Next Article in Journal
A Combined Control Scheme of Direct Torque Control and Field-Oriented Control Algorithms for Three-Phase Induction Motor: Experimental Validation
Next Article in Special Issue
Predicting Angle of Internal Friction and Cohesion of Rocks Based on Machine Learning Algorithms
Previous Article in Journal
Joint Resource and Trajectory Optimization for Energy Efficiency Maximization in UAV-Based Networks
Previous Article in Special Issue
Effect of Different Tunnel Distribution on Dynamic Behavior and Damage Characteristics of Non-Adjacent Tunnel Triggered by Blasting Disturbance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanical Properties and Strength Evolution Model of Sandstone Subjected to Freeze–Thaw Weathering Process: Considering the Confining Pressure Effect

1
School of Resources and Safety Engineering, Central South University, Changsha 410083, China
2
Research Center for Mining Engineering and Technology in Cold Regions, Central South University, Changsha 410083, China
*
Author to whom correspondence should be addressed.
Submission received: 4 September 2022 / Revised: 11 October 2022 / Accepted: 13 October 2022 / Published: 17 October 2022
(This article belongs to the Special Issue Mathematical Problems in Rock Mechanics and Rock Engineering)

Abstract

:
Freeze-and-thaw (F&T) weathering cycles induced by day–night and seasonal temperature changes cause a large number of rock mass engineering disasters in cold areas. Investigating the impact of F&T weathering process on the strength and deformation characteristics of frozen–thawed rocks is therefore of critical scientific importance for evaluating the stability and optimizing the design of rock mass engineering in these areas. In this research, the evolution characteristics of F&T damage were analyzed based on T2 spectrum distribution curves of sandstone specimens before and after F&T weathering cycles. The coupling impact of the quantity of F&T weathering cycles and confining pressure on pre-peak and post-peak deformation behaviors of sandstone specimens were analyzed in detail. By introducing the confining pressure increase factor (CPIF), the impact of confining pressure on the triaxial compressive strength (TCS) of sandstone specimens after undergoing different quantities of F&T weathering cycles was further investigated. A novel strength evolution model was proposed that could effectively describe the coupling impact of the quantity of F&T weathering cycles and confining pressure on TCS of rocks after undergoing the F&T weathering process. The proposed strength evolution model was cross-verified with experimental data from the published literature and all correlation coefficients were above 0.95, which proved that the strength evolution model proposed in this paper was reasonable; in addition, this model has strong applicability.

1. Introduction

During the construction of rock mass engineering (such as mines, roads, and tunnels) in cold regions, the recurrence of F&T weathering processes induced by day–night and seasonal temperature changes causes rapid damage and the deterioration of rock masses and has initiated a large number of F&T disasters such as rock falls, landslides [1,2,3], and the cracking of rock surrounding tunnels [4,5], which has a major impact on the design, construction, and operation of rock mass engineering in cold areas. The strength and deformation behaviors of rocks are the theoretical basis for evaluating the stability and optimizing the engineering design in rock mass engineering [6]. Therefore, investigating the mechanical properties of rocks after undergoing F&T weathering cycles has great significance in evaluating the stability and optimizing the design in rock mass engineering in cold areas.
The impact of F&T weathering process on rock mechanical properties have been investigated by a considerable number of scholars. These investigation results indicated that after undergoing the F&T weathering process, the elastic modulus [7,8,9,10,11], uniaxial compressive strength (UCS) [7,8,10,11,12,13,14], Brazilian tensile strength (BTS) [15,16,17,18], point load strength (PLS) [15,16], dynamic uniaxial compressive strength (UCSd) [7,8,10,19,20,21,22] and dynamic tensile strength (BTSd) [19,23] of rocks such as sandstone, tuff, gneiss, granite and shale all decreased as the quantity of F&T weathering cycles increased, but to different extents. To reveal the change laws in these mechanical parameters with the quantity of F&T weathering cycles, a sequence of regression analysis models [24,25] and exponential decay models [15,16,26,27,28] were proposed based on the relationships between these mechanical parameters and the quantity of F&T weathering cycles. In addition, Liu et al. [29] regarded F&T weathering cycles as a kind of fatigue damage and thus established a fatigue damage model of rocks after undergoing F&T weathering cycles. A prediction model of UCS of rocks after undergoing F&T weathering cycles was obtained based on this fatigue damage model. Gao et al. [30] established a UCS evolution model of frozen–thawed rocks based on the energy evolution characteristics of rock failure. However, these research findings mainly focused on the uniaxial mechanical properties of rocks after undergoing F&T weathering cycles and did not consider the impact of the confining pressure on the mechanical properties of rocks after undergoing F&T weathering process. There is no doubt that actual rock mass engineering is always in a certain stress field [31,32]. To be closer to the actual engineering, it is necessary to study the triaxial mechanical properties of rocks after undergoing F&T weathering cycles.
Recently, scholars have increasingly studied the triaxial mechanical properties of rocks after undergoing F&T weathering cycles. Tan et al. [9] used granite as the research object to investigate the impact of the quantity of F&T weathering cycles on uniaxial and triaxial mechanical properties according to uniaxial compression tests (UCTs) and triaxial compression tests (TCTs) after granite specimens being underwent F&T weathering process. They found that both the UCS and TCS decreased exponentially as the quantity of F&T weathering cycles increased, as did the elastic modulus and cohesion. The relationships between these mechanical properties and the quantity of F&T weathering cycles were built according to an exponential function. Wang et al. [31] and Hosseini and Khodayari [33] also carried out similar studies but took sandstone as the research object; the research results for both were in accordance with those of Tan et al. [9]. In addition, Hosseini and Khodayari [33] also found that the rate of reduction in the TCS was less than that of the UCS at the equivalent quantity of F&T weathering cycles and that the higher the confining pressure was, the lower the rate of reduction. It was evident that the confining pressure had a significant impact on the TCS of rocks after undergoing F&T weathering cycles. To this end, Fu et al. [34] operated a sequences of TCTs after transversely isotropic rocks underwent an F&T weathering process and found that the quantity of F&T weathering cycles, bedding plane orientation, and confining pressure had a significant impact on the TCS of slate. A TCS prediction model for transversely isotropic rocks after undergoing an F&T weathering process was proposed based on the single discontinuity theory and the functional relationships between cohesion and internal friction angle and the quantity of F&T weathering cycles. Seyed Mousavi et al. [35] also carried out similar studies taking calc-schist rock specimens as the research objects. Finally, an empirical expression among the TCS, the quantity of F&T weathering cycles, and the confining pressure was obtained according to experimental results and the prediction model suggested by Fu et al. [34]. Although the prediction model proposed by Fu et al. [34] could effectively reflect the coupling impact of the quantity of F&T weathering cycles and confining pressure on TCS of rocks after undergoing an F&T weathering process, this prediction model was based on transversely isotropic rocks and its applicability was not strong. Therefore, it is necessary to establish a more applicable strength evolution model to reveal the coupling impact of the quantity of F&T weathering cycles and confining pressure on TCS of rocks after undergoing an F&T weathering process. In addition, the above-mentioned studies mainly focused on investigating the strength-deterioration characteristics of rocks after undergoing an F&T weathering process and did not deeply investigate the deformation behaviors of rocks, especially the coupling impact of the quantity of F&T weathering cycles and confining pressure on deformation behaviors of rocks during the entire loading procedure.
In this study, saturated sandstone specimens were first subjected to different quantities of F&T weathering cycles, then change laws in the T2 spectrum distribution curves of sandstone specimens before and after the F&T weathering cycles were investigated. UCTs and TCTs of sandstone specimens after undergoing different quantities of F&T weathering cycles were conducted to obtain the UCS and TCS and the corresponding stress–strain curves. The coupling impact of the quantity of F&T weathering cycles and confining pressure on the deformation behaviors of rocks were investigated. A novel strength evolution model that considered the coupling impact of the quantity of F&T weathering cycles and confining pressure on the TCS of rocks after undergoing the F&T weathering process was established; the proposed strength evolution model was cross-verified with experimental data from the published literature.

2. Experimental Materials and Methods

2.1. Rock Specimen Preparation

In this paper, rock specimens of sandstone were taken from the Jiama open pit copper mine located in the Tibet Autonomous Region of China. The sampling site is shown in Figure 1a. Figure 1b shows the distribution map for the frozen soil of China. It can be seen in Figure 1b that the sampling site was on the boundary between the permafrost regions and seasonal frozen regions; therefore, the rock was subjected to repeated F&T weathering cycles. Table 1 shows the mineral composition of sandstone obtained via an X-ray diffraction (XRD) technique.
Based on the method suggested by the International Society for Rock Mechanics (ISRM), all rock specimens were processed into a cylinder with a diameter of 50 mm; we ensured that the flatness of the end surfaces was less than 0.05 mm [36]. Careful preparations ensured that the maximum deviations of the diameter and height were less than 0.30 mm and the vertical variance was less than 0.25° [36]. In this study, 15 and 60 rock specimens were used in UCTs and TCTs, respectively, and the length/diameter ratio of the rock specimens was 2.0 [36]. These specimens were divided into five groups (labeled A, B, C, D, and E); each group comprised 15 rock specimens (labeled 1, 2 … 14, 15). The sandstone specimens from groups A, B, C, D, and E were treated in 0 cycles, 10 cycles, 20 cycles, 30 cycles, and 40 cycles, respectively. As shown in Figure 2a, the sandstone specimens labeled 1–3 in each group were used in the UCTs. The confining pressures were 3, 6, 9, and 12 MPa in the TCTs, corresponding to the sandstone specimens labeled 4–6, 7–9, 10–12, and 13–15 in each group, as shown in Figure 2b.

2.2. Test Procedures and Experimental Apparatus

All sandstone specimens were divided into five groups and dried in an oven for 48 h at 65 °C. The sandstone specimens that did not undergo F&T weathering cycles were used directly in the UCTs and TCTs. The other sandstone specimens were placed in a vacuum pump at a pressure of 0.1 MPa for 4 h and then soaked in distilled water for 24 h. The saturated specimens then went through the specified quantity of F&T weathering cycles in a TDS-300 automatic F&T testing machine (as shown in Figure 3a). When the quantity of F&T weathering cycles reached the specified quantities, the corresponding sandstone specimens were removed and UCTs and TCTs were conducted using an MTS815 electrohydraulic servo-controlled rock-testing machine (as shown in Figure 3c).

2.3. F&T Weathering Cycle Tests

Designed based on the local climate of the sampling site, one F&T weathering cycle in our tests included freezing the saturated rock specimens at −20 °C for four hours and then thawing in water at +20 °C for four hours. The temperature variation curve of the F&T weathering process is shown in Figure 4. In this study, four groups of sandstone specimens were subjected to F&T weathering tests corresponding to 10 cycles, 20 cycles, 30 cycles and 40 cycles. To reveal the evolution characteristics of the F&T damage, the MesoMR23-060H-I NMR system (as shown in Figure 3b) was used to conduct nuclear magnetic resonance (NMR) tests to obtain T2 spectrum distribution curves before and after the F&T weathering cycles [37,38,39,40,41,42,43].
Figure 5 displays T2 spectrum distribution curves of the sandstone specimens before and after 20 (as shown in Figure 5a) and 40 (as shown in Figure 5b) F&T weathering cycles. In the T2 spectrum distribution curves, the T2 relaxation time (horizontal axis) is a measurement of the internal pore sizes and the porosity component (vertical axis) is the proportion of the corresponding pore sizes [20]. As can be seen in Figure 5a, after the sandstone specimens underwent 20 cycles, the peak values of the T2 spectrum distribution curve increased, indicating that the sizes of the internal pores increased. As shown in Figure 5b, when the sandstone specimens underwent 40 cycles, the increases in the peak values were even more significant. In addition, the expansion of the curve to the left suggested that the sizes of some small pores increased. This demonstrated that in the early stage of the F&T weathering cycles, the original internal pores and microcracks were constantly developing. With the increase in the quantity of F&T weathering cycles, in the later stage of the cycles, in addition to the expansion of original internal pores and the constant extension of microcracks, new pores and microcracks were generated; i.e., with the increase in the quantity of the F&T weathering cycles, the accumulated damage to the inner sandstone specimens constantly increased.

2.4. Uniaxial and Triaxial Compression Tests

UCTs and TCTs were conducted on an MTS815 electrohydraulic servo-controlled rock-testing machine; axial and circumferential extensometers were used in our experiment to measure the axial and lateral strains. In our experiment, the displacement-control loading mode was used at a loading rate of 0.1 mm/min. The measured UCSs and TCSs of the sandstone specimens are listed in Table 2.

3. Experimental Results and Analysis

3.1. Uniaxial Mechanical Properties Variation Characteristics for Sandstone after Undergoing F&T Weathering Cycles

3.1.1. Stress–Strain Curve

Figure 6 displays the stress–strain curves for the sandstone specimens after undergoing different quantities of F&T weathering cycles under uniaxial compression conditions. All of the stress–strain curves had the same variation patterns and could be separated into five stages during the entire loading process; that is, the compaction stage, elastic deformation stage, yield stage, failure stage, and strain softening stage. As the quantity of F&T weathering cycles increased, the stress–strain curves showed three obvious features: (1) the compaction stage became longer; (2) the slope at the linear deformation stage decreased, as did the UCS; and (3) the stress-dropping rate of the post-peak decreased. The main reason was that the length of the compaction stage and the slope of the linear deformation stage were proportional to the number of microdefects inside the rock [44]. Under the F&T weathering process, water migration and transformation from water to ice caused microdefects to gradually develop; the number of microdefects increased with the increase in the quantity of F&T weathering cycles [45,46]. The main reason for the reduction in the stress-dropping rate of the post-peak was that the F&T weathering cycles caused the cohesion between the particles to gradually decrease, which caused the sandstone specimens to become soft and the plasticity to increase.

3.1.2. UCS

Figure 7 shows the changes in the UCS and its reduction ratio for different quantities of F&T weathering cycles. The reduction ratio for UCS is defined as follows:
η = σ 0 σ N σ 0 × 100 %
where σ 0 is the UCS of sandstone specimens without any F&T weathering cycles, σ N is the UCS of sandstone specimens after undergoing N quantity of F&T weathering cycles, and η is the reduction ratio of the UCS.
As shown in Figure 7, compared with the original average UCS (27.50 MPa), the reduction ratios were 11.33% (24.38 MPa), 24.41% (20.79 MPa), 33.94% (18.17 MPa), and 40.40% (16.39 MPa), corresponding to 10 cycles, 20 cycles, 30 cycles, and 40 cycles, respectively. The reason can be explained as follows: the water migration and transformation from water to ice under the F&T weathering process caused microdefects to gradually develop and the sandstone specimens to become more fragmented. It was noticeable that the UCS exponentially decayed as the quantity of F&T weathering cycles increased, similar to laws found in the literature [16,26,28]. The experimental data were fitted by the decay model suggested by Mutlutürk et al. [26]; this decay model is defined as follows:
I N = I 0 e λ N
where I is the rock integrity, λ is the decay coefficient, and N is the quantity of F&T weathering cycles.
In this paper, the UCS was regarded as the rock integrity; therefore, the decay model became as follows:
σ N = σ 0 e λ N
The fitting curves of our test are shown in Figure 7. The model fit well with the experimental data: the fitting coefficient of determination (R2) was greater than 0.99.

3.1.3. Failure Modes

Figure 8 shows the failure modes of the sandstone specimens after undergoing different quantities of F&T weathering cycles in uniaxial compression conditions. The failure mode was single inclined plane shear failure when the sandstone specimens did not undergo F&T weathering cycles. The failure modes became tension-shear comprehensive failure and splitting failure as the quantity of F&T weathering cycles increased. Tension-shear comprehensive failure occurred in the sandstone specimens after undergoing 10 cycles and 20 cycles and the sandstone specimens were fragmented. Splitting failure occurred in the sandstone specimens when the quantity of F&T weathering cycles was 30 and 40. The increase in the number of macroscopic cracks on the surfaces of sandstone specimens caused them to become more fragmented. The reason why the failure modes changed with the quantity of F&T weathering cycles was that the F&T weathering process causes microdefects to gradually develop inside the sandstone specimens, which caused cracks to be more likely to expand in the axial direction under uniaxial compression conditions.

3.2. Triaxial Mechanical Properties’ Variation Characteristics for Sandstone after Undergoing F&T Weathering Cycles

3.2.1. Stress–Strain Curve

Figure 9 displays the stress–strain curves for the sandstone specimens after undergoing different quantities of F&T weathering cycles under different confining pressures. Compared with the stress–strain curves of sandstone specimens after undergoing different quantities of F&T weathering cycles under uniaxial compression conditions, all of the stress–strain curves had obvious residual strength characteristics. In addition, the variation characteristics of the stress–strain curves were similar under different confining pressures as the quantity of F&T weathering cycles increased. All of the stress–strain curves could be separated into six stages: compaction stage, elastic deformation stage, yield stage, failure stage, strain softening stage, and residual strength stage. When the confining pressure was constant, with the increase in the quantity of the F&T weathering cycles, except for a gradual decrease in the residual strength, the other variation characteristics were similar to those in uniaxial compression conditions: (1) the compaction stage became longer; (2) the slope at the linear deformation stage decreased, as did the TCS; and (3) the stress-drop rate in the post-peak stress decreased. The main reason for these characteristics was the same as that under uniaxial compression conditions.
Figure 10 displays the stress–strain curves of the sandstone specimens after undergoing 20 F&T weathering cycles at different confining pressures. With the increase in confining pressure, the stress–strain curves showed four obvious features: (1) the compaction stage became shorter; (2) the slope at the linear deformation stage increased, as did the TCS; (3) the stress-dropping ratio of the post-peak decreased; and (4) the residual strength increased. The main reason was that the length of the compaction stage and the slope of the linear deformation stage are proportional to the number of microdefects in the rock. Under the confining pressure, microdefects inside the rock caused by the F&T weathering process were pre-compression, which caused a decrease in the number of microdefects; the higher the confining pressure was, the larger the number of microdefects that the pre-compression had. The main reason for the decrease in the stress-dropping ratio of the post-peak and the increase in residual strength was that the confining pressure could make the post-peak deformation behavior of the rocks transition from brittleness to plasticity [47,48,49,50].
These variation characteristics indicated that the impact of the quantity of F&T weathering cycles on the pre-peak deformation behavior and residual strength was the opposite of the confining pressure while the post-peak deformation behaviors were similar to the confining pressure.

3.2.2. TCS

Figure 11 displays the relationship between the TCS and its reduction ratio and the quantity of F&T weathering cycles at different confining pressures. It can be seen that the TCS decreased with an increase in the quantity of F&T weathering cycles when the confining pressure was constant, which was similar to laws in the uniaxial compression conditions. Compared with the original average TCS, the reduction ratios were: 9.55%, 18.71%, 26.09%, and 33.45% for 10 cycles, 20 cycles, 30 cycles, and 40 cycles, respectively, when the confining pressure was 3 MPa; 9.02%, 15.86%, 22.12%, and 30.57% for 10 cycles, 20 cycles, 30 cycles, and 40 cycles, respectively, when the confining pressure was 6 MPa; 7.53%, 13.90%, 20.38%, and 29.91% for 10 cycles, 20 cycles, 30 cycles, and 40 cycles, respectively, when the confining pressure was 9 MPa; and 5.82%, 12.70%, 20.95%, and 30.57% for 10 cycles, 20 cycles, 30 cycles, and 40 cycles, respectively, when the confining pressure was 12 MPa. At different confining pressures, the variation characteristics between the TCS of the sandstone specimens and the quantity of F&T cycles were similar to the UCS. Therefore, the experimental data could be used to fit the model suggested by Mutlutürk et al. [26]; the fitting results of our tests are shown in Figure 11. The model fit well with experimental data: the fitting coefficient of determination (R2) was greater than 0.98. The decay coefficient was 0.01018, 0.00881, 0.00819, and 0.00778 for 3, 6, 9, and 12 MPa, respectively. The results indicated that the decay coefficient decreased as the confining pressure increased and that the higher the confining pressure was, the larger the reduction in the decay coefficient. Therefore, the TCS of the sandstone after undergoing F&T weathering process was impacted by the confining pressure. Further study of the impact of confining pressure on the TCS of sandstone after undergoing the F&T weathering process referred to the definition of the dynamic increase factor [51,52]. The confining pressure increase factor (CPIF) could be defined as TCS/UCS.
Figure 12 displays the CPIF curves of the sandstone specimens after undergoing different quantities of F&T weathering cycles in different confining pressures. As shown in Figure 12, the variation characteristics of the CPIF curves were similar under different confining pressures as the quantity of F&T weathering cycles increased; that is, the CPIF increased as the quantity of F&T weathering cycles increased. However, the higher the confining pressure was, the larger the increased amplitude of the CPIF. For example, when the confining pressure was 3 MPa, the value of the CPIF was 1.79 without F&T weathering cycles; the CPIF is 2.31 after the sandstone underwent F&T weathering cycles, an increase ratio of 28.97%; when the confining pressure was 12 MPa, the value of the CPIF was 3.09 without F&T weathering cycles and the CPIF was 4.83 after the sandstone underwent F&T weathering cycles, an increase ratio of 56.27%. This indicated that sandstone specimens were more sensitive to the confining pressure after undergoing more F&T weathering cycles.
These results demonstrated that the UCS of the sandstone after undergoing F&T weathering process was mainly controlled by the quantity of F&T weathering cycles under uniaxial compression conditions. The TCS of the sandstone after undergoing the F&T weathering process was impacted by the quantity of F&T weathering cycles and confining pressure under triaxial compression conditions; these two factors had opposite impacts. The main reason was that the water–ice phase and water migration under the F&T weathering cycles caused microdefects to gradually develop inside the sandstone specimens, which induced F&T damage. However, under confining pressure conditions, some microdefects inside the sandstone specimens were closed, which lessened the damage induced by the F&T weathering process. Therefore, for the slope engineering of open-pit mines in cold areas, the application of anchor reinforcement technology to provide pre-stress could lessen the damage induced by the F&T weathering process and improve the stability of the slope.

3.2.3. Failure Modes

Under different quantities of F&T weathering cycles, the failure modes of the sandstone specimens had the same evolution characteristics with the change in confining pressure. Therefore, Figure 13 only shows the failure modes under different confining pressures after the sandstone specimens underwent 20 F&T weathering cycles. As shown in Figure 13, under different confining pressures, the failure modes of the sandstone specimens all were single incline plane shear failure; however, the length of the shear failure plane became shorter with the increase in the confining pressure. The main reason was that the lateral deformation was limited under the confining pressure, so the sandstone specimens only exhibited single inclined plane shear failure. In addition, the higher the confining pressure was, the more severe the limiting effect, so the length of the shear failure plane became shorter with the increase in the confining pressure.

4. Strength Evolution Model of Rock Specimens Considering the Freeze–Thaw Weathering Process and Confining Pressure

The experimental results demonstrated that the TCS of rocks after undergoing the F&T weathering process was impacted by the quantity of F&T weathering cycles and the confining pressure under triaxial compression conditions. The model suggested by Mutlutürk et al. [26] could only describe the change in the peak compressive strength of rocks after undergoing the F&T weathering process at a specific confining pressure, which did not consider the coupling impact of the quantity of F&T weathering cycles and the confining pressure. Therefore, it was necessary to establish a novel model that could reflect the evolution laws of the TCS of rocks after undergoing the F&T weathering process. The impact of the confining pressure on the TCS could be described by the rock-strength criterion. The Hoek–Brown strength criterion proposed by Hoek–Brown could describe the failure of the broken rock mass. The expression of the Hoek–Brown strength criterion is as follows [53,54]:
σ 1 = σ 3 + m i σ c i σ 3 + σ c i 2
where σ c i is the UCS, σ 1 is the TCS, m i is the material constant, and σ 3 is the confining pressure.
Under the F&T weathering process, the water migration and transformation from water to ice caused microdefects to gradually develop, which caused the rock specimens to become more fragmented. Therefore, the Hoek–Brown strength criterion could be adopted to describe the impact of the confining pressure on the TCS of rocks after undergoing the F&T weathering process. According to the expression of the Hoek–Brown strength criterion, the relationship between the TCS σ 1 and the confining pressure σ 3 could be simplified into a quadratic function expression. Based on this simplified relationship and the model suggested by Mutlutürk et al. [26], this paper proposed a novel strength evolution model to describe the coupling impact of the quantity of F&T weathering cycles and the confining pressure on the TCS. The expression of the strength evolution model is as follows:
σ 1 = a + b σ 3 + c σ 3 2 exp d + e σ 3 + f σ 3 2 N
where a, b, c, d, e, and f are fitting parameters determined by the properties of rocks.
Using Equation (5), MATLAB was adopted to fit the experimental data of the TCS of the sandstone specimens after undergoing different quantities of F&T weathering cycles and different confining pressures; the fitting surface is shown in Figure 14 and the fitting parameters are shown in Table 3. As shown in Figure 14 and Table 3, the fitting surface agreed well with the experimental data: the correlation coefficient was up to 0.992, which indicated that the proposed model could effectively describe the coupling impact of the quantity of F&T weathering cycles and the confining pressure on the TCS of the sandstone after undergoing the F&T weathering process.
To further validate the rationality of the proposed model in this paper, Equation (5) was adopted to fit the tested value from the published literature [32,33,34,35]. The fitting surfaces are shown in Figure 15 and the fitting parameters are shown in Table 4. As shown in Figure 15 and Table 4, the fitting surfaces agreed well with the tested value and all fitting correlation coefficients were above 0.95, which proved that the strength evolution model proposed in this paper was reasonable; in addition, this model has strong applicability.

5. Conclusions

In this paper, the evolution characteristics of the F&T damage of sandstone were analyzed based on NMR techniques. Uniaxial and triaxial compression tests of sandstone after undergoing different quantities of F&T weathering cycles were conducted to investigate the coupling impact of the quantities of F&T weathering cycles and the confining pressure on the mechanical properties and failure modes. A novel strength evolution model was proposed to describe the coupling impact of the quantity of F&T weathering cycles and the confining pressure on the TCS of rocks after undergoing the F&T weathering process. The following main conclusions could be drawn from this research:
(a).
In the early stage of the F&T weathering cycles, original internal pores and microcracks were constantly developing. As the quantity of F&T weathering cycles increased, in the later stage of the F&T weathering cycles, in addition to the expansion of the original internal pores and the constant extension of microcracks, new pores and microcracks were generated; i.e., the accumulated damage inside the sandstone specimens constantly increased as the quantity of F&T weathering cycles increased.
(b).
The impact of the quantity of F&T weathering cycles on the pre-peak deformation behaviors, peak compressive strength, and residual strength was the opposite of the confining pressure while the post-peak deformation behaviors were similar to confining pressure. When the confining pressure was constant, with an increase in the quantity of F&T weathering cycles, the compaction stage became longer while the slope at the linear deformation stage, the peak compressive strength, the residual strength, and the stress-dropping rate of the post-peak all decreased. When the quantity of the F&T weathering cycles was constant, with an increase in the confining pressure, the compaction stage became shorter and the slope at the linear deformation stage, the peak compressive strength, and the residual strength all increased, but the stress-dropping rate of the post-peak decreased.
(c).
In the uniaxial compression tests, the failure mode of the sandstone specimens changed as the quantity of F&T cycles increased. The failure mode was a single inclined plane shear failure when the sandstone specimens did not undergo F&T weathering cycles. In the early and later stages of the F&T weathering cycles, the failure modes became tension–shear comprehensive failure and splitting failure, respectively. In the triaxial compression tests, the failure mode of the sandstone specimens under different confining pressures was a single inclined plane shear failure regardless of the quantity of F&T weathering cycles experienced. However, the length of the shear failure plane became shorter as the confining pressure increased.
(d).
The variation characteristics of the CPIF curves were similar under different confining pressures as the quantity of F&T weathering cycles increased; that is, the CPIF increased as the quantity of F&T weathering cycles increased. However, the higher the confining pressure was, the larger the increased amplitude of the CPIF. This indicated that the sandstone specimens were more sensitive to the confining pressure after undergoing more F&T weathering cycles.
(e).
A novel strength evolution model that could describe the coupling impact of the quantity of F&T weathering cycles and the confining pressure on the TCS of rocks after undergoing the F&T weathering process was proposed. The proposed model was cross-verified with tested values from the published literature. Fitting surfaces agreed well with the tested value and all fitting correlation coefficients were above 0.95, which proved that the strength evolution model proposed in this paper was reasonable; in addition, this model has strong applicability.

Author Contributions

Conceptualization, K.Z. and C.Y.; Data curation, X.X. and F.G.; Funding acquisition, X.X., F.G., K.Z. and J.L.; Investigation, X.X. and C.Y.; Methodology, X.X.; Supervision, K.Z.; Writing—original draft, X.X.; Writing—review & editing, F.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 51774323), the Hunan Provincial Natural Science Foundation of China (Grant Nos. 2020JJ4704 and 2020JJ4712), and the Fundamental Research Funds for the Central Universities of Central South University (Grant No. 2021zzts0279).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Matsuoka, N. Direct observation of frost wedging in alpine bedrock. Earth Surf. Process. Landf. 2001, 26, 601–614. [Google Scholar] [CrossRef]
  2. Nicholson, D.T.; Nicholson, F.H. Physical deterioration of sedimentary rocks subjected to experimental freeze-thaw weathering. Earth Surf. Process. Landf. 2000, 25, 1295–1307. [Google Scholar] [CrossRef]
  3. Matsuoka, N. Frost weathering and rockwall erosion in the southeastern Swiss Alps: Long-term (1994–2006) observations. Geomorphology 2008, 99, 353–368. [Google Scholar] [CrossRef]
  4. Lai, Y.; Wu, H.; Wu, Z.; Liu, S.; Den, X. Analytical viscoelastic solution for frost force in cold-region tunnels. Cold Reg. Sci. Technol. 2000, 31, 227–234. [Google Scholar] [CrossRef]
  5. Zhang, S.; Lai, Y.; Zhang, X.; Pu, Y.; Yu, W. Study on the damage propagation of surrounding rock from a cold-region tunnel under freeze–thaw cycle condition. Tunn. Undergr. Space Technol. 2004, 19, 295–302. [Google Scholar] [CrossRef]
  6. Kang, K.; Hu, N.; Sin, C.; Rim, S.; Han, E.; Kim, C. Determination of the mechanical parameters of rock mass based on a GSI system and displacement back analysis. J. Geophys. Eng. 2017, 14, 939–948. [Google Scholar] [CrossRef]
  7. Zhang, J.; Deng, H.; Taheri, A.; Ke, B.; Liu, C. Deterioration and strain energy development of sandstones under quasi-static and dynamic loading after freeze-thaw cycles. Cold Reg. Sci. Technol. 2019, 160, 252–264. [Google Scholar] [CrossRef]
  8. Wang, P.; Xu, J.; Liu, S.; Wang, H.; Liu, S. Static and dynamic mechanical properties of sedimentary rock after freeze-thaw or thermal shock weathering. Eng. Geol. 2016, 210, 148–157. [Google Scholar] [CrossRef] [Green Version]
  9. Tan, X.; Chen, W.; Yang, J.; Cao, J. Laboratory investigations on the mechanical properties degradation of granite under freeze–thaw cycles. Cold Reg. Sci. Technol. 2011, 68, 130–138. [Google Scholar] [CrossRef]
  10. Wang, P.; Xu, J.; Liu, S.; Liu, S.; Wang, H. A prediction model for the dynamic mechanical degradation of sedimentary rock after a long-term freeze-thaw weathering: Considering the strain-rate effect. Cold Reg. Sci. Technol. 2016, 131, 16–23. [Google Scholar] [CrossRef]
  11. Fang, X.; Xu, J.; Wang, P. Compressive failure characteristics of yellow sandstone subjected to the coupling effects of chemical corrosion and repeated freezing and thawing. Eng. Geol. 2018, 233, 160–171. [Google Scholar] [CrossRef]
  12. Khanlari, G.; Sahamieh, R.Z.; Abdilor, Y. The effect of freeze–thaw cycles on physical and mechanical properties of Upper Red Formation sandstones, central part of Iran. Arab. J. Geosci. 2014, 8, 5991–6001. [Google Scholar] [CrossRef]
  13. Momeni, A.; Abdilor, Y.; Khanlari, G.R.; Heidari, M.; Sepahi, A.A. The effect of freeze–thaw cycles on physical and mechanical properties of granitoid hard rocks. Bull. Eng. Geol. Environ. 2016, 75, 1649–1656. [Google Scholar] [CrossRef]
  14. Yavuz, H. Effect of freeze–thaw and thermal shock weathering on the physical and mechanical properties of an andesite stone. Bull. Eng. Geol. Environ. 2010, 70, 187–192. [Google Scholar] [CrossRef]
  15. Altindag, R.; Alyildiz, I.S.; Onargan, T. Mechanical property degradation of ignimbrite subjected to recurrent freeze–thaw cycles. Int. J. Rock Mech. Min. Sci. 2004, 41, 1023–1028. [Google Scholar] [CrossRef]
  16. Jamshidi, A.; Nikudel, M.R.; Khamehchiyan, M. Predicting the long-term durability of building stones against freeze–thaw using a decay function model. Cold Reg. Sci. Technol. 2013, 92, 29–36. [Google Scholar] [CrossRef]
  17. Jia, H.; Xiang, W.; Krautblatter, M. Quantifying Rock Fatigue and Decreasing Compressive and Tensile Strength after Repeated Freeze-Thaw Cycles. Permafr. Periglac. Process. 2015, 26, 368–377. [Google Scholar] [CrossRef]
  18. Zhang, J.; Fu, H.; Huang, Z.; Wu, Y.; Chen, W.; Shi, Y. Experimental study on the tensile strength and failure characteristics of transversely isotropic rocks after freeze-thaw cycles. Cold Reg. Sci. Technol. 2019, 163, 68–77. [Google Scholar] [CrossRef]
  19. Weng, L.; Wu, Z.; Taheri, A.; Liu, Q.; Lu, H. Deterioration of dynamic mechanical properties of granite due to freeze-thaw weathering Considering the effects of moisture conditions. Cold Reg. Sci. Technol. 2020, 176, 103092. [Google Scholar] [CrossRef]
  20. Li, J.; Kaunda, R.B.; Zhou, K. Experimental investigations on the effects of ambient freeze-thaw cycling on dynamic properties and rock pore structure deterioration of sandstone. Cold Reg. Sci. Technol. 2018, 154, 133–141. [Google Scholar] [CrossRef]
  21. Ke, B.; Zhou, K.; Xu, C.; Deng, H.; Li, J.; Bin, F. Dynamic Mechanical Property Deterioration Model of Sandstone Caused by Freeze—Thaw Weathering. Rock Mech. Rock Eng. 2018, 51, 2791–2804. [Google Scholar] [CrossRef]
  22. Ma, Q.; Ma, D.; Yao, Z. Influence of freeze-thaw cycles on dynamic compressive strength and energy distribution of soft rock specimen. Cold Reg. Sci. Technol. 2018, 153, 10–17. [Google Scholar] [CrossRef]
  23. Liu, C.; Deng, H.; Zhao, H.; Zhang, J. Effects of freeze-thaw treatment on the dynamic tensile strength of granite using the Brazilian test. Cold Reg. Sci. Technol. 2018, 155, 327–332. [Google Scholar] [CrossRef]
  24. Bayram, F. Predicting mechanical strength loss of natural stones after freeze–thaw in cold regions. Cold Reg. Sci. Technol. 2012, 83–84, 98–102. [Google Scholar] [CrossRef]
  25. İnce, İ.; Fener, M. A prediction model for uniaxial compressive strength of deteriorated pyroclastic rocks due to freeze–thaw cycle. J. Afr. Earth Sci. 2016, 120, 134–140. [Google Scholar] [CrossRef]
  26. Mutlutürk, M.; Altindag, R.; Türk, G. A decay function model for the integrity loss of rock when subjected to recurrent cycles of freezing–thawing and heating–cooling. Int. J. Rock Mech. Min. Sci. 2004, 41, 237–244. [Google Scholar] [CrossRef]
  27. Gao, F.; Xiong, X.; Zhou, K.; LI, J.; SHI, W. Strength deterioration model of saturated sandstone under freeze-thaw cycles. Rock Soil Mech. 2019, 40, 926–932. [Google Scholar] [CrossRef]
  28. Ghobadi, M.H.; Taleb Beydokhti, A.R.; Nikudel, M.R.; Asiabanha, A.; Karakus, M. The effect of freeze–thaw process on the physical and mechanical properties of tuff. Environ. Earth Sci. 2016, 75, 846. [Google Scholar] [CrossRef]
  29. Liu, Q.; Huang, S.; Kang, Y.; Liu, X. A prediction model for uniaxial compressive strength of deteriorated rocks due to freeze–thaw. Cold Reg. Sci. Technol. 2015, 120, 96–107. [Google Scholar] [CrossRef]
  30. Gao, F.; Cao, S.; Zhou, K.; Lin, Y.; Zhu, L. Damage characteristics and energy-dissipation mechanism of frozen–thawed sandstone subjected to loading. Cold Reg. Sci. Technol. 2020, 169, 102920. [Google Scholar] [CrossRef]
  31. Wang, L.; Li, N.; Qi, J.; Tian, Y.; Xu, S. A study on the physical index change and triaxial compression test of intact hard rock subjected to freeze-thaw cycles. Cold Reg. Sci. Technol. 2019, 160, 39–47. [Google Scholar] [CrossRef]
  32. Zhang, H.; Yuan, C.; Yang, G.; Wu, L.; Peng, C.; Ye, W.; Shen, Y.; Moayedi, H. A novel constitutive modelling approach measured under simulated freeze–thaw cycles for the rock failure. Eng. Comput. 2021, 37, 779–792. [Google Scholar] [CrossRef]
  33. Hosseini, M.; Khodayari, A.R. Effect of freeze-thaw cycle on strength and rock strength parameters (A Lushan sandstone case study). J. Min. Environ. 2019, 10, 257–270. [Google Scholar] [CrossRef]
  34. Fu, H.; Zhang, J.; Huang, Z.; Shi, Y.; Chen, W. A statistical model for predicting the triaxial compressive strength of transversely isotropic rocks subjected to freeze–thaw cycling. Cold Reg. Sci. Technol. 2018, 145, 237–248. [Google Scholar] [CrossRef]
  35. Seyed Mousavi, S.Z.; Tavakoli, H.; Moarefvand, P.; Rezaei, M. Assessing the effect of freezing-thawing cycles on the results of the triaxial compressive strength test for calc-schist rock. Int. J. Rock Mech. Min. Sci. 2019, 123, 104090. [Google Scholar] [CrossRef]
  36. Ulusay, R. The ISRM Suggested Methods for Rock Characterization, Testing and Monitoring 2007–2014; Springer International Publishing: Cham, Switzerland, 2015; p. 292. [Google Scholar]
  37. Li, J.; Zhou, K.; Liu, W.; Deng, H. NMR research on deterioration characteristics of microscopic structure of sandstones in freeze–thaw cycles. Trans. Nonferrous Met. Soc. China 2016, 26, 2997–3003. [Google Scholar] [CrossRef]
  38. Gao, F.; Wang, Q.; Deng, H.; Zhang, J.; Tian, W.; Ke, B. Coupled effects of chemical environments and freeze–thaw cycles on damage characteristics of red sandstone. Bull. Eng. Geol. Environ. 2016, 76, 1481–1490. [Google Scholar] [CrossRef]
  39. Li, J.; Zhou, K.; Zhang, Y.; Xu, Y. Experimental study of rock porous structure damage characteristics under condition of freezing-thawing cycles based on nuclear magnetic resonance technique. Chin. J. Rock Mech. Eng. 2012, 31, 1208–1214. [Google Scholar]
  40. Liu, C.; Deng, J.; Yu, S.; Li, P.; Lin, Y. Effect of Freezing and Thawing on Microstructure Damage and Dynamic Flexural Tension of Granite. Rock Mech. Rock Eng. 2020, 53, 3853–3858. [Google Scholar] [CrossRef]
  41. Shen, Y.; Wang, Y.; Wei, X.; Jia, H.; Yan, R. Investigation on meso-debonding process of the sandstone–concrete interface induced by freeze–thaw cycles using NMR technology. Constr. Build. Mater. 2020, 252, 118962. [Google Scholar] [CrossRef]
  42. Sun, Y.; Zhai, C.; Xu, J.; Cong, Y.; Qin, L.; Zhao, C. Characterisation and evolution of the full size range of pores and fractures in rocks under freeze-thaw conditions using nuclear magnetic resonance and three-dimensional X-ray microscopy. Eng. Geol. 2020, 271, 105616. [Google Scholar] [CrossRef]
  43. Wang, F.; Cao, P.; Wang, Y.; Hao, R.; Meng, J.; Shang, J. Combined effects of cyclic load and temperature fluctuation on the mechanical behavior of porous sandstones. Eng. Geol. 2020, 266, 105466. [Google Scholar] [CrossRef]
  44. Xie, S.; Han, Z.; Lin, H. A quantitative model considering crack closure effect of rock materials. Int. J. Solids Struct. 2022, 251, 111758. [Google Scholar] [CrossRef]
  45. Pak, J.; Hyun, C.; Park, H. Changes in microstructure and physical properties of rocks caused by artificial freeze–thaw action. Bull. Eng. Geol. Environ. 2015, 74, 555–565. [Google Scholar] [CrossRef]
  46. Li, B.; Zhu, Z.; Ning, J.; Li, T.; Zhou, Z. Viscoelastic–plastic constitutive model with damage of frozen soil under impact loading and freeze–thaw loading. Int. J. Mech. Sci. 2022, 214, 106890. [Google Scholar] [CrossRef]
  47. Rummel, F.; Fairhurst, C. Determination of the post-failure behavior of brittle rock using a servo-controlled testing machine. Rock Mech. Rock Eng. 1970, 2, 189–204. [Google Scholar] [CrossRef]
  48. Zhang, F.; Sheng, Q.; Zhu, Z.; Zhang, Y. Study on post-peak mechanical behaviour and strain-softening model of three gorges granite. Chin. J. Rock Mech. Eng. 2008, 27, 2651–2655. [Google Scholar]
  49. Yao, M.; Rong, G.; Zhou, C.; Peng, J. Effects of Thermal Damage and Confining Pressure on the Mechanical Properties of Coarse Marble. Rock Mech. Rock Eng. 2016, 49, 2043–2054. [Google Scholar] [CrossRef]
  50. Zhao, J.; Feng, X.-T.; Zhang, X.-W.; Zhang, Y.; Zhou, Y.-Y.; Yang, C.-X. Brittle-ductile transition and failure mechanism of Jinping marble under true triaxial compression. Eng. Geol. 2018, 232, 160–170. [Google Scholar] [CrossRef]
  51. Wang, P.; Xu, J.; Liu, S.; Wang, H. Dynamic mechanical properties and deterioration of red-sandstone subjected to repeated thermal shocks. Eng. Geol. 2016, 212, 44–52. [Google Scholar] [CrossRef]
  52. Cadoni, E. Dynamic Characterization of Orthogneiss Rock Subjected to Intermediate and High Strain Rates in Tension. Rock Mech. Rock Eng. 2010, 43, 667–676. [Google Scholar] [CrossRef] [Green Version]
  53. Hoek, E.; Brown, E.T. Practical Estimates of Rock Mass Strength. Int. J. Rock Mech. Min. Sci. 1997, 34, 1165–1186. [Google Scholar] [CrossRef]
  54. Hoek, E.; Carranza-Torres, C.; Corkum, B. Hoek-brown failure criterion—2002 edition. In Proceedings of the 5th North American Rock Mechanics Symposium and the 17th Tunnelling Association of Canada Conference—NARMS-TAC 2002, Toronto, ON, Canada, 7–10 July 2002; pp. 267–273. [Google Scholar]
Figure 1. Location of the sampling site: (a) location of the Jiama open pit copper mine; (b) distribution map for frozen soil of China.
Figure 1. Location of the sampling site: (a) location of the Jiama open pit copper mine; (b) distribution map for frozen soil of China.
Mathematics 10 03841 g001
Figure 2. Rock Specimens: (a) specimens for uniaxial compression tests; (b) specimens for triaxial compression tests.
Figure 2. Rock Specimens: (a) specimens for uniaxial compression tests; (b) specimens for triaxial compression tests.
Mathematics 10 03841 g002
Figure 3. Experiment instrument: (a) TDS-300 automatic freeze–thaw test machine; (b) MesoMR23-060H-I NMR system; (c) MTS815 electrohydraulic servo-controlled rock-testing machine.
Figure 3. Experiment instrument: (a) TDS-300 automatic freeze–thaw test machine; (b) MesoMR23-060H-I NMR system; (c) MTS815 electrohydraulic servo-controlled rock-testing machine.
Mathematics 10 03841 g003
Figure 4. Temperature variation curves for one F&T weathering cycle.
Figure 4. Temperature variation curves for one F&T weathering cycle.
Mathematics 10 03841 g004
Figure 5. T2 spectrum distribution curves for sandstone specimens at different quantities of F&T weathering cycles: (a) 20 F&T weathering cycles; (b) 40 F&T weathering cycles.
Figure 5. T2 spectrum distribution curves for sandstone specimens at different quantities of F&T weathering cycles: (a) 20 F&T weathering cycles; (b) 40 F&T weathering cycles.
Mathematics 10 03841 g005
Figure 6. Stress–strain curves for sandstone specimens after undergoing different quantities of F&T weathering cycles in uniaxial compression tests.
Figure 6. Stress–strain curves for sandstone specimens after undergoing different quantities of F&T weathering cycles in uniaxial compression tests.
Mathematics 10 03841 g006
Figure 7. UCS and its reduction ratio for sandstone specimens after undergoing different quantities of F&T weathering cycles.
Figure 7. UCS and its reduction ratio for sandstone specimens after undergoing different quantities of F&T weathering cycles.
Mathematics 10 03841 g007
Figure 8. Failure modes of sandstone specimens after undergoing different quantities of F&T weathering cycles in uniaxial compression tests.
Figure 8. Failure modes of sandstone specimens after undergoing different quantities of F&T weathering cycles in uniaxial compression tests.
Mathematics 10 03841 g008
Figure 9. Stress–strain curves of sandstone specimens after undergoing different quantities of F&T weathering cycles under triaxial compression tests: (a) 3 MPa; (b) 6 MPa; (c) 9 MPa; (d) 12MPa.
Figure 9. Stress–strain curves of sandstone specimens after undergoing different quantities of F&T weathering cycles under triaxial compression tests: (a) 3 MPa; (b) 6 MPa; (c) 9 MPa; (d) 12MPa.
Mathematics 10 03841 g009
Figure 10. Stress–strain curves for sandstone specimens at different confining pressures (20 F&T weathering cycles).
Figure 10. Stress–strain curves for sandstone specimens at different confining pressures (20 F&T weathering cycles).
Mathematics 10 03841 g010
Figure 11. TCS and its reduction ratio for sandstone specimens after undergoing different quantities of F&T weathering cycles: (a) 3 MPa; (b) 6 MPa; (c) 9 MPa; (d) 12 MPa.
Figure 11. TCS and its reduction ratio for sandstone specimens after undergoing different quantities of F&T weathering cycles: (a) 3 MPa; (b) 6 MPa; (c) 9 MPa; (d) 12 MPa.
Mathematics 10 03841 g011
Figure 12. Confining pressure increase factor curves at different confining pressures.
Figure 12. Confining pressure increase factor curves at different confining pressures.
Mathematics 10 03841 g012
Figure 13. Failure modes of sandstone specimens at different confining pressures (20 F&T weathering cycles).
Figure 13. Failure modes of sandstone specimens at different confining pressures (20 F&T weathering cycles).
Mathematics 10 03841 g013
Figure 14. Fitting results of TCS of sandstone specimens for different quantities of F&T weathering cycles and different confining pressures.
Figure 14. Fitting results of TCS of sandstone specimens for different quantities of F&T weathering cycles and different confining pressures.
Mathematics 10 03841 g014
Figure 15. Fitting results of TCS of rocks for different quantities of F&T weathering cycles and different confining pressures: (a) from [33]; (b) from [34]; (c) from [32]; (d) from [35].
Figure 15. Fitting results of TCS of rocks for different quantities of F&T weathering cycles and different confining pressures: (a) from [33]; (b) from [34]; (c) from [32]; (d) from [35].
Mathematics 10 03841 g015
Table 1. Mineral composition of sandstone specimens.
Table 1. Mineral composition of sandstone specimens.
Rock TypeMineral Composition
Quartz (%)Kaolinite (%)Feldspar (%)Mica (%)
Sandstone88.147.443.051.37
Table 2. Uniaxial and triaxial compression test results for sandstone specimens after undergoing different quantities of F&T weathering cycles.
Table 2. Uniaxial and triaxial compression test results for sandstone specimens after undergoing different quantities of F&T weathering cycles.
Quantity of F&T Weathering CyclesConfining Pressure (MPa)Specimen IDDiameter (mm)Height (mm)Peak Compressive Strength (Mpa)
Tested ValueAverage Value
00A148.54100.2626.9427.50
A249.12100.2827.51
A348.9699.8828.05
3A449.06100.0248.7849.34
A549.1499.8447.99
A649.02100.1651.26
6A748.94100.1062.8763.28
A849.06100.1464.75
A949.00100.2862.23
9A1048.74100.2272.1274.61
A1148.88100.0875.46
A1248.96100.2076.25
12A1349.4899.8483.7685.08
A1448.9499.8885.24
A1548.04100.1686.25
100B149.12100.0225.6024.38
B249.46100.0224.11
B348.9699.8423.44
3B448.86100.0444.8844.58
B548.74100.2845.58
B648.56100.0643.28
6B748.52100.1256.8657.57
B849.10100.1257.22
B949.04100.6658.64
9B1049.02100.1870.3368.99
B1148.9899.9467.40
B1248.96100.2469.24
12B1349.1499.9481.8880.13
B1448.8299.7279.55
B1548.84100.1678.96
200C149.04100.0219.6820.79
C248.88100.1020.62
C349.1899.9022.06
3C448.96100.2439.1240.11
C549.04100.1042.89
C649.04100.1238.33
6C749.08100.1852.1153.25
C848.7699.8254.48
C948.52100.2253.15
9C1048.7499.9865.4064.24
C1148.7099.8664.06
C1249.12100.1463.25
12C1349.20100.2074.5974.28
C1449.7899.8672.66
C1548.8899.9675.58
300D149.52100.1619.0418.17
D248.8499.8818.2
D349.04100.0217.26
3D448.36100.0235.1836.47
D549.1299.7838.97
D648.76100.0635.26
6D748.82100.1648.8049.29
D849.0699.7850.81
D949.92100.0448.25
9D1048.8099.8658.2859.41
D1149.06100.0260.16
D1249.8299.9059.78
12D1349.02100.2266.6267.26
D1449.28100.1665.37
D1549.1299.9469.78
400E148.8499.8415.8816.39
E248.84100.0416.41
E348.60100.2016.88
3E449.02100.1833.0632.84
E548.8699.8831.78
E648.68100.3033.68
6E748.7699.8445.8243.94
E848.98100.0842.75
E948.68100.1443.25
9E1048.9610.1452.7352.29
E1148.88100.0249.9
E1249.08100.1454.25
12E1348.92100.1861.7761.13
E1449.1699.9259.38
E1548.90100.2062.25
Table 3. Fitting parameters of strength evolution model of sandstone specimens.
Table 3. Fitting parameters of strength evolution model of sandstone specimens.
Fitting ParametersabcdefR2RMSE
Value27.3587.412−0.2161.101 × 10−23.257 × 10−4−6.043 × 10−60.9921.731
Table 4. Fitting parameters of the strength evolution model of rock specimens from the published literature.
Table 4. Fitting parameters of the strength evolution model of rock specimens from the published literature.
Data SourceFitting Parameters
abcdefR2RMSE
[33]32.24211.493−0.5809.190 × 10−31.570 × 10−3−1.309 × 10−40.9882.249
[34]34.4444.333−0.0422.314 × 10−23.788 × 10−4−2.546 × 10−70.9774.341
[32]4.3875.155−0.3138.880 × 10−31.500 × 10−3−9.298 × 10−50.9920.688
[35]31.6805.9070.299−2.328 × 10−21.810 × 10−3−8.714 × 10−50.9655.817
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiong, X.; Gao, F.; Zhou, K.; Yang, C.; Li, J. Mechanical Properties and Strength Evolution Model of Sandstone Subjected to Freeze–Thaw Weathering Process: Considering the Confining Pressure Effect. Mathematics 2022, 10, 3841. https://0-doi-org.brum.beds.ac.uk/10.3390/math10203841

AMA Style

Xiong X, Gao F, Zhou K, Yang C, Li J. Mechanical Properties and Strength Evolution Model of Sandstone Subjected to Freeze–Thaw Weathering Process: Considering the Confining Pressure Effect. Mathematics. 2022; 10(20):3841. https://0-doi-org.brum.beds.ac.uk/10.3390/math10203841

Chicago/Turabian Style

Xiong, Xin, Feng Gao, Keping Zhou, Chun Yang, and Jielin Li. 2022. "Mechanical Properties and Strength Evolution Model of Sandstone Subjected to Freeze–Thaw Weathering Process: Considering the Confining Pressure Effect" Mathematics 10, no. 20: 3841. https://0-doi-org.brum.beds.ac.uk/10.3390/math10203841

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