Next Article in Journal
Examining the Spatiotemporal Evolution of Land Use Conflicts from an Ecological Security Perspective: A Case Study of Tianshui City, China
Previous Article in Journal
Livestock and Water Resources: A Comparative Study of Water Footprint in Different Farming Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Geochemical Characteristics of Composite Geothermal Reservoirs for Sustainable Utilization: A Case Study of the Northwestern Shandong Geothermal Area in China

1
China Renewable Energy Engineering Institute, Beijing 100120, China
2
Institute of Hydrogeology and Environmental Geology, Chinese Academy of Geological Sciences, Shijiazhuang 050061, China
3
Technical Innovation Center for Geothermal and HDR Exploration and Development, Ministry of Natural Resources, Shijiazhuang 050800, China
4
Guiyang Engineering Corporation Limited, Guiyang 553304, China
*
Author to whom correspondence should be addressed.
Sustainability 2025, 17(5), 2252; https://doi.org/10.3390/su17052252
Submission received: 12 December 2024 / Revised: 28 February 2025 / Accepted: 28 February 2025 / Published: 5 March 2025

Abstract

Presently, geothermal resources have been globally recognized as an indispensable component of the energy system due to their sustainability. However, previous studies on geothermal reservoirs focus primarily on single reservoirs, lacking a systematic investigation of composite geothermal reservoirs. The geothermal reservoirs in the northwestern Shandong geothermal area in China are primarily of sandstone and karst types, characterized by extensive distributions, shallow burial depths, high water temperatures, and high water abundance, holding considerable potential for exploitation. This study explored the hydrochemical, isotopic, and circulation characteristics of geothermal fluids in the composite geothermal reservoirs in the study area using methods like hydrogeochemistry and geothermal geology. The purpose is to determine the geochemical differences in geothermal fluids across the composite geothermal reservoirs and provide scientific support for subsequently efficient and sustainable exploitation and utilization of geothermal resources in the study area. The composite geothermal reservoirs in the study area are composed of porous sandstone geothermal reservoirs (also referred to as sandstone reservoirs) in the upper part and karst-fissured geothermal reservoirs (also referred to as karst reservoirs) in the lower part. The results show that the geothermal fluids in the sandstone and karst reservoirs are primarily of Na-Cl-SO4 and Na-Ca-Cl-SO4 types, respectively. The hydrochemical composition of geothermal fluids in the karst reservoirs is principally influenced by the precipitation–dissolution equilibrium of carbonate and sulfate minerals, while that in the sandstone reservoirs is predominantly influenced by the precipitation–dissolution equilibrium of carbonate and silicate minerals, as well as cation exchange reactions. The temperatures of the karst reservoirs were calculated at 52.9–82.09 °C using geothermometers. Given the cold-water mixing ratios range from 89% to 96%, the corrected reservoir temperatures vary from 200 to 225 °C. In contrast, the temperatures of the sandstone reservoirs were calculated at 60.54–85.88 °C using geothermometers. These reservoirs exhibit cold water mixing ratios ranging from 85% to 90%, and their corrected reservoir temperatures vary from 150 to 200 °C accordingly. The circulation depths of geothermal fluids in the karst and sandstone reservoirs range from 1107.28 to 1836.69 m and from 1366.60 to 2102.29 m, respectively. The study area is primarily recharged by meteoric water from Mount Tai and the Lushan and Yishan mountains (collectively referred to as the Tai-Lu-Yi mountains) to the southeast of the study area. Investigating the differences in geochemical characteristics of geothermal fluids in composite geothermal reservoirs in the study area is significant for balancing the exploitation and supply of geothermal resources, optimizing the exploitation and utilization modes, and promoting the efficient and sustainable exploitation and utilization of geothermal resources in the study area.

1. Introduction

With rapid economic and social advancement, the world is facing increasingly severe challenges such as environmental degradation, climate change, and shortages of renewable energy, thus showing a growing demand for green renewable energy [1,2,3,4,5]. Geothermal energy, as a clean renewable energy source, has gradually attracted global attention in recent decades. China, located at the junction of the circum-Pacific and Mediterranean-Himalayan geothermal belts, boasts abundant geothermal resources [6,7,8]. Shandong Province in China exhibits diverse and widely distributed geothermal resources. Within Shandong Province, the sedimentary-basin geothermal resources can be found primarily in the study area and the southwestern Shandong geothermal area [9,10,11]. The study area consists of the Jiyang depression, the Linqing depression, and the piedmont loose sedimentary cover area in the northern Luzhong uplift. It manifests relatively shallow Moho discontinuity, high heat flow values, and aquifers with strong water yield properties, establishing it as the area with the highest exploitation potential for geothermal resources in Shandong Province [12]. In addition to abundant geothermal resources, the study area predominantly develops stratified geothermal reservoirs, which are characterized by extensive distribution areas, shallow burial depths, high water temperatures, and high water yields from individual geothermal wells. These characteristics facilitate exploitation and are particularly suitable for decentralized heating exploitation in towns, villages, enterprises, and public institutions. Currently, the geothermal resources in the study area are primarily exploited for heating, bathing, swimming, and aquaculture, with geothermal heating reaching a certain scale. Considering various geothermal reservoirs in the study area, it is essential to balance and regulate their exploitation and replenishment to ensure a long-term stable supply of geothermal resources. Many studies delved into the hydrochemical characteristics and genetic mechanisms of geothermal fluids in the dominant sandstone and karst reservoirs in the study area [13,14]. They found that the geothermal fluids in sandstone reservoirs originated primarily from the infiltration and recharge of meteoric water and were subjected to minor influence of tectonic activity [15]. In contrast, the geothermal fluids in karst reservoirs were principally sourced from a mixture of meteoric water and deep circulating groundwater and were significantly affected by tectonic processes and water–rock interactions [16].
The hydrochemical and especially isotopic characteristics of geothermal fluids are often used to analyze the circulation characteristics, water–rock interactions, and hydrogeochemical processes of geothermal fluids [17,18,19,20,21,22]. Isotopic characteristics can identify various physicochemical processes in geothermal systems, such as the origin of geothermal fluids, water–rock interactions, circulation depths, recharge elevations, and the mixing of shallow groundwater with geothermal water [23,24,25,26,27]. Classic geothermometers based on the chemical compositions of geothermal fluids are effective in assessing whether a hydrothermal system is a high-temperature system [28,29,30,31,32,33,34]. Based on the hydrochemical and isotopic characteristics of geothermal fluids in sandstone and karst reservoirs in the study area, this study comprehensively analyzes the origin, water–rock interactions, circulation characteristics, and mixing processes of geothermal fluids in different reservoir layers in the study area. It reveals the differences in geothermal fluids among various reservoir layers in the study area, providing a theoretical basis for the balanced regulation and exploitation of various reservoir layers in the study area.

2. Geological Setting

Shandong Province is located on the third major step of China’s topography. Based on geomorphic characteristics and genetic types, it is zoned into three parts: the low mountain and hilly area dominated by denudation structures in eastern Shandong, the medium-low hilly area with tectonic erosion in south-central Shandong, and the accumulation plain in northwestern Shandong. The sedimentary-basin geothermal resources in Shandong Province are primarily distributed in the study area and the southwestern Shandong geothermal area. The study area consists of the Jiyang depression, the Linqing depression, and the piedmont loose sedimentary cover area in the northern Luzhong uplift. It manifests relatively shallow Moho discontinuity, high heat flow values, and aquifers with strong water yield properties, establishing it as the area with the highest exploitation potential for geothermal resources in Shandong Province. The study area, characterized by alternating uplifts and sags, resides in the North China depression in terms of tectonic unit and the accumulation plain in northwestern Shandong in terms of geomorphology. The uplift zone in the study area is primarily composed of Archean metamorphic rocks and Paleozoic limestones or sandstones. The sag zone develops a faulted basin with the Archean Taishan metamorphic rock series as the basement, on which extremely thick Paleozoic–Cenozoic sedimentary rocks were deposited. The study area primarily includes northern Ji’nan City, western Liaocheng and Heze cities, and most parts of Dezhou, Dongying, and Binzhou cities [9].
The study area is located in a Mesozoic–Cenozoic faulted basin developed in the North China platform. In terms of tectonic unit, it is situated in the North China depression within the North China plate. From the Mesozoic to the Cenozoic, the Himalayan orogeny and Yanshanian movement resulted in fault structures and concave–convex tectonic units like the Jiyang and Linqing depressions in the study area. Under the influence of the Neocathaysian tectonic system, the bedrock fault structures were intensely developed in the study area. The fault structures in the study area are all of the hidden type, with the Qihe–Guangrao, Liaocheng–Lankao, and Cangdong faults being major deep heat-controlling faults [35,36].
The study area develops stratified geothermal reservoirs and extremely thick Paleozoic and Cenozoic strata. The extremely thick Cenozoic loose sedimentary strata cover the metamorphic and magmatic rocks of the basement while providing insulation for hot water in the underlying reservoirs. Additionally, pores and fractures in the Cenozoic sandstones manifest high water-bearing properties and a certain water-retaining function. The dominant geothermal reservoirs comprise the Neogene sandstone reservoirs and the Cambrian–Ordovician karst reservoirs. The Neogene sandstone reservoirs primarily include the sandstone reservoirs of the Minghuazhen and Guantao formations. The sandstone reservoirs of the Minghuazhen formation are distributed throughout the study area, with floor burial depths generally ranging from 500 to 1100 m. Typically, they display thicknesses ranging from 30 to 120 m, with slender reservoirs in the uplift zone and thick reservoirs in the sag zone. They show single-well water yields ranging from 40 to 80 m3/d and wellhead water temperatures ranging from 30 to 48 °C. The sandstone reservoirs of the Guantao formation are distributed in most of the study area except for a local absence in northern Weifang City, with floor burial depths generally ranging from 1000 to 1700 m and up to 2300 m. They display thicknesses generally ranging from 250 to 400 m, with single-well water yields ranging from 40 to 80 m3/h and water temperatures ranging from 49 to 78 °C. The Cambrian–Ordovician geothermal limestone reservoirs for karst water are primarily distributed in the Ningjin, Yihezhuang, Chenjiazhuang, Chengdong, Qingtuozi, Guangrao, and Shouguang uplifts, and the Guantao and Gaotang uplifts in the Linqing depression. Their burial depths are principally subjected to the basement structure, with roof burial depths generally ranging from 500 to 1700 m. With thicknesses ranging from about 100 to 700 m, they are principally composed of limestones and dolomites. They show single-well water yields generally ranging from 100 to 1000 m3/d and up to over 3000 m3/d and wellhead temperatures ranging from 50 to 90 °C [37].
The study area is part of the upper mantle uplift zone in North China, with Moho depths ranging from approximately 33 to 37 km. It displays concave-downward Moho discontinuity in the uplift zone, suggesting thicker crust, and convex-upward Moho discontinuity in the sag zone, indicating thinner crust. The terrestrial heat flow in the study area shows values ranging from about 55 to 75 MW/m2, exceeding the average value (47.16 mW/m2) of terrestrial heat flow in the North China region. The geothermal gradients in the study area show lower values in the inland and higher values along the coast, and lower and higher values in the west and east, respectively. The geothermal gradients range from 3.0 to 3.5 °C/100m in the southwestern part of the study area, and from 3.5 to 5.0 °C/100m in the northern and eastern parts. At a depth of 1000 m below the surface, the formation temperatures range from 40 to 50 °C in the southwestern part and 45 to 60 °C along the coast and in the eastern and northeastern parts. In terms of thermal conductivity in the study area, the Paleozoic strata show the highest values, followed by the Paleogene and Neogene strata, with the Quaternary strata showing the lowest thermal conductivity. Specifically, the Quaternary strata exhibit high insulation performance due to the low heat conduction efficiency. Despite relatively low geothermal gradients, the Paleogene and Neogene strata exhibit high heat conduction performance, facilitating the upward conduction of terrestrial heat flow. The geothermal reservoirs in the study area exhibit extensive distributions, considerable thicknesses, shallow burial depths, high temperatures, and high water yield properties, laying a solid thermal source foundation for the exploitation and utilization of geothermal resources [37]. The distributions of geothermal reservoirs in the study area are shown in Figure 1.

3. Materials and Methods

A total of 43 groups of geothermal fluid samples were collected from the study area, including 29 groups from the Neogene sandstone reservoirs and 14 groups from the karst reservoirs. The sampling sites are shown in Figure 1. The data of samples S8–S14 and S18–S19 were obtained from references [12,13], respectively.
The samples for water quality analysis from the study area were tested at the laboratory of the Groundwater, Mineral Water, and Environmental Monitoring Center of the Ministry of Land and Resources, supported by the Institute of Hydrogeology and Environmental Geology of the Chinese Academy of Geological Sciences. The concentrations of cations like K+, Na+, Ca2+, and Mg2+ in the samples were determined using an inductively coupled plasma—optical emission spectrometer (Thermo Fisher iCAP6300, Waltham, MA, USA), while the concentrations of anions like Cl, SO42−, NO3, and F were tested using a liquid-phase ion chromatograph. HCO3 and CO32− were measured using the double indicator titration method. The relative errors of all test analyses were controlled within 5%. Additionally, water temperatures and pH values were determined on site using a portable tester, the Easy Probe water quality analyzer, with resolution of 0.01. Hydrogen and oxygen isotopes were tested using a water isotope analyzer (Picaro2140-i) with a test accuracy of 0.1‰.
Concerning hydrochemical data, this study employed the Shukarev classification method to analyze the hydrochemical types of the samples [40] and software including IBM SPSS Statistics 26 (for correlation and principal component analyses), Aquachem 3.70 (for hydrochemical type analysis), and Excel to analyze the hydrochemical data.

4. Results

Table 1 shows the statistics of hydrochemical components in geothermal fluid samples from the study area. Specifically, samples from the Neogene sandstone reservoirs exhibited water temperatures ranging from 42 to 65 °C, pH values between 7.6 and 9.2, and total dissolved solids (TDS) concentrations primarily ranging from 3676 to 6788 mg/L. In contrast, samples from karst reservoirs manifested water temperatures ranging from 33 to 82 °C, pH values between 6.9 and 8.0, and TDS concentrations primarily ranging from 3314 to 7416 mg/L.
According to the Shukarev classification method, the data of cations and anions in all samples were projected onto the Piper diagram using the Aquachem software (Figure 2). Figure 3 and Figure 4 show the concentrations and percentages of major ions in geothermal fluid samples. The results indicate that the samples from sandstone reservoirs contained Na+ as major cations and Cl and SO42− as major anions, suggesting hydrochemical types of Na-Cl-SO4 (63%), Na-SO4-Cl, Na-Cl, and Na-SO4. In contrast, samples from karst reservoirs contained Na+ and Ca2+ as major cations and Cl and SO42− as major anions, suggesting hydrochemical types of Na-Ca-Cl-SO4 (50%), Ca-Mg-SO4, and Ca-Na-SO4-Cl.

5. Discussion

5.1. Hydrogeochemical Processes for the Formation of Major Ions

5.1.1. Dissolution–Precipitation Processes

The hydrogeochemical processes that form major ions in geothermal fluids primarily include the dissolution–precipitation processes of carbonate, sulfate, and silicate minerals. This study analyzed the dissolution–precipitation processes in geothermal fluids in the study area based on the ion ratios describing the relationships between different ions (Figure 5) [41]. The results show that the geothermal fluid samples from sandstone reservoirs were distributed near the line of 2Ca2+ = HCO3, leaning towards the Ca2+ side, suggesting that ions Ca2+ and HCO3 in geothermal fluids from sandstone reservoirs originated primarily from the dissolution of carbonate minerals like calcite and dolomite. In addition, leaning towards the Ca2+ side indicates that in addition to carbonates, silicate minerals like feldspar also experienced dissolution. In contrast, geothermal fluid samples from karst reservoirs showed much higher Ca2+ concentration compared to the HCO3 concentration, suggesting other major sources for Ca2+. The HCO3 concentration in the geothermal fluid samples from two types of reservoirs, especially karst reservoirs, decreased with an increase in the Ca2+ concentration. This is because in a high-temperature environment, high-concentration Ca2+ ions are prone to interact with HCO3 ions to form CaCO3 precipitates, leading to decreased HCO3 ions in geothermal fluids.
The Mg2+/Ca2+ ratios were all below 1 in the geothermal fluid samples from two types of reservoirs (Figure 5), indicating that magnesium-deficient minerals like calcite had minor impacts on the hydrochemical composition of deep geothermal fluids. Concerning the Na+/Ca2+ ratios, the geothermal fluid samples from sandstone reservoirs showed relatively high Na+/Ca2+ ratios, indicating that compared to carbonate dissolution, silicate dissolution showed more significant impacts on the hydrochemical composition of geothermal fluids in sandstone reservoirs. Moreover, the continuous increase in Na+/Ca2+ ratios suggests more intense carbonate precipitation. In contrast, the geothermal fluid samples from karst reservoirs exhibited relatively low Na+/Ca2+ ratios, near the line of Na+/Ca2+ = 1, indicating that silicate dissolution exerted minor impacts on the hydrochemical composition of geothermal fluids in karst reservoirs.
Using the (SO42− + HCO3) vs. (Mg2+ + Ca2+) diagram, this study analyzed the hydrogeochemical processes experienced by the principal chemical components of geothermal fluids. Figure 6 shows that the geothermal fluid samples from karst reservoirs exhibited Mg2+ + Ca2+/SO42− + HCO3 ratios close to 1, indicating that the precipitation–dissolution equilibrium of carbonate, silicate, and sulfate minerals in karst reservoirs jointly affected the formation of major ions in geothermal fluids. Combined with the finding that silicate dissolution exerted minor impacts on the hydrochemical composition of geothermal fluids in karst reservoirs, it further indicates that geothermal fluids in karst reservoirs were primarily affected by the precipitation and dissolution of carbonate and sulfate minerals. In contrast, the geothermal fluid samples from sandstone reservoirs manifested Mg2+ + Ca2+/SO42− + HCO3 ratios far below 1, with relatively stable Ca2+ + Mg2+ values, indicating that geothermal fluids in sandstone reservoirs were principally affected by the dissolution of carbonate and silicate minerals. The interrelationship between Na+ + K+ and Cl concentrations can be employed to analyze the possibility of a source of silicate mineral dissolution. Figure 6 illustrates that most geothermal fluid samples from karst reservoirs showed similar Na+ + K+ concentrations to Cl concentrations, indicating that the related ions originated primarily from halite dissolution. However, the geothermal fluid samples from sandstone reservoirs showed Na+ + K+ concentrations higher than Cl concentrations, reflecting that they were also affected by silicate mineral dissolution in addition to halite dissolution. In line with the earlier findings, silicate dissolution exerted minor impacts on the hydrochemical composition of geothermal fluids in karst reservoirs but significant impacts on that of geothermal fluids in sandstone reservoirs.

5.1.2. Cation Exchange Reactions

In the study of the hydrochemical evolutionary process of geothermal fluids, it is found that the ionic equivalent ratio rNa+/rCl can effectively reflect the extent of cation exchange. Figure 7 shows that the rNa+/rCl ratios of geothermal fluid samples from sandstone reservoirs all exceeded 1. With an increase in Cl concentration, they gradually decreased and tended to stabilize, suggesting intense cation exchange reactions. In contrast, the rNa+/rCl ratios of geothermal fluid samples from karst reservoirs were distributed near value 1, suggesting less intense cation exchange reactions. Additionally, Mg2+ in geothermal fluids of sandstone reservoirs also participated in cation exchange reactions, but its exchange intensity was lower than that of Na+. The exchange effect of Mg2+ was weaker in the geothermal fluids of karst reservoirs.
Moreover, chloro-alkaline indices CAI1 and CAI2 can be employed to illustrate the likelihood of cation exchange reactions [42]. They are calculated as follows:
CAI 1 = Cl ( Na + + K + ) / Cl
CAI 2 = Cl ( Na + + K + ) / ( SO 4 2 + HCO 3 + NO 3 + CO 3 2 )
When Na+ and K+ in geothermal fluids exchange with adsorbed Ca2+ and Mg2+, the values of CAI1 and CAI2 will be positive. Conversely, when Ca2+ and Mg2+ in geothermal fluids exchange with adsorbed Na+ and K+, the values of CAI1 and CAI2 will be negative. Moreover, higher absolute values of CAI1 and CAI2 suggest more significant impacts from cation exchange reactions [43]. Figure 8 illustrates that both indices for geothermal fluids in sandstone reservoirs were negative, with higher absolute values. In contrast, the two indices for geothermal fluids in karst reservoirs were primarily distributed near zero, indicating relatively minor impacts from cation exchange reactions. The results align with the previous analysis.

5.2. Circulation Characteristics of Geothermal Fluids

Stable isotopes in geothermal fluids can identify various physicochemical processes in geothermal systems, such as the origin of geothermal fluids, water–rock interactions, circulation depths, recharge elevations, and the mixing of shallow groundwater with geothermal water. In this study, O and H stable isotopes were employed to comprehensively analyze the origin, water–rock interactions, circulation characteristics, and mixing processes of geothermal fluids in sandstone and carbonate reservoirs in the study area.

5.2.1. Estimation of Reservoir Temperatures

Reservoir temperatures, a significant parameter for geothermal systems, play a critical role in the assessment, exploitation, and utilization of geothermal resources. However, it is difficult to obtain reservoir temperatures directly due to the deep burial depths of geothermal reservoirs. Moreover, single geothermal drilling can merely reveal local reservoir temperatures. Hence, geothermometers were employed to predict the approximate temperature ranges of reservoirs. The commonly used geothermometers include cationic and SiO2 geothermometers [28,44]. Cationic geothermometers can be used when geothermal fluids reach an equilibrium state. The Na-K-Mg equilibrium diagram can assist in determining the equilibrium state of geothermal fluids [45]. Figure 9 shows that the geothermal fluid samples from the karst reservoirs all fell within the zone of immature water at the bottom right corner of the equilibrium diagram. In contrast, the geothermal fluid samples from sandstone reservoirs all fell within the zone of partially equilibrated water. Therefore, the cationic geothermometers are inapplicable to the estimation of reservoir temperatures of geothermal fluids in karst reservoirs. Among the cationic geothermometers, the Na-K geothermometer applies only to geothermal water above 150 °C, especially in boreholes. The Na-K-Ca geothermometer is specifically used for calcium-rich geothermal water. The K-Mg geothermometer is suitable for low-temperature geothermal water, with estimated temperatures generally higher than the water temperatures at the wellheads. Therefore, this study calculated the reservoir temperatures of geothermal fluids in sandstone reservoirs based on the K-Mg geothermometer, and the reservoir temperatures of geothermal fluids in karst reservoirs based on the quartz and chalcedony geothermometers. The calculation equations are as follows [46]:
Quartz geothermometer (non-destructive):
T = 1309 5.19 lgSiO 2 273.15
Chalcedony geothermometer (non-destructive):
T = 1302 4.69 lgSiO 2 273.15
Cationic K-Mg geothermometer:
T = 4410 13.95 lg ( K ) 2 / Mg 273.15
Table 2 shows the estimation results of reservoir temperatures. The calculation results of SiO2 geothermometers are recommended for karst reservoirs. The measured temperatures of geothermal fluid samples from karst reservoirs ranged from 27 to 56 °C. However, the calculation results of the chalcedony geothermometer were generally lower than the measured temperatures at the wellheads, failing to conform to the actual conditions. Therefore, the calculation results (52.97–82.09 °C) of the quartz geothermometer were regarded as the reservoir temperatures of geothermal fluids in karst reservoirs. For sandstone reservoirs, the SiO2 (quartz and chalcedony) and cationic (K-Mg) geothermometers were used to calculate the reservoir temperatures of geothermal fluids. The measured temperatures of geothermal fluid samples from sandstone reservoirs ranged from 42 to 65 °C. The chalcedony geothermometer yielded too low calculated values, which are unrealistic. Since geothermal fluid samples from sandstone reservoirs were partially equilibrated water, major cations were in a relatively active state and failed to fully reflect the final temperatures of geothermal fluids. As a result, the cationic K-Mg geothermometer is inapplicable to sandstone reservoirs. Therefore, it is more appropriate to select the calculation results (60.54–85.88 °C) of the quartz geothermometer as the reservoir temperatures of geothermal fluids in sandstone reservoirs [41]. Table 2 shows the final temperature selection results, involving the measured temperatures (T (Measured)), quartz thermometer calculated temperatures (T(Quartz)), chalcedony geothermometer calculated temperatures (T(Chalcedony)), K-Mg geothermometer calculated temperatures (T(K-Mg)), Na-K geothermometer calculated temperatures (T(Na-K)), and Na-K-Ca geothermometer calculated temperatures (T(Na-K-Ca)).

5.2.2. Circulation Depths of Geothermal Fluids

The circulation depths of geothermal fluids are positively correlated with reservoir temperatures. This correlation can be described by the following equation [47]
H = (t1 − t2)/G + h
where H is the circulation depth of geothermal fluids; t1 is the reservoir temperature (°C); t2 is the average local temperature (°C), and h is the thickness of the constant temperature zone (m).
Table 3 shows that the circulation depths of geothermal fluids in karst and sandstone reservoirs ranged from 1107.28 to 1836.69 m and 1366.60 to 2102.29 m, respectively.

5.2.3. Fluid Sources

δ18O and δD values can serve as effective tracers to determine the recharge sources and elevations of geothermal fluids. Figure 10 shows the δ18O vs. δD plot of geothermal fluids in the study area. The geothermal fluid samples from sandstone reservoirs yielded δ18O values ranging from −7.4‰ to −10.2‰ and δD values ranging from −59‰ to −80.94‰, while those from karst reservoirs yielded δ18O values ranging from −8.0‰ to −10.6‰ and δD values ranging from −62.8‰ to −76‰. The δ18O and δD values of geothermal fluids in the study area were correlated with the global, Chinese, and local meteoric water lines (i.e., GMWL, CMWL, and LMWL). The equations for the GMWL and CMWL are δD = 8δ18O + 10 and δD = 7.9δ18O + 8.2, respectively. For LMWLs, the meteoric water lines of the East China monsoon region and the Yiyuan area were selected for a comparative study [48]. Figure 10 illustrates that the geothermal fluid samples from karst reservoirs fell near the MWLs, with some exhibiting oxygen isotope shifts, indicating that the geothermal fluids in karst reservoirs in the study area originated principally from meteoric water. However, most geothermal fluid samples from sandstone reservoirs manifested significant δ18O shifts, with some showing severe shifts. The dissolution–precipitation processes reveal that the significant impacts of silicate mineral dissolution on geothermal fluids in sandstone reservoirs might be due to isotopic exchange between geothermal fluids and oxygen-bearing rocks (limestones or silicate rocks), which increased the 18O concentration in groundwater, causing δ18O values to deviate from the MWLs.
According to the elevation effect principle of stable hydrogen and oxygen isotopes, δD and δ18O values decrease with an increase in groundwater recharge elevation [47]. Based on this, the recharge areas and elevations of geothermal fluids can be determined. Considering oxygen isotope shifts in geothermal fluids in the study area, δD was selected to calculate the recharge elevations of geothermal fluids [47].
H = Hr + (D − Dr)/grad D
where H is the elevation of the geothermal fluid recharge area (m); Hr is the surface elevation of the geothermal fluid sampling site (m); D is the δD value of meteoric water (−55.02‰); Dr is the δD value of geothermal fluids (‰), and grad D is the decreasing gradient with elevation (3‰/100 m).
The temperatures of geothermal fluid recharge sources can be estimated using hydrogen and oxygen isotopes. Due to oxygen isotope shifts in geothermal fluids in the study area, δD was also employed to calculate the temperatures of geothermal fluid recharge sources. The specific calculation equations are as follows [49]:
δD = 5.6T − 100
δD = 3T − 92
where T is the recharge source temperature (°C).
Table 4 shows that the geothermal fluid samples from sandstone reservoirs exhibited overall recharge elevations ranging from 651 to 1522 m and recharge source temperatures ranging from 4.81 to 9.16 °C. Specifically, their recharge elevations ranged from 1021 to 1378 m in the Ji’nan area, 1042 to 1541 m in the Liaocheng area, 1061 to 1522 m in the Dezhou area, 651 to 1353 m in the Dongying area, and 760 to 1262 m in the Binzhou area. Their recharge source temperatures ranged from 5.58 to 7.37 °C in the Ji’nan area, 4.81 to 7.37 °C in the Liaocheng area, 4.81 to 7.11 °C in the Dezhou area, 5.58 to 9.16 °C in the Dongying area, and 6.09 to 8.65 °C in the Binzhou area. In contrast, the geothermal fluid samples from karst reservoirs manifested overall recharge elevations ranging from 634 to 1526 m and recharge source temperatures ranging from 4.81 to 8.03 °C. Specifically, they showed a recharge elevation of 634 m and a recharge source temperature of 6.86 °C in the Ji’nan area. They showed recharge elevations ranging from 862 to 1526 m and recharge source temperatures ranging from 4.81 to 8.03 °C in the Ji’nan area. The topography of Shandong Province is centered around the Tai-Lu-Yi mountains, gradually descending towards the surrounding areas. The Tai-Lu-Yi mountains show elevations of 1545, 1108, and 1032 m, respectively, with the main peak of Mount Tai being the highest point in Shandong Province. The study area is primarily located to the northwest of the Tai-Lu-Yi mountains, highly coinciding with the recharge elevations of geothermal fluids in the study area, indicating that the estimation results are relatively reasonable. In summary, the primary recharge source for geothermal fluids in the study area is distributed near the Tai-Lu-Yi mountains to the southeast of the study area.

5.2.4. Mixing Processes of Geothermal Fluids

Using the silica-enthalpy mixing model and equations can eliminate the influence of cold water mixing and analyze the proportion of cold water mixing and the reservoir temperature before mixing [50].
S c x + S h 1 x = S s S i O 2 c x + S i O 2 h 1 x = S i O 2 s
where S and SiO2 are the enthalpy (J/g) and SiO2 concentration (mg/L) of water, respectively; c is the initial content for near-surface cold water; h is the initial content for hot water; s is the final content for spring water; and x is the mixing ratio of underground cold water. Table 5 shows the relationships between temperature, enthalpy, and SiO2 concentration.
The silica-enthalpy mixing model was employed to calculate the mixing ratios of cold water in geothermal fluids in sandstone and karst reservoirs in the study area. Figure 11 shows that the cold water mixing ratios in geothermal fluids of sandstone reservoirs ranged from 85% to 90%, resulting in corrected reservoir temperatures ranging from 150 to 200 °C. In contrast, Figure 12 shows that the cold water mixing ratios in geothermal fluids of karst reservoirs ranged from 89% to 96%, resulting in corrected reservoir temperatures ranging from 200 to 225 °C.

6. Conclusions

The geothermal fluids in the Neogene sandstone reservoirs show measured temperatures ranging from 42 to 65 °C, pH values from 7.6 to 9.2, and TDS concentrations from 3676 to 6788 mg/L, with a predominant hydrochemical type of Na-Cl-SO4. In contrast, the geothermal fluids in the karst reservoirs show measured temperatures ranging from 33 to 82 °C, pH values from 6.9 to 8.0, and TDS concentrations from 3314 to 7416 mg/L, with a primary hydrochemical type of Na-Ca-Cl-SO4. The hydrochemical composition of geothermal fluids in the karst reservoirs is primarily influenced by the precipitation–dissolution equilibrium of carbonate and sulfate minerals, with minor impacts of silicate dissolution. The hydrochemical composition of geothermal fluids in the sandstone reservoirs is primarily affected by the precipitation–dissolution processes of carbonate minerals and especially silicate minerals. Compared to those in the karst reservoirs, geothermal fluids in the sandstone reservoirs exhibit more intense cation exchange reactions. Hydrogen and oxygen isotope analyses indicate that geothermal fluids in the sandstone reservoirs have experienced significant oxygen isotope shifts. This aligns with the finding that the hydrochemical composition of geothermal fluids in the sandstone reservoirs is significantly influenced by the dissolution and precipitation of silicates. The isotopic exchange between geothermal fluids and oxygen-bearing rocks (limestones or silicate rocks) increases the 18O concentration in the geothermal fluids, leading to the deviation between δ18O values and the MWLs. In contrast, the hydrochemical composition of geothermal fluids in the karst reservoirs is less affected by the dissolution and precipitation of silicates, resulting in slight oxygen isotope shifts.
The calculations of reservoir temperatures using geothermometers reveal that the quartz geothermometer is more effective in calculating the temperatures of the karst and sandstone reservoirs, obtaining temperatures ranging from 52.97 to 82.09 °C and 60.54 to 85.88 °C, respectively. The geothermal fluids in the sandstone reservoirs manifest circulation depths ranging from 1366.60 to 2102.29 m and cold water mixing ratios from 85% to 90%. Accordingly, the reservoir temperatures corrected using the silica-enthalpy mixing model range from 150 to 200 °C. The geothermal fluids in the karst reservoirs display circulation depths ranging from 1107.28 to 1836.69 m and cold water mixing ratios ranging from 89% to 96%. Accordingly, the corrected reservoir temperatures range from 200 to 225 °C. The geothermal fluids in the sandstone reservoirs feature overall recharge elevations ranging from 651 to 1522 m and recharge source temperatures from 4.81 to 9.16 °C. In contrast, the geothermal fluids in karst reservoirs manifest overall recharge elevations ranging from 634 to 1526 m and recharge source temperatures from 4.81 to 8.03 °C. The primary recharge source for geothermal fluids in the study area is distributed near the Tai-Lu-Yi mountains to the southeast of the study area.
To exploit and utilize geothermal resources in the study area efficiently and sustainably, it is recommended to consider the geochemical differences in geothermal fluids across the composite geothermal reservoirs, ecological characteristics, and anthropogenic factors in the subsequent exploitation and utilization.

Author Contributions

Conceptualization, M.L. and Y.Q.; methodology, Y.Q.; software, Y.Q.; formal analysis, M.L.; investigation, Y.Q. and H.Z.; resources, W.Z.; data curation, Y.Q.; writing—original draft preparation, M.L.; writing—review and editing, M.L., Y.Q., and H.Z.; visualization, L.C.; supervision, Y.Q.; project administration, W.Z.; funding acquisition, W.Z. and M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Wei Zhang and Man Li, the research and engineering application project of heat transfer and control technology in heat storage (grant number [DJ-ZDXM-2022-11]) and the key R&D program of Henan Province (grant number [241111321000]).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

All the data have been presented in the tables and figures.

Conflicts of Interest

Author Long Chen was employed by the company Guiyang Engineering Corporation Limited, Guiyang. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Spittler, N.; Shafiei, E.; Davidsdottir, B.; Juliusson, E. Modelling geothermal resource utilization by incorporating resource dynamics, capacity expansion, and development costs. Energy 2020, 190, 116407. [Google Scholar] [CrossRef]
  2. Jalilinasrabady, S.; Itoi, R. Classification of Geothermal Energy Resources in Japan Applying Exergy Concept. Int. J. Energy Res. 2013, 37, 1842–1850. [Google Scholar] [CrossRef]
  3. Barbacki, A. Classification of geothermal resources in Poland by exergy analysis-Comparative study. Renew. Sustain. Energy Rev. 2012, 16, 123–128. [Google Scholar] [CrossRef]
  4. Bina, S.M.; Jalilinasrabady, S.; Fujii, H.; Pambudi, N.A. Classification of geothermal resources in Indonesia by applying exergy concept. Renew. Sustain. Energy Rev. 2018, 93, 499–506. [Google Scholar] [CrossRef]
  5. Górecki, W.; Sowizdzal, A.; Hajto, M.; Wachowicz-Pyzik, A. Atlases of geothermal waters and energy resources in Poland. Environ. Earth Sci. 2015, 74, 7487–7495. [Google Scholar] [CrossRef]
  6. Wang, G.L.; Lin, W.J. Main hydro-geothermal systems and their genetic models in China. Acta Geol. Sin. 2020, 94, 1923–1937. (In Chinese) [Google Scholar]
  7. Karlsdottir Marta, R.; Heinonen, J.; Palsson, H.; Palsson, O. Life cycle assessment of a geothermal combined heat and power plant based on high temperature utilization. Geothermics 2020, 84, 101727. [Google Scholar] [CrossRef]
  8. Jolie, E.; Scott, S.; Faulds, J.; Chambefort, I.; Axelsson, G.; Guti’errez-Negrín, L.C.; Regenspurg, S.; Ziegler, M.; Ayling, B.; Richter, A.; et al. Geological controls on geothermal resources for power generation. Nat. Rev. Earth Environ. 2021, 2, 324–339. [Google Scholar] [CrossRef]
  9. Mou, L.K. Study of Occurrence Characteristics and Formation Model of Geothermal Water in Shandong Province; Shandong University of Science and Technology: Qingdao, China, 2017. (In Chinese) [Google Scholar]
  10. Gao, Z.J.; Sun, Z.J.; Yang, Y.H.; Mou, L.K.; Cui, Y.C. Occurrence Characteristics and Hydrochemical Characteristics of Geothermal Water in Shandong Province. Sci. Technol. Eng. 2019, 19, 85–90. (In Chinese) [Google Scholar]
  11. Li, M.; Zhang, W. Characteristics and mechanisms of fluorine enrichment in the geothermal water of south central Shandong Province. Bull. Geol. Sci. Technol. 2024, 43, 36–47. (In Chinese) [Google Scholar]
  12. Zhang, B.J. Hydrogeochemical Characteristics and Formation Conditions of the Geothermal Water in Northwestern Shandong Province; China University of Geosciences: Beijing, China, 2011. (In Chinese) [Google Scholar]
  13. Cui, Y.; Kang, F.X.; Zhong, Z.N.; Yang, X.C.; Sui, H.B.; Zhao, Q. Gas Isotope Constraints on the Geothermal Heat Source Mechanism in Northwest Shandong Plain. Acta Geosci. Sin. 2023, 44, 93–106. [Google Scholar]
  14. Cui, R.; Wang, X.P.; Feng, B.; Liu, X.Y.; Feng, S.T.; Liu, S. Comparative analysis of the genesis models of different geothermal reservoirs in Chengning uplift area in northwest Shandong based on hydrochemical isotope technology. Carsologica Sin. 2023, 42, 969–981, 994. (In Chinese) [Google Scholar]
  15. Yang, X.C.; Kang, F.X.; Wang, X.P. Hydrochemical features of geothermal reservoir geotemperature field in sandstone porosity and enrichment mechanism of geothermal water:a case study of geothermal reservoir of Guantao Formation in the Lubei. Acta Geol. Sin. 2019, 93, 738–750. (In Chinese) [Google Scholar]
  16. Hu, C.P.; Wang, N.S. Thermal anomaly mechanism of carbonate rock thermal storage in the shallow buried areas of northwestern Jinan. Acta Geol. Sin. 2019, 93 (Suppl. 1), 178–183. (In Chinese) [Google Scholar]
  17. Bullen, T.D.; Krabbenhoft, D.P.; Kendall, C. Kinetic and mineralogic controls on the evolution of groundwater chemistry and 87Sr/86Sr in a sandy silicate aquifer, northern Wisconsin, USA. Geochim. Cosmochim. Acta 1996, 60, 1807–1821. [Google Scholar] [CrossRef]
  18. Peiffer, L.; Taran, Y.A.; Lounejeva, E.; Solis-Pichardo, G.; Rouwet, D.; Bemard-Romero, R.A. Tracing thermal aquifers of El Chichón volcano-hydrothermal system (México) with 87Sr/86Sr, Ca/Sr and REE. J. Volcanol. Geotherm. Res. 2011, 205, 55–66. [Google Scholar] [CrossRef]
  19. Huang, L.; Liu, J.; Zhang, F.; Dontsov, E.; Damjanac, B. Exploring the influence of rock inherent heterogeneity and grain size on hydraulic fracturing using discrete element modeling. Int. J. Solids Struct. 2019, 176, 207–220. [Google Scholar] [CrossRef]
  20. Temizel, E.H.; Gültekin, F.; Ersoy, A.F. (Eds.) Multi-isotopic (O, H, C, S, Sr, B, Li) characterization of waters in a low-enthalpy geothermal system in Havza (Samsun), Turkey. Geothermics 2021, 97, 102240. [Google Scholar] [CrossRef]
  21. Li, M.; Xing, L.X.; Wang, G.L. Distribution characteristics of fluorine in deep geothermal water in Jizhong Depression and its risk assessment and development utilization suggestions. Geol. China 2023, 50, 1857–1870. (In Chinese) [Google Scholar] [CrossRef]
  22. Tan, P.; Fu, S.; Huang, L.; Chen, Z.; Cao, J. Effects of orthogonal cleat structures on hydraulic fracture evolution behavior. Geoenergy Sci. Eng. 2024, 241, 213119. [Google Scholar] [CrossRef]
  23. Gültekin, F.; Hatipoglu, E.; Ersoy, A.F. Hydrogeochemistry, environment-al isotopes and the origin of the Hamamayagi-Ladik thermal spring (Samsun, Turkey). Environ. Earth Sci. 2011, 62, 1351–1360. [Google Scholar] [CrossRef]
  24. Sack, A.L.; Sharma, S. A multi-isotope approach for understanding sources of water, carbon and sulfur in natural springs of the Central Appalachian region. Environ. Earth Sci. 2014, 71, 4715–4724. [Google Scholar] [CrossRef]
  25. Guo, Q.H.; He, T.; Wu, Q.F.; Liu, M.L. Constraints of major ions and arsenic on the geological genesis of geothermal water: Insight from a com-parison between Xiong’an and Yangbajain, two hydrothermal systems in China. Appl. Geochem. 2020, 117, 104589. [Google Scholar] [CrossRef]
  26. Rahayudin, Y.; Kashiwaya, K.; Tada, Y.; Iskandar, I.; Koike, K.; Atmaja, R.W.; Herdianita, N.R. On the origin and evolution of geothermal fluids in the Patuha geothermal field, Indonesia based on geochemical and stable isotope data. Appl. Geochem. 2020, 114, 104530. [Google Scholar] [CrossRef]
  27. Fu, H.; Huang, L.; Hou, B.; Weng, D.; Guan, B.; Zhong, T.; Zhao, Y. Experimental and numerical investigation on interaction mechanism between hydraulic fracture and natural fracture. Rock Mech. Rock Eng. 2024, 57, 10571–10582. [Google Scholar] [CrossRef]
  28. Fournier, R.O. Chemical geothermometers and mixing models for geothermal systems. Geothermics 1977, 5, 41–50. [Google Scholar] [CrossRef]
  29. Giggenbach, W.F.; Gonfiantini, R.; Jangi, B.L.; Truesdell, A.H. Isotopic and chemical composition of Parbati valley geothermal discharges, north-west Himalaya, India. Geothermics 1983, 12, 199–222. [Google Scholar] [CrossRef]
  30. Fournier, R.O.; Potter, R.W. Magnesium correction to the Na-K-Ca chemical geothermometer. Geochim. Cosmochim. Acta 1979, 43, 1543–1550. [Google Scholar] [CrossRef]
  31. Sanjuan, B.; Millot, R.; Asmundsson, R.; Brach, M.; Giroud, N. Use of two new Na/Li geothermometric relationships for geothermal fluids in volcanic environments. Chem. Geol. 2014, 389, 60–81. [Google Scholar] [CrossRef]
  32. Kharaka, Y.K.; Lico, M.S.; Law, L.M. Chemical geothermometers applied to formation waters, Gulf of Mexico and California Basins (abstract). Am. Ass. Petrol. Geol. Bull. 1982, 66, 588. [Google Scholar]
  33. Michard, G. Behaviour of major elements and some trace elements (Li, Rb, Cs, Fe, Mn, W, F) in deep hot waters from granitic areas. Chem. Geol. 1990, 89, 117–134. [Google Scholar] [CrossRef]
  34. Verma, S.P.; Santoyo, E. New improved equations for Na/K, Na/Li and SiO2 geothermometers by outlier detection and rejection. J. Volcanol. Geoth. Res. 1997, 79, 9–23. [Google Scholar] [CrossRef]
  35. Kang, F.X.; Yang, X.C.; Wang, X.P.; Zheng, T.T.; Bai, T.; Liu, Z.T.; Sui, H.B. Hydrothermal Features of a Sandstone Geothermal Reservoir in the North Shandong Plain, China. Lithosphere 2022, 5, 1675798. [Google Scholar] [CrossRef]
  36. Tan, P.; Chen, Z.; Huang, L.; Zhao, Q.; Shao, S. Evaluation of the combined influence of geological layer property and in-situ stresses on fracture height growth for layered formations. Pet. Sci. 2024, 21, 3222–3236. [Google Scholar] [CrossRef]
  37. Wang, G.L. Geothermal Records of China; Science Press: Beijing, China, 2019. (In Chinese) [Google Scholar]
  38. Long, H.; Zhu, Q.; Tian, P.; Hu, W. Technologies and Applications of Geophysical Exploration in Deep Geothermal Resources in China; World Geothermal Congress: Melbourne, Australia, 2015. [Google Scholar]
  39. Zhang, L.; Chen, S.; Zhang, C. Geothermal power generation in China: Status and prospects. Energy Sci. Eng. 2019, 7, 1428–1450. [Google Scholar] [CrossRef]
  40. Gao, Z.; Chen, C. The classification method of water chemical types based on the principle of Kurllov’s Formula and Shoka Lev classification. Ground Water 2018, 4, 6–11. [Google Scholar]
  41. Xiao, W. Formation Conditions and Hydrogeochemical Characteristics of the Geothermal Water in Typical Coastal Geothermal Field with Deep Faults, Guangdong Province; China University of Geosciences: Wuhan, China, 2018. (In Chinese) [Google Scholar]
  42. Scholler, H. Qualitative evaluation of groundwater resource: Methods and techniques of groundwater investigation and development. Water Res. 1967, 33, 4452. [Google Scholar]
  43. Mao, R.; Guo, H.; Jia, Y.; Jiang, Y.; Cao, Y.; Zhao, W.; Wang, Z. Distribution Characteristics and Genesis of Fluoride Groundwater in the Hetao Basin, Inner Mongolia. Earth Sci. Front. 2016, 23, 260–268, (In Chinese with English Abstract). [Google Scholar]
  44. Nicholson, K. Geothermal Fluids; Springer: Berlin/Heidelberg, Germany, 1993; pp. 141–149. [Google Scholar] [CrossRef]
  45. Giggenbach, W.F. Geothermal solute equilibria. Derivation of Na-K-Mg-Ca geoindicators. Geochim. Cosmochim. Acta 1988, 52, 2749–2765. [Google Scholar] [CrossRef]
  46. DZ/T 0331-2020; Specification for Estimation and Evaluation of Geothermal Resources. Ministry of Natural Resources, People’s Republic of China: Beijing, China; Geology Press: Beijing, China, 2020. (In Chinese)
  47. Sun, Z.X.; Li, X.L.; Shi, W.J. Isotopic hydrogeochemistry of medium-low temperature geothermal water in Jiangxi. J. East. China Inst. Geol. 1992, 15, 243–248. (In Chinese) [Google Scholar]
  48. Liu, J.R.; Song, X.F. Characteristics of δ18O in precipitation over Eastern Monsoon China and the water vapor sources. Chin. Sci. Bull. 2009, 54, 3521–3531. [Google Scholar]
  49. Wang, H.C. Introduction to Isotope Hydrogeology; Geological Publishing House: Beijing, China, 1991. (In Chinese) [Google Scholar]
  50. Truesdell, A.H.; Nathenson, M.; Rye, R.O. The effects of subsurface boiling and dilution on the isotopic compositions of Yellowstone thermal waters. J. Geophys. Res. Part B Solid Earth 1977, 82, 3694–3704. [Google Scholar] [CrossRef]
Figure 1. Map showing the distribution of fluorine in geothermal fluids and sampling sites in the study area, (a) conceptual model of the geothermal system in the study area (b) [9], and map showing the distribution of geothermal resources in China (c) [38,39].
Figure 1. Map showing the distribution of fluorine in geothermal fluids and sampling sites in the study area, (a) conceptual model of the geothermal system in the study area (b) [9], and map showing the distribution of geothermal resources in China (c) [38,39].
Sustainability 17 02252 g001
Figure 2. Piper diagram of geothermal fluid samples from the study area.
Figure 2. Piper diagram of geothermal fluid samples from the study area.
Sustainability 17 02252 g002
Figure 3. Schoeller diagram of geothermal fluid samples from the study area.
Figure 3. Schoeller diagram of geothermal fluid samples from the study area.
Sustainability 17 02252 g003
Figure 4. Bar chart showing the percentages of major ions in the geothermal fluid samples from the study area.
Figure 4. Bar chart showing the percentages of major ions in the geothermal fluid samples from the study area.
Sustainability 17 02252 g004
Figure 5. HCO3 vs. Ca2+ and Na+/Ca2+ vs. Mg2+/Ca2+ concentrations in geothermal fluid samples.
Figure 5. HCO3 vs. Ca2+ and Na+/Ca2+ vs. Mg2+/Ca2+ concentrations in geothermal fluid samples.
Sustainability 17 02252 g005
Figure 6. (SO42− + HCO3) vs. (Mg2+ + Ca2+) and (Na+ + K+) vs. Cl concentrations in geothermal fluid samples.
Figure 6. (SO42− + HCO3) vs. (Mg2+ + Ca2+) and (Na+ + K+) vs. Cl concentrations in geothermal fluid samples.
Sustainability 17 02252 g006
Figure 7. Cl vs. rNa+/r(Na+ + Ca2+ + Mg2+) and rNa+/rCl of geothermal fluid samples.
Figure 7. Cl vs. rNa+/r(Na+ + Ca2+ + Mg2+) and rNa+/rCl of geothermal fluid samples.
Sustainability 17 02252 g007
Figure 8. CAI1 vs. CAI2 of geothermal fluid samples.
Figure 8. CAI1 vs. CAI2 of geothermal fluid samples.
Sustainability 17 02252 g008
Figure 9. Na-K-Mg equilibrium diagram of geothermal fluids.
Figure 9. Na-K-Mg equilibrium diagram of geothermal fluids.
Sustainability 17 02252 g009
Figure 10. δ18O vs. δD plot of geothermal fluids.
Figure 10. δ18O vs. δD plot of geothermal fluids.
Sustainability 17 02252 g010
Figure 11. Results derived from the silica-enthalpy mixing model for karst reservoirs.
Figure 11. Results derived from the silica-enthalpy mixing model for karst reservoirs.
Sustainability 17 02252 g011
Figure 12. Results derived from the silica-enthalpy mixing model for sandstone reservoirs.
Figure 12. Results derived from the silica-enthalpy mixing model for sandstone reservoirs.
Sustainability 17 02252 g012
Table 1. Statistics of hydrochemical components in geothermal fluid samples from the study area.
Table 1. Statistics of hydrochemical components in geothermal fluid samples from the study area.
Reservoir Type pHDepth
(m)
T (°C)TDS
(mg/L)
Concentrations of Major Ions (mg/L)
KNaCaMgSO4HCO3ClSrF
Sandstone reservoirMinimum7.601050.0042.003676.035.201037.5014.568.83631.64134.24417.410.001.05
Maximum9.201649.0265.006788.4749.002400.00280.5665.002734.35369.913578.6810.887.50
Average8.031416.5455.145079.1612.971675.42136.7926.001260.83221.741829.105.661.66
Karst reservoirMinimum7.10576.9033.003314.9118.50152.50609.40124.461626.94103.94173.4011.752.60
Maximum8.002400.0056.007416.2680.001450.00868.30159.572337.87176.692393.8117.504.50
Average7.561204.8543.985307.0346.11805.39727.57142.292030.42150.151374.2314.873.39
Table 2. Calculated reservoir temperatures in the study area.
Table 2. Calculated reservoir temperatures in the study area.
Reservoir TypeSample No.AreaT
(Measured)
(°C)
T
(Quartz)
(°C)
T
(Chalcedony)
(°C)
T
(K-Mg)
(°C)
T
(Na-K)
(°C)
T
(Na-K-Ca)
(°C)
Selected Temperature (°C)
Karst reservoirTR1Liaocheng56.0082.0950.8876.72180.78105.0882.09
TR2Liaocheng50.2082.0950.8879.07190.41106.0582.09
TR3Liaocheng52.0078.0146.5976.79186.73102.6578.01
TR4Dezhou/76.5845.0952.59234.1250.4476.58
TR5Dezhou27.0067.0635.1355.54226.6656.5867.06
TR6Ji’nan/52.9720.5466.49159.7883.7452.97
TR7Ji’nan41.8076.5845.0985.23170.80117.2476.58
TR8Ji’nan36.2061.5329.3863.93159.6680.8161.53
TR9Ji’nan33.0059.5327.3162.22154.0179.4959.53
TR10Ji’nan40.6068.7836.9279.22156.39108.8868.78
TR11Ji’nan42.0073.6041.9682.18163.99113.0573.60
TR12Ji’nan55.0075.1143.5558.81217.8862.5975.11
TR13Ji’nan43.0075.1143.5555.43220.2056.4975.11
TR14Ji’nan39.9073.6041.9681.47163.99111.6473.60
Sandstone reservoirSR1Liaocheng/75.1143.5557.3760.7280.0575.11
SR2Liaocheng/75.1143.5551.7570.1872.8775.11
SR3Liaocheng52.0067.0635.1353.0263.8575.6067.06
SR4Liaocheng50.0082.0950.8851.1059.2879.6882.09
SR5Liaocheng55.0082.0950.8856.3361.9076.6382.09
SR6Liaocheng53.0085.8854.8972.9769.2994.0985.88
SR7Liaocheng54.00//64.4568.1778.39/
SR8Dezhou62.5084.6453.5970.3072.0294.2884.64
SR9Dezhou/84.6453.5967.3171.2093.2584.64
SR10Dezhou55.0079.4148.0662.7362.3187.7279.41
SR11Dezhou58.5079.4148.0663.9965.1989.7679.41
SR12Dezhou56.0080.7649.4964.3664.5789.8580.76
SR13Dezhou58.0082.0950.8865.5258.7684.7782.09
SR14Dezhou56.0079.4148.0665.9759.6686.6879.41
SR15Dezhou/84.6453.5955.8152.9379.9484.64
SR16Dezhou51.0079.4148.0663.6664.7891.8379.41
SR17Dezhou47.0078.0146.5955.1048.6577.3578.01
SR18Dezhou58.00//61.9368.4991.46/
SR19Dezhou59.00//63.7563.0988.50/
SR20Binzhou/76.5845.0951.2142.9771.7676.58
SR21Binzhou/76.5845.0954.5945.0074.2476.58
SR22Binzhou/60.5428.3635.8538.4863.5060.54
SR23Ji’nan54.0081.0849.8364.8267.5289.7981.08
SR24Ji’nan55.0085.8854.8966.0368.0090.2585.88
SR25Ji’nan56.0072.0440.3352.0749.2876.4272.04
SR26Dongying65.0082.0950.8865.2463.7787.6382.09
SR27Dongying60.0076.5845.0965.0160.7686.3376.58
SR28Dongying54.0078.0146.5967.1766.8090.1078.01
Table 3. Estimated circulation depths of geothermal fluids in the reservoirs.
Table 3. Estimated circulation depths of geothermal fluids in the reservoirs.
Reservoir TypeAreaSample No.H (m)t1 (°C)t2 (°C)G (°C/100 m)h (m)
Karst reservoirLiaochengTR11742.2182.0913.2420
LiaochengTR21742.2182.0913.2420
LiaochengTR31640.3578.0113.2420
DezhouTR41297.6876.5812.7520
DezhouTR51107.2867.0612.7520
Ji’nanTR61161.8952.97133.520
Ji’nanTR71836.6976.58133.520
Ji’nanTR81406.4561.53133.520
Ji’nanTR91349.3959.53133.520
Ji’nanTR101613.7168.78133.520
Ji’nanTR111751.5073.60133.520
Ji’nanTR121794.7075.11133.520
Ji’nanTR131794.7075.11133.520
Ji’nanTR141751.5073.60133.520
Sandstone reservoirLiaochengSR11567.8775.1113.2420
LiaochengSR21567.8775.1113.2420
LiaochengSR31366.6067.0613.2420
LiaochengSR41742.2182.0913.2420
LiaochengSR51742.2182.0913.2420
LiaochengSR61837.0185.8813.2420
LiaochengSR7//13.2420
DezhouSR82075.5784.6412.73.520
DezhouSR92075.5784.6412.73.520
DezhouSR101925.9079.4112.73.520
DezhouSR111925.9079.4112.73.520
DezhouSR121964.6980.7612.73.520
DezhouSR132002.5282.0912.73.520
DezhouSR141925.9079.4112.73.520
DezhouSR152075.5784.6412.73.520
DezhouSR161925.9079.4112.73.520
DezhouSR171886.1278.0112.73.520
DezhouSR18//12.73.520
DezhouSR19//12.73.520
BinzhouSR201853.8476.5812.43.520
BinzhouSR211853.8476.5812.43.520
BinzhouSR221395.3760.5412.43.520
Ji’nanSR231965.2881.08133.520
Ji’nanSR242102.2985.88133.520
Ji’nanSR251707.0072.04133.520
DongyingSR262016.8182.0912.23.520
DongyingSR271859.5576.5812.23.520
DongyingSR281900.4078.0112.23.520
Table 4. Estimation results of δD and δ18O values and elevations for geothermal fluid samples from the study area.
Table 4. Estimation results of δD and δ18O values and elevations for geothermal fluid samples from the study area.
Reservoir TypeAreaδDVSMOW
(‰)
δ18OVSMOW
(‰)
Elevation of the Sampling SiteElevation
(D)
Elevation
(O)
Recharge Source Temperature
Karst reservoirLiaocheng−68−9.434634.00649.386.86
Ji’nan−72−9.834834.00803.235.83
−74−9.625925.00717.315.32
−67.4−9291099.001067.467.01
−63.4−8.821891.00982.548.03
−62.8−9.222862.001137.388.19
−68.1−9.6251130.001294.236.83
−76−10.5261526.001641.384.81
−71.3−10.1221287.001483.546.01
−68.2−9.7211131.001328.696.81
−69.4−9.8251195.001371.156.50
−68.3−8.7301145.00953.086.78
−63.4−834904.00687.858.03
−72−9.8341334.001380.155.83
−70−9.6341234.001303.236.35
−70−9.6251225.001294.236.35
−71−9.3251275.001178.856.09
−68−9.7341134.001341.696.86
−69−9.3251175.001178.856.60
−73−9.8251375.001371.155.58
−73−10.6251375.001678.855.58
−76−10.1251525.001486.544.81
Sandstone reservoirJi’nan−73−10.1281378.001489.545.58
−68−9.2171117.001132.386.86
−70.39−8.53171236.50874.696.25
−66−8.9211021.001021.007.37
−70.39−8.53151234.50872.696.25
Liaocheng−71−9.4361286.001228.316.09
−69−9.4321182.001224.316.60
−70−9.4341234.001226.316.35
−71−9.7331283.001340.696.09
−73−9.4421392.001234.315.58
−66−8.8421042.001003.547.37
−70−9.5341234.001264.776.35
−76−9.9411541.001425.624.81
−72−9.4401340.001232.315.83
−73−9.7331383.001340.695.58
−72−9.5311331.001261.775.83
−71−10.2361286.001536.006.09
−70.37−8.03371255.50702.386.25
Dezhou−75.05−9.59221474.501287.385.05
−73−10241374.001447.085.58
−74−10221422.001445.085.32
−75−10261476.001449.085.07
−74−9.9261426.001410.625.32
−68−8.6271127.00911.626.86
−67−9.2111061.001126.387.11
−75−9.7261476.001333.695.07
−70−8.8221222.00983.546.35
−76−9.7221522.001329.694.81
−71−9.3111261.001164.856.09
−72−9.4261326.001218.315.83
−74−9.4201420.001212.315.32
−70−8.5131213.00859.156.35
−73−8.7181368.00941.085.58
−70−8.7231223.00946.086.35
Dongying−59−8.71651.00924.089.16
−66−831003.00656.857.37
−65−8.21951.00731.777.63
−68−7.441104.00427.086.86
−64−8.11901.00693.317.88
−68−8.211101.00731.776.86
−73−9.431353.001195.315.58
−67−7.461056.00429.087.11
Binzhou−71−9.3121262.001165.856.09
−64−7.512912.00473.547.88
−66−9.1131013.001089.927.37
−61−7.810760.00586.928.65
Weifang−68−9531153.001091.466.86
Table 5. Relationships between temperature, enthalpy, and SiO2 concentration.
Table 5. Relationships between temperature, enthalpy, and SiO2 concentration.
T (°C)Enthalpy (J/g)SiO2
(mg/L)
T (°C)Enthalpy (J/g)SiO2
(mg/L)
T (°C)Enthalpy
(J/g)
SiO2
(mg/L)
5050.013.5150151.0125.0250259.2486.0
7575.026.6175177.0185.0275289.0614.0
100100.148.0200203.6265.0300321.0692.0
125125.180.0225230.9365.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Qiao, Y.; Li, M.; Chen, L.; Zhang, H.; Zhang, W. Exploring Geochemical Characteristics of Composite Geothermal Reservoirs for Sustainable Utilization: A Case Study of the Northwestern Shandong Geothermal Area in China. Sustainability 2025, 17, 2252. https://doi.org/10.3390/su17052252

AMA Style

Qiao Y, Li M, Chen L, Zhang H, Zhang W. Exploring Geochemical Characteristics of Composite Geothermal Reservoirs for Sustainable Utilization: A Case Study of the Northwestern Shandong Geothermal Area in China. Sustainability. 2025; 17(5):2252. https://doi.org/10.3390/su17052252

Chicago/Turabian Style

Qiao, Yong, Man Li, Long Chen, Hanxiong Zhang, and Wei Zhang. 2025. "Exploring Geochemical Characteristics of Composite Geothermal Reservoirs for Sustainable Utilization: A Case Study of the Northwestern Shandong Geothermal Area in China" Sustainability 17, no. 5: 2252. https://doi.org/10.3390/su17052252

APA Style

Qiao, Y., Li, M., Chen, L., Zhang, H., & Zhang, W. (2025). Exploring Geochemical Characteristics of Composite Geothermal Reservoirs for Sustainable Utilization: A Case Study of the Northwestern Shandong Geothermal Area in China. Sustainability, 17(5), 2252. https://doi.org/10.3390/su17052252

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