scholar] [CrossRef] [Green Version]
  52. Sağır, Ç.; Kurtuluş, B.; Razack, M. Hydrodynamic Characterization of Mugla Karst Aquifer Using Correlation and Spectral Analyses on the Rainfall and Springs Water-Level Time Series. Water 2019, 12, 85. [Google Scholar] [CrossRef] [Green Version]
  53. Li, Y.; Liu, H.; Li, G.; Luo, W.; Sun, Y. Manure digestate storage under different conditions: Chemical characteristics and contaminant residuals. Sci. Total Environ. 2018, 639, 19–25. [Google Scholar] [CrossRef] [PubMed]
  54. Appiah–Effah, E.; Duku, G.A.; Dwumfour–Asare, B.; Manu, I.; Nyarko, K.B. Toilet chemical additives and their effect on faecal sludge characteristics. Heliyon 2020, 6, e04998. [Google Scholar] [CrossRef] [PubMed]
  55. Lee, C.L.; Chin, K.L.; H’ng, P.S.; Rashid, U.; Maminski, M.; Khoo, P.S. Effect of pretreatment conditions on the chemical-structural characteristics of coconut and palm kernel shell: A potentially valuable precursor for eco-efficient activated carbon production. Environ. Technol. Innov. 2020. [Google Scholar] [CrossRef]
  56. Santana, N.A.; Jacques, R.J.S.; Antoniolli, Z.I.; Martínez-Cordeiro, H.; Domínguez, J. Changes in the chemical and biological characteristics of grape marc vermicompost during a two-year production period. Appl. Soil Ecol. 2020, 154, 103587. [Google Scholar] [CrossRef]
  57. Chen, H.; Zhang, W.; Zhou, Y.; Li, J.; Zhao, H.; Xu, S.; Xia, W.; Cai, Z.; Li, Y. Characteristics of exposure to multiple environmental chemicals among pregnant women in Wuhan, China. Sci. Total Environ. 2021, 754, 142167. [Google Scholar] [CrossRef]
  58. Li, T.; Liu, B.; Bi, X.; Wu, J.; Zhang, Y.; Feng, Y. Size and chemical characteristics of particles emitted from typical rural biomass cookstoves in North China. Atmos. Res. 2021, 249, 105295. [Google Scholar] [CrossRef]
  59. Li, C.; Gao, X.; Wang, W.; Zhang, X.; Zhang, X.; Jiang, C.; Wang, Y. Hydro-biogeochemical processes of surface water leakage into groundwater in large scale karst water system: A case study at Jinci, northern China. J. Hydrol. 2020. [Google Scholar] [CrossRef]
  60. Valiente, N.; Gil-Márquez, J.M.; Gómez-Alday, J.J.; Andreo, B. Unraveling groundwater functioning and nitrate attenuation in evaporitic karst systems from southern Spain: An isotopic approach. Appl. Geochem. 2020, 123, 104820. [Google Scholar] [CrossRef]
  61. Yue, F.J.; Li, S.L.; Waldron, S.; Wang, Z.J.; Oliver, D.M.; Chen, X.; Liu, C.-Q. Rainfall and conduit drainage combine to accelerate nitrate loss from a karst agroecosystem: Insights from stable isotope tracing and high-frequency nitrate sensing. Water Res. 2020, 186, 116388. [Google Scholar] [CrossRef]
  62. Choi, B.Y.; Yun, S.T.; Mayer, B.; Hong, S.Y.; Kim, K.H.; Jo, H.-Y. Hydrogeochemical processes in clastic sedimentary rocks, South Korea: A natural analogue study of the role of dedolomitization in geologic carbon storage. Chem. Geol. 2012, 306–307, 103–113. [Google Scholar] [CrossRef]
  63. Rameil, N. Early diagenetic dolomitization and dedolomitization of Late Jurassic and earliest Cretaceous platform carbonates: A case study from the Jura Mountains (NW Switzerland, E France). Sediment. Geol. 2008, 212, 70–85. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Hydrogeological map of Niangziguan spring catchment.
Figure 1. Hydrogeological map of Niangziguan spring catchment.
Water 13 00390 g001
Figure 2. Box chart of ionic concentration in karst groundwater of Niangziguan spring catchment.
Figure 2. Box chart of ionic concentration in karst groundwater of Niangziguan spring catchment.
Water 13 00390 g002
Figure 3. Piper diagram of ionic concentrations in the karst groundwater.
Figure 3. Piper diagram of ionic concentrations in the karst groundwater.
Water 13 00390 g003
Figure 4. Box chart of ionic concentration in Acidic Mine Drainage (AMD).
Figure 4. Box chart of ionic concentration in Acidic Mine Drainage (AMD).
Water 13 00390 g004
Figure 5. Box chart of ionic concentrations in the river.
Figure 5. Box chart of ionic concentrations in the river.
Water 13 00390 g005
Figure 6. Piper diagram of ionic concentrations in the AMD or river.
Figure 6. Piper diagram of ionic concentrations in the AMD or river.
Water 13 00390 g006
Figure 7. Box chart of stable isotopes in the AMD, river, or karst groundwater.
Figure 7. Box chart of stable isotopes in the AMD, river, or karst groundwater.
Water 13 00390 g007
Figure 8. Box chart of K+, Na+, and Cl.
Figure 8. Box chart of K+, Na+, and Cl.
Water 13 00390 g008
Figure 9. Relationships between (a) K+ vs. Cl in karst groundwater (b) Na+ vs. Cl in karst groundwater.
Figure 9. Relationships between (a) K+ vs. Cl in karst groundwater (b) Na+ vs. Cl in karst groundwater.
Water 13 00390 g009
Figure 10. (a) Relationships between Mg2+ vs. Ca2+ (b) Relationships between SO42+ vs. Ca2+.
Figure 10. (a) Relationships between Mg2+ vs. Ca2+ (b) Relationships between SO42+ vs. Ca2+.
Water 13 00390 g010
Figure 11. Relationships between SO42− vs. Mg2+.
Figure 11. Relationships between SO42− vs. Mg2+.
Water 13 00390 g011
Figure 12. Ion relationship involving karst groundwater in the Niangziguan Spring Catchment (a) Relationships between Mg2+ vs. SO42−, (b) SO42−+HCO3 vs. Ca2++Mg2+, (c) Na++K+-Cl vs. Ca2++Mg2+-SO42+-CO3, (d) Ca2+/Na+ vs. HCO3/Na+.
Figure 12. Ion relationship involving karst groundwater in the Niangziguan Spring Catchment (a) Relationships between Mg2+ vs. SO42−, (b) SO42−+HCO3 vs. Ca2++Mg2+, (c) Na++K+-Cl vs. Ca2++Mg2+-SO42+-CO3, (d) Ca2+/Na+ vs. HCO3/Na+.
Water 13 00390 g012
Figure 13. Relationships between TDS vs. SO42−.
Figure 13. Relationships between TDS vs. SO42−.
Water 13 00390 g013
Figure 14. Relationships between δD vs. δ18O for the karst groundwater.
Figure 14. Relationships between δD vs. δ18O for the karst groundwater.
Water 13 00390 g014
Figure 15. Box chart of SO42−.
Figure 15. Box chart of SO42−.
Water 13 00390 g015
Figure 16. Box chart of δ34S values of SO42− forming the dissolution of gypsum and forming the oxidation of FeS2.
Figure 16. Box chart of δ34S values of SO42− forming the dissolution of gypsum and forming the oxidation of FeS2.
Water 13 00390 g016
Table 1. Statistical summary of water quality parameters and stable isotopes in karst groundwater.
Table 1. Statistical summary of water quality parameters and stable isotopes in karst groundwater.
Sample NO.TypeT/℃pHMajor Elements (mg·L−1)
Ca2+Mg2+K+Na+ClSO42−HCO3CO3D18O34S
N01karst well207.8100.934.11.3662.919.5304.2212.5n.d.−70−9.3−6.11
N02karst well217.5512019.32.5222.117.7221.2188.2n.d−49−5.8−3.54
N03karst well167.7329.168.54.23131.470.91029176.1n.d−63−8.5−0.21
N04karst well157.7126.221.22.3265.736.5231.6248.9n.d−65−8.50.77
N05karst well187.21223.448.56.56193.6164.9623.5273.2n.d−60−84.03
N06karst well147.887.816.50.7546.414.2106261.1n.d−69−9.44.24
N07karst well19.97.53220.344.35.45149.5152.5536.2261.1n.d−62−8.24.84
N08karst well14.57.871.515.30.7533.312.4104191.2n.d−68−9.24.97
N09karst well167.5183.937.53.4399.886.9427261.1n.d−63−8.45.98
N10karst well19.27.867.418.21.0533.21697.1200.4n.d−68−9.46.14
N11karst well17.57.577.923.91.0410.6621.7303.6n.d−69−9.36.37
N12karst well17.17.998.227.21.3812.616.7118.4245.9n.d−70−9.57.22
N13karst well187.5329.979.13.45125.265.6975.2297.5n.d−67−8.97.43
N14karst well19.27.569.530.91.348.48.9202206.4n.d−70−9.47.49
N15karst well15.97.14262.655.63.1678.3101595.1297.5n.d−63−8.58.56
N16karst well25.88220.746.55.01115.6119.5664.6112.3n.d−52−6.29.41
N17karst well177.57185.340.11.210.49.6357.8270.8n.d−73−1010.31
N18karst well177.8696.424.91.2610.416112.5230.7n.d−70−9.410.43
N19karst well177.666.821.30.7650.413.1111.5236.8n.d−71−9.711.86
N20karst well197.678.823.11.286.39.641.7276.2n.d−71−9.812.24
N21karst well20.77.87164.346.52.1850.458.5396.9212.5n.d−71−9.612.31
N22karst well19.47.8141.439.22.8974.469.1320.3242.9n.d−69−9.213.55
N23karst well19.77.881.428.31.123512.4175.7203.4n.d−70−9.514.1
N24karst well157.7133.736.51.0935.221.6286.6224.6n.d−67−8.914.6
N25karst well17884.825.51.2640.123.8177.6194.3n.d−65−8.514.66
N26karst well167.562.122.10.8210.17.124.4264.1n.d−72−1015.12
N27karst well19.47.5207.952.21.272226.6495.6227.7n.d−68−9.316.45
N28karst well15.87.872.520.31.0751.27.1117.3267.1n.d−70−9.516.87
N29karst well17.57.789.921.21.2740.912.4147.5236.8n.d−68−9.217
N30karst well177.7566.724.41.066.7629.1276.2n.d−73−9.917.02
N31karst well20.56.96375.871.91.7520.430.11043236.8n.d−67−917.3
N32karst well207.568.424.80.95138.964.2255n.d−72−9.917.46
N33karst well177.512540.51.1535.512.4320.8218.6n.d−68−9.117.98
N34karst well177.8115.828.70.9736.59.2252.8225.9n.d−74−9.618.03
N35karst well177.599.636.41.3813.18.9193.6239.8n.d−73−10.119.08
N36karst well16.17.6112.427.90.95812.4139.5267.1n.d−74−10.119.16
N37karst well18.17.5119.5301.098.37.1176.6261.1n.d−72−9.920.52
N38karst well17.68141.534.24.9935.647.9349.8145.7n.d−69−9.421.15
N39karst well177.3127.542.31.4614.28.9293.7242.9n.d−73−10.121.18
N40karst well20.77.18346.665.91.3216.812.4916.5252n.d−70−9.722.06
N41karst well197.4132.2410.8613.29.6257.7256.2n.d−72−9.722.6
N42karst well207.6112.636.91.212.47.1272.6206.4n.d−73−1022.65
N43karst well227.8119.724.90.949.716236.2145.7n.d−66−8.924.47
N44karst well197.411531.81.2612.68.9222224.6n.d−74−10.124.52
N45karst well25.97.7100.232.80.9911.37.1181.6236.8n.d−73−10.124.62
N46karst well20.57.86122.951.30.9410.78.9307.4245.9n.d−71−9.724.87
N47karst well20.87.2633470.51.346.816918.3276.2n.d−70−9.524.97
N48karst well197.616035.41.24467.8386.4236.8n.d−71−9.626.06
N49karst well20.97.3166.838.51.53519.6431.7233.7n.d−70−9.327.68
N50Chengxi spring167.6118.431.62.0542.853.2185.3255n.d−68−9.18.31
N51Wulong spring197.66127.235.61.9271.158.5315.1233.7n.d−71−9.617.26
N52Jiquanzhan 197.45101.8362.0374.460.3326.9157.9n.d−71−9.718.54
Table 2. The stable isotopes of AMD and River samples.
Table 2. The stable isotopes of AMD and River samples.
Sample NO.TypespHT/°CDatemg·L−1
TDSδ34SδDδ18O
C01_13AMD6.011318 April 20134610−3.3//
C02_13AMD7.4815.226 April 20133292−6.6//
C03_13AMD3.2816.52 May 20138272−4.9//
C01_14AMD7.252425 June 20144240///
C02_14AMD7.272223 June 20143518///
C03_14AMD3.3927.514 June 20145970−4.6//
C01_15AMD7.252111 September 20153901///
C02_15AMD7.1720.817 September 20153310///
C03_15AMD6.7617.515 September 20151717///
C04_19AMD2.831930 April 201918936−0.6//
C05_19AMD2.7516.530 April 2019190940.3//
S01River3.332725 June 20146155///
S02River7.2518.214 June 20141702−5.40−66−8.9
S03River7.272316 June 20141496///
S04River7.2728.417 June 2014748.3///
S05River7.322020 June 2014522.5///
S06River6.992922 June 201413164.79−57−7.4
S07River7.0326.322 June 201416793.94−62−8.3
Table 3. Ratio of sulfate ion from pyrite oxidation in the calculation of water/%.
Table 3. Ratio of sulfate ion from pyrite oxidation in the calculation of water/%.
Sample NO.δ34S (‰)From FeS2(%)Sample NO.δ34S (‰)From FeS2(%)Sample NO.δ34S (‰)From FeS2(%)
N01−6.11100N1911.8654.7 N3720.5219.3
N02−3.54100N2012.2453.4 N3821.1517.1
N03−0.2189.5 N2112.3153.2 N3921.1817.0
N040.7786.2 N2213.5549.0 N4022.0614.0
N054.0375.2 N2314.147.1 N4122.612.2
N064.2474.4 N2414.645.4 N4222.6512.0
N074.8472.4 N2514.6645.2 N4324.475.9
N084.9772.0 N2615.1243.7 N4424.525.7
N095.9868.5 N2716.4539.2 N4524.625.4
N106.1468.0 N2816.8737.7 N4624.874.5
N116.3767.2 N291737.3 N4724.974.2
N127.2264.3 N3017.0237.2 N4826.060.5
N137.4363.6 N3117.336.3 N4927.680.0
N147.4963.4 N3217.4635.7 N508.3160.6
N158.5659.8 N3317.9834.0 N5117.2630.3
N169.4156.9 N3418.0333.8 N5218.5426.0
N1710.3153.9 N3519.0830.2
N1810.4353.5 N3619.1630.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tang, C.; Jin, H.; Liang, Y. Using Isotopic and Hydrochemical Indicators to Identify Sources of Sulfate in Karst Groundwater of the Niangziguan Spring Field, China. Water 2021, 13, 390. https://doi.org/10.3390/w13030390

AMA Style

Tang C, Jin H, Liang Y. Using Isotopic and Hydrochemical Indicators to Identify Sources of Sulfate in Karst Groundwater of the Niangziguan Spring Field, China. Water. 2021; 13(3):390. https://doi.org/10.3390/w13030390

Chicago/Turabian Style

Tang, Chunlei, Hua Jin, and Yongping Liang. 2021. "Using Isotopic and Hydrochemical Indicators to Identify Sources of Sulfate in Karst Groundwater of the Niangziguan Spring Field, China" Water 13, no. 3: 390. https://doi.org/10.3390/w13030390

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