Next Article in Journal
CO2-Enhanced Gas Recovery (EGR) in Offshore Carbon-Rich Gas Reservoirs—Part 2: EGR Performance and Its Dependency
Next Article in Special Issue
Simulating Horizontal CO2 Plume Migration in a Saline Aquifer: The Effect of Injection Depth
Previous Article in Journal
Development of an LC-MS Method for the Quantitative Determination of Six Organic Acids Produced by Bacterial Fermentation In Vitro and In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Liquid Nitrogen on Mechanical Deterioration and Fracturing Efficiency in Hot Dry Rock

1
Xinjiang Yaxin Coalbed Methane Investment and Development (Group) Co., Ltd., Urumqi 830063, China
2
CNPC Engineering Technology R&D Company Limited, Beijing 102206, China
3
YunLong Lake Laboratory of Deep Underground Science and Engineering, Xuzhou 221006, China
4
School of Science, Qingdao University of Technology, Qingdao 266520, China
*
Author to whom correspondence should be addressed.
Processes 2025, 13(3), 696; https://doi.org/10.3390/pr13030696
Submission received: 7 February 2025 / Revised: 18 February 2025 / Accepted: 26 February 2025 / Published: 28 February 2025

Abstract

:
Geothermal energy can be obtained from hot dry rock (HDR). The target temperatures for heat extraction from HDR range from 100 to 400 °C. Artificial fracturing is employed to stimulate HDR and create a network of fractures for geothermal resource extraction. Liquid nitrogen (LN2) is environmentally friendly and shows better performance in reservoir stimulation than does conventional fracturing. In this study, triaxial compression experiments and acoustic emission location techniques were used to evaluate the impacts of temperatures and confining pressures on the mechanical property deterioration caused by LN2 cooling. The numerical simulation of LN2 fracturing was performed, and the results were compared with those for water and nitrogen fracturing. The results demonstrate that the confining pressure mitigated the deterioration effect of LN2 on the crack initiation stress, crack damage stress, and peak stress. From 20 to 60 MPa, LN2-induced reductions in these three stress parameters ranged between 7.73–18.51%, 3.46–12.15%, and 2.51–8.50%, respectively. Cryogenic LN2 increased the number and complexity of cracks generated during rock failure, further enhancing the fracture performance. Compared with those for water and nitrogen fracturing, the initiation pressures of LN2 fracturing decreased by 61.54% and 68.75%, and the instability pressures of LN2 fracturing decreased by 20.00% and 29.41%, respectively. These results contribute to the theoretical foundation for LN2 fracturing in HDR.

1. Introduction

Geothermal energy, as a low-carbon and stable source of energy, stands out as a highly promising option among sustainable energy alternatives for achieving low-carbon energy transitions [1,2,3,4,5]. Geothermal energy can be obtained from hot dry rock (HDR) that has a temperature of 180–650 °C and is located at a depth of more than 3 km [4,5,6]. The target temperatures for heat extraction from HDR are in the range of 100–400 °C [4,5,7]. The abundance of energy in HDR provides a sustainable solution to energy supply [4,5,8,9]. Artificial fracturing is used to stimulate HDR and form a network of fractures, thereby establishing a circulation system between injection and production wells that allows for the economical extraction of deep geothermal energy from HDR [10,11]. The utilization of liquid nitrogen (LN2) has been developed for various new fracturing techniques that have good prospects for fracturing performance enhancement [12,13,14,15].
Unlike normal rock, the temperature of HDR can promote the deteriorating effects of LN2 on rock properties. Wu et al. [16], Kang et al. [17], and Sha et al. [18] evaluated the influences of air, water, and LN2 cooling on granite mechanical performance at different temperatures and found that LN2 cooling can significantly damage granites. Wu et al. [19,20] investigated the LN2 effects on the wave velocity, permeability, and compression strength of rocks (granite, shale, and sandstone) at different temperatures (25, 150, and 260 °C) and found that increasing the rock temperature can strengthen the LN2’s thermal effect. Hou et al. [21] and Su et al. [22] analyzed the LN2 effect on the mechanical performance of marble at different temperatures (25–300 °C) and also found temperature enhancement. Kang et al. [17] and Sha et al. [18] found that the mechanical property deterioration is more prominent above 400 °C. Wu et al. [23] and Rong et al. [24] found that increasing the cycle number of LN2 treatments caused the continuous deterioration of rocks. Besides focusing on the mechanical parameters, Sha et al. [18], Ge et al. [25], and Rong et al. [24] analyzed the impact of LN2 treatment on the acoustic emission (AE) parameters during rock failure. In addition, Wu et al. [20] found that confining pressure weakens the LN2 effect. Most studies have failed to considered the confining pressure effect. For HDR under high temperatures and high in-situ stress, one must consider the confining pressure to reveal LN2’s deterioration effect. Further research on fracturing performance is needed for LN2 to be applied to the fracturing of HDR reservoirs. Laboratory-conducted LN2 fracturing is carried out under low confining pressures [26]. Experiments with high confining pressures are difficult to implement. Numerical simulations can solve this difficulty by setting up conditions closer to site status. Therefore, this study investigated the mechanical performance of granite after LN2 cooling under various temperatures and confining pressures via experiments and the fracturing performance of LN2 via numerical simulations.
To explore the effectiveness of LN2, triaxial compression experiments were performed on heated granite after air and LN2 cooling. The impacts of different temperatures and confining pressures on mechanical property deterioration with LN2 cooling were investigated. The numerical simulation of LN2 fracturing was performed to analyze the fracturing performance enhancement.

2. Experiment and Simulation

2.1. Thermal Treatments and Mechanical Property Experiments

The granite was taken from an open-pit mining site in Rizhao, China. The mineral content was quartz (25%), plagioclase (39%), potassium feldspar (22%), and biotite (14%), based on the X-ray diffraction test results. The granite was prepared in 50 × 100 mm cylindrical samples (Figure 1a), according to the standards of the International Society for Rock Mechanics. By density and ultrasonic testing, samples with a low dispersion of test results were selected for the experiment. The samples were subjected to a thermal process (Figure 1b) before testing their mechanical properties. The heating temperatures were 25, 200, and 400 °C. The samples were first heated to the predetermined temperatures at a slow rate (2–4 °C/min) and then held for 2 h to ensure that the internal temperature of the sample was uniform. After the conclusion of the maintenance time, two cooling treatments were used, i.e., the air-cooling treatment and the LN2-cooling treatment. The samples in the air-cooling group underwent natural cooling under ambient air conditions. The samples in the LN2-cooling group were cooled by immersion in LN2 tanks for 1 h. After the heating and cooling treatments, all samples were left under ambient air conditions for 24 h prior to mechanical testing, ensuring that the samples returned to room temperature.
For the samples after thermal treatment, triaxial compression experiments were performed using the TAW-2000 experimental system (Changchun Chaoyang Test Instrument Co., Ltd., Changchun, China) (Figure 1c). The confining pressure was initially applied at 0.05 MPa/s until it reached predetermined levels (0, 20, 40, and 60 MPa), after which axial displacement loading was initiated at 0.2 mm/min. Synchronized with the loading progress, the AE locations were collected via the PCI-2 AE test system (Figure 1d).
In the experiment, for the sample deformation process, the stress parameters for crack initiation and damage were identified using the stress–strain curve, according to the crack volume strain method [4,5] (Figure 1e). Crack initiation represents the stress level at which cracks begin and is marked according to the reversal points on the curve of crack volumetric strain [4,5]. Crack damage represents the stress level at which unstable crack growth occurs and is marked according to the reversal points on the curve of volumetric strain [4,5].

2.2. Numerical Simulation of Liquid Nitrogen Fracturing

The LN2 fracturing for HDR reservoirs is a typical multi-physical field coupling problem. The cryogenic LN2 injection process causes changes in the rock deformation, along with the damage, temperature, and fluid flow, and each physical field influences the other during the change process. Multi-physical field coupling problems can be solved by numerical methods. In this study, the fracturing process was simulated using COMSOL Multiphysics 5.6 software.

2.2.1. Governing Equations

If LN2 is applied to fracturing for HDR reservoirs, one must consider the physical changes in the LN2. The equation of state for LN2 can be expressed as [27]
A ρ , T R T = α δ , τ = α 0 δ , τ + α r δ , τ
where A is the Helmholtz energy, α is the dimensionless Helmholtz energy, α 0 is the ideal gas contribution to the Helmholtz energy [27], α r is the residual Helmholtz energy [27], ρ is the density, T is the temperature, R = 8.314510 J/(mol·K), δ = ρ / ρ c , τ = T c / T , ρ c is 11.1839 mol/dm3, and T c is 126.192 K.
The density and isobaric heat capacity equations of Span et al. [27], as well as the viscosity and thermal conductivity equations of Lemmon and Jacobsen [28], were adopted for LN2.
Darcy’s law was adopted to calculate the flow of fracturing media in the HDR reservoirs. The governing equation of fluid flow during the fracturing process is derived using the work of Zhou et al. [29]. For slightly compressible fluids, the fluid flow equation can be expressed as [29]
ϕ β f + α B ϕ 1 + S 1 K s p t ϕ α f + α B ϕ 1 + S α T T t k μ f p + α B ϕ 1 + S ε V t = 0
For compressible fluids, the fluid flow equation can be expressed as [29]
ϕ β f + α B ϕ 1 + S 1 K s p t ϕ α f + α B ϕ 1 + S α T T t k μ f p β f k μ f p p + α f k μ f T p + α B ϕ 1 + S ε V t = 0
where ϕ is the porosity, β f is the fluid compressibility coefficient, α B is Biot’s coefficient, S = ε V + p / K s α T T T 0 , ε V is the volumetric strain, p is the pore pressure, K s is the modulus of the solid matrix, α T is the volumetric thermal expansion coefficient, T 0 is the initial temperature, t is the time, α f is the fluid thermal expansion coefficient, k is the permeability, and μ f is the dynamic viscosity of a fluid.
The energy conservation equation for heat transfer is expressed as [30]
ρ C p eff T t + ρ f C p , f q T λ eff T = Q T T K α T ε V t
where ρ C p eff = ϕ ρ f C p , f + 1 ϕ ρ s C p , s , ρ f is the density of the fluids, C p , f is the isobaric heat capacity of the fluids, ρ s is the rock density, C p , s is the isobaric heat capacity of the rock matrix, λ eff = ϕ λ f + 1 ϕ λ s , λ f is the thermal conductivity of the fluids, λ s is the thermal conductivity of the rock matrix, Q T is the energy sources, and K is the bulk modulus.
The mechanical equilibrium equation can be given as follows (tensile is positive) [31]:
G u i , j j + G 1 2 ν u j , j i α B p , i K α T T , i + f i = 0
where G is the shear modulus, u i is the displacement component, ν is Poisson’s ratio, and f i is the body force component.
Two strength criteria were adopted using the following formula [32]:
F 1 = σ 1 f t = 0 F 2 = σ 3 + σ 1 1 + sin φ 1 sin φ f c = 0
where σ 1 and σ 3 are the principal stresses, f t is the uniaxial tensile strength, f c is the uniaxial compressive strength, and φ is the internal frictional angle.
The damage variable D is defined by [32]
D = 0 F 1 < 0 , F 2 < 0 ε t 0 ε 1 2 1 F 1 = 0 , d ε 1 > 0 1 ε c 0 ε 3 2 F 2 = 0 , d ε 3 > 0
where ε t 0 and ε c 0 are the maximum tensile and compressive principal strains when damage is initiated, respectively, and ε 1 and ε 3 are the principal strains. The tensile damage is the priority.
When a rock element is damaged, its elastic modulus, permeability, and thermal conductivity change accordingly, as follows [32,33]:
E = 1 D E 0
k = k 0 ϕ ϕ 0 3 exp α k D
λ s = λ 0 exp α λ D
where E 0 is the initial elastic modulus, k 0 is the initial permeability, ϕ 0 is the initial porosity, λ 0 is the initial thermal conductivity, and α k and α λ are the amplification factors.
To account for the inherent heterogeneity in the physical and mechanical properties of the rock constituents at the mesoscale, the Weibull distribution [34,35] was implemented to statistically characterize the probabilistic strength distribution across rock elements:
P χ = m χ 0 χ χ 0 m 1 exp χ χ 0 m
where χ is the parameter of mesoscopic elements, χ 0 is the scale parameter, and m is the heterogeneity coefficient.

2.2.2. Model Descriptions

A plane strain model (Figure 2) of 1 × 1 m with a 0.1 m borehole diameter was adopted in this study. The mechanical parameters are presented in Table 1. The initial temperature and initial pore pressure were 200 °C and 30 MPa, respectively. The initial values were maintained for both temperature and pore pressure at the outer boundary. The in-situ stresses were 75 MPa (for a 3000 m depth, according to 25 MPa/km [11]), applied to the outer boundaries. For LN2 fracturing, the borehole boundary was pressurized at 0.5 MPa/s [26,36], and the time step was 20 s/step. The LN2 temperature was 100 K [37]. For water or nitrogen fracturing, both water and nitrogen were injected at 25 °C. The pressurization rate was the same as that used for the liquid nitrogen fracturing.

3. Mechanical Property Deterioration

3.1. Crack Initiation Stress

Figure 3 shows the crack initiation stress of heated granite after air and LN2 cooling under various temperatures and confining pressures. For all samples (Figure 3a), high temperatures led to a reduction in the crack initiation stress, especially above 200 °C, and the temperature sensitivity was weakened by the confining pressure. For instance, the crack initiation stress of the LN2-cooled samples decreased by 46.11% when the sample temperature escalated from 25 to 400 °C under unconfined conditions (0 MPa); in contrast, the crack initiation stress decreased by 6.1% from 25 to 400 °C under the confining pressure of up to 60 MPa. The confining pressure caused the cracks to close and made crack initiation more difficult, thus mitigating the temperature-induced differences [20,38]. In addition, the crack initiation stresses of all samples significantly increased with the increase in confinement pressure, suggesting that the presence of confinement pressure makes rock cracking difficult. The crack initiation stress of the LN2-cooled samples at 200 °C increased by 5.97 times when the confinement pressure was increased from 0 to 60 MPa. Moreover, the crack initiation stresses in the LN2-cooled samples were lower than those of their air-cooled counterparts under different confinement pressures, highlighting the efficacy of LN2 treatment in thermally induced mechanical weakening across varying confining pressures. Figure 3b shows the crack initiation stress reduction rate induced by LN2. The reduction rate induced by LN2 was the highest under unconfined conditions. The reduction rate significantly decreased when the pressure exceeded 20 MPa, indicating that the deterioration effect of LN2 on crack initiation stress was weakened by confining pressure due to crack closure caused by the confining pressure. The reduction rate resulting from LN2 was 7.73–18.51% in the range of 20 to 60 MPa, indicating that the LN2 still had a weakening effect on the rock cracking threshold under high confining pressures.

3.2. Crack Damage Stress

Figure 4 shows the crack damage stress of heated granite after air and LN2 cooling under various temperatures and confining pressures. The temperature effect on the crack damage stress was weakened by the confinement pressure. The crack damage stress of the LN2-cooled samples from 25 to 400 °C decreased by 51.77% under unconfined conditions (Figure 4a) and by 5.33% under the confining pressure of 60 MPa. Under different conditions, the LN2-cooled samples exhibited a reduction in crack damage stress compared to the results for their air-cooled counterparts, attributable to the preferential coalescence of cryogenically induced microcracks under loading conditions. Figure 4b shows the reduction rate of crack damage stress induced by LN2. The reduction rate in crack damage stress caused by LN2 cooling was significant under 0 MPa. The reduction rate caused by LN2 cooling was lower when high confining pressures were applied, which indicates that the presence of confining pressure weakened the deterioration of crack damage stresses induced by LN2. The reduction rate induced by LN2 cooling was 3.46–12.15% when the confining pressure was in the range of 20 to 60 MPa, indicating that LN2 remained effective in reducing the crack damage stress, while the confining pressure generally enhanced the resistance to cracking.

3.3. Peak Stress

Figure 5 presents the peak stress of heated granite after air and LN2 cooling under various conditions. Similar to the crack initiation and damage stresses, the peak stress tended to decrease with temperature, and the confining pressure suppressed the temperature effect. The peak stress of the LN2-cooled samples from 25 to 400 °C decreased by 30.52% under 0 MPa and by 5.5% under 60 MPa (Figure 5a). Additionally, the LN2-cooled samples exhibited reduced peak stress magnitudes compared to those of their air-cooled counterparts, with this weakening effect attributed to thermally induced microcracks and localized stress concentration under cryogenic treatment [22,39]. Figure 5b shows the reduction rate in the peak stress induced by LN2. The reduction rate in the peak stress caused by LN2 cooling was the highest under unconfined conditions and decreased when it exceeded 20 MPa, which indicated that the presence of confinement pressure weakened the deterioration effect of LN2. When the rock is unconstrained, LN2 cooling can cause more stress reduction due to the thermal-induced cracks. The confining pressure limited this thermal-induced damage by restricting the expansion of cracks and maintaining the structural integrity of the material, thus weakening the overall degrading effect of LN2 at higher stress levels. In addition, the weakening of the confinement pressure became stable when the confinement pressure exceeded 20 MPa. In the range of 20–60 MPa, the reduction rate in the peak stress caused by LN2 cooling was in the range of 2.51–8.50%, showing that LN2 cooling could decrease the peak stress under high confining pressures. This indicates that the thermal cracking due to LN2 remained effective in reducing strength, even though the confining pressure led to crack closure and rock densification.

4. Acoustic Emission Event Statistics

In the triaxial compression experiments, the AE events during sample failure were collected using an AE testing system. The AE event distribution can reflect the crack distribution during sample failure. The LN2 effect on the crack distribution of heated granite during sample failure was evaluated in terms of the statistical characteristics of the AE events.

4.1. Event Count

Figure 6 presents the count of AE events of heated granite after air and LN2 cooling under various conditions. Each datum in Figure 6a is the average value of four scenarios (0, 20, 40, and 60 MPa) at the same temperature. Each datum in Figure 6b is the average value of three scenarios (25, 200, and 400 °C) at the same confining pressure.
For all samples, the count of AE events showed an ascending trend with the sample temperature (Figure 6a) and a descending trend with the confining pressure (Figure 6b). The LN2-cooled samples consistently generated a higher count of AE events compared to their air-cooled counterparts across varying temperature and confining pressure regimes. At 25, 200, and 400 °C, the AE event counts after LN2 cooling increased by 19.47%, 129.87%, and 25.66%, respectively, compared with the counts after air cooling. Under 0 and 60 MPa, the AE event counts after LN2 cooling increased by 29.38% and 18.98%, respectively, compared with the counts after air cooling. The AE event counts were an indicator of the number of cracks formed during sample failure. These results revealed that the cryogenic effect of LN2 contributed to an increase in crack formation across various conditions. The cryogenic LN2 promoted crack initiation and propagation, leading to a higher number of cracks during failure and enhancing the rock’s susceptibility to cracking. This behavior indicated the role of cryogenic treatments in altering the rock’s mechanical response, weakening its structural integrity under specific conditions.

4.2. Fractal Dimension

Figure 7 presents the fractal dimension (box counting) of AE events in heated granite after air cooling and LN2 cooling under various conditions. Each datum in Figure 7a is the average of four scenarios (0, 20, 40, and 60 MPa) at the same temperature. Each datum in Figure 7b is the average of three scenarios (25, 200, and 400 °C) at the same confining pressure. The fractal dimension of all samples tended to increase with the sample temperature (Figure 7a) and decrease with confining pressure (Figure 7b). The fractal dimension of the LN2-cooled samples exceeded that of their air-cooled counterparts across varying temperature and confining pressure regimes, indicating that the cryogenic LN2 enhanced crack complexity under diverse conditions. The higher fractal dimension indicates a more intricate crack network, resulting from the LN2-induced thermal damage. This increased crack complexity under LN2 cooling suggested that the cryogenic effect not only influenced the initiation of cracks but also their distribution.

5. Comparison of Liquid Nitrogen Fracturing and Conventional Fracturing

Experimental analysis showed that LN2 increased the number and complexity of cracks and decreased the mechanical strength of high-temperature granite. LN2 fracturing was compared with water and nitrogen fracturing via numerical simulation to reveal the advantages of LN2 applied to fracturing for HDR reservoirs. Figure 8 displays the evolution of the damage distribution during water, nitrogen, and LN2 fracturing, with positive values representing shear damage and negative values representing tensile damage. In the three fracturing models, the damage first appeared around the borehole and gradually extended outward, which is consistent with the fracture propagation during the fracturing process. When the fracturing continued, the fracture zones formed by LN2 fracturing were more uniformly distributed in the radial directions, and the number of fracture zones was more than that of the other two modes. These results indicate that LN2 fracturing contributed to the formation of multi-fractures, which is in agreement with the findings of Yang et al. [6] and proves the model’s correctness.
Figure 9 presents the cumulative number of damage elements and the fractal dimension (box counting) of the damage distribution during water, nitrogen, and LN2 fracturing. The number and fractal dimension of damage due to LN2 fracturing were higher than those of the other two methods, indicating that LN2 fracturing promoted fracture complexity. The temperature difference between the reservoir and LN2 led to significant thermal stresses, combined with the injection pressure, resulting in the formation of a greater number of more complex fractures. There was zero damage at a lower injection pressure. The number of damages increased with the increase in the injection pressure, and the corresponding injection pressure was determined as the initiation pressure [40]. The number and fractal dimension of damages rapidly increased as the injection pressure further increased, indicating that unstable damage occurred. The corresponding injection pressure was determined as the instability pressure [40].
Figure 10 presents the initiation pressure and instability pressure of the water, nitrogen, and liquid nitrogen fracturing models. The initiation and instability pressures of nitrogen fracturing exhibited high values, indicating that a higher breakdown pressure was needed for deep geothermal reservoirs with a high temperature and in-situ stress via mere injection pressure. This also reflected the fracturing difficulty regarding deep geothermal reservoirs. The initiation and instability pressures of water fracturing were lower than those of nitrogen fracturing because thermal stresses were generated to reduce breakdown pressures. However, the breakdown pressures of water fracturing still displayed high values under the effect of thermal stress, which was one of the limiting factors for hydraulic stimulation in deep geothermal reservoirs. Compared with water and nitrogen fracturing, the initiation pressures of LN2 fracturing decreased by 61.54% and 68.75%, and the instability pressures of LN2 fracturing decreased by 20% and 29.41%, respectively. These results suggest that LN2 fracturing offers advantages by reducing the breakdown pressures for the geothermal reservoirs under high temperatures and high in-situ stress conditions.

6. Conclusions

The stress parameters and AE event statistics of heated granite after air cooling and LN2 cooling were compared. Additionally, numerical simulation of LN2 fracturing was performed and compared with the results for water and nitrogen fracturing. The following conclusions are summarized:
(1)
Confining pressure mitigated the effect of temperature on crack initiation stress, crack damage stress, and peak stress. It also reduced the deterioration effect of LN2 cooling on these stress parameters.
(2)
LN2 still reduced stress parameters under high confinement pressures. From 20 to 60 MPa, the reduction rates of crack initiation stress, crack damage stress, and peak stress induced by LN2 cooling were 7.73–18.51%, 3.46–12.15%, and 2.51–8.50%, respectively.
(3)
Cryogenic LN2 increased the counts and fractal dimensions of AE events during sample failure under different conditions.
(4)
LN2 fracturing led to a greater number and complexity of fractures and reduced the breakdown pressures. Compared with water and nitrogen fracturing, the initiation pressures of LN2 fracturing decreased by 61.54% and 68.75%, while the instability pressures of LN2 fracturing decreased by 20.00% and 29.41%, respectively.
This study investigated the impacts of LN2 on the mechanical degradation of granite. Future work will systematically explore the influence of LN2 cycling on the rock’s mechanical behavior and its application in enhancing fracturing efficiency.

Author Contributions

Conceptualization, H.W. and C.Z.; methodology, C.C.; validation, H.W., Y.H. and N.L.; investigation, Y.H.; data curation, N.L.; writing—original draft preparation, H.W.; writing—review and editing, C.Z.; supervision, C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Yunlong Lake Laboratory of Deep Underground Science and Engineering Project (grant number: 104024008).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Authors Hu Wang and Yong Hu were employed by the company Xinjiang Yaxin Coalbed Methane Investment and Development (Group) Co., Ltd. Author Na Luo was employed by the company CNPC Engineering Technology R&D Company Limited. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Xie, H.; Li, C.; Zhou, T.; Chen, J.; Liao, J.; Ma, J.; Li, B. Conceptualization and evaluation of the exploration and utilization of low/medium-temperature geothermal energy: A case study of the Guangdong-Hong Kong-Macao Greater Bay Area. Geomech. Geophys. Geo-Energy Geo-Resour. 2020, 6, 18. [Google Scholar] [CrossRef]
  2. Kong, G.; Wu, D.; Wei, Y. Experimental and numerical investigations on the energy and structural performance of a full-scale energy utility tunnel. Tunn. Undergr. Space Tech. 2023, 139, 105208. [Google Scholar] [CrossRef]
  3. Chen, R.; Wu, D.; Zhang, T.; Kong, G.; Fang, J. Thermo-mechanical behavior of energy piles equipped with PCM tubes. Comput. Geotech. 2025, 179, 107005. [Google Scholar] [CrossRef]
  4. Martin, C.D.; Chandler, N.A. The progressive fracture of Lac du Bonnet granite. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 1994, 31, 643–659. [Google Scholar] [CrossRef]
  5. Li, C.; Xie, H.; Wang, J. Anisotropic characteristics of crack initiation and crack damage thresholds for shale. Int. J. Rock. Mech. Min. 2020, 126, 104178. [Google Scholar] [CrossRef]
  6. Yang, R.; Hong, C.; Liu, W.; Wu, X.; Wang, T.; Huang, Z. Non-contaminating cryogenic fluid access to high-temperature resources: Liquid nitrogen fracturing in a lab-scale Enhanced Geothermal System. Renew. Energ. 2021, 165, 125–138. [Google Scholar] [CrossRef]
  7. Tomac, I.; Sauter, M. A review on challenges in the assessment of geomechanical rock performance for deep geothermal reservoir development. Renew. Sustain. Energy Rev. 2018, 82, 3972–3980. [Google Scholar] [CrossRef]
  8. Pan, S.; Gao, M.; Shah, K.J.; Zheng, J.; Pei, S.; Chiang, P. Establishment of enhanced geothermal energy utilization plans: Barriers and strategies. Renew. Energ. 2019, 132, 19–32. [Google Scholar] [CrossRef]
  9. Li, B.; Xie, H.; Sun, L.; Gao, T.; Xia, E.; Liu, B.; Wang, J.; Long, X. Advanced exergy analysis and multi-objective optimization of dual-loop ORC utilizing LNG cold energy and geothermal energy. Renew. Energ. 2025, 239, 122164. [Google Scholar] [CrossRef]
  10. Su, X.; Zhou, L.; Li, H.; Lu, Y.; Song, X.; Shen, Z. Effect of mesoscopic structure on hydro-mechanical properties of fractures. Environ. Earth Sci. 2020, 79, 146. [Google Scholar] [CrossRef]
  11. McClure, M.W.; Horne, R.N. An investigation of stimulation mechanisms in Enhanced Geothermal Systems. Int. J. Rock. Mech. Min. 2014, 72, 242–260. [Google Scholar] [CrossRef]
  12. Huang, Z.; Zhang, S.; Yang, R.; Wu, X.; Li, R.; Zhang, H.; Hung, P. A review of liquid nitrogen fracturing technology. Fuel 2020, 266, 117040. [Google Scholar] [CrossRef]
  13. Wang, L.; Yao, B.; Cha, M.; Alqahtani, N.B.; Patterson, T.W.; Kneafsey, T.J.; Miskimins, J.L.; Yin, X.; Wu, Y. Waterless fracturing technologies for unconventional reservoirs-opportunities for liquid nitrogen. J. Nat. Gas. Sci. Eng. 2016, 35, 160–174. [Google Scholar] [CrossRef]
  14. Cai, C.; Tao, Z.; Ren, K.; Liu, S.; Yang, Y.; Feng, Y.; Su, S.; Hou, P. Experimental investigation on the breakdown behaviours of sandstone due to liquid nitrogen fracturing. J. Petrol. Sci. Eng. 2021, 200, 108386. [Google Scholar] [CrossRef]
  15. Zhang, S.; Huang, Z.; Huang, P.; Wu, X.; Xiong, C.; Zhang, C. Numerical and experimental analysis of hot dry rock fracturing stimulation with high-pressure abrasive liquid nitrogen jet. J. Petrol. Sci. Eng. 2018, 163, 156–165. [Google Scholar] [CrossRef]
  16. Wu, X.; Huang, Z.; Song, H.; Zhang, S.; Cheng, Z.; Li, R.; Wen, H.; Huang, P.; Dai, X. Variations of physical and mechanical properties of heated granite after rapid cooling with liquid nitrogen. Rock. Mech. Rock. Eng. 2019, 52, 2123–2139. [Google Scholar] [CrossRef]
  17. Kang, F.; Jia, T.; Li, Y.; Deng, J.; Tang, C.; Huang, X. Experimental study on the physical and mechanical variations of hot granite under different cooling treatments. Renew. Energ. 2021, 179, 1316–1328. [Google Scholar] [CrossRef]
  18. Sha, S.; Rong, G.; Chen, Z.; Li, B.; Zhang, Z. Experimental evaluation of physical and mechanical properties of geothermal reservoir rock after different cooling treatments. Rock. Mech. Rock. Eng. 2020, 53, 4967–4991. [Google Scholar] [CrossRef]
  19. Wu, X.; Huang, Z.; Li, R.; Zhang, S.; Wen, H.; Huang, P.; Dai, X.; Zhang, C. Investigation on the damage of high-temperature shale subjected to liquid nitrogen cooling. J. Nat. Gas. Sci. Eng. 2018, 57, 284–294. [Google Scholar] [CrossRef]
  20. Wu, X.; Huang, Z.; Zhang, S.; Cheng, Z.; Li, R.; Song, H.; Wen, H.; Huang, P. Damage analysis of high-temperature rocks subjected to LN2 thermal shock. Rock. Mech. Rock. Eng. 2019, 52, 2585–2603. [Google Scholar] [CrossRef]
  21. Hou, P.; Su, S.; Zhang, Y.; Gao, F.; Gao, Y.; Liang, X.; Ding, R.; Cai, C. Effect of liquid nitrogen thermal shock on structure damage and brittleness properties of high-temperature marble. Geomech. Geophys. Geo-Energy Geo-Resour. 2022, 8, 69. [Google Scholar] [CrossRef]
  22. Su, S.; Hou, P.; Gao, F.; Liang, X.; Ding, R.; Cai, C. Changes in mechanical properties and fracture behaviors of heated marble subjected to liquid nitrogen cooling. Eng. Fract. Mech. 2022, 261, 108256. [Google Scholar] [CrossRef]
  23. Wu, X.; Huang, Z.; Cheng, Z.; Zhang, S.; Song, H.; Zhao, X. Effects of cyclic heating and LN2-cooling on the physical and mechanical properties of granite. Appl. Therm. Eng. 2019, 156, 99–110. [Google Scholar] [CrossRef]
  24. Rong, G.; Sha, S.; Li, B.; Chen, Z.; Zhang, Z. Experimental investigation on physical and mechanical properties of granite subjected to cyclic heating and liquid nitrogen cooling. Rock. Mech. Rock. Eng. 2021, 54, 2383–2403. [Google Scholar] [CrossRef]
  25. Ge, Z.; Sun, Q.; Yang, T.; Luo, T.; Jia, H.; Yang, D. Effect of high temperature on mode-I fracture toughness of granite subjected to liquid nitrogen cooling. Eng. Fract. Mech. 2021, 252, 107834. [Google Scholar] [CrossRef]
  26. Yang, R.; Hong, C.; Huang, Z.; Wen, H.; Li, X.; Huang, P.; Liu, W.; Chen, J. Liquid nitrogen fracturing in boreholes under true triaxial stresses: Laboratory investigation on fractures initiation and morphology. Spe J. 2021, 26, 135–154. [Google Scholar] [CrossRef]
  27. Span, R.; Lemmon, E.W.; Jacobsen, R.T.; Wagner, W.; Yokozeki, A. A reference equation of state for the thermodynamic properties of nitrogen for temperatures from 63.151 to 1000 K and pressures to 2200 MPa. J. Phys. Chem. Ref. Data 2000, 29, 1361–1433. [Google Scholar] [CrossRef]
  28. Lemmon, E.W.; Jacobsen, R.T. Viscosity and thermal conductivity equations for nitrogen, oxygen, argon, and air. Int. J. Thermophys. 2004, 25, 21–69. [Google Scholar] [CrossRef]
  29. Zhou, C.; Su, S.; Liang, X.; Xue, Y.; Cai, C.; Gao, F. Liquid nitrogen pre-injection assisted fracturing in hot dry rock reservoirs. Phys. Fluids 2025, 37, 012007. [Google Scholar] [CrossRef]
  30. Guo, T.; Tang, S.; Liu, S.; Liu, X.; Zhang, W.; Qu, G. Numerical simulation of hydraulic fracturing of hot dry rock under thermal stress. Eng. Fract. Mech. 2020, 240, 107350. [Google Scholar] [CrossRef]
  31. Xia, T.; Zhou, F.; Liu, J.; Kang, J.; Gao, F. A fully coupled hydro-thermo-mechanical model for the spontaneous combustion of underground coal seams. Fuel 2014, 125, 106–115. [Google Scholar] [CrossRef]
  32. Zhu, W.C.; Wei, C.H. Numerical simulation on mining-induced water inrushes related to geologic structures using a damage-based hydromechanical model. Environ. Earth Sci. 2011, 62, 43–54. [Google Scholar] [CrossRef]
  33. Zhang, W.; Wang, C.; Guo, T.; He, J.; Zhang, L.; Chen, S.; Qu, Z. Study on the cracking mechanism of hydraulic and supercritical CO2 fracturing in hot dry rock under thermal stress. Energy 2021, 221, 119886. [Google Scholar] [CrossRef]
  34. Zhu, W.C.; Tang, C.A. Micromechanical model for simulating the fracture process of rock. Rock. Mech. Rock. Eng. 2004, 37, 25–56. [Google Scholar] [CrossRef]
  35. Tang, C.A.; Yang, W.T.; Fu, Y.F.; Xu, X.H.; Pusch, R. A new approach to numerical method of modelling geological processes and rock engineering problems; continuum to discontinuum and linearity to nonlinearity. Eng. Geol. 1998, 49, 207–214. [Google Scholar] [CrossRef]
  36. Wang, J.; Elsworth, D.; Wu, Y.; Liu, J.; Zhu, W.; Liu, Y. The Influence of fracturing fluids on fracturing processes: A comparison between water, oil and SC-CO2. Rock. Mech. Rock. Eng. 2018, 51, 299–313. [Google Scholar] [CrossRef]
  37. Wu, X.; Huang, Z.; Dai, X.; Song, H.; Zhang, S. Thermo-coupled FSI analysis of LN2 jet impinging on hot dry rock. Appl. Therm. Eng. 2020, 165, 114621. [Google Scholar] [CrossRef]
  38. Hou, P.; Su, S.; Gao, F.; Liang, X.; Wang, S.; Gao, Y.; Cai, C. Influence of Liquid Nitrogen Freeze–Thaw Cycles on Mechanical Behaviors and Permeability Properties of Coal Under Different Confining Pressures. Rock. Mech. Rock. Eng. 2024, 57, 2625–2644. [Google Scholar] [CrossRef]
  39. Alameedy, U.; Al-Behadili, A. An Overview of How the Petrophysical Properties of Rock Influenced After Being Exposed to Cryogenic Fluid. J. Eng. 2023, 29, 1–16. [Google Scholar] [CrossRef]
  40. Lu, Y.L.; Elsworth, D.; Wang, L.G. Microcrack-based coupled damage and flow modeling of fracturing evolution in permeable brittle rocks. Comput. Geotech. 2013, 49, 226–244. [Google Scholar] [CrossRef]
Figure 1. (ad) Samples and equipment and (e) stress identification.
Figure 1. (ad) Samples and equipment and (e) stress identification.
Processes 13 00696 g001
Figure 2. Reservoir fracturing model.
Figure 2. Reservoir fracturing model.
Processes 13 00696 g002
Figure 3. (a) Crack initiation stress under various conditions and (b) crack initiation stress reduction rate.
Figure 3. (a) Crack initiation stress under various conditions and (b) crack initiation stress reduction rate.
Processes 13 00696 g003
Figure 4. (a) Crack damage stress under various conditions and (b) crack damage stress reduction rate.
Figure 4. (a) Crack damage stress under various conditions and (b) crack damage stress reduction rate.
Processes 13 00696 g004
Figure 5. (a) Peak stress under various conditions and (b) peak stress reduction rate.
Figure 5. (a) Peak stress under various conditions and (b) peak stress reduction rate.
Processes 13 00696 g005
Figure 6. Count of acoustic emission events under various (a) granite temperatures and (b) confining pressures.
Figure 6. Count of acoustic emission events under various (a) granite temperatures and (b) confining pressures.
Processes 13 00696 g006
Figure 7. Fractal dimension of acoustic emission events under various (a) granite temperatures and (b) confining pressures.
Figure 7. Fractal dimension of acoustic emission events under various (a) granite temperatures and (b) confining pressures.
Processes 13 00696 g007
Figure 8. Damage distribution during fracturing processes with (a) water, (b) nitrogen, and (c) liquid nitrogen.
Figure 8. Damage distribution during fracturing processes with (a) water, (b) nitrogen, and (c) liquid nitrogen.
Processes 13 00696 g008
Figure 9. (a) Cumulative number and (b) fractal dimension of damage elements during water, nitrogen, and liquid nitrogen fracturing.
Figure 9. (a) Cumulative number and (b) fractal dimension of damage elements during water, nitrogen, and liquid nitrogen fracturing.
Processes 13 00696 g009
Figure 10. Initiation and instability pressures of water, nitrogen, and liquid nitrogen fracturing models.
Figure 10. Initiation and instability pressures of water, nitrogen, and liquid nitrogen fracturing models.
Processes 13 00696 g010
Table 1. Mechanical parameters for the mesoscopic element.
Table 1. Mechanical parameters for the mesoscopic element.
ParameterValue
Average   value   of   initial   permeability   ( m = 20 ) (m2)4 × 10−19
Average value of the coefficient of thermal expansion (K−1)8 × 10−6
Average value of modulus of elasticity (GPa)45 *
Average compressive strength (MPa)400 *
Average compressive strength/average tensile strength16
Internal friction angle (°)54
Heterogeneity coefficient (m)4 *
* For heterogeneous models, the parameter values of mesoscopic elements are not equal to the macroscopic values and need to be determined via trial calculations.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Hu, Y.; Luo, N.; Zhou, C.; Cai, C. Effects of Liquid Nitrogen on Mechanical Deterioration and Fracturing Efficiency in Hot Dry Rock. Processes 2025, 13, 696. https://doi.org/10.3390/pr13030696

AMA Style

Wang H, Hu Y, Luo N, Zhou C, Cai C. Effects of Liquid Nitrogen on Mechanical Deterioration and Fracturing Efficiency in Hot Dry Rock. Processes. 2025; 13(3):696. https://doi.org/10.3390/pr13030696

Chicago/Turabian Style

Wang, Hu, Yong Hu, Na Luo, Chunbo Zhou, and Chengzheng Cai. 2025. "Effects of Liquid Nitrogen on Mechanical Deterioration and Fracturing Efficiency in Hot Dry Rock" Processes 13, no. 3: 696. https://doi.org/10.3390/pr13030696

APA Style

Wang, H., Hu, Y., Luo, N., Zhou, C., & Cai, C. (2025). Effects of Liquid Nitrogen on Mechanical Deterioration and Fracturing Efficiency in Hot Dry Rock. Processes, 13(3), 696. https://doi.org/10.3390/pr13030696

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