Next Article in Journal
Evaluation of the Perception and Experience of Rural Natural Landscape among Youth Groups: An Empirical Analysis from Three Villages around Hefei
Next Article in Special Issue
Application of Long-Range Cross-Hole Acoustic Wave Detection Technology in Geotechnical Engineering Detection: Case Studies of Tunnel-Surrounding Rock, Foundation and Subgrade
Previous Article in Journal
Rural Tourism—Viable Alternatives for Preserving Local Specificity and Sustainable Socio-Economic Development: Case Study—“Valley of the Kings” (Gurghiului Valley, Mureș County, Romania)
Previous Article in Special Issue
Modified RMR Rock Mass Classification System for Preliminary Selection of Potential Sites of High-Level Radioactive Waste Disposal Engineering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Investigation and Micromechanical Modeling of Hard Rock in Protective Seam Considering Damage–Friction Coupling Effect

1
Pingdingshan Tianan Coal. Mining Co., Ltd., No. 12 Mine, Pingdingshan 467099, China
2
Key Laboratory of Ministry of Education for Geomechanics and Embankment Engineering, Hohai University, Nanjing 210024, China
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(23), 16296; https://doi.org/10.3390/su142316296
Submission received: 5 October 2022 / Revised: 25 November 2022 / Accepted: 2 December 2022 / Published: 6 December 2022
(This article belongs to the Special Issue Sustainability in Geology and Civil Engineering)

Abstract

:
The hard rock in the protective coal seam of the Pingdingshan Mine in China is a typical quasi-brittle material exhibiting complex mechanical characteristics. According to available research on the mechanical property, the inelastic deformation and development of damage are considered related with crack initiation and propagation, which are main causes of the material degradation. In the present study, an original experimental investigation on the rock sample of the Pingdingshan coal mine is firstly carried out to obtain the basic mechanical responses in a conventional triaxial compression test. Based on the homogenization method and thermodynamic theory, a damage–friction coupled model is proposed to simulate the non-linear mechanical behavior. In the framework of micromechanics, the hard rock in a protective coal seam is viewed as a heterogeneous material composed of a homogeneous solid matrix and a large number of randomly distributed microcracks, leading to a Representative Elementary Volume (REV), i.e., the matrix–cracks system. By the use of the Mori–Tanaka homogenization scheme, the effective elastic properties of cracked material are obtained within the framework of micromechanics. The expression of free energy on the characteristic unitary is derived by homogenization methods and the pairwise thermodynamic forces associated with the inelastic strain and damage variables. The local stress tensor is decomposed to hydrostatic and deviatoric parts, and the effective tangent stiffness tensor is derived by considering both the plastic yield law and a specific damage criterion. The associated generalized Coulomb friction criterion and damage criterion are introduced to describe the evolution of inelastic strain and damage, respectively. Prepeak and postpeak triaxial response analysis is carried out by coupled damage–friction analysis to obtain analytical expressions for rock strength and to clarify the basic characteristics of the damage resistance function. Finally, by the use of the returning mapping procedure, the proposed damage–friction constitutive model is applied to simulate the deformation of Pingdingshan hard rock in triaxial compression with respect to different confining pressures. It is observed that the numerical results are in good agreement with the experimental data, which can verify the accuracy and show the obvious advantages of the micromechanic-based model.

1. Introduction

An increasing number of large rock projects such as the Sichuan-tibet Railway tunnel project [1] and Beishan high-level nuclear waste geological disposal underground project [2] are currently being planned and developed worldwide. In order to guarantee the safety and stability of the rock mass engineering, the mechanical properties of the engineering rocks need to be mastered. The high ground stress, high gas and low permeability conditions that exist in protruding mines at a depth of 1000 meters make protrusion accidents a frequent problem [3,4]. The rock (as shown in Figure 1) here is drilled and blasted out of the coal mine tunnels and made into standard rock samples. The corresponding mechanical parameters can be obtained by experiment and theoretical simulation. This method can provide reference for the calculation of deformation and stability of the coal mine [5,6]. In this paper, the hard rock in the protective coal seam of the Pingdingshan mine is studied experimentally and analytically in order to provided a unified constitutive model for predicting its mechanical behaviors based on the obtained experimental data.
The horizontal shafts of the Pingdingshan coal mine are located a thousand meters deep underground. The hard rock is collected from the protective coal seam, which is a typical discontinuous, anisotropic and quasi-brittle material. It has a complex microstructure, with a high number of fractures and microcracks in its internal structure. In recent years, to obtain the basic mechanical properties of various rocks, the conventional triaxial compression tests are accepted as a powerful method by many scholars on this subject [7,8,9]. Over the past three decades, significant progress has been made in the modeling of plastic degradation in quasi-brittle materials [10,11,12]. Several theoretical frameworks have also been proposed for plastic damage models, including [13,14,15,16,17], only to name a few. For concrete and other related materials, isotropic and anisotropic damage models have been developed with or without the plastic coupling, for example, [18,19,20,21,22,23]. For rock materials, certain models [24,25,26,27] have been developed. On the other hand, certain discrete plasticity–damage models [28,29,30,31] have been developed with the help of the micro-plane theory and the discrete thermodynamics formulation in order to better understand the consequences of anisotropic distribution of microcracks in brittle materials [32,33,34]. Furthermore, certain discrete plasticity–damage models have been developed with the help of the micro-plane theory and the discrete thermodynamics formulation [35,36,37] in order to better understand the consequences of anisotropic distribution of microcracks in brittle materials [38,39,40]. Recently, within the framework of micromechanics, elastoplastic damage modeling has been successfully applied on rock materials by considering penny-shaped cracks [41,42].
In this paper, based on the available micromechanic-based modeling method [43,44,45], a new multiscale constitutive model simulating the mechanical properties of the hard rock in the protective coal seam of the Pingdingshan coal mine is constructed by combining homogenization theory and irreversible thermodynamics. The discussion focuses on strength prediction and parameter determination based on coupled damage–friction analysis, and new damage criteria are proposed based on the characterization of the damage evolution resistance function. The model is illustrated by a returning mapping procedure to simulate the mechanical behavior of the hard rock in conventional triaxial compression tests with respect to different confining pressures. It is emphasized that for the evolution of inelastic strains, the deformation can be still accurately predicted by the constructed model despite the application of the associated flow law. As one of the prominent advantages of the multiscale model, by using the homogenization method to obtain the full expression of the free energy on the characteristics of a unit cell to establish the basic pattern of damage friction coupling, the strengthening of the local stress contains a similar back stress weakening function, and the model parameters are greatly reduced. The physical meaning is clear and provides a convenient process for the parameter calibration and the engineering application, which are also demonstrated in this paper.

2. Experimental Investigation of the Hard Rock in Protective Coal Seam of Pingdingshan Mine

2.1. Sample Preparation and Testing Producure

The tested samples in this study are obtained from the protective coal seam of the Pingdingshan Mine in China. The type of rock is diorite. The overall rock samples are light gray surrounded by black particles (as shown in Figure 1). Rock test specifications require standard specimens to be cylindrical, with a diameter of 50 mm and a range of 48 to 52 mm allowed. The height is 100 mm and the allowable range is 95 to 105 mm. For rocks with a heterogeneous coarse-grained structure, non-standard samples are allowed, but the ratio of height to diameter should be within 2 to 2.5. After mining by blasting method, the rock is processed into standard samples with a diameter of 50 mm and a height of 100 mm. The non-parallelism error of the two end faces of the rock sample is less than 0.05 mm, the end faces are perpendicular to the axis of the rock sample, and the deviation is less than 0 . 25 . No obvious cracks are observed on the outer surface, and a good homogeneity is seen. The average density of the samples is 2.6 g/cm3.
The conventional triaxial compression tests were carried out at the Multiscale Multi-Field Coupled Rock Mechanics Laboratory of Hohai University, using a triaxial rheometer manufactured by TOP Industrie France (see Figure 2). The equipment mainly consists of a triaxial pressure chamber, an axial pressure servo pump, a perimeter pressure servo pump and a computer control system, which can realize conventional triaxial compression tests on rocks, with a wide range of application and high measurement accuracy. Pressure control adopts a high-precision electronic control servo high pressure pump, and the measurement accuracy can reach 0.01 MPa; two highly sensitive displacement sensors are used for axial displacement measurement, which can directly output the axial displacement value of the tested sample. The utilized loading technique is strain-prescribed load. The strain-prescribed load requires the instrument’s indenter to press down on the rock at a rate of 0.02 mm/min. The lateral strain deformation measuring device (the right subfigure of Figure 2) here is directly placed on the rubber sleeve outside the rock sample when in use. During the test, the computer can directly read the corresponding deformation of the measuring device. The measurement range is 20 mm, and the measurement accuracy is 10 3 mm. The system can produce the pressure in many ways, among which the axial pressure can be carried out by axial displacement control, pressure control, flow control, lateral displacement control and other loading methods.

2.2. Experimental Results

Figure 3 illustrates the curves of deviatoric stress ( σ 1 σ 3 ) versus the axial and lateral strains at various confining pressures (confining pressure P c = 0, 10, 20 and 30 MPa). The rock of the protective coal seam is noted to exhibit the characteristic mechanical properties of brittle solids. According to the linear portion of the curves, the mean Young’s modulus E = 29 GPa and the Poisson’s ratio ν = 0.08 are calculated.

3. Construction of the Damage–Friction Coupling Constitutive Model

3.1. Theoretical Framework of Elastic Damage Model

The initiation, propagation and connection of microcracks within the solid matrix are the main mechanisms leading to the regression of mechanical property and material failure. Quasi-brittle rocks with microcracks can be regarded as a kind of composite material and can be therefore studied as a Representative Elementary Volume (REV) containing a solid matrix and a large number of penny-shaped microcracks. By experimental observation, the fracture of the rock sample in conventional compression tests is mainly caused by the crack initiation and propagation. Thus, only the reduction of effective properties by cracks is considered and that by pores is ignored. Based on this, the REV is viewed as a heterogeneous system with scattered microcracks and a solid matrix that has been weakened by pores (Figure 4). The following derivations are studied within the framework of homogenization methods.
Consider an isotropic rock matrix where the fourth-order elastic stiffness tensor is expressed as:
C m = 3 k m J + 2 μ m K
where k m and μ m are the compressive bulk modulus and shear modulus of the matrix, respectively. By introducing the second-order unit tensor δ , the fourth-order tensor operators J and K , they are denoted as J i j k l = δ i j δ k l / 3 , K i j k l = ( δ i k δ j l + δ i l δ j k ) / 2 δ i j δ k l / 3 .
The existence of microcracks leads to the discontinuity of the displacement field in rock material. Therefore, the macroscopic strain ε can be decomposed into two components, namely, the elastic strain ε m and the inelastic strain ε c , corresponding to the solid matrix component and the microcrack component, respectively:
ε = ε m + ε c
Then, the relationship between macroscopic stress σ and strain ε is obtained as follows:
σ = C m : ( ε ε c )
The degree of shear expansion of the microcrack and the relative slip of the microcrack face are expressed in terms of the scalar β and the second-order tensor T , respectively. Thereby, one has the inelastic strain tensor ε c as the sum of the hydrostatic and deviatoric parts:
ε c = T + 1 3 β δ , β = t r ε c
The construction of a constitutive model in terms of damage mechanics generally includes three steps: firstly, select a suitable damage variable φ to describe the damage state of the material. Budiansky and Connell [46] proposed that the damage variable is related to the fracture density, namely, φ = N d 3 (where N is the number of microcracks per unit volume and d is the radius of the coin crack surface). Secondly, an effective elastic tensor or the expression for free energy of the REV is established, and the thermodynamic forces associated with the internal variables are derived; finally, a suitable damage criterion is proposed and the evolution equation for the damage variables is determined [2].
According to Zhu et al. [40], for damaged solids with microcracks, the free energy W of a single unit can be expressed in the following general form:
W = 1 2 ( ε ε c ) : C m : ( ε ε c ) + 1 2 ε c : C b : ε c
where W represents the elastic free energy of a solid matrix, and the term of the right side of the equation is the free energy stored in a solid matrix caused by inelastic strain related to crack, and C b is the fourth-order back stress modulus. We are taking into account the internal relationship [37] between the Mori–Tanaka homogenization method (MT) and linear elastic fracture mechanics in dealing with crack problems. In the case of isotropic and open cracks, the following effective elastic tensors can be obtained by the application of the MT method:
C h o m = 1 1 + α 1 φ 3 k m J + 1 1 + α 2 φ 2 μ m K
where α 1 , α 2 are constants only related to the Poisson coefficient ν m of the rock matrix, α 1 = 16 9 1 ( ν m ) 2 1 2 ν m , and α 2 = 32 45 ( 1 ν m ) ( 5 ν m ) ) 2 ν m . Considering the expression of C h o m , the system free energy is expressed as W = ε : C h o m ε / 2 . By combining Equation (5), we can obtain:
C b = 1 α 1 φ 3 k m J + 1 α 2 φ 2 μ m K
It should be pointed out that according to the research of Zhu et al. [40], Equation (5) is also suitable for crack closure. In this case, the energy dissipation mechanism of crack propagation and sliding friction coupling exist in the REV. The analytical relationship between inelastic strain and macroscopic strain caused by microcracks is no longer valid.
The thermaldynamic force σ c associated with the inelastic strain ε c can be determined from the system free energy, that is, the local stress acting on the crack:
σ c = W ε c = σ C b : ε c
Furthermore, σ c is decomposed into two portions: deviatoric stress s c and hydrostatic part: s c = K : σ c and p c = t r σ c / 3 . In order to describe the inelastic strain due to the sliding friction of closed cracks and to capture the compressive shear damage pattern of quasi-brittle rock materials under compression, a Coulomb-type yielding criterion based on local stress is adopted in the present study:
f s ( σ c ) = s c + α p c 0
where α is the coefficient of friction of the cracked surface of the rock material. The local stress tensor can be decomposed into deviatoric and hydrostatic parts:
s = K : σ = σ C b : ε c : K = s 1 α 2 φ 2 μ m T
p = t r σ / 3 = 1 3 σ C b : ε c : δ = p 1 α 1 φ k m β
Combined with the above equations, Equation (9) can be rewritten as:
f s = s 1 α 2 φ 2 μ m T + α ( p 1 α 1 φ k m β ) 0
where s = K : σ and p = t r σ / 3 .
Considering the expression of free energy and the theory of irreversible thermodynamics, the thermaldynamic force associated with the damage variable φ , namely the damage driving force F φ , can be derived by following equation:
F φ = W φ = 1 2 ε c : C b φ : ε c
Substituting Equations (4), (5) and (7) into Equation (13), the explicit form of damage driving force can be obtained as follows:
F φ = 1 2 α 1 φ 2 k m β 2 + 1 α 2 φ 2 μ m T : T
The damage evolution criterion based on the strain energy release rate can also be derived
f φ ( F φ , φ ) = F φ R ( φ ) 0
where R ( φ ) is the current damage evolution resistance force.

3.2. Coupled Friction–Damage Effect and the Strength Criterion

As mentioned above, in the process of rock deterioration, there are two fundamental pathways for energy dissipation in compressive stress-dominated loading condition. One is damage evolution caused by the development of microcracks, and the other is friction caused by sliding along the fissure surface accompanied by dilatation (volume expansion) caused by the non-smooth crack surface. Damage and inelastic strain constantly rise upon loading in the damage–friction coupling process.
In this study, the evolution of damage variable ω and inelastic strain ε c are determined by the associated flow rule, and the directions of the evolution are determined by the orthogonalization criteria:
φ ˙ = λ φ f φ F φ = λ φ
ε ˙ c = λ c f s σ c = λ c ( V + 1 3 α δ )
where
V = s c / s c
in which V represents the flow direction of the partial portion of inelastic strain ε c . λ φ and λ c are damage and plasticity multipliers, respectively.
Comparing the expression of Equation (18) with that of Equation (4), one has
T ˙ = λ c V , β ˙ = λ c α
Under standard triaxial loading circumstances, the evolution direction V of inelastic shear strain does not change, so the cumulative damage parameter is Λ φ = ε λ φ and cumulative inelastic variable Λ c = ε λ c . Thus,
ε c = Λ c ( V + 1 3 α δ ) , φ = Λ φ
Rock material in engineering is mainly subjected to compression, and the strength needs to be determined within the damage–friction coupling framework. It is generally believed that inelastic strain causes the strengthening behavior of the material, while damage causes the strain softening after the peak stress. Therefore, two competing non-linear mechanical mechanisms occur in the coupling process of inelastic strain and damage. It is difficult to determine the analytical form of material strength for plastic damage coupled models.
For the conventional triaxial compression loading path, in the principal stress space, the stress tensor is σ = σ 1 σ 2 σ 3 . At the same time, assuming that σ 1 < σ 2 = σ 3 , the deviatoric stress s is
s = σ 1 σ 3 3 2 1 1
When the applied stress increases monotonously, the flow direction V can be written as V = s / s
V = 1 6 2 1 1
According to the relation s c = K : σ c and p c = t r σ c / 3 , we can obtain the following equations
s c = 2 3 ( σ 1 σ 3 ) 2 μ m α 2 Λ c φ
p c = 1 3 ( σ 1 + 2 σ 3 ) k m α α 1 Λ c φ
Substituting Equation (20) into Equations (13) and (15), the damage criteria in the case of crack closure can be obtained as follows:
f φ = k m α 2 2 α 1 + μ m α 2 Λ c φ 2 R ( ω ) 0
For simplicity, let us define χ = k m α 2 2 α 1 + μ m α 2 . Substituting Equations (23) and (24) into Equation (9), one obtains
f s = 2 3 ( σ 1 σ 3 ) + 1 3 α ( σ 1 + 2 σ 3 ) 2 χ Λ c φ 0
From f φ = 0 , it follows that
Λ c φ = R ( φ ) χ
Substituting Equation (27) into Equation (26) and using the sign convention in geotechnics, it is derived:
f s = σ 1 2 α + 6 6 α σ 3 6 R ( φ ) χ 6 α 0
Noting that p = ( σ 1 + 2 σ 3 ) / 3 and q = ( σ 1 σ 3 ) , Equation (28) can be reformed into a loading function on the p-q surface as follows:
f s = q 3 2 α p 6 R ( φ ) χ 0
When the damage variable reaches its critical value φ = φ c , the damage resistance function R ( φ ) reaches its maximum value R ( φ c ) , at which point the axial stress reaches its maximum value and the material reaches its peak strength. Based on the above analysis, the rock strength envelope predicted by the coupled friction-damage model is expressed in the following form:
f s = q 3 2 α p 6 R ( φ c ) χ = 0
As mentioned above, geotechnical materials have hardening and softening properties. So, for a given loading path, R ( φ ) reaches its maximum value R ( φ c ) when the damage factor reaches its critical value φ c , namely R ( φ ) < R ( φ c ) , and the rock is in the strain-hardening phase. Meanwhile, for R ( φ ) > R ( φ c ) , the rock is in the strain-softening phase. The mathematical description of the above properties is R ( φ ) > 0 when 0 < φ < φ c ; R ( φ ) = 0 when φ = φ c ; R ( φ ) < 0 when φ < φ c . For this purpose, the dimensionless parameter ξ = φ / φ c is defined, and the following expression for the damage resistance R ( φ ) is used:
R ( φ ) = R ( φ c ) 2 ξ 1 + ξ 2

4. Numerical Simulation of Damage–Friction Coupling Model and Validation by Experimental Results

4.1. Description of the Returning Mapping Procedure

After the constitutive model and strength criterion are established, the inelastic strain ε c and damage variable ω are calculated iteratively according to the loading criterion.
The values of the k 1 loading step variables ε k 1 , T k 1 , β k 1 , φ k 1 and σ k 1 are known. The flow for calculating the kth loading step ε k , T k , β k , φ k , σ k using strain loading is shown in Figure 5.
1.
The strain increment d ε k was superimposed onto ε k 1 to estimate the strain ε k at step k, and the macroscopic stress σ k was preliminarily calculated according to Equation (3).
2.
The loading and unloading conditions are judged by the yield function f s ( σ c ) . If f s ( σ c ) > 0 , then by the consistency condition f ˙ ( ε , β , T ) = 0 and Equation (19), we find λ c and hence T k , β k ; f s ( σ c ) 0 ; then, we update the stress σ k according to ε k only.
3.
Calculate the damage driving force F φ ( k ) according to the updated T k , β k , based on Equation (14).
4.
Examine the damage yield function Equation (15). If f φ ( F φ , φ ) > 0 , apply the Newton–Raphson algorithm to calculate φ k , and if f φ ( F φ , φ ) 0 , then there is no damage increment.
5.
Update the stress σ k from Equation (3).
6.
Substitute the updated ε k , T k , β k , φ k , σ k into the next loading loop.

4.2. Determination of Model Parameters

The coupled damage–friction model proposed in this paper contains only five parameters, E m , ν m , α , R ( φ c ) and φ c , which can all be determined by a set of conventional triaxial mechanical tests. The Young’s modulus and Poisson’s coefficient are taken as E m = 30 GPa and ν m = 0.1 for the hard rock samples from the Pingdingshan coal mine. The remaining parameters were determined as follows:
The parameter φ c included in the damage criterion is the damage threshold value, the value of which corresponds to the damage value at the peak stress. According to Lockner [47], the value of φ c is approximately linearly related to the perimeter pressure. For simplicity, a constant value of φ c = 1.5 is taken here. The effect of the value of φ c on the numerical simulation results will be discussed through parametric sensitivity analysis. As mentioned above, when ω = ω c that is, R ( φ c ) = R ( φ c ) , the material reaches its maximum axial stress (the intensity envelope on the p-q surface is shown in Figure 6). Based on the protected hard rock test data, the parameter values can be determined by applying Equation (30): α 0 = 1.3, R ( φ c ) = 5.34 × 10 2 MPa.

4.3. Numerical Simulations for Diorite in Pingdingshan Coal Mine

Utilizing the values of the model parameters established in the preceding subsection, the stress–strain data from conventional triaxial compression tests are numerically simulated for different envelope pressure conditions (envelope pressure P c = 0, 10, 20 and 30 MPa), as shown in Figure 7, Figure 8, Figure 9 and Figure 10. The model more accurately simulates the main macro-mechanical behavior of the rock, and it better describes the strength and stress-softening characteristics of the rock under different circumferential pressure conditions, especially in the elastic phase. Both in the axial and lateral directions, the stress–strain relationship is more accurately described; after the material enters non-linear deformation, before the peak strength, the model also fits the lateral deformation of the rock relatively well.
According to the above simulated results, it can be seen that the constructed damage–friction coupling model in this paper can capture the main properties of the stress–strain curves of Pingdingshan hard rock under different confining pressures (envelope pressure P c = 0, 10, 20 and 30 MPa). Especially for confining pressures of 0 MPa, 10 MPa, and 20 MPa, this model can better simulate the peak pressure and related damage. However, when the confining pressure is 30 MPa, the simulation result of the height strength is smaller than the experimental data. This is because the rock sample in this experiment is obtained by the blasting method, so some microcracks exist inside the body, resulting in a certain dispersion of its peak strength. It must be pointed out that small but noticeable differences on the lateral strain are observed in the above figures. This is due to the fact that the tested tock is collected close to the coal seam, so there is more coal inclusions distributed in the rock solids, resulting in a large discrete property of the rock samples. Since the rock samples in the Pingdingshan Coal mine are obtained by an explosive method, unnatural cracks appear in the rock samples, which may lead to irregular lateral deformation. For most geomatrials, non-associated plastic flow rules are usually adopted to study the lateral and volumetric strains. However, in the present study, for the sake of simplicity, an associated flow law is adopted, which may result in a large difference between the experimental and analytical results of lateral strain.

4.4. Sensitivity Analysis of Parameter φ c

In order to determine the effect of the parameter φ c on the numerical simulation results, a conventional test with an envelope pressure of 10 MPa is used as the subject of the study. Figure 11 compares the results of the simulations for φ c values of 1.5, 2.5 and 3.5, respectively. It can be seen that the higher the φ c is, the greater the axial and lateral strains at the peak stress point are and the smoother the hardening and softening curve, namely, the slower the rate of hardening and softening, due to the larger the critical value of the damage. In addition, the number of microcracks within the characteristic cell or their radius is greater when the critical value is reached, which is the same for the material deterioration and the axial and lateral strains. The strengthening speed becomes smaller when the inelastic strain between the elastic stage and the peak stress point of REV become larger.
As mentioned before, the magnitude of the parameter φ c is approximately linearly related to the surrounding pressure. For the sake of simplicity, a constant value was taken in the numerical simulations, which need further improvement of the simulation results for different envelope pressures. At the same time, due to the associated flow rule, the magnitude of lateral strain is measured when the stress peak point is reached at the stress-softening rate. The smaller φ c is, the faster the lateral strain reaches the peak, and the faster the stress-softening rate.

4.5. Numerical Simulations for Dagangshan Diabase

For further validation of the proposed model, comparisons between experimental results of the Dagangshan diabase and the numerical simulations are carried out in this subsection. The experimental data are provided by an earlier work [48]. Now, we will provide numerical simulations of standard triaxial compression tests conducted with confining pressures of p c = 10, 15, 20, 30, 50 MPa. The specific model parameters used are listed in Table 1 below. The results are illustrated in Figure 12, Figure 13, Figure 14, Figure 15 and Figure 16 with respect to confining pressures. It can be seen that the predicted numerical solutions for both low and high confining pressure levels match quite well with the experimental data.

5. Conclusions

Based on homogenization and irreversible thermodynamics theories, a coupled damage–friction model for Pingdingshan hard rock in a protective coal seam and Dagangshan diabase are constructed by the use of an associated flow law. The model successfully simulates the main mechanical behaviors of the tested rock in a conventional triaxial compression test. In order to assess the proposed coupled plastic–damage model, numerical simulations of triaxial compression tests on Pingdingshan rock and Dagangshan diabase have been carried out. By comparing with the experimental data, the proposed model can well describe the peak strengths and the mechanical responses from low to high levels of confining pressure for two types of rocks. However, the lateral deformation around the peak stress is not accurately predicted by the model due to the associated plastic flow rule. Determination of the critical damage parameter is also crucial, which determines the peak strength.
Determination of the free energy expressions for the RVE of cracked rock based on homogenization theory is the key point of the research results, which are combined with thermodynamic theory to the closed damage-friction coupling and strain strengthening and weakening processes. The proposed multiscale intrinsic model has the advantage of having only five parameters, which have specific physical meanings, and can be easily determined by the experiments. Inelastic deformation and the propagation of cracks are two of the fundamental mechanisms of material damage and destruction. Therefore, the coupled damage–friction model is more suitable for describing the mechanical behavior of such quasi-brittle material. In geotechnics, non-associated flow law is generally considered to be necessary. This model is suitable for quasi-brittle rock and closed fracture. Secondly, the selection of damage parameters in this model is related to the fracture density. After rock is compressed, cracks in its interior will be derived and developed, which is consistent with the material failure mechanism described above. However, in this study, the associated flow law of inelastic strain is adopted to predict the deformation of the tested rock.

Author Contributions

Software, C.S.; Validation, S.H.; Investigation, S.H.; Data curation, W.L.; Writing—original draft, C.Z. and W.L.; Writing—review & editing, J.Z.; Visualization, C.Z. and C.S.; Supervision, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the project “Research of high efficiency precracking technology for working face in rock protective coal seam of kilometer-deep mine”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The experimental data used to support the findings of this study are available from the corresponding author upon request.

Acknowledgments

The authors would like to thank the project "Research of high efficiency precracking technology for working face in rock protective coal seam of kilometer-deep mine" for the funding support.

Conflicts of Interest

The writers affirm that they do not have any conflict of interest, either financial or otherwise.

References

  1. Zhang, D.; Sun, Z.; Fang, Q. Scientific problems and research proposals for Sichuan–Tibet railway tunnel construction. Undergr. Space 2022, 7, 419–439. [Google Scholar] [CrossRef]
  2. Zhu, Q.; Liu, H.; Wang, W.; Shao, J. A micromechanical constitutive damage model for beishan granite. Chin. J. Rock Mech. Eng. 2015, 34, 433–439. [Google Scholar]
  3. Kim, M.M.; Ko, H.Y. Multistage triaxial testing of rocks. Geotech. Test. J. 1979, 2, 98–105. [Google Scholar]
  4. Xie, H.; He, C. Study of the unloading characteristics of a rock mass using the triaxial test and damage mechanics. Int. J. Rock Mech. Min. Sci. 2004, 41, 74–80. [Google Scholar] [CrossRef]
  5. Liu, A.; Liu, S. A fully-coupled water-vapor flow and rock deformation/damage model for shale and coal: Its application for mine stability evaluation. Int. J. Rock Mech. Min. Sci. 2021, 146, 104880. [Google Scholar] [CrossRef]
  6. Wei, X.; Sun, Q. Damage mechanism and constitutive model of surface paste disposal of a coal mine collapsed pit in a complicated environment. Constr. Build. Mater. 2021, 304, 124637. [Google Scholar] [CrossRef]
  7. Chang, C.; Haimson, B. A failure criterion for rocks based on true triaxial testing. In The ISRM Suggested Methods for Rock Characterization, Testing and Monitoring: 2007–2014; Springer: Berlin/Heidelberg, Germany, 2012; pp. 259–262. [Google Scholar]
  8. Höfer, K.; Thoma, K. Triaxial tests on salt rocks. In International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts; Elsevier: Amsterdam, The Netherlands, 1968; Volume 5, pp. 195–196. [Google Scholar]
  9. Lane, K.; Heck, W. Triaxial testing for strength of rock joints. In Proceedings of the 6th US Symposium on Rock Mechanics (USRMS). OnePetro, Rolla, MI, USA, 28–30 October 1964. [Google Scholar]
  10. Dragon, A.; Mroz, Z. A continuum model for plastic-brittle behaviour of rock and concrete. Int. J. Eng. Sci. 1979, 17, 121–137. [Google Scholar] [CrossRef]
  11. Frantziskonis, G.; Desai, C. Elastoplastic model with damage for strain softening geomaterials. Acta Mech. 1987, 68, 151–170. [Google Scholar] [CrossRef]
  12. Maugin, G.A. The Thermomechanics of Plasticity and Fracture; Cambridge University Press: Cambridge, UK, 1992; Volume 7. [Google Scholar]
  13. Hansen, N.; Schreyer, H. A thermodynamically consistent framework for theories of elastoplasticity coupled with damage. Int. J. Solids Struct. 1994, 31, 359–389. [Google Scholar] [CrossRef]
  14. Lubarda, V.; Krajcinovic, D. Some fundamental issues in rate theory of damage-elastoplasticity. Int. J. Plast. 1995, 11, 763–797. [Google Scholar] [CrossRef]
  15. Lawn, B.R.; Marshall, D.B. Nonlinear stress-strain curves for solids containing closed cracks with friction. J. Mech. Phys. Solids 1998, 46, 85–113. [Google Scholar] [CrossRef]
  16. de Sciarra, F.M. Hardening plasticity with nonlocal strain damage. Int. J. Plast. 2012, 34, 114–138. [Google Scholar] [CrossRef]
  17. Paliwal, B.; Ramesh, K. An interacting micro-crack damage model for failure of brittle materials under compression. J. Mech. Phys. Solids 2008, 56, 896–923. [Google Scholar] [CrossRef]
  18. Yazdani, S.; Schreyer, H. Combined plasticity and damage mechanics model for plain concrete. J. Eng. Mech. 1990, 116, 1435–1450. [Google Scholar] [CrossRef]
  19. Abu-Lebdeh, T.M.; Voyiadjis, G.Z. Plasticity-damage model for concrete under cyclic multiaxial loading. J. Eng. Mech. 1993, 119, 1465–1484. [Google Scholar] [CrossRef]
  20. Luccioni, B.M.; Rougier, V.C. A plastic damage approach for confined concrete. Comput. Struct. 2005, 83, 2238–2256. [Google Scholar] [CrossRef]
  21. Jason, L.; Huerta, A.; Pijaudier-Cabot, G.; Ghavamian, S. An elastic plastic damage formulation for concrete: Application to elementary tests and comparison with an isotropic damage model. Comput. Methods Appl. Mech. Eng. 2006, 195, 7077–7092. [Google Scholar] [CrossRef] [Green Version]
  22. Grassl, P.; Jirásek, M. Plastic model with non-local damage applied to concrete. Int. J. Numer. Anal. Methods Geomech. 2006, 30, 71–90. [Google Scholar] [CrossRef]
  23. Wu, J.Y.; Li, J.; Faria, R. An energy release rate-based plastic-damage model for concrete. Int. J. Solids Struct. 2006, 43, 583–612. [Google Scholar] [CrossRef] [Green Version]
  24. Khan, A.S.; Xiang, Y.; Huang, S. Behavior of Berea sandstone under confining pressure part I: Yield and failure surfaces, and nonlinear elastic response. Int. J. Plast. 1991, 7, 607–624. [Google Scholar] [CrossRef]
  25. Chiarelli, A.S.; Shao, J.F.; Hoteit, N. Modeling of elastoplastic damage behavior of a claystone. Int. J. Plast. 2003, 19, 23–45. [Google Scholar] [CrossRef]
  26. Salari, M.; Saeb, S.a.; Willam, K.; Patchet, S.; Carrasco, R. A coupled elastoplastic damage model for geomaterials. Comput. Methods Appl. Mech. Eng. 2004, 193, 2625–2643. [Google Scholar] [CrossRef]
  27. Conil, N.; Djeran-Maigre, I.; Cabrillac, R.; Su, K. Thermodynamics modelling of plasticity and damage of argillite. Comptes Rendus Mec. 2004, 332, 841–848. [Google Scholar] [CrossRef]
  28. Shao, J.F.; Jia, Y.; Kondo, D.; Chiarelli, A.S. A coupled elastoplastic damage model for semi-brittle materials and extension to unsaturated conditions. Mech. Mater. 2006, 38, 218–232. [Google Scholar] [CrossRef]
  29. Chen, L.; Shao, J.; Zhu, Q.Z.; Duveau, G. Induced anisotropic damage and plasticity in initially anisotropic sedimentary rocks. Int. J. Rock Mech. Min. Sci. 2012, 51, 13–23. [Google Scholar] [CrossRef]
  30. Lai, Y.; Jin, L.; Chang, X. Yield criterion and elasto-plastic damage constitutive model for frozen sandy soil. Int. J. Plast. 2009, 25, 1177–1205. [Google Scholar] [CrossRef]
  31. Parisio, F.; Samat, S.; Laloui, L. Constitutive analysis of shale: A coupled damage plasticity approach. Int. J. Solids Struct. 2015, 75, 88–98. [Google Scholar] [CrossRef]
  32. Klisiński, M.; Mroz, Z. Description of inelastic deformation and degradation of concrete. Int. J. Solids Struct. 1988, 24, 391–416. [Google Scholar] [CrossRef]
  33. Lubliner, J.; Oliver, J.; Oller, S.; Oñate, E. A plastic-damage model for concrete. Int. J. Solids Struct. 1989, 25, 299–326. [Google Scholar] [CrossRef]
  34. Zhu, Q.; Shao, J.F.; Mainguy, M. A micromechanics-based elastoplastic damage model for granular materials at low confining pressure. Int. J. Plast. 2010, 26, 586–602. [Google Scholar] [CrossRef]
  35. Gambarotta, L.; Lagomarsino, S. A microcrack damage model for brittle materials. Int. J. Solids Struct. 1993, 30, 177–198. [Google Scholar] [CrossRef]
  36. Prat, P.C.; Bažant, Z.P. Tangential stiffness of elastic materials with systems of growing or closing cracks. J. Mech. Phys. Solids 1997, 45, 611–636. [Google Scholar] [CrossRef]
  37. Pensée, V.; Kondo, D.; Dormieux, L. Micromechanical analysis of anisotropic damage in brittle materials. J. Eng. Mech. 2002, 128, 889–897. [Google Scholar] [CrossRef]
  38. François, B.; Dascalu, C. A two-scale time-dependent damage model based on non-planar growth of micro-cracks. J. Mech. Phys. Solids 2010, 58, 1928–1946. [Google Scholar] [CrossRef]
  39. Zhu, Q.; Shao, J.F. A refined micromechanical damage–friction model with strength prediction for rock-like materials under compression. Int. J. Solids Struct. 2015, 60, 75–83. [Google Scholar] [CrossRef]
  40. Zhu, Q.Z.; Kondo, D.; Shao, J. Micromechanical analysis of coupling between anisotropic damage and friction in quasi brittle materials: Role of the homogenization scheme. Int. J. Solids Struct. 2008, 45, 1385–1405. [Google Scholar] [CrossRef]
  41. Yuan, S.S.; Zhu, Q.Z.; Zhao, L.Y.; Chen, L.; Shao, J.F.; Zhang, J. Micromechanical modelling of short-and long-term behavior of saturated quasi-brittle rocks. Mech. Mater. 2020, 142, 103298. [Google Scholar] [CrossRef]
  42. Hu, K.; Zhu, Q.; Chen, L.; Shao, J.; Liu, J. A micromechanics-based elastoplastic damage model for rocks with a brittle–ductile transition in mechanical response. Rock Mech. Rock Eng. 2018, 51, 1729–1737. [Google Scholar] [CrossRef]
  43. Cicekli, U.; Voyiadjis, G.Z.; Al-Rub, R.K.A. A plasticity and anisotropic damage model for plain concrete. Int. J. Plast. 2007, 23, 1874–1900. [Google Scholar] [CrossRef]
  44. Voyiadjis, G.Z.; Taqieddin, Z.N.; Kattan, P.I. Anisotropic damage–plasticity model for concrete. Int. J. Plast. 2008, 24, 1946–1965. [Google Scholar] [CrossRef]
  45. Zhu, Q.; Kondo, D.; Shao, J.; Pensee, V. Micromechanical modelling of anisotropic damage in brittle rocks and application. Int. J. Rock Mech. Min. Sci. 2008, 45, 467–477. [Google Scholar] [CrossRef]
  46. Budiansky, B.; O’connell, R.J. Elastic moduli of a cracked solid. Int. J. Solids Struct. 1976, 12, 81–97. [Google Scholar] [CrossRef]
  47. Lockner, D.A. A generalized law for brittle deformation of Westerly granite. J. Geophys. Res. Solid Earth 1998, 103, 5107–5123. [Google Scholar] [CrossRef]
  48. Chen, F. Rheological Strength Aging Model of Hard Brittle Fracture Rock in High Dam Zone and Its Engineering Application. Ph.D. Thesis, Shandong University, Jinan, China, 2009. [Google Scholar]
Figure 1. Hard rock samples obtained from the protective coal seam of the Pingdingshan Mine.
Figure 1. Hard rock samples obtained from the protective coal seam of the Pingdingshan Mine.
Sustainability 14 16296 g001
Figure 2. Autonomous and auto-compensated multi-field coupling testing system.
Figure 2. Autonomous and auto-compensated multi-field coupling testing system.
Sustainability 14 16296 g002
Figure 3. The stress–strain curves of the hard rock in a protective coal seam with respect to different confining pressures.
Figure 3. The stress–strain curves of the hard rock in a protective coal seam with respect to different confining pressures.
Sustainability 14 16296 g003
Figure 4. Representative Elementary Volume of hard rock considering microcracks.
Figure 4. Representative Elementary Volume of hard rock considering microcracks.
Sustainability 14 16296 g004
Figure 5. Flow chart of the numerical returning mapping procedure.
Figure 5. Flow chart of the numerical returning mapping procedure.
Sustainability 14 16296 g005
Figure 6. Strength envelope of the Pingdingshan hard rock under triaxial compressions.
Figure 6. Strength envelope of the Pingdingshan hard rock under triaxial compressions.
Sustainability 14 16296 g006
Figure 7. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 0 MPa.
Figure 7. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 0 MPa.
Sustainability 14 16296 g007
Figure 8. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 10 MPa.
Figure 8. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 10 MPa.
Sustainability 14 16296 g008
Figure 9. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 20 MPa.
Figure 9. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 20 MPa.
Sustainability 14 16296 g009
Figure 10. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 30 MPa.
Figure 10. Simulation of conventional triaxial compression test on hard rock protection layer rock in the Pingdingshan coal mine with the confining pressure of 30 MPa.
Sustainability 14 16296 g010
Figure 11. Sensitivity analysis for parameter φ c .
Figure 11. Sensitivity analysis for parameter φ c .
Sustainability 14 16296 g011
Figure 12. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 10 MPa.
Figure 12. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 10 MPa.
Sustainability 14 16296 g012
Figure 13. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 15 MPa.
Figure 13. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 15 MPa.
Sustainability 14 16296 g013
Figure 14. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 20 MPa.
Figure 14. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 20 MPa.
Sustainability 14 16296 g014
Figure 15. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 30 MPa.
Figure 15. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure of 30 MPa.
Sustainability 14 16296 g015
Figure 16. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure 50 MPa.
Figure 16. Simulation of conventional triaxial compression test on Dagangshan diabase with the confining pressure 50 MPa.
Sustainability 14 16296 g016
Table 1. Identified parameters in the numerical simulations for Dagangshan diabase.
Table 1. Identified parameters in the numerical simulations for Dagangshan diabase.
E m (MPa) ν m α 0 φ c R ( φ c )
55,0000.21.655.5 1.3 × 10 2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, C.; Liu, W.; Shi, C.; Hu, S.; Zhang, J. Experimental Investigation and Micromechanical Modeling of Hard Rock in Protective Seam Considering Damage–Friction Coupling Effect. Sustainability 2022, 14, 16296. https://doi.org/10.3390/su142316296

AMA Style

Zhang C, Liu W, Shi C, Hu S, Zhang J. Experimental Investigation and Micromechanical Modeling of Hard Rock in Protective Seam Considering Damage–Friction Coupling Effect. Sustainability. 2022; 14(23):16296. https://doi.org/10.3390/su142316296

Chicago/Turabian Style

Zhang, Chuangye, Wenyong Liu, Chong Shi, Shaobin Hu, and Jin Zhang. 2022. "Experimental Investigation and Micromechanical Modeling of Hard Rock in Protective Seam Considering Damage–Friction Coupling Effect" Sustainability 14, no. 23: 16296. https://doi.org/10.3390/su142316296

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