Next Article in Journal
Using Animated Videos to Enhance Vocabulary Learning at the Noble Private Technical Institute (NPTI) in Northern Iraq/Erbil
Previous Article in Journal
Influence of Freeze–Thaw Cycles on Physical and Mechanical Properties of Cement-Treated Silty Sand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization Design of Multi-Factor Combination for Power Generation from an Enhanced Geothermal System by Sensitivity Analysis and Orthogonal Test at Qiabuqia Geothermal Area

1
Powerchina HuaDong Engineering Corporation Limited, Hangzhou 311122, China
2
Key Laboratory of Efficient Utilization of Low and Medium Grade Energy (Tianjin University), MOE, Tianjin 300350, China
3
College of Water Resources Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(12), 7001; https://doi.org/10.3390/su14127001
Submission received: 24 April 2022 / Revised: 25 May 2022 / Accepted: 2 June 2022 / Published: 8 June 2022

Abstract

:
In order to explore the optimal mining strategy of a fractured Enhanced Geothermal System (EGS) reservoir, we numerically investigated the influence of seven factors on heat production and conducted an optimization analysis of a multi-factor and multi-level combination by an orthogonal test based on the geological data at the Qiabuqia geothermal field. Seven factors were considered, including five reservoir factors (fracture spacing, fracture permeability, fracture permeability anisotropy, matrix permeability, and heat conductivity) and two operation factors (injected section length and injection rate). The results show that injection rate and fracture permeability have the greatest influence on production performance. Different factor combinations have a great influence on the productivity. The multi-factor and multi-level combination optimization is needed, and the optimization scheme of the EGS can be achieved through the orthogonal test and range analysis. The order of influence degree on the power generation is injection rate > fracture permeability > fracture permeability anisotropy > injected section length > matrix permeability > fracture spacing > heat conductivity. The order of influence degree on the coefficient of performance of the EGS is fracture permeability > injection rate > injected section length > fracture permeability anisotropy > matrix permeability > fracture spacing > heat conductivity. For reservoir stimulation, the stratum with dense natural fractures should be selected as the target EGS reservoir. It is not advisable to acidify the EGS reservoir too much to widen the apertures of the natural fractures. Fracture permeability anisotropy will increase pump energy consumption, but this adverse effect can be greatly reduced if the other parameters are well matched. Matrix permeability and heat conductivity may not be used as indicators in selecting a target reservoir. For project operation, the injected section length should be as long as possible. The injection rate plays a major role in all factors. Special attention should be paid to the value of the injection rate, which should not be too large. The appropriate injection temperature should be determined in accordance with the water source condition and the engineering requirement. If a commercial rate (100 kg/s) is to be obtained, the permeability of the reservoir fracture network needs to be stimulated to be higher. Meanwhile, in order to ensure that the production temperature is both high and stable, it is necessary to further increase the volume of the EGS reservoir.

1. Introduction

In 2020, the Chinese government announced that China’s carbon emissions would peak by 2030 and that China would strive to achieve carbon neutrality by 2060. Large-scale non-carbon energy is urgently needed to improve China’s energy structure. Geothermal energy is a competitive clean and renewable energy in coping with global climate change and emission reduction [1]. Hot Dry Rock (HDR) resources denote high temperature geothermal energy occurring in metamorphic or crystalline rock with low permeability and porosity. Its temperature ranges between 150 °C and 650 °C [1]. The first EGS project was prototyped at the Los Alamos National Laboratory at Fenton Hill in the 1970s, and subsequently, more than a dozen research and industrial projects have been carried out in various countries [1,2]. A systematic review of over 30 worldwide EGS projects can be found in Li et al. [2]. For the United States, the HDR resource reserve within a 3~10 km depth amounts to 14 × 106 EJ [1]. Located at the junction of three tectonic plates with a vast territory, China possesses abundant HDR geothermal energy. It is estimated that the HDR resource reserve within a 3~10 km depth in China’s mainland amounts to 20.9 × 106 EJ, which is larger than that of the United States [3]. If 2% of the reserve was exploited, it would be equivalent to 4400 times China’s total energy consumption in 2010 [3]. Since 2015, Chinese provincial governments have been actively investing in the exploration of HDR resources. Many potential HDR targets have been found, mainly including those in the Songliao Basin, southern Hainan, Wendeng, Datong, Zhangzhou, the Gonghe Basin, and the Subei Basin [4]. The Gonghe Basin is expected to be the first HDR pilot in China.
Although the HDR resource reserve is huge, it must be artificially stimulated to obtain economical geothermal fluids due to the low natural permeability of the reservoir. The technology integration for developing HDR energy is called the Enhanced Geothermal System (EGS) [1]. An EGS involves the creation of a fracture system in the target geological formation used for the injected geothermal fluid circulation. The geothermal structure of the target formation determines the upper limit of the HDR resource reserve. The characteristics of the target formation determine the property of the created EGS reservoir after stimulation. The operation conditions determine the amount of thermal energy actually extracted from the EGS subsurface system. Therefore, the scientific combination of the target reservoir factors and the operation factors has a great significance in the optimization of the EGS mining and in reducing the risks.
Hydraulic, chemical, and thermal stimulation techniques are currently available to stimulate existing fracture networks and/or create new fractures [2]. The cost of one reservoir treatment is approximately USD 0.5 M [4,5]. Multiple treatments can be more expensive. Numerical simulation has become an important research method for analyzing the production performance of EGS reservoirs. The numerical simulation needs to characterize and represent the subsurface fracture system. Natural fracture networks significantly control the hydrodynamic and thermodynamic behaviors in EGS reservoirs. The fracture–matrix interactions are treated with dual-continuum approaches, including equivalent porous media (EPM), the double-porosity method (DPM) [6], and the more rigorous multiple interacting continua (MINC) [7]. EPM is mainly used to model a densely fractured reservoir in which the fracture density is high and the fracture spacing is small; normally, the average fracture spacing is less than 2~3 m [8]. The DPM method is more reasonable when the average fracture spacing is higher than 10 m [8]. The MINC yields intermediate behaviors in between a porous medium and a single fracture. Suzuki et al. [9] conducted fractional diffusion modeling of heat transfer in porous and fractured media. They concluded that the porous model characterized the result of the MINC for the smaller fracture spacing (<5 m), while the single-fracture model agreed with the results of the MINC for larger fracture spacing (>20 m). In addition, the MINC method is able to describe gradients of pressures, temperatures, or concentrations inside the matrix by subdividing individual matrix blocks. It can be used to model the coupled hydrologic-thermal processes within the reservoirs. Fracture networks are the dominate channels for water to transfer mass and heat in an HDR reservoir. The discrete fracture network (DFN) model could simulate the fracture–matrix interactions based on the detailed fracture orientation, size, and spacing and other mechanical properties. Unfortunately, most of the measurement data are obtained from limited samples and do not accurately describe a fractured medium.
Parameter optimizations are the key to maximizing the heat extraction efficiency of the EGS. Many scholars have studied the conditions affecting the heat-production performance of the fractured EGS reservoir. Gong et al. [10] proposed a three-dimensional thermo-hydraulic model with multiple fracturing horizontal wells to evaluate the efficiency of the heat extraction in the EGS. Zhou et al. [11] studied the influence of randomly distributed fracture apertures in a fracture network on heat production from an EGS. Pandey et al. [12] researched the influence of fracture alteration on the heat production through the single-fracture THM model. The fractured EGS reservoir contains three porous media: rock matrix, natural fractures, and hydraulic fractures. Many EGS pilot production cases have revealed that the thermal reservoir is a highly heterogeneous fractured medium [13]. Compared to the fluid flow in a uniformly distributed porous medium, a fractured medium has an obvious preferential flow direction. Guo et al. [14] developed a 3D THM model with a 2D fracture layer to study the heterogeneity of the fractures on the heat-production efficiency and the service life of an EGS. Huang et al. [15] established an equivalent continuum 3D THM model to consider the heterogeneous permeability and porosity. Guo et al. [16] considered the THM coupled processes on the heterogeneity of fracture aperture and studied the influence of aperture heterogeneity on the flow behavior of fluid within a single-fracture system.
Although the numerical simulations of the EGS reservoirs have made great progress, there is still a lack of systematic study on the multi-factor combination conditions affecting the heat-production performance. These preliminary studies do not compare every condition and do not uncover the order of relative importance of the conditions affecting the heat production. Most productivity-optimization studies are focused on the effect of a single factor. Asai et al. [17] analyzed the effect of seven different flow schemes on the rate of heat production and obtained the optimum case when considering the rise in energy demand. Patterson et al. [18] optimized geothermal production in fractured rock reservoirs with a combination of simple analytical models, while considering reservoir structural uncertainty. Wu et al. [19] established a three-dimensional hydraulic fracturing model and used heat extraction as an evaluation indicator to optimize fracturing spacing and geothermal mining. Different reservoir factors and operation factors are involved in the optimization design of geothermal projects. Decision making in geothermal projects requires consideration of an optimal combination of different factors, rather than a single factor. There is little research on multi-factor combination analysis to optimize HDR productivity based on a comprehensive production index. The orthogonal test method is a kind of method to analyze the multi-factor and multi-level influence on system objectives and was proposed by Fisher in 1920 [20]. The orthogonal test method is a kind of approach that can compare every condition and rank in the order of relative importance of every condition. This method has been widely used in test design in science, engineering, and business [21,22]. Xie and Wang [23] conducted a performance optimization of a pinnate horizontal well in geothermal energy utilization with an orthogonal test. Yu et al. [24] proposed a modified zipper fracturing in an EGS reservoir and analyzed heat extraction optimization via an orthogonal test.
EGS projects are expensive and risky. All the factors of the project should be optimized as much as possible before operation. This makes engineering sense when improving production performance, reducing investment, and maintaining longer reservoir life. In this paper, the primary objective is to answer the following questions. (1) Which factors have the greatest impact on the productivity of an EGS reservoir? (2) How do these factors affect productivity over a long time, given that the EGS project is expected to serve for decades? (3) The EGS operation involves many factors and needs to meet several production indicators. The effect of different factors on the indicators may be the opposite. How can the combination of multi-factors be optimized to make all the indicators reach the best level? (4) Which factors should be considered in target reservoir selection and stimulation? In order to answer the above questions and explore the optimal mining strategy of a fractured EGS reservoir, we numerically investigated the influence of seven single factors on heat production and conducted an optimization analysis of multi-factor and multi-level combinations by orthogonal test based on the geological data at the Qiabuqia geothermal field. The seven factors include five reservoir factors (fracture spacing, fracture permeability, fracture permeability anisotropy, matrix permeability, and heat conductivity) and two operation factors (injected section length and injection rate). The results could provide a scientific guidance for future fracturing design and reservoir operation at the Qiabuqia geothermal field.

2. Methodology

In this study, the geological and geothermal characteristics of the study area are studied first. Then, a three-dimensional hydraulic-thermal coupled numerical model is established for the fractured EGS reservoir using the MINC method. Adopting the selected reservoir factors and operation factors, the sensitivity of a single factor to the EGS productivity is analyzed by using a control variable method. According to the results of a single-factor sensitivity analysis, the most influential factors are screened out. Subsequently, the effects of a multi-factor and multi-level combination on EGS productivity are analyzed. Through the orthogonal test and range analysis, the order of influence degree of each factor on each performance indicator is determined. Finally, we conclude the optimal multi-factor and multi-level combination for EGS mining at the Qiabuqia area. The methodological flowchart is illustrated in Figure 1.

2.1. Governing Equation of Hydraulic-Thermal Coupled Numerical Model

TOUGH2 software was adopted for numerical simulation in this study. It is a well-accepted numerical simulation program for the multi-dimensional fluid and heat flows of multiphase, multicomponent fluid mixtures in porous and fractured media. The TOUGH2-EOS1 codes use the integral finite difference method to solve the conservation equations of mass, momentum, and energy, and their accuracy and reliability have been widely proven [8,9,25,26]. The theoretical background of the software is provided by [25]. In this study, water is used as the working fluid through the EGS reservoir. The mass–energy balance equation during the simulation can be written as:
d d t V n M κ d V n = Γ n F κ n d Γ n + V n q κ d V n ,
where the quantity M represents mass or energy per volume; t is time; κ denotes mass components; Vn is an arbitrary subdomain of the flow system; F denotes mass or heat flux; q denotes sinks and sources; and n is a normal vector on surface element dΓn, pointing inward into Vn.
The mass accumulation term is:
M κ = φ Σ S β ρ β X β κ ,
where φ denotes the porosity of the rock; Sβ is the saturation of phase β (i.e., the fraction of the pore volume occupied by phase β); ρβ is the density of phase β; and X β κ is the mass fraction of component κ present in phase β.
The heat accumulation term is:
M N K + 1 = ( 1 φ ) ρ R C R T + φ Σ β S β ρ β u β ,
where ρR and CR are, respectively, the density and specific heat of the rock; T is the temperature of the rock; and uβ represents specific internal energy in phase β.
The mass flux is given by:
F κ = Σ β X β κ F β ,
The heat flux is calculated by:
F N K + 1 = λ T + Σ β h β F β ,
where λ is the thermal conductivity and hβ is the specific enthalpy in phase β.
The governing equation of the β-phase Darcy flow in a fractured reservoir is:
F β = ρ β v β = k k r β ρ β μ β ( P β ρ β g ) ,
where vβ is the Darcy velocity (volume flux) in phase β; k is the absolute permeability; k is the permeability relative to phase β; and μβ denotes the dynamic coefficient of viscosity.

2.2. Orthogonal Test Method

For the analysis of multi-factor and multi-level combinations, the orthogonal test method can represent complete tests through very limited combination tests and greatly reduces the workload [20,21,22]. It considers the characteristics of uniform dispersion and uniform comparability. Taking 3 factors (A, B, C) and 3 levels as an example, each factor has 3 levels (a1, a2, a3), (b1, b2, b3) and (c1, c2, c3), as shown in Table 1.
A complete test would conduct 3 × 3 × 3 = 27 trials in total, while only 9 trials are required according to the orthogonal test. The specific tests are listed in Table 2.

2.3. Range Analysis

Based on the results of the orthogonal test, a range analysis is required to determine the relative importance of every factor and obtain the optimal combination [27]. The calculation formula of range, R, is
R = max { k 1 , k 2 , k 3 , , k n } min { k 1 , k 2 , k 3 , , k n } ,
k i = K i s ,
where R is range; Ki is the summation of the corresponding test results when the level number in each column is i; ki is the average of Ki; and s is the sum of each level number in each column.
A larger R represents the greater importance of the factor. The optimal combination of factors and levels can be obtained according to the optimal k value of each factor.

3. Geothermal Characteristics of the Gonghe Basin

3.1. Geological Setting

The Gonghe Basin (GHB), located on the northeastern edge of the Qinghai-Tibet Plateau, is a rhombus-shaped intermountain basin extending to the NWW [28]. It is about 210 km long from east to west, 90 km wide from north to south, and has a total area of 21,186 km2. It is the third largest basin in Qinghai Province, and the Yellow River runs across the short axis of the basin (Figure 2a). The GHB is a faulted basin formed since the Cenozoic. It is located at the intersection of the East Kunlun and West Qinling orogenic belts in the Qinling-Qilian-Kunlun fold system. It is bounded by the uplift of the fault fold belt. Affected by the fault activity on the boundary of the basin, the surrounding mountain ranges have been uplifting and rising recently, and the interior of the whole basin has relatively subsided. A set of huge Cenozoic sedimentary layers are formed in the basin, mainly composed of Quaternary alluvial deposits, river-lacust deposits, and Neo-Paleogene lacustrine deposits. The base of the basin mainly consists of intrusive rocks dominated by the Triassic strata and granite diorite (zircon dating from 222 to 245 Ma). The Yanshan intrusive magmatic rocks (zircon dating from 180 to 195 Ma) are exposed on a small scale on the surface of the basin [26].
The Wahongshan fault in the western part of the basin is a right-lateral strike-slip fault zone with a strike of NNW. Three hot springs (42.5 °C, 30.0 °C~62.0 °C, and 61.0 °C) are exposed along the fault. In the east of the basin, there are also three hot springs on the front line of the right-lateral strike-slip fault zone of the Waligongshan Mountain, and the temperatures are 86.0 °C, 93.5 °C, and 62.0 °C, respectively [29].
The China Geological Survey selected the Qiabuqia geothermal area as the target HDR pilot. The geothermal abnormality in this area is obvious, and there are two groups of fault zones in the Neogene mudstone formation along the Shangtamai-Ayihai area. Both sets of fault zones cut through the base of the basin. Deep thermal energy surges up along the fracture zone (F2, F4), forming a high temperature abnormal zone (Figure 2b). There are seven geothermal wells in the Qiabuqia site, four of which are HDR exploration wells (DR3, DR4, GR1, and GR2) [28].

3.2. Geothermal Features

The terrestrial heat flow in the GHB is 65~78 mW/m2, with an average of 72.58 mW/m2, which is much higher than the average heat flow in mainland China (62.53 mW/m2) [29]. It was reported from the DR3 and GR2 boreholes that the average heat flow is 119.3 mW/m2. The core record of the GR1 hole shows that the strata of the basin can be divided into the following three parts (Figure 3):
(1)
Part I (0~−505 m): The upper of Part I is a thin mid-late Pleistocene sand and gravel layer. Its grain gradually becomes finer downwards; the middle of Part I is mainly the lacustrine deposits of the Gonghe formation. The bottom of Part I is composed of sandy mudstone.
(2)
Part II (−505~−1350 m): It is composed of the Linxia formation and the Xianshuihe formation. The lithology is mainly medium-thick mudstone and medium-thick siltstone. The integrity of the mudstone is better. The sandstone particles are finer.
(3)
Part III (−1350~−3705 m): It is mid-late Triassic granite; the main lithology is granite and granodiorite. The radioactive heat generation rate of the granite ranges from 1.73 to 4.48 μW/m3, with an average of 3.04 μW/m3, and there is no abnormal high radioactive heat generation rate.
If the granite with a temperature greater than 150 °C is classified as HDR, the interface between the HDR and the overlying strata is shown in Figure 3. The HDR reservoir in the Qiabuqia area is located below about 2500 m.
Table 3 shows the temperature curves of four HDR exploration wells in the Qiabuqia area. The temperature data were measured using a DP-SJSW deep well thermometer. Within 0~−1500 m, the temperature gradients of GR1, GR2, DR3, and DR4 are 54.7, 46.7, 66.7, and 21.3 °C/km, respectively. Below −1500 m (to the bottom of each hole), the temperature gradients of GR1, GR2, DR3, and DR4 are 40.6, 50, 45, and 50.7 °C/km, respectively. The geothermal gradient in the HDR is 40.6~50.7 °C/km, and the formation temperature increases steadily with the depth, which belongs to the conductive geothermal energy [30]. The abnormal temperature of DR4 in the shallow part indicates that the basin is rich in hydrothermal resources in the shallow stratum (Part I and Part II).
The geothermal structure includes heat sources, heat reservoirs, caps, and channels. Gao et al. [28] conducted magnetotelluric sounding interpretation and found that there may be a molten layer in the depth interval of 15~35 km. This molten layer may be the heat source for high-temperature geothermal resources in the GHB. The widely distributed huge-thick high-temperature granite formation is an ideal target reservoir for HDR. The cap rocks are mainly mudstone (Part II) and have the characteristics of low heat conductivity (average 1.58 W/(m·K), Table 4). The thermal physical parameters were obtained by testing the core samples with a LFA447 laser thermal conductivity instrument in a laboratory. Therefore, it can effectively prevent the rapid loss of deep heat. The complex geological structure of the GHB has produced many deep, large faults, which can transfer the heat of the deep melt to the relatively shallow granite, thereby forming a relatively shallow high-temperature HDR reservoir [30].

4. EGS Reservoir Production Simulation Analysis

4.1. Target Reservoir and Project Design

The ideal target formation is one with a high temperature and well-developed natural fractures. Because reservoir stimulation has not been carried out in the study area, the geometry of the EGS reservoir is assumed in reference to the international EGS projects. The horizontal length of the Soultz EGS reservoir is about 2000 m, and the well spacing (GPK3-GPK2) is about 500 m [31]. The horizontal length of the Desert Peak EGS reservoir is about 2500 m, and the well spacing (27–15~22–22) is about 1000 m [32]. In this paper, the horizontal length of the created EGS reservoir in the Qiabuqia area is assumed to be 600 m, which is achievable according to current fracturing technology. Due to the subduction and compression of the Indian Ocean plate on the Eurasian continent, the internal horizontal tectonic stress in the GHB is relatively high. The dominant direction of the maximum compressive stress is in the NE direction, at about 51° [33]. It likely belongs to the stress mechanism of σH > σV > σh. Therefore, the stimulated EGS reservoir may be horizontal and may strip along the NE direction. Based on fracturing experience, the height and width of the created EGS reservoir are both assumed to be half its length, i.e., 300 m. Thus, the layer between the depths of 3700~4000 m is selected as the target reservoir, and the stimulated EGS reservoir geometry is assumed to be 600 m (length) × 300 m (width) × 300 m (height) (Figure 4). The natural fracture characteristics of the target reservoir are not clear; so, sensitivity studies are subsequently conducted to analyze the impact of the different fracture characteristics on productivity.
The vertical-doublet-well mode is the basis of another complex production mode. The injected section (Linj) and the produced section (Lpro) are arranged diagonally, considering the buoyancy of the high-temperature fluid (Figure 4). The local surface water (i.e., river or lake) or groundwater is used as the injected water and makeup water. It is assumed that the injected water is heated to 10 °C when it reaches the injected section of the reservoir. After surface power generation and stepped heat exchange, the circulated water is reduced to 10 °C and then pumped back into the injection well. In order to study the variation of the heat transfer effect of the EGS during a long operation, the project lifetime is set to 50 years.

4.2. Power System

In designing a power generation system for the EGS project, a binary system utilizing an organic working fluid called ORC (Organic Rankine Cycle) technology was assumed to be the best choice. It is widely applied in EGS power plants [34,35]; thus, we also adopted it as the surface power generating system this time. Figure 4 shows the principle of operation of an ORC geothermal power plant. The working fluid receives energy from the geothermal fluid through a heat exchanger unit and evaporates; it then expands in the turbine, condenses, and is returned to the heat exchanger by a pump. The organic working medium should be selected according to the heat-source temperature. The temperature between the depths of 3700~4000 m in the Qiabuqia area is about 224 °C, which is similar to that of the Soultz reservoir [31]. Therefore, isobutene was also adopted for the ORC system in Qiabuqia area.

4.3. Numerical Procedure

In this work, we used the TOUGH2-EOS1 code to simulate the thermal production performance of the EGS in the Qiabuqia area. The MINC approach was used to model the flow in the fractured EGS reservoir. It is a generalization of the classical double-porosity concept developed by Warren and Root [7]. In the typical use of the dual-porosity option, the communication throughout the model is through the fractures, and communication between the fractures and matrix is through the interporosity flow defined by nested elements. The matrix blocks are discretized into a sequence of nested volume elements, which are defined on the basis of distance from the fractures (Figure 5).
Figure 5 illustrates the 3D hydraulic-thermal coupled numerical model corresponding to the conceptual model of Figure 4. The geometric size of the model is 1000 m (x) × 700 m (y) × 700 m (z). The grid number is 24 (x) × 21 (y) × 16 (z), with a total of 8064 primary cells. In order to get reliable results, the grid was refined to 5 m × 5 m × 30 m near the wells to accommodate the high pressure and thermal gradients. The matrix blocks were divided into four subgrids with volume fractions of 0.08, 0.2, 0.35, and 0.35. The fracture domain occupied a volume fraction of 0.02. After this partition, there was a total of 40,320 subgrid blocks nested in the 8064 gridblocks. This mesh scheme was verified by grid sensitivity analysis to ensure adequate accuracy of the solutions.

4.4. Model Parameters and Single-Factor Sensitivity Scheme

Table 5 lists the values of the reservoir factors and the operation factors in the simulations. The matrix porosity, density, and specific heat are all based on the field data and laboratory experiments. A total of seven factors have been selected to study their influence on the thermal performance of the EGS. The factors consist of fracture spacing, fracture permeability, fracture permeability anisotropy, matrix permeability, heat conductivity, injected section length, and injection rate. We only changed one factor value in each simulation. The choice of factor is given as the following.
The larger the fracture spacing, the better the heat transfer effect of single fracture, but the tolerable total injection rate decreases for a certain length of injection section. In the Soultz project, the reservoir granite was pervasively fractured with a mean fracture spacing of about 3 m [31]. Therefore, the fracture spacing of the target reservoir was also assumed to 3 m in this paper and was selected as the base value for simulation. Meanwhile, a fracture spacing of 50 m was selected to represent a reservoir with sparse natural fractures. Suzuki et al. [9] concluded that the single-fracture model agreed with the results of the MINC for larger fracture spacing (>20 m). Thus, the MINC method is also suitable for a fracture spacing of 50 m.
The natural permeability of the deep granite in Qiabuqia is extremely low (about 1 × 10−16 m2~1 × 10−17 m2). After stimulation, the fracture permeability is generally increased by more than 1000 times [32]. In addition, through simulations, we found that when the fracture permeability was set as 1 × 10−14 m2, the flow impedance of the reservoir still increased significantly (see Section 5.1). Therefore, the fracture permeability is assumed to be 1 × 10−13 m2 as the base value. Through a long time of chemical stimulation, the fractures will be subjected to continuous corrosion, and the fracture permeability also increases. In order to evaluate whether excessive fracture permeability has a positive impact on productivity, we selected 1 × 10−11 m2 for the simulation analysis.
As mentioned above, the horizontal tectonic stress in the GHB is relatively high, and an EGS reservoir dominated by horizontal fractures will likely be formed after fracturing. This will make the horizontal permeability of the reservoir larger than the vertical permeability [36]. Therefore, this paper considers the impact of fracture permeability anisotropy on productivity. α = kx/kz = 10 denotes that the fracture permeability in the z (vertical) direction is 10 times higher than that in the x (horizontal) direction.
Matrix permeability is related to porosity; so, only the influence of matrix permeability on production is studied. The granite matrix is almost impermeable at about 1 × 10−17 m2. However, some lithology may contain abundant micro-fractures or large dissolved pores, which may lead to higher permeability. Water would also flow through these pore spaces to extract heat [37]. Therefore, we studied the heat transfer effect under the condition of matrix permeability of 1 × 10−14 m2.
According to Table 4, the heat conductivity of granite in the deep reservoir of Qiabuqia is between 2.10 W/(m·K) and 3.17 W/(m·K). In order to study the effect of heat conductivity on heat transfer, we slightly expanded the range, using 2.0 W/(m·K) as the base value and 3.5 W/(m·K) as the maximum value. In general, the heat conductivity of most crystalline rocks falls within this range.
In the EGS, both the injection sections and the pumping sections circulate water with the reservoir through casing perforations or grooves. Considering that this reservoir is only 300 m high in this study, the injected section lengths of 60 m and 120 m were chosen, respectively, for the simulations.
Injection rates are limited by temperature and pressure criteria. The simulations showed that 30 kg/s was the maximum injectable flow rate in this formation condition (see Section 4.6.1). For comparison, the effect of an injection rate of 20 kg/s on productivity was studied.

4.5. Initial and Boundary Conditions

In this study, all the boundaries were assumed to be no-flow for mass and heat transfer [4,8]. The wall rock, cap, and base layers were modeled thick enough to avoid boundary effects. The initial pressure conditions were assigned based on the hydrostatic pressure. As the height of the reservoir is only 300 m, the entire reservoir adopts a uniform temperature field. The initial reservoir temperature is 224 °C at a depth of 3850 m. For simplification, the fracture was assumed to be unchanged during the simulation process. Water losses were ignored, and the injection rate was assumed to be equal to the production rate.

4.6. Simulation Results of Single-Factor Sensitivity Analysis

4.6.1. Maximum Injection Rate under Base Case Condition

During the water circulation stage, when the cold water is injected into the stimulated reservoir, the reservoir pressure will grow and fluctuate with time. If the bottom-hole pressure of the injection well (Pinj) exceeds the reservoir minimum principal stress (σh), the created fracture will dilate, giving rise to second reservoir growth and water losses; so, Pinj must be lower than σh. According to Ref. [33], the σh at a depth of 3700 m in the Qiabuqia area is about 66 MPa. Thus, the maximum injection rate can be determined by the pressure criterion:
Pinj < 66 MPa,
The HDR resource development of Qiabuqia is mainly for electricity generation. If isobutene is chosen as the organic working medium, the production water temperature for the ORC generating electricity should meet the temperature criterion [38,39]:
Tpro > 105.36 °C,
where Tpro is the production water temperature.
Figure 6 shows the evolutions of Tpro and Pinj at different injection rates (qinj) over the 50 years. When qinj was 30 kg/s, the Pinj slightly increased from 37.92 MPa up to 38.38 MPa after 50 years of production. Meanwhile, Tpro decreased from 225 °C to a final 104 °C (down 53.8%). Within 50 years, the injection rate of 30 kg/s meets the pressure criterion and almost meets the temperature criterion. The injection pressure experienced three stages: slight rise, rapid rise, and slight rise. Tpro has gone through three stages: stable, rapid decline, and slight decline. When the injection rate became large, the water was pumped out before it fully exchanged heat with the reservoir, resulting in a rapid drop in the production temperature. When the heat-supplying reservoir from the wall rock is approximately equal to that taken away by water circulation, the production temperature tends to be stable again. From the 30th year to the 50th year, Tpro decreased by about 10 °C, while from the 45th year to the 50th year, Tpro decreased by only about 1 °C.
When qinj is 80 kg/s, Pinj increases from the initial 38.62 MPa to 65.89 MPa in the 50th year. Tpro decreased from 225 °C to the final 31 °C. Within 50 years, the injection rate of 80 kg/s meets the pressure criterion but does not meet the temperature criterion. To ensure continuous power generation for the EGS, the maximum qinj should be set to 30 kg/s.

4.6.2. Production Temperature

Figure 7 shows the evolutions of Tpro under different parameters over the 50 years. The curve of Tpro could be divided into two stages: the stable period and the decline period. In the early stage of the EGS operation, the reservoir and wall rock contain sufficient heat; so, Tpro can be equal to the reservoir temperature. This stage belongs to the stable period. As cold water is continuously injected into the reservoir, the reservoir heat is continuously extracted. Tpro would enter decline period at some point. Among the seven selected parameters, Pf, α, and qinj have the greatest influence on Tpro. The second is the Sf. While Pr, λr, and Linj all have a slight effect on Tpro. The curve patterns of Tpro are in accordance with previous studies of Refs. [4,8,35].

4.6.3. Power Generation

Garg et al. derived a new expression for the available power generation, We, for the binary cycle with the following equation [38]:
W e = η e m w ( h w h r ) ( T R T P ) ( T R T r ) h s f g l ( T b ) { h s f g l ( T b ) h s f l ( T c ) T c k [ s s f g ( T b ) s s f l ( T c ) ] V s f ( T C , P s f b ) [ P i n P s f b ( T c ) ] } ,
where ηe, mw, T, h, s, P, and V denote conversion efficiency, production rate, temperature, specific enthalpy, entropy, pressure, and specific volume, respectively. The subscripts w, r, R, P, b, c, g, l, in, sf, and k denote wellhead, reference, reservoir, pinch point, boiling, condenser, gas, liquid, inlet, secondary fluid, and Kelvin, respectively. More detailed descriptions of Equation (11) can be found in Ref. [40].
Figure 8 shows the evolutions of We under different parameters over the 50 years. The most influential factors are Pf, α, and qinj. Comparing Figure 7 and Figure 8, it can be seen that except for qinj, the variation trends of We under the other six factors are the same as the trends of Tpro. Although Tpro declined much more slowly when qinj decreased from 30 kg/s to 20 kg/s, the early We at 20 kg/s is lower than that at 30 kg/s due to the flow rate limitation. In the base case condition, the We in 50 years decreases from 2.53 MW to 0.16 MW, and the stable period is about 10 years. When qinj is 20 kg/s, the We in 50 years is 1.69 MW~0.91 MW, and the stable period is about 13 years.
Over the whole 50 years, λr has the least influence on We, followed by Linj and Sf. If the lifetime of the EGS project is 30 years, Pr also has little influence on We, and the influence of Sf should be considered. It should be noted that the qinj is set to be constant in the numerical simulations of case Linj and case Sf. However, in a practical project, the longer the Linj and the smaller the Sf, the more conducive they are to the increase of qinj.

4.6.4. Bottom-Hole Pressure of Injection Well

Figure 9 shows the evolutions of Pinj under different parameters over the 50 years. It illustrates that Pr, Sf, and λr have the least influence on Pinj. However, α, qinj, and Linj have great influence on Pinj. Pf has the greatest effect.
In the base case, Pinj slightly increased from 37.92 MPa up to 38.38 MPa after the 50-year production, increasing by 0.46 MPa (1.2%). When α become 10, the Pinj increased from 37.92 MPa to 41.70 MPa over the 50 years, increasing by 3.78 MPa (9.1%). When qinj was 20kg/s, Pinj increased from 32.78 MPa to 34.70 MPa, with an increase of 1.92 MPa (5.9%). Pinj increased by 0.49 MPa (1.6%), from 30.70 MPa to 31.19MPa, over the 50 years when Pf increased by 100 times. Among all the parameters, Pinj increased the most when α become 10 due to a great decrease in vertical permeability compared to the base case. Under the same qinj, the longer the Linj, the lower the Pinj. Pinj among all the cases met the pressure criterion.

4.6.5. Reservoir Injectivity

The reservoir injectivity, Iinj, reflects the stimulation effect of the reservoir permeability. An ideal EGS typically requires an Iinj greater than 10 kg/s/MPa. The higher the Iinj, the easier it is for water to flow from the injection well to the pumping well at the same pressure difference. It reflects the average flow impedance of the reservoir. Iinj is calculated by [1]:
Iinj = qinj/(PinjPpro),
Figure 10 shows the evolutions of Iinj under the different parameters over the 50 years. Pf, Linj, and α are the most important factors. In the base case, Iinj ranged from 9.0 to 4.2 kg/s/MPa over the 50 years. In the Pf case, Iinj was 1028.2~584.7 kg/s/MPa over the 50 years, which was about 100 times that of the other cases. When the injection length doubled, Iinj was 19.7~7.7 kg/s/MPa, which also nearly doubled. In the α case, Iinj, 7.6–2.9 kg/s/MPa, was lower than that of the base case.
Pf is directly related to Sf. Assuming that the fracture is flat and smooth, a Pf of 1 × 10−11m2 is equivalent to the fracture aperture of 0.05 mm [41]. At present, a fracture aperture of this magnitude mainly depends on the self-supporting of the shear fracture after stimulation. Additionally, in this model, qinj has little influence on Iinj. The main reason is that the Pf of 1 × 10−13 m2 can still smoothly accommodate a 20 kg/s injection flow. If the Pf decreases further, an excess qinj can cause a rapid rise in Pinj, resulting in a significant decrease in Iinj. Basically, Iinj is mainly dependent on the overall permeability of the stimulated reservoir; so, λr and Pr have little influence on Iinj.

4.6.6. Pump Power

The internal energy consumption, Wp, mainly includes the energy consumption of the injection pump and that of the suction pump [1,8]. Wp mainly refers to the electric energy consumed to drive water flow through the reservoir. It does not include the energy consumed by injecting water from the surface into the reservoir and pumping water from the reservoir to the surface. If the energy loss due to duct friction and water internal friction is neglected, the pump efficiency ηp is 80%, and the internal energy consumption Wp is Equation (13), based on [8]:
W P = q ( P inj     P pro )     ρ qg ( h 1   h 2 ) ρ η P ,
Figure 11 shows the evolutions of Wp under different parameters over the 50 years. Pf is the most important factor, followed by α, qinj, and Linj. Other factors have a slight impact on Wp. Among them, α causes Wp to be higher than that of the base case, while Pf, qinj, and Linj make Wp lower. In the base case, Wp is 0.14~0.30 MW over the 50 years. In the α case, Wp ranges from 0.16 MW to 0.43MW. In the Pf case, Wp is 0.001~0.002 MW. When Pf increases, the reservoir has little resistance to water flow. Water can flow through the reservoir only driven by gravity; so, Wp is extremely low.

4.6.7. Coefficient of Performance

The coefficient of performance, η, is defined as the ratio of the total produced power energy to the internal energy consumption [1,33]. It represents the net power generation. Based on Equations (11) and (13), η can be calculated as Equation (14):
η = We/Wp,
Figure 12 shows the evolutions of η under the different parameters over the 50 years. It can be seen that the η of all cases decreased over the entire 50-year operation. Pf is the most important factor affecting η, followed by qinj, Linj, and α. The other factors are less important.
In the base case, η is 18.64~0.64 over the 50 years. In the Pf case, η is 570.19~0, which decreases the fastest mainly because We declines too fast. Over the 50 years, the η of the α case was lower than that of the other cases because its We declined the most, while Wp increased the most. The decrease of qinj can improve system η, but the curve trend of η is almost unchanged. In the ground source heat pump industry, η is usually expected to be greater than 4. In this simulation, only the case of qinj can meet η larger than 4 during the whole 50 years. The curve patterns of η are in accordance with the previous studies of Refs. [4,8,35].

4.7. Discussions of Single-Factor Sensitivity Analysis

4.7.1. Effect of Reservoir Factors on Production Performance

The EGS reservoir factors mainly consist of fracture spacing (Sf), fracture permeability (Pf), fracture permeability anisotropy (α), matrix permeability (Pr), and heat conductivity (λr).
Sf reflects the development degree of the reservoir fractures. The smaller the Sf, the more developed the fractures. More seepage channels can be selected for water flow. When the qinj is constant and Sf increases from 3 m to 50 m, the flux that each fracture needs to bear increases. Thus, increasing Sf leads to a drop in the produced water temperature and power generation. The larger the Sf is, the more quickly the produced water temperature and power generation decrease, and their stable periods are both shortened. Sf has little effect on the other four indicators. Therefore, the stratum with dense natural fractures should be selected as the target EGS reservoir in practical engineering.
The influence of Pf on each indicator is very large. The greater the Pf, the easier it is for water to pass through the reservoir. According to Darcy’s law, when the hydraulic gradient is constant, the water seepage velocity is proportional to the permeability of the reservoir. Thus, when Pf is 100 times greater than that of the base case, the water velocity in the reservoir also increases significantly. The water circulation process would be completed quickly before the cold water absorbed enough heat from the reservoir. The produced water temperature and power generation dropped quickly compared to the base case, and the pump consumption and injectivity increased rapidly. Therefore, it is not advisable to acidify the EGS reservoir too much to widen the apertures of natural fractures. This is likely to lead to a rapid decline in net power generation.
α (kx/kz = 10) means that the vertical permeability is 10 times lower than that of the base case. When water flows through such fractured reservoirs, it will preferentially reach the production well in the horizontal direction. Thus, the water flow in the vertical direction is greatly restricted. Therefore, the water flow path and the heat exchange area are also smaller than that of the isotropic reservoir. Meanwhile, the water flow through the horizontal direction is greater than that of the base case. In general, the produced water temperature and power generation of an anisotropic reservoir are much lower than that of the base case. In addition, fracture permeability anisotropy increases pump energy consumption greatly.
The greater the Pr, the larger the pores, and therefore the less heat storage per unit volume of rock. When Pr increased by 100 times, the heat storage in the matrix also greatly decreased. In the dual-porous medium model, the matrix only exchanges heat with the fractures, and there is no heat exchange between the matrixes. In the early stage of the water circulation, the heat in the matrix is continuously transferred to the water flow, and the produced water temperature during this period is almost the same as that of the base case. With the continuous operation of the project, the matrix heat becomes less and less. However, the heat recharge of the wall rock to the reservoir is limited. Therefore, the produced water temperature and power generation both drop faster than that of the base case in the later stage of the project (about 24th year). Pr has little effect on the other four indicators.
The greater the λr, the faster the heat conduction from the far field to the reservoir and from the reservoir matrix to the water flow. When λr of the rock becomes larger, the stable period of the power generation also becomes longer, indicating that the heat of the reservoir rock is transferred to the fracture water flow more than that of the base case. When entering the descending period, the power generation drops faster than that of the base case due to the excessive extraction of heat from the rock in the early stage. At this time, the temperature difference between the wall rock and the reservoir is not very large; so, the wall rock supplies less heat to the reservoir. As the water continues to circulate, the reservoir with a large λr will cool down faster. The heat transfer from the wall rock to the reservoir will increase, and the drop of power generation will begin to slow down. Therefore, the power generation with a large λr is slightly higher than that of the base case in the 50th year. Therefore, rock heat conductivity may not be used as an indicator in selecting a target reservoir.

4.7.2. Effect of Operation Factors on Production Performance

The operation factors have a great impact on the EGS production performance. They include injected section length (Linj) and injection rate (qinj).
The lengthening of Linj means the flow rate of a single fracture near the injection well decreases, and the water temperature will become higher. Therefore, the power generation after increasing the length of Linj is a little higher than that of the base case at first. When entering the descending period, because the heat near the injection well is extracted prematurely, the power generation drops faster than that of the base case. It makes the amount of heat supplied to the reservoir by the wall rock higher. The power generation of the extended Linj is a little larger than that of the base case in the 50th year. However, Linj has a great influence on the pump power and the coefficient of performance. The longer the Linj, the lower the pump power and the higher the coefficient of performance. Therefore, in practice, the Linj should be as long as possible.
The lower the qinj, the longer the heat exchange time between the water flow and the reservoir, and therefore the longer the stable period of power generation. The stable periods of power generation under qinj = 20 kg/s and qinj = 30 kg/s are 18 years and 10 years, respectively. Moreover, the smaller the qinj, the slower the drop of power generation during the descending period. qinj has a great influence on the pump power and the coefficient of performance. However, the smaller the qinj, the lower the pump power and the higher the coefficient of performance. Therefore, in practical engineering, the most appropriate qinj should be determined by comprehensive simulations.

4.7.3. Main Factors Affecting Production Performance

According to the above simulation results, it can be seen that the seven parameters selected this time have different effects on the production performance of the EGS. In order to identify their influence degree, a sensitivity analysis was performed based on the above simulation results. The sensitivity, φ, is calculated by:
φ = V sele     V base V base ,
where Vsele and Vbase, respectively, denote the 50-year average of each index under the selected case and the base case. Figure 13 shows the sensitivity analysis results of the different factors on each index.
For Tpro, the order of influence degree is as follows: qinj > Pf > α > Sf > Pr > Linj > λr. In the long run, the decrease of qinj is beneficial to the increase of Tpro. The increase of Sf, Pf, α, and Pr will result in the decrease of Tpro. Linj andλr have little effect on Tpro.
For We, the order of influence degree is as follows: Pf > α > Pr > qinj > Sf > Linj > λr. In the long run, the increase of Pf and α will obviously reduce We. Although a higher temperature may be obtained when the injection flow rate decreases, the average annual power generation still decreases.
For Pinj, the order of influence degree is as follows: Pf > α > qinj> Linj > λr >Pr > Sf. In the long run, the increase of Pf, qinj, and Linj will reduce Pinj. The increase of α will obviously increase Pinj.
For Iinj, the order of influence degree is as follows: Pf > Linj > α > qinj > λr > Pr > Sf. In the long run, the increase of Pf and Linj will significantly increase Iinj. The increase of α will decrease Iinj.
For Wp, the order of influence degree is as follows: Pf > qinj > α > Linj > λr > Pr > Sf. In the long run, the increase of Pf, qinj, and Linj can significantly reduce Wp. The increase of α will increase Wp.
For η, the order of influence degree is as follows: Pf > qinj > Linj > α >Pr > Sf > λr. In the long run, the increase of Pf, qinj, and Linj can significantly increase η. The increase of α will reduce η.
On the whole, the four factors of fracture permeability, fracture permeability anisotropy, injected section length, and injection rate have the greatest influence on the EGS production. We and η together determine the net generating capacity. It can be seen that only α acts in the same direction on the two indexes, while the other three parameters (Pf, qinj, and Linj) act in opposite directions on them. Therefore, multi-factor combination analysis is needed to achieve the optimal combination of We and η.

5. Multi-Factor Combination Analysis

In practical engineering, the performance index of the EGS is determined by the combination of multi-factors rather than a single factor. Each factor has a certain value range (i.e., multi-level). Therefore, multi-factor and multi-level combination analysis is needed to quantitatively obtain the optimal combination scheme. Generally, the statistical factor analysis (SFA) method needs to match each level of each factor to conduct complete tests, which requires a lot of time. In addition, when the number of the varying parameters in the model becomes large, a large number of numerical samples have to be produced using the SFA, making the optimization study challenging. Compared to the SFA technique, the orthogonal test method can avoid the production of a large number of numerical samples and is more efficient than the SFA technique [21]. The orthogonal test could be used to scientifically arrange and analyze multi-factor and multi-level tests. A standard orthogonal test table is selected based on the number of levels of multi-factors. It can evenly select the representative groups of data in all the test schemes. Through the orthogonal test, the number of experiments greatly is reduced without decreasing the feasibility of the test. Therefore, the orthogonal test is conducted in this part for multi-factor and multi-level combination analysis.

5.1. Orthogonal Test Scheme

As mentioned above, the four factors that had the greatest impact on the EGS productivity were screened out, including fracture permeability, fracture permeability anisotropy, injected section length, and injection rate. For the tests of four factors involving four levels, a complete experiment would test 4 × 4 × 4 × 4 = 256 trials in total. It is a significant amount of work. In this study, a four-factor and four-level orthogonal test table was adopted. Then, we only need do 16 trials to determine the relative importance of every factor. The distribution of the levels of each factor is shown in Table 6. When Pf is 1 × 10−14 m2, the bottom-hole pressure increased rapidly from 44.1 MPa to 80.24 MPa in 6 days with the injection rate of 20 kg/s. Thus, the minimum Pf was set to 5 × 10−14 m2. Linj was set from 30 m to the reservoir height of 300 m. The maximum injection rate was set to 110 kg/s, exceeding the commercial flow rate of 100 kg/s. The maximum level of anisotropy was set as 1000 to consider the influence of a strong anisotropy geology condition. These four levels cover almost all possible scenarios. The orthogonal test scheme is shown in Table 7.

5.2. Orthogonal Test Results

Table 8 lists the simulation results of the orthogonal test according to Table 7. According to the single-factor simulation curve in Section 4, it can be seen that the variation path of We may be different in different situations. Even if both scenarios produce the same amount of electricity in the 50th year, they may produce very different amounts over the entire 50 years. So, the 50-year average of each index is needed. In addition, the initial values of some indexes, such as η and Iinj, may differ greatly from the final values in the 50th year. The project may be no longer operational before 50 years. Only using the 50-year average cannot accurately represent the influence of the test. Therefore, the average annual value is mainly used for comparison, and the value at the end of the simulation is used for auxiliary judgment.
From Table 8, based on the average annual value of the 16 tests, We ranges from −3.47 MW to 1.46 MW; η ranges from −439.38 to 420.36; Tpro ranges from 49.37 °C to 223.83 °C; Iinj ranges from 1.17 to 892.88; Pinj ranges from 30.61 to 68.63; and Wp ranges from 0.0019 MW to 3.77 MW. It can be seen that the different combinations have great influence on the heat transfer performance, and it is difficult to meet all six indexes at the same time. For example, although the We of test 1 is good, its η and Iinj are low. For We, the better combinations include test 7, test 1, and test 12. For η, the better combinations include test 7 and test 14. Their common characteristic is an injection rate of 20 kg/s. All six indexes of test 7 are good, indicating that the level of all the factors in scheme 7 is better combined at this time. The negative value of We in Table 8 is because the selected organic working medium isobutene cannot generate power when its temperature is lower than 105.36 °C, as mentioned in Section 4.6.1. In the practical engineering, We and η cannot be negative. However, in order to directly reflect the influence of each factor, their theoretical negative values are adopted this time. In addition, for test 6, the injection rate of 110 kg/s was too high for the fracture permeability of 1 × 10−13 m2, resulting in a rapid increase of Pinj; so, test 6 has no data. Similarly, test 3 and test 8 only run to 15.8 years and 25.2 years, respectively, due to the excessive pressure growth.

5.3. Discussions of Multi-Factor Combination Analysis

5.3.1. Effect of Main Factors on Production Performance

The ultimate purpose of heat extraction is to generate electricity; so, net power generation is the first benefit indicator to be concerned with and is determined by power generation (We), pump power (Wp), and the coefficient of performance (η). Injection pressure is allowed to grow as long as it is not so high as to exceed the strength capacity of the pump and pipeline. In order to further explore which factor plays a dominant role in the multi-parameter combination, we sorted out the results in Table 8 according to We, Wp, and η (Figure 14).
For the fracture permeability (Figure 14a), it can be seen that We varies greatly even though it is for the same level. This means that even when the fracture permeability is selected to the optimal value, the resulting We may be very low if the other parameters do not match the fracture permeability. The variation range of η at the same level may be very different, indicating that a different fracture permeability may have a great impact on η. The Wp of the other three levels, except that of level 1, does not change much. Considering the three indicators, the best cases at levels 1~4 are G1, G3, G4, and G2, respectively. From level 1 to level 4, We, η, and Wp do not increase or decrease as the fracture permeability increases, indicating that it is not necessary to increase fracture permeability too much in the actual reservoir stimulation. Depending on other parameters, even small fracture permeability can generate significant We (i.e., G1 at level 1 and G3 at level 2). However, when the fracture permeability is low (level 1), Wp is higher in most cases than that at the other levels, and overall η is higher than that at the other levels. When fracture permeability is large (level 4), We is also small. Therefore, the overall fracture permeability should not be too low or too high.
For the fracture permeability anisotropy (Figure 14b), it can be seen that We varies greatly even though it is for the same level, indicating that fracture permeability anisotropy does not dominate. If the values of the other factors do not match, productivity will also be affected. With the increase in fracture permeability anisotropy (from level 1 to level 4), there is no obvious increase or decrease of each indicator. Even if the anisotropy is very large, it is possible to obtain larger We as long as the other parameters are well matched. Considering the three indicators, the best cases at level 1~4 are G1, G4, G2, and G3, respectively. It also shows that even though the actual fracture permeability anisotropy varies greatly, we only need to find other parameters matching it. It is not necessary to stimulate the reservoir permeability to be isotropic.
For the injected section length (Figure 14c), it can be seen that We varies greatly even though it is for the same level. The injected section length does not dominate. The fluctuation of the three indicators of level 1 is relatively small. This means that when the injected section length is small, the other parameters have difficulty playing a large role. At this time, if the injected section length is not increased, the EGS production performance cannot be improved. The three indicators at level 3 and level 4 fluctuate greatly, indicating more possibility of combination with the other factors. Therefore, in practical engineering, the injected section length should be larger. Ref. [42] shows that a larger Linj enhances heat extraction when Linj is less than the reservoir height. The regularity of our results is consistent with Ref. [42].
For the injection rate (Figure 14d), We and η obviously decrease as injection rate increases (from level 1 to level 4), regardless of whether the other factors are matched. The lower the injection rate, the lower the Wp. Moreover, as the injection rate increases, We and η drop rapidly to negative values. It indicates that the produced water temperature drops too fast with the increase in the injection rate. All three indicators of level 1 are the best of the four levels. Thus, the injection rate determines in all factors, and the smaller the better. Therefore, in practical engineering, special attention should be paid to the value of the injection rate, which should not be too large.

5.3.2. Order of Main Factors and Optimal Combination

In order to quantitatively determine the order of influence degree of each factor, a range analysis was carried out based on the orthogonal test results. Table 9 lists the range analysis results. The greater the range, the more important the influence of the factors on the production evaluation index [22,27].
For Tpro, the order of influence degree is as follows: qinj > Pf > α > Linj. This result is consistent with that of the single-factor sensitivity analysis.
For We, the order of influence degree is as follows: qinj > Pf > α > Linj. In comparison with the result of the single-factor sensitivity analysis, only the order of qinj and Pf has changed.
For Pinj, the order of influence degree is as follows: qinj > Pf > α > Linj. In comparison with the result of the single-factor sensitivity analysis, only the order of qinj and Pf has changed.
For Iinj, the order of influence degree is as follows: Linj > Pf > α > qinj. In comparison with the result of the single-factor sensitivity analysis, only the order of Linj and Pf has changed.
For Wp, the order of influence degree is as follows: Pf > qinj >α > Linj. This result is consistent with that of the single-factor sensitivity analysis.
For η, the order of influence degree is as follows: Pf > qinj > Linj > α. This result is consistent with that of the single-factor sensitivity analysis.
On the whole, the rank results of the orthogonal test are almost the same as those of single-factor sensitivity. It proves the accuracy of the simulation results. In the actual EGS project, the two most concerned indicators are We and η. According to Table 9, the better level combination of these two indicators is A2B3C1D1 and A2B2C3D1, respectively. From the orthogonal results listed in Table 8, the optimal combination is test 7 (A2B3C4D1). The comparison of these three combinations and the results are shown in Figure 15. The We of A2B3C1D1, A2B2C3D1, and A2B3C4D1 over the 50 years are 1.69 MW~1.04 MW, 1.69 MW~0.68 MW, and 1.69 MW~1.11 MW. Between 0 and 42.3 years, the We of A2B3C1D1 was consistently higher than that of A2B3C4D1. The η of A2B3C1D1, A2B2C3D1, and A2B3C4D1 over the 50 years are 12.53~1.70, 14.59~12.70, and 120.26~35.26. Although the We of A2B3C1D1 is overall higher than that of A2B3C4D1, the η of A2B3C1D1 is much lower than that of A2B3C4D1. Thus, considering We and η comprehensively, the optimal scheme is A2B3C4D1. Based on the above results, this indicates that the optimal EGS reservoir should have moderate fracture permeability and moderate fracture permeability anisotropy; the injected section length should be large; and the injection rate should be moderate.

5.3.3. Optimal Scheme Considering Fracture Permeability Anisotropy

In practice, it is difficult to control and determine fracture permeability anisotropy (α) because it is greatly affected by in situ stress and natural fracture distribution. To study the effect of fracture permeability anisotropy on the production performance, a single-factor three-level test was specially arranged. The results are listed in Table 10. With the increase of α, We and Tpro increase first and then decrease, while Pinj and Wp keep growing. So, η decreases downwards. All the cases in Table 10 comply with the temperature and pressure criteria. If permeability anisotropy can be controlled during reservoir stimulation, the optimal strategy is still test A2B3C4D1 (α = 100). If permeability anisotropy is not controlled, the 50-year average We, η, and Iinj with A2C4D1 are likely to be 1.23 MW~1.46 MW, 11.42~67.50, and 14.28~20.92, respectively. In addition, it can be seen that the 6 indicators of these 4 tests in Table 10 are all good and are all better than that of the 15 tests, except test 7 in Table 8. Therefore, this proves that the optimization combinatorial scheme of EGS can be achieved through the orthogonal test and range analysis.

5.3.4. Optimal Scheme Considering Injection Rate

As mentioned above, the injection rate is decisive even in a multi-factor combination. Meanwhile, the lower the injection rate, the greater the average power generation over 50 years. In order to explore the lower limit of the optimal injection rate, we conducted simulations of net power generation (WeWp) under different injection rates based on test A2B3C4D1, as shown in Figure 16. The results show that with the decrease in the injection rate, the stable period of net power generation gradually lengthens. The decreasing rate of the net power generation during the descending period also slows down. However, in the initial period, the net power generation is still higher when the injection rate is higher. Thus, the net power generation at 30 kg/s is greater than that at 20 kg/s up to 34.2 years. The net power generation at 20 kg/s was consistently greater than that at 15 kg/s for the entire 50 years. The intersection of the two curves occurred 50 years later. The net power generation at 10 kg/s is almost constant over a 50-year period. Therefore, the choice of optimal injection rate also depends on the life of the project. The injection rate should not be excessive if the EGS project life is required to be longer.
The commercial injection rate of the EGS is expected more than 100 kg/s. According to the above simulation results, when the injection rate reaches 100 kg/s, the produced water temperature and the power generation both decline quickly. Meanwhile, the injection pressure will also rise rapidly, which can easily cause high pump pressure, causing great difficulties to the operation. If a commercial rate (100 kg/s) is to be obtained, the permeability of the reservoir fracture network needs to be stimulated to be higher. Meanwhile, in order to ensure that the production temperature is both high and stable, it is necessary to further increase the volume of the EGS reservoir.

5.3.5. Well Layout Mode

The EGS reservoir volume in this paper is 0.54 × 108 m3. Sanyal and Butler [43] proposed that the stimulated volume should exceed 1 × 108 m3 for the EGS. The EGS reservoir volume needs to be increased further. If the created EGS reservoir is too long, a large flow resistance needs to be overcome in order to drive water through the reservoir. Considering the symmetry of hydraulic fracturing, the triplet-well straight-line mode with one injector and two producers is recommended according to case A2B3C4D1 (Figure 17). Well spacing is 600 m, and the total EGS reservoir volume reaches 1.08 × 108 m3. During reservoir stimulation, fractures mostly extend parallel to the direction of maximum principal stress. Thus, the direction of the well connection line is set in the same direction as the maximum principal stress in the study area (N51°). The injection rate is 40 kg/s and the flow rate from each production well is 20 kg/s. In actual engineering, more fluid may need to be injected to compensate for the fluid loss. The 50-year average of We, η, and Iinj with this triplet-well mode can reach 2.46 MW~2.92 MW, 11.42~67.50, and 14.28~20.92, respectively.

5.3.6. Optimal Scheme Considering Injection Water Temperature

In this study, groundwater is assumed to be used as the injection water, with an injection temperature of 10 °C. In practical projects, the reinjection water may have different temperatures after surface power generation and stepped heat exchange. To further optimize the scheme, the effect of different injection water temperatures on productivity was studied. Figure 18 shows the comparison of the production index under different values of injection temperature in the 50th year, based on test A2B3C4D1.
The temperature difference between the reservoir and the injected water of the four cases is higher than 100 °C. It results in a small difference in their stable periods of Tpro and We (Figure 18a). During the decline stage, a high injection temperature causes a slow drop of Tpro and We. Increasing the injection temperature will definitely increase Tpro and We; however, the increase in the injection temperature also requires extra energy. The appropriate injection temperature should be determined in accordance with the water source condition and the engineering requirement.
With the increase in the injection temperature, Pinj and Wp both decrease. Pinj is the function of (μ/ρ), which is also the function of temperature. As heat production continues, the reservoir temperature continually declines; (μ/ρ) significantly increases when the temperature declines (Figure 18d). The higher injection temperature causes the production temperature to drop more slowly. Thus, Pinj and Wp both increase slowly when the injection temperature is higher (Figure 18b,c). Finally, Iinj and η are both improved according to Equations (12) and (14).

5.3.7. Space Variation of Reservoir Temperature Field

Figure 19 shows the profiles of the space variations of the reservoir temperature fields over the 50 years (y = 350m) (a~e) and the temperature plan in the 50th year (z = 350m) (f). The results are based on test A2B3C4D1 with an injection temperature of 70 °C. As can be seen from the profile, with the cold water gradually injected into the reservoir, the cold front continues to move towards the pumping well. Generally, the density of the injected cold water is large; so, its downward speed is high under the gravity drive. Because the water injection section and pumping section are located at the same depth in this study, the vertical diffusion velocity of the cold halo is nearly the same as the horizontal diffusion velocity (Figure 19a–e). Thus, the cold front almost uniformly moves horizontally from the injection well to the pumping well, and the heat extracted from the whole reservoir is sufficient. It can be seen from Figure 19f that after 50 years of operation, the lateral influence range of thermal extraction on wall rock is about 99 m on each side and that of the injection well and pumping well sides is 148 m and 24 m, respectively. Cooling does not affect the boundary of the model in this simulation, indicating that the size and boundary conditions of this model are reasonable.

6. Conclusions

Based on the geothermal geological characteristics of the Qiabuqia area, Gonghe Basin, a hydraulic-thermal coupled numerical model was established by using the MINC method. In order to explore the optimal mining strategy of the fractured EGS reservoir, we numerically investigated the influence of seven single factors on heat production and conducted an optimization of multi-factor combinations by an orthogonal test. The following conclusions were drawn:
(1)
The Gonghe Basin possess a good geothermal structure. The molten layer in the depth interval of 15~35 km may be the heat source. The widely distributed huge-thick high-temperature granite formation is an ideal target reservoir for HDR. The cap rocks are mainly mudstone and sandstone, which have the characteristics of low heat conductivity. The complex geological structure has produced many deep large faults, which can serve as heat conduction channels. Therefore, there has formed a relatively shallow high-temperature HDR reservoir in the Gonghe Basin.
(2)
For single-factor sensitivity analysis, when qinj was constant, the increase of Sf had a slight influence on Tpro and We. Pf had the greatest influence on the production performance among all the factors. For the condition with α = 10, We declined the fastest, while Wp rose the fastest, resulting in the lowest η. The increase of Pr had a slight influence on Tpro and We in the later stage of the project. λr was the least influential of all the parameters. Increasing Linj had little influence on Tpro and We but had great influence on Pinj and Wp. When qinj decreased, the early We decreased due to the flow rate limitation although Tpro declined much more slowly. Meanwhile, Wp decreased and η increased significantly. Therefore, the biggest influence factor on the We value was the overall permeability of the fractured reservoir. For η, Pf was the most important factor, followed by qinj, Linj, and α.
(3)
The four factors of fracture permeability, fracture permeability anisotropy, injected section length, and injection rate had the greatest influence on the EGS production. For the multi-factor sensitivity analysis, the order of influence degree on We was qinj > Pf > α > Linj. The order of influence degree on η was Pf > qinj > Linj > α. On the whole, the rank results of the orthogonal test are almost the same as those of single-factor sensitivity.
(4)
Different factor combinations have great influence on the heat transfer performance. The multi-factor and multi-level combination optimization is needed and the optimization scheme of the EGS can be achieved through the orthogonal test and range analysis.
(5)
For reservoir stimulation, the stratum with dense natural fractures should be selected as the target EGS reservoir. It is not advisable to acidify the EGS reservoir too much to widen the apertures of natural fractures. This is likely to lead to a rapid decline in net power generation. Fracture permeability anisotropy will increase pump energy consumption, but this adverse effect can be greatly reduced if the other parameters are well matched. Matrix permeability and heat conductivity may not be used as an indicator in selecting a target reservoir.
(6)
For project operation, the injected section length should be as long as possible. The injection rate plays a major role in all factors. Special attention should be paid to the value of the injection rate, which should not be too large. The appropriate injection temperature should be determined in accordance with the water source condition and the engineering requirement. If a commercial rate (100 kg/s) is to be obtained, the permeability of the reservoir fracture network needs to be stimulated to be higher. Meanwhile, in order to ensure that the production temperature is both high and stable, it is necessary to further increase the volume of the EGS reservoir.
In practical EGS projects, we suggest that one should be very careful when choosing which formation to use as the EGS target reservoir. Firstly, the high-temperature reservoir should be selected, as far as possible, and the reservoir size should be created as large as possible. This can ensure the lower limit of the EGS reservoir productivity. As for optimization, the best way is to obtain as much reservoir information as possible and conduct sufficient numerical simulations. In this study, considering the complexity of the actual project, we simplified the numerical model. Only the coupling of the hydraulic-thermal effect was considered and water losses in the reservoir were neglected. However, the rock mechanical deformation and the water–rock chemical reaction must be considered during the long-term operation of the EGS. Additionally, the logging data of the distribution characteristics of the facture spacing, aperture, and orientation should be collected to establish a more realistic hydraulic–thermal–mechanical–chemical model in the future.

Author Contributions

Conceptualization, Y.Z. and L.G.; methodology, L.G.; software, J.Z.; validation, L.S.; writing—original draft preparation, S.C.; writing—review and editing, Y.Z.; supervision, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by project of Energy Acquisition, Conversion, and Efficient Utilization of Hot Dry Rock (No. 20211101); China Postdoctoral Science Foundation (Nos. 2020M681799, 2019M661053 and 2020T130390).

Institutional Review Board Statement

Not applicable. The study did not involve humans or animals.

Informed Consent Statement

Not applicable. The study did not involve humans.

Data Availability Statement

Some or all of the data and the models generated or used during the study are available from the corresponding author by request.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

CRRock specific heat, J/(kg·K)hβSpecific enthalpy in phase β, J/kg
FMass or heat flux, kg/m2 or W/m2tTime, s
IinjInjectivity, kg/s/MPauβSpecific internal energy in phase β, J/kg
LinjInjected section length, mαFracture permeability anisotropy
MMass or energy per volume, kg/m3 or J/m3ΦPorosity
PfFracture permeability, m2φSensitivity
PinjBottom-hole pressure of injection well, MPaκMass components
PrMatrix permeability, m2ηCoefficient of performance
qinjInjection rate, kg/sηeConversion efficiency
RRangeηpPump efficiency
SfFracture spacing, mλrHeat conductivity, W/(m·K)
SβSaturation of phase βρRRock density, kg/m3
TTemperature, °CρβDensity of phase β, kg/m3
TproProduction water temperature, °CvβDarcy velocity in phase β, m/s
VnSubdomain of the flow systemμβDynamic coefficient of viscosity, Pa·s
WePower generation, MWσHMaximum horizontal stress, MPa
WpPump power, MWσVVertical principal stress, MPa
zDepth, mσhMinimum horizontal stress, MPa

References

  1. Tester, J.; Livesay, B.; Anderson, B.; Moore, M.; Bathchelor, A.; Nichols, K.; Blackwell, D.; Petty, S.; DiPippo, D.; Toksoz, M.; et al. The Future of Geothermal Energy: Impact of Enhanced Geothermal Systems (EGS) on the United States in the 21st Century; An Assessment by an MIT-Led Interdisciplinary Panel; Massachusetts Institute of Technology: Cambridge, MA, USA, 2006. [Google Scholar]
  2. Li, S.; Wang, S.; Tang, H. Stimulation mechanism and design of enhanced geothermal systems: A comprehensive review. Renew. Sustain. Energy Rev. 2022, 155, 111914. [Google Scholar] [CrossRef]
  3. Wang, J.; Hu, S.; Pang, Z.; He, L.; Zhao, P.; Zhu, C.; Rao, S.; Tang, X.; Kong, Y.; Luo, L.; et al. Estimate of geothermal resources potential for hot dry rock in the continental area of China. Sci. Technol. Rev. 2012, 30, 25–31. (In Chinese) [Google Scholar]
  4. Zhang, Y.; Zhang, Y.J.; Zhou, L.; Lei, Z.H.; Guo, L.L.; Zhou, J. Reservoir stimulation design and evaluation of heat exploitation of a two-horizontal-well enhanced geothermal system (EGS) in the Zhacang geothermal field, Northwest China. Renew. Energy 2022, 183, 330–350. [Google Scholar] [CrossRef]
  5. Hofmann, H.; Babadagli, T.; Zimmermann, G. Hot water generation for oil sands processing from enhanced geothermal systems: Process simulation for different hydraulic fracturing scenarios. Appl. Energy 2016, 113, 524–547. [Google Scholar] [CrossRef]
  6. Barenblatt, G.; Zheltov, I.; Kochina, I. Basic concepts in the theory of seepage of homogeneous liquids in fissured rocks. J. Appl. Math. Mech. 1960, 24, 1286–1303. [Google Scholar] [CrossRef]
  7. Warren, J.E.; Root, J.P. The Behavior of Naturally Fractured Reservoirs. Soc. Pet. Eng. J. 1963, 228, 245–255. [Google Scholar] [CrossRef] [Green Version]
  8. Zeng, Y.; Su, C.Z.; Wu, N.Y. Numerical simulation of heat production potential from hot dry rock by water circulating through two horizontal wells at Desert Peak geothermal field. Energy 2013, 10, 92–107. [Google Scholar] [CrossRef]
  9. Suzuki, A.; Fomin, S.A.; Chugunov, V.A.; Niibori, Y.; Hashida, T. Fractional diffusion modeling of heat transfer in porous and fractured media. Int. J. Heat Mass Transf. 2016, 103, 611–618. [Google Scholar] [CrossRef]
  10. Gong, F.; Guo, T.; Sun, W.; Li, Z.; Yang, B.; Chen, Y.; Qu, Z. Evaluation of geothermal energy extraction in Enhanced Geothermal System (EGS) with multiple fracturing horizontal wells (MFHW). Renew. Energy 2020, 151, 1339–1351. [Google Scholar] [CrossRef]
  11. Zhou, D.J.; Tatomir, A.; Niemi, A.; Tang, C.F.; Sauter, M. Study on the influence of randomly distributed fracture aperture in a fracture network on heat production from an enhanced geothermal system (EGS). Energy 2022, 250, 123–781. [Google Scholar] [CrossRef]
  12. Pandey, S.N.; Chaudhuri, A.; Kelkar, S. A coupled thermo-hydro-mechanical modeling of fracture aperture alteration and reservoir deformation during heat extraction from a geothermal reservoir. Geothermics 2017, 65, 17–31. [Google Scholar] [CrossRef]
  13. Soltani, M.; Kashkooli, F.M.; Souri, M.; Rafiei, B.; Jabarifar, M.; Gharali, K.; Nathwani, J.S. Environmental, economic, and social impacts of geothermal energy systems. Renew. Sustain. Energy Rev. 2021, 140, 110–750. [Google Scholar] [CrossRef]
  14. Guo, T.; Gong, F.; Wang, X.; Lin, Q.; Qu, Z.; Zhang, W. Performance of enhanced geothermal system (EGS) in fractured geothermal reservoirs with CO2 as working fluid. Appl. Eng. 2019, 152, 215–230. [Google Scholar] [CrossRef]
  15. Huang, W. Heat extraction performance of EGS with heterogeneous reservoir: A numerical evaluation. Int. J. Heat Mass Transf. 2017, 13, 645–657. [Google Scholar] [CrossRef] [Green Version]
  16. Guo, B.; Fu, P.; Hao, Y.; Peters, C.A.; Carrigan, C.R. Thermal drawdown-induced flow channeling in a single fracture in EGS. Geothermics 2016, 61, 46–62. [Google Scholar] [CrossRef] [Green Version]
  17. Asai, P.; Panja, P.; McLennan, J.; Deo, M. Effect of different flow schemes on heat recovery from Enhanced Geotherm al Systems (EGS). Energy 2019, 175, 667–676. [Google Scholar] [CrossRef]
  18. Patterson, J.R.; Cardiff, M.; Feig, K.L. Optimizing geothermal production in fractured rock reservoirs under uncertainty. Geothermics 2020, 88, 101–906. [Google Scholar] [CrossRef]
  19. Wu, X.T.; Li, Y.C.; Tang, C.A. Fracture spacing in horizontal well multi-perforation fracturing optimized by heat extraction. Geothermics 2022, 101, 102–376. [Google Scholar] [CrossRef]
  20. Fisher, A. A mathematic examination of the methods of determining the accuracy of an observation by the mean error and by the mean square error. Mon. Not. R Astron. Soc. 1920, 80, 758–770. [Google Scholar] [CrossRef] [Green Version]
  21. Han, Y.S.; He, J.T.; Lu, Y.T. Sensitivity of the properties of the graduated compression stocking and soft tissues on the lower limb-stocking interfacial pressure using the orthogonal simulation test. Med. Eng. Phys. 2021, 95, 84–89. [Google Scholar] [CrossRef]
  22. Lin, R.; Diao, X.Y.; Ma, T.C.; Tang, S.H.; Chen, L.; Liu, D.C. Optimized microporous layer for improving polymer exchange membrane fuel cell performance using orthogonal test design. Appl. Energy 2019, 254, 113–714. [Google Scholar] [CrossRef]
  23. Xie, J.X.; Wang, J.S. Performance optimization of pinnate horizontal well in geothermal energy utilization with orthogonal test. Appl. Therm. Eng. 2022, 209, 118–321. [Google Scholar] [CrossRef]
  24. Yu, L.K.; Wu, X.T.; Hassan, N.S.; Wang, Y.D.; Ma, W.W.; Liu, G. Modified zipper fracturing in enhance d geothermal system reservoir and analyzed heat extraction optimization via orthogonal design. Renew. Energy 2020, 161, 373–385. [Google Scholar] [CrossRef]
  25. Pruess, K.; Oldenburg, C.; Moridis, G. TOUGH2 User’s Guide, Version 2.0; Lawrence Berkeley National Laboratory: Berkeley, CA, USA, 1999. [Google Scholar]
  26. Pruess, K.; Faybishenho, B.; Bodvarsson, G.S. Alternative concepts and approaches for modeling flow and transport in thick unsaturated zones of fractured rocks. J. Contam. Hydrol. 1999, 38, 281–322. [Google Scholar] [CrossRef]
  27. Wu, L.; Liu, J.; Zhou, J.; Zhang, Q.; Song, Y.; Du, S.; Tian, W. Evaluation of tar from the microwave co-pyrolysis of low-rank coal and corncob using orthogonal-test-based grey relational analysis (GRA). J. Clean. Prod. 2022, 337, 130–362. [Google Scholar] [CrossRef]
  28. Gao, J.; Zhang, H.J.; Zhang, S.Q.; Chen, X.B.; Cheng, Z.P.; Jia, X.F. Three-dimensional magnetotelluric imaging of the geothermal system beneath the Gonghe Basin, Northeast Tibetan Plateau. Geothermics 2018, 76, 15–25. [Google Scholar] [CrossRef]
  29. Zhang, C.; Jiang, G.; Shi, Y.; Wang, Z.; Wang, Y.; Li, S.; Jia, X.; Hu, S. Terrestrial heat flow and crustal thermal structure of the Gonghe-Guide area, northeastern Qinghai-Tibetan plateau. Geothermics 2018, 72, 182–192. [Google Scholar] [CrossRef]
  30. Zhang, S.Q.; Li, X.F.; Song, J.; Wen, D.G.; Li, Z.W.; Li, D.M.; Cheng, Z.; Fu, L.; Zhang, L.; Feng, Q.; et al. Analysis on Geophysical Evidence for Existence of Partial Melting Layer in Crust and Regional Heat Source Mechanism for Hot Dry Rock Resources of Gonghe Basin. Earth Sci. 2021, 46, 1416–1436. [Google Scholar]
  31. Genter, A.; Fritsch, D.; Cuenot, N.; Baumgärtner, J.; Graff, J.J. Overview of the current activities of the European EGS Soultz project: From exploration to electricity production. In Proceedings of the Thirty-Fourth Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, CA, USA, 9–11 February 2009. SGP-TR-187. [Google Scholar]
  32. U.S. Department of Energy. Geothermal Technologies Program 2013 Peer Review; U.S. Department of Energy: Washington, DC, USA, 2013.
  33. Lei, Z.H.; Zhang, Y.J. Investigation on the effect of symmetrical multi-well layout on geothermal energy extraction from a fractured granitic reservoir: A case study in the Gonghe Basin, Northwestern China. Energy Rep. 2021, 7, 7741–7758. [Google Scholar] [CrossRef]
  34. Huang, X.X.; Zhu, J.L.; Niu, C.K.; Li, J.; Hu, X.; Jin, X.P. Heat extraction and power production forecast of a prospective enhanced geothermal system site in Songliao Basin. China Energy 2014, 75, 360–370. [Google Scholar] [CrossRef]
  35. Zhang, Y.J.; Li, Z.W.; Yu, Z.W.; Guo, L.L.; Jin, X.P.; Xu, T.F. Evaluation of developing an enhanced geothermal heating system in northeast China: Field hydraulic stimulation and heat production forecast. Energy Build. 2015, 88, 1–14. [Google Scholar] [CrossRef]
  36. Zoback, M.D. Reservoir Geomechanics, 1st ed.; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2007; p. 464. [Google Scholar]
  37. Hofmann, H.; Weides, S.; Babadagli, T.; Zimmermann, G.; Moeck, I.; Majorowicz, J.; Unsworth, M. Potential for enhanced geothermal systems in Alberta, Canada. Energy 2014, 69, 578–591. [Google Scholar] [CrossRef]
  38. Garg, S.K.; Combs, J. A reformulation of USGS volumetric “heat in place” resource heat in place” resource estimation method. Geothermics 2015, 55, 150–158. [Google Scholar] [CrossRef]
  39. NIST. Thermophysical Properties of Fluid Systems. 2010. Available online: http://webbook.nist.gov/chemistry/fluid/ (accessed on 1 January 2022).
  40. Guo, L.L.; Zhang, Y.B.; Zhang, Y.J.; Yu, Z.W.; Zhang, J.N. Experimental investigation of granite properties under different temperatures and pressures and numerical analysis of damage effect in enhanced geothermal system. Renew. Energy 2018, 126, 107–125. [Google Scholar] [CrossRef]
  41. Li, Z.W.; Feng, X.T.; Zhang, Y.J.; Xu, T.F. Feasibility study of developing a geothermal heating system in naturally fractured formations: Reservoir hydraulic properties determination and heat production forecast. Geothermics 2018, 73, 1–15. [Google Scholar] [CrossRef]
  42. Cheng, W.L.; Wang, C.L.; Nian, Y.L.; Han, B.B.; Liu, J. Analysis of influencing factors of heat extraction from enhanced geothermal systems considering water losses. Energy 2016, 115, 274–288. [Google Scholar] [CrossRef]
  43. Sanyal, S.K.; Butler, S.J. An analysis of power generation prospects from enhanced geothermal systems. In Proceedings of theWorld Geothermal Congress, Antalya, Turkey, 24–29 April 2005; pp. 1–6. [Google Scholar]
Figure 1. Methodological flowchart.
Figure 1. Methodological flowchart.
Sustainability 14 07001 g001
Figure 2. Location of Gonghe Basin (a). Distribution of geothermal wells and faults in the Qiabuqia area (b).
Figure 2. Location of Gonghe Basin (a). Distribution of geothermal wells and faults in the Qiabuqia area (b).
Sustainability 14 07001 g002
Figure 3. Geological section in the Qiabuqia area.
Figure 3. Geological section in the Qiabuqia area.
Sustainability 14 07001 g003
Figure 4. Schematic of EGS (left); operation diagram of surface ORC system (right up); conceptual model of stimulated EGS reservoir (right down).
Figure 4. Schematic of EGS (left); operation diagram of surface ORC system (right up); conceptual model of stimulated EGS reservoir (right down).
Sustainability 14 07001 g004
Figure 5. Subgridding in the MINC method (left) and the 3D numerical model (right).
Figure 5. Subgridding in the MINC method (left) and the 3D numerical model (right).
Sustainability 14 07001 g005
Figure 6. Evolutions of Tpro and Pinj at different injection rates (qinj) over the 50 years.
Figure 6. Evolutions of Tpro and Pinj at different injection rates (qinj) over the 50 years.
Sustainability 14 07001 g006
Figure 7. Evolutions of Tpro under different parameters over the 50 years.
Figure 7. Evolutions of Tpro under different parameters over the 50 years.
Sustainability 14 07001 g007
Figure 8. Evolutions of We under different parameters over the 50 years.
Figure 8. Evolutions of We under different parameters over the 50 years.
Sustainability 14 07001 g008
Figure 9. Evolutions of Pinj under different parameters over the 50 years.
Figure 9. Evolutions of Pinj under different parameters over the 50 years.
Sustainability 14 07001 g009
Figure 10. Evolutions of Iinj under different parameters over the 50 years.
Figure 10. Evolutions of Iinj under different parameters over the 50 years.
Sustainability 14 07001 g010
Figure 11. Evolutions of Wp under different parameters over the 50 years.
Figure 11. Evolutions of Wp under different parameters over the 50 years.
Sustainability 14 07001 g011
Figure 12. Evolutions of η under different parameters over the 50 years.
Figure 12. Evolutions of η under different parameters over the 50 years.
Sustainability 14 07001 g012
Figure 13. Sensitivity analysis results of different factors on each index.
Figure 13. Sensitivity analysis results of different factors on each index.
Sustainability 14 07001 g013
Figure 14. Variations of We, Wp, and η at four levels of different factors based on the results in Table 8. G1–G4 represent the results of the four tests at each level: (a) fracture permeability; (b) fracture permeability anisotropy; (c) injected section length; and (d) injection rate.
Figure 14. Variations of We, Wp, and η at four levels of different factors based on the results in Table 8. G1–G4 represent the results of the four tests at each level: (a) fracture permeability; (b) fracture permeability anisotropy; (c) injected section length; and (d) injection rate.
Sustainability 14 07001 g014
Figure 15. Simulation results of three combinations of A2B3C1D1, A2B2C3D1, and A2B3C4D1.
Figure 15. Simulation results of three combinations of A2B3C1D1, A2B2C3D1, and A2B3C4D1.
Sustainability 14 07001 g015
Figure 16. Comparison of net power generation (WeWp) under different of injection rates based on test A2B3C4D1.
Figure 16. Comparison of net power generation (WeWp) under different of injection rates based on test A2B3C4D1.
Sustainability 14 07001 g016
Figure 17. Schematic diagram of optimal well layout designed according to case A2B3C4D1.
Figure 17. Schematic diagram of optimal well layout designed according to case A2B3C4D1.
Sustainability 14 07001 g017
Figure 18. Comparison of production index under different injection temperature in the 50th year based on test A2B3C4D1. The production index include Tpro and We (a), Iinj and Pinj (b), η and Wp (c), and production water density, ρwater (d).
Figure 18. Comparison of production index under different injection temperature in the 50th year based on test A2B3C4D1. The production index include Tpro and We (a), Iinj and Pinj (b), η and Wp (c), and production water density, ρwater (d).
Sustainability 14 07001 g018aSustainability 14 07001 g018b
Figure 19. Profiles of space variations of reservoir temperature fields over 50 years (y = 350 m) (ae) and the temperature plan in the 50th year (z = 350 m) (f). The results are based on test A2B3C4D1 with injection temperature of 70 °C.
Figure 19. Profiles of space variations of reservoir temperature fields over 50 years (y = 350 m) (ae) and the temperature plan in the 50th year (z = 350 m) (f). The results are based on test A2B3C4D1 with injection temperature of 70 °C.
Sustainability 14 07001 g019
Table 1. Orthogonal test with 3 factors and 3 levels.
Table 1. Orthogonal test with 3 factors and 3 levels.
LevelFactor
ABC
Level la1b1c1
Level 2a2b2c2
Level 3a3b3c3
Table 2. Specific trials of orthogonal test for 3 factors and 3 levels.
Table 2. Specific trials of orthogonal test for 3 factors and 3 levels.
Test NumberFactor
ABC
1a1b1c1
2a1b2c2
3a1b3c3
4a2b1c3
5a2b2c1
6a2b3c2
7a3b1c2
8a3b2c3
9a3b3c1
Table 3. Measured temperature data of four geothermal wells.
Table 3. Measured temperature data of four geothermal wells.
WellsDepth, z (km)Measured Temperature, T (°C)
GR10.1~1.0T = 66.061z + 18.467 (R2 = 0.9958)
1.1~2.8T = 40.289z + 48.992 (R2 = 0.9967)
2.9~3.6T = 57.738z − 5.0238 (R2 = 0.9861)
GR20.1~3.0T = 50.154z + 33.795 (R2 = 0.9996)
DR30.1~1.4T = 72.879z + 14.198 (R2 = 0.9988)
1.5~2.9T = 44.357z + 55.081 (R2 = 0.996)
DR40.1~0.5T = 7z + 77.3 (R2 = 0.9423)
0.6~1.5T = 8.303z + 99.782 (R2 = 0.91)
1.6~3.1T = 44.456z + 46.466 (R2 = 0.9983)
Table 4. Heat conductivity of different lithologies with depth.
Table 4. Heat conductivity of different lithologies with depth.
LithologyDepth (km)Heat Conductivity (W/(m·K))
Mudstone0.20~1.401.25~1.99 (average is 1.58)
Igneous Rock1.50–3.632.10–3.17 (average is 2.53)
Table 5. EGS reservoir properties and operation parameters.
Table 5. EGS reservoir properties and operation parameters.
ItemsParametersBase Case ValueSelected Case Value
ReservoirFracture spacing, Sf3 m50 m
Fracture permeability (kx = ky = kz), Pf1 × 10−13 m21 × 10−11 m2
Fracture permeability anisotropy, α
(α = kx/kz, kx = ky),
110
Matrix permeability, Pr1 × 10−17 m21 × 10−14 m2
Matrix porosity0.025No changed
Matrix density2360 kg/m3No changed
Heat conductivity, λr2.0 W/(m·K)3.5 W/(m·K)
Specific heat754.4 J/(kg·K)No changed
Initial pressureP = 4 × 10−7–10,000z (Pa)No changed
Initial temperature225 °CNo changed
OperationInjected section length, Linj60 m120 m
Injection rate, qinj30 kg/s20 kg/s
Injection water temperature10 °CNo changed
Injection water specific enthalpy78.77 kJ/kgNo changed
Productivity index5.4 × 10−12 m3No changed
Production bottom-hole pressure30 MPaNo changed
Table 6. Four-factor (A~D) and four-level (1~4).
Table 6. Four-factor (A~D) and four-level (1~4).
LevelFactors
ABCD
Fracture Permeability, Pf (m2)Fracture Permeability Anisotropy, αInjected Section Length, Linj (m)Injection Rate, qinj (kg/s)
15 × 10−1413020
21 × 10−131012050
31 × 10−1210021080
41 × 10−111000300110
Table 7. Orthogonal test scheme.
Table 7. Orthogonal test scheme.
Test NumberFactors
ABCD
Fracture Permeability, Pf (m2)Fracture Permeability Anisotropy, αInjected Section Length, Linj (m)Injection Rate, qinj (kg/s)
1A1 (5 × 10−14)B1 (1)C1 (30)D1 (20)
2A1 (5 × 10−14)B2 (10)C2 (120)D2 (50)
3A1 (5 × 10−14)B3 (100)C3 (210)D3 (80)
4A1 (5 × 10−14)B4 (1000)C4 (300)D4 (110)
5A2 (1 × 10−13)B1 (1)C2 (120)D3 (80)
6A2 (1 × 10−13)B2 (10)C1 (30)D4 (110)
7A2 (1 × 10−13)B3 (100)C4 (300)D1 (20)
8A2 (1 × 10−13)B4 (1000)C3 (210)D2 (50)
9A3 (1 × 10−12)B1 (1)C3 (210)D4 (110)
10A3 (1 × 10−12)B2 (10)C4 (300)D3 (80)
11A3 (1 × 10−12)B3 (100)C1 (30)D2 (50)
12A3 (1 × 10−12)B4 (1000)C2 (120)D1 (20)
13A4 (1 × 10−11)B1 (1)C4 (300)D2 (50)
14A4 (1 × 10−11)B2 (10)C3 (210)D1 (20)
15A4 (1 × 10−11)B3 (100)C2 (120)D4 (110)
16A4 (1 × 10−11)B4 (1000)C1 (30)D3 (80)
Table 8. Orthogonal test results. VEnd and VAve. denote the value at the end of simulation and the average annual value of each index, respectively.
Table 8. Orthogonal test results. VEnd and VAve. denote the value at the end of simulation and the average annual value of each index, respectively.
Test NumberIndex
Power Generation, WeCoefficient of Performance, ηProduction Temperature, TproInjectivity, IinjBottom-Hole Pressure, PinjPump Power, Wp
1VEnd
VAve.
1.00
1.45
2.38
3.63
176.46
208.11
1.13
1.17
30.69
30.61
0.42
0.40
2VEnd
VAve.
−1.60
0.33
−0.88
0.21
52.87
110.62
1.79
2.12
60.27
55.82
1.81
1.54
3VEnd
VAve.
−1.24
0.49
−0.28
0.20
77.53
110.07
1.96
2.32
74.69
68.63
4.38
3.77
4VEnd
VAve.
−3.35
−1.44
−1.01
−0.37
54.78
81.09
5.01
5.62
55.83
53.46
3.30
2.99
5VEnd
VAve.
−2.62
−0.36
−1.55
0.13
51.67
93.97
4.68
5.42
49.62
46.90
1.69
1.48
6VEnd
VAve.
------
7VEnd
VAve.
1.11
1.46
35.26
51.99
183.40
208.79
17.43
19.28
33.65
33.44
0.03
0.03
8VEnd
VAve.
−0.86
0.67
−0.31
0.57
222.41
223.83
1.22
1.69
74.70
64.89
−0.86
0.67
9VEnd
VAve.
−4.96
−3.74
−20.56
−16.04
32.41
49.37
45.66
48.45
36.71
36.00
0.24
0.22
10VEnd
VAve.
−2.29
-0.88
−24.98
−7.97
57.69
84.32
95.71
110.12
34.60
34.12
0.09
0.08
11VEnd
VAve.
−1.27
0.88
−2.08
2.80
62.2
126.6
4.39
5.15
45.92
42.59
0.61
0.50
12VEnd
VAve.
0.43
1.14
5.55
12.41
132.38
185.98
2.65
3.63
38.82
36.93
0.16
0.11
13VEnd
VAve.
−1.68
−0.91
−377.26
−439.38
50.34
73.60
897.44
892.88
33.10
32.92
0.0038
0.0038
14VEnd
VAve.
0.05
0.52
18.77
420.36
104.76
140.23
22.34
22.90
32.68
32.58
0.0024
0.0019
15VEnd
VAve.
−4.28
−1.71
−25.12
−6.44
41.83
77.13
40.07
41.85
36.94
35.94
0.17
0.15
16VEnd
VAve.
−3.17
−0.88
−2.67
0.05
39.48
83.07
6.04
7.23
53.71
48.87
1.19
0.99
Table 9. Range analysis results of the orthogonal test. For simplicity, only all ki values of We and η are listed in the table. The other indexes only give R and rank results.
Table 9. Range analysis results of the orthogonal test. For simplicity, only all ki values of We and η are listed in the table. The other indexes only give R and rank results.
IndexValueFactors
ABCD
Wek10.21−0.890.481.26
k20.59−0.01−0.030.24
k3−0.530.28−0.52−0.41
k4−0.75−0.01−0.44−2.30
R1.341.170.103.56
Rank2341
Better levelA2B3C1D1
ηk1−112.920.922.16124.52
k2137.5317.564.00−108.95
k312.140.22101.27−1.8975
k45.59−6.35−98.93−7.62
R250.4523.92200.21233.47
Rank1432
Better levelA2B2C3D1
TproR82.0245.6927.31125.04
Rank2341
IinjR238.41232.21252.46213.49
Rank2314
PinjR15.3013.8512.0416.82
Rank2341
WpR1.960.660.541.45
Rank1342
Table 10. Results of single-factor (α) three-level test based on test A2B3C4D1.
Table 10. Results of single-factor (α) three-level test based on test A2B3C4D1.
TestWeηTproIinjPinjWp
A2B1C4D1 (α = 1)VEnd0.9769.29173.0719.6333.030.014
VAve.1.2364.74192.2820.9233.030.019
A2B2C4D1 (α = 10)VEnd1.0263.75177.1819.2933.10.016
VAve.1.3567.50201.0220.6933.070.02
A2B3C4D1 (α = 100)VEnd1.1135.26183.417.4333.650.03
VAve.1.4651.99208.7919.2833.440.03
A2B4C4D1 (α = 1000)VEnd1.016.20174.8812.3538.830.163
VAve.1.3711.42202.7714.2837.060.12
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, Y.; Shu, L.; Chen, S.; Zhao, J.; Guo, L. Optimization Design of Multi-Factor Combination for Power Generation from an Enhanced Geothermal System by Sensitivity Analysis and Orthogonal Test at Qiabuqia Geothermal Area. Sustainability 2022, 14, 7001. https://doi.org/10.3390/su14127001

AMA Style

Zhao Y, Shu L, Chen S, Zhao J, Guo L. Optimization Design of Multi-Factor Combination for Power Generation from an Enhanced Geothermal System by Sensitivity Analysis and Orthogonal Test at Qiabuqia Geothermal Area. Sustainability. 2022; 14(12):7001. https://doi.org/10.3390/su14127001

Chicago/Turabian Style

Zhao, Yuan, Lingfeng Shu, Shunyi Chen, Jun Zhao, and Liangliang Guo. 2022. "Optimization Design of Multi-Factor Combination for Power Generation from an Enhanced Geothermal System by Sensitivity Analysis and Orthogonal Test at Qiabuqia Geothermal Area" Sustainability 14, no. 12: 7001. https://doi.org/10.3390/su14127001

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