Next Article in Journal
Impact Properties of Novel Natural Fibre Metal Laminated Composite Materials
Next Article in Special Issue
Applications of Electromagnetic Induction and Electrical Resistivity Tomography for Digital Monitoring and Assessment of the Soil: A Case Study of Al-Ahsa Oasis, Saudi Arabia
Previous Article in Journal
Ecosystem Service Assessments within the EU Water Framework Directive: Marine Mussel Cultivation as a Controversial Measure
Previous Article in Special Issue
An Improved Aeromagnetic Compensation Method Robust to Geomagnetic Gradient
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Anisotropic Effective Elastic Properties for Multi-Dimensional Fractured Models

1
Shandong Provincial Key Laboratory of Deep Oil and Gas, China University of Petroleum (East China), Qingdao 266580, China
2
School of Geosciences, China University of Petroleum (East China), Qingdao 266580, China
3
Laboratory for Marine Mineral Resources, Qingdao National Laboratory for Marine Science and Technology, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(4), 1873; https://doi.org/10.3390/app12041873
Submission received: 27 December 2021 / Revised: 8 February 2022 / Accepted: 9 February 2022 / Published: 11 February 2022
(This article belongs to the Special Issue Advances in Applied Geophysics)

Abstract

:

Featured Application

The proposed method in the paper has significance in calculating the effective elastic properties of multi-dimensional fracture models and interaction mechanisms.

Abstract

The size, distribution, and orientation of fractures are generally multiscale and multi-dimensional in nature, leading to complex anisotropic characteristics. Theoretical or semi-analytical methods to determine the effective elastic properties depend on several assumptions, including the absence of the stress interaction and idealized fractures. On the basis of finite-element models, we conduct numerical oscillatory relaxation tests for determining the effective elastic properties of fractured rocks. The numerical approach for calculating equivalent stiffness tensors in two-dimensions is compared to the theoretical models for different fracture densities. Due to fracture interactions at high fracture densities, the suggested model makes a physical prediction. The effective elastic properties obtained from the application to a real fractured model, established from an outcrop, obviously disperse at different frequencies, which can be used to investigate fracture interactions and dynamic stress disturbances. The algorithm is extended to three-dimensional cases and also validated by using conventional effective medium theories. It is found that the fracture density obviously impacts the effective anisotropy properties, and the proposed method gives a reasonable prediction for high-fracture density. This work is significant because it enables the calculation of effective elastic properties of multi-dimensional fractured models and the fracture interaction mechanism.

1. Introduction

Cracks are the key factor in determining the deformability of rocks, properties of fluid movement, and tectonic history. We need to enhance ways for detecting and characterizing fractures. Due to the fact that natural fractures in geological formations frequently have a complicated shape and distribution, their interactions cause the complex anisotropic features of effective elastic properties. Thus, knowing the natural fractures’ elastic behavior is essential for interpreting and predicting seismic responses in terms of rock properties and enables us to bridge the gaps in the spatial and temporal scale [1,2]. The overall effective elastic coefficient of a fractured rock is frequently composed of the stiffness tensor of the background host rock and fractures [3], which attract considerable attention because of their contribution to the elastic property. The theoretical medium models (TMM) have been widely applied to calculate the effective elastic properties of fractured rocks analytically [4,5,6,7,8]. However, due to the dilute assumption of fracture densities, these models ignore fracture interactions. The second-order Hudson’s model takes into account the pair-wise interaction among cracks [9]. Chapman [10] proposed an equivalent-media theory that was sensitive to fracture length and studied the frequency-dependent anisotropy of fractured porous rocks using the squirt flow mechanism. Guo et al. [11] united linear-slip theory and anisotropic Gassmann equations to analyze equivalent stiffness coefficients. The self-consistent approximation (SCA) and differential effective medium (DEM) were applied to control the concentration of fractures to investigate the consequences of fracture interactions [12,13,14]. Jakobsen et al. [15] applied the T-matrix method to characterize the pores–fracture interaction. Meanwhile, the idealization and oversimplification of real rock severely limit the utility of theoretical or semi-analytical models.
In an attempt to circumvent certain assumptions inherent in effective medium theories [16,17,18,19,20,21,22,23], numerical simulations have been extensively applied for studying the elastic properties of complex rocks. Numerical methods can be used to characterize real fractures with different shapes, sizes, partial interconnections, and intersections, as well as provide independent validation of theoretical methods and laboratory measurements as an efficient means of determining their general applicability. Masson et al. [24] used numerical experiments to analyze the correlation between material structure and seismic attenuation anisotropy in porous media. Rubino et al. [25,26] employed 2D numerical oscillatory relaxation experiments to study the frequency-dependent effective stiffness matrix using the least-squares optimal approach. The method is used to characterize the frequency- and incidence-angle- dependent attenuation and velocity dispersion in cracked porous media for the P wave [27,28,29] and P-SV wave [30]. Jian et al. [31] expanded Rubino’s relaxation tests to a three-dimensional (3D) case for computing the equivalent stiffness matrix.
We apply Rubino’s numerical method to explore the elastic and seismic anisotropy characteristics of real fractures in this paper. Particular attention should be paid to the anisotropy variation of fractures generated by different frequencies and crack densities. The following is the outline. In the first part of the technique, we demonstrate the novel form of Rubino’s 2D algorithm using three oscillatory relaxation tests with different fracture densities. Additionally, six oscillatory relaxation tests are used to spread Rubino’s algorithm to three-dimensional fractured media. Following that, the effective stiffness of the 2D numerical method is compared with that of the classical Hudson’s model and the Eshelby–Cheng model with different fracture densities. Then, 3D fractured media are compared with TMM in elastic responses to different fracture densities. Three 2D fracture models are presented to illustrate the primary mechanism of stress interreaction and their effects on the entire effective elasticity. Finally, the method is applied to a real fractured medium generated from an outcrop, demonstrating that frequency influences the stiffness tensor anisotropy. The anisotropic effective elastic properties have been discovered to have a direct effect on fracture density. This study aims to analyze the anisotropic effective elastic properties of multi-dimensional (2D and 3D) fracture models at different fracture densities.

2. Methodology

This section mainly introduces the methodology of studying anisotropic effective elastic properties of multi-dimensional fracture models. The entire modeling methodology is illustrated in Figure 1. First, we demonstrate the simulation equations and three oscillatory relaxation tests. Next, the 2D equivalent stiffness matrix is calculated by solving the simulation equations using the finite-element method (FEM). Following that, the effective stiffness of the 2D numerical method is compared with that of Hudson’s model and the Eshelby–Cheng model with different fracture densities. Additionally, six oscillatory relaxation tests are used to spread to three-dimensional fractured media.

2.1. Anisotropic Effective Elastic Properties for 2D Fractured Models

2.1.1. Simulation Equations

We have to use numerical methods because the TMM for predicting the elasticity of multi-scale real intersecting cracks is limited by its own assumptions. We here apply Rubino’s 2D numeric upscaling procedure in an attempt to pattern natural cracks. Biot’s consolidation system of equations in the space-frequency domain is expressed as [26].
· σ = 0 ,
P f = i ω η κ W .
where represents the spatial derivatives; σ ( σ i j ) is total stress tensor; Pf denotes the fluid pressure, κ the permeability and ω the angular frequency; and W denotes average relative fluid–solid displacement tensor. η indicates the viscosity of pore fluid. Thus, the stress–strain relationships [32] state that
σ i j = 2 μ e i j + δ i j ( λ u U α N ξ ) ,
e i j = 1 2 ( U i / x j + U j / x x ) ,
P f = α N U + N ξ ,  
where δij is the Kronecker delta, and ξ = W denotes the measurement of the local variations in the fluid content. U is the solid displacement vector; eij denotes strain tensor; and μ denotes the shear modulus of the rock with saturated water, which is equivalent to that of a dry frame μ m . The Biot constant α , fluid storage coefficient N, and Lame parameter λ u are expressed as
N = ( α ϕ K s + ϕ K f ) 1 ,
λ u = K m + α 2 N 2 3 μ m ,
α = 1 K m K s ,
where ϕ is the porosity; Ks denotes the bulk moduli of solid grains; Km is the bulk moduli of dry frame; and Kf denotes the bulk moduli of fluid phase.
Substituting Equation (7) to Equation (3) and taking Equation (5), we have a new form of 2D/3D total stress tensor in porous fractured medium
σ i j = 2 μ e i j + δ i j ( λ U α P f ) ,
where λ = K m 2 μ m / 3 . When i , j = x , y , it is the 2D case. When i , j = x , y , z , it is the 3D case. Substituting ξ = W into Equation (5), we have
P f = α N U N W ,
Taking Equations (2) and (10), the Pf is given by
P f = α N U N ( i k η ω P f ) .
In order to obtain the solid displacement fields, we conduct relaxation tests and solve Equations (1) and (11) with the corresponding boundary conditions applying the finite-element method (FEM).

2.1.2. 2D Oscillatory Relaxation Tests

As shown in Figure 2, we need to conduct three simulation oscillatory relaxation tests for calculating the elements of an equivalent stiffness tensor of a 2D sample of interest according to the 2D stress–strain correlation. Let D = ( 0 , L x ) × ( 0 , L y ) be an area that represents the interest model. Performing the first oscillatory relaxation test (see Figure 2a), a uniform vertical displacement is imposed on the bottom and top boundaries, with two lateral boundaries being confined and no shear stress being applied. As a result, the x- and y-directions’ boundary conditions are as follows:
{ U · ν = Δ U ,     ( x , y ) Γ b Γ t , U · ν = 0 ,                   ( x , y ) Γ l Γ r , ( σ ν ) · χ = 0 ,       ( x , y ) Γ ,
where U is the solid displacement; Δ U is the displacement shift; σ represents the total stress tensor; Γ is the model boundary; Γb, Γt, Γl, and Γr are the bottom, top, left, and right boundary, respectively, on Γ; χ is a unit tangent; and ν is the unit outer normal so that { ν , χ } is an orthonormal system on Γ.
Figure 2b illustrates the second test that is analogous to the first, but with a uniform horizontal displacement to the right and left boundaries and the two other boundaries confined. Thus, the corresponding boundary conditions are expressed as
{ U · ν = Δ U ,     ( x , y ) Γ l Γ r , U · ν = 0 ,                   ( x , y ) Γ b Γ t ( σ ν ) · χ = 0 ,       ( x , y ) Γ , ,
Figure 2c shows the third test, a shear test, applied to the interested model, where the top and bottom boundaries are loaded with displacement. The corresponding boundary conditions are
{ U · χ = Δ U ,     ( x , y ) Γ b Γ t , U · χ = 0 ,                   ( x , y ) Γ l Γ r ( σ ν ) · ν = 0 ,       ( x , y ) Γ . ,

2.1.3. Calculation of 2D Equivalent Stiffness Matrix

After computing the displacement field, the stress and strain components of each element are computed using Equations (3) and (4). The corresponding stress and strain average over the model’s volume can be calculated using these parameters
e i j m ¯ = 1 S D e i j m d S ,
σ i j m ¯ = 1 S D σ i j m d S ,
where m = 1–3 denotes the mth test, and S is the surface of domain D.
Taking the prerequisite that an equivalent homogeneous anisotropic elastic solid can stand for the average response of the interest model, the average stress and strain components can be related through the Voigt stiffness matrix MMij. In the case of general anisotropy and under plane strain conditions, this relation is given by
( σ x x m ¯ σ y y m ¯ σ x y m ¯ ) ( M 11 M 12 M 16 M 12 M 22 M 26 M 16 M 26 M 66 ) ( e x x m ¯ e y y m ¯ 2 e x y m ¯ )
where σ i j m ¯ and e i j m ¯ are cells of the averaging stress and strain vector, respectively, and m = 1–3 denotes the mth simulation test. It is worthy to note that the matrix M considers the greatest extent of anisotropy.
Thus, the six unknown components of the stiffness matrix possibly are calculated utilizing a least-squares procedure. That is, the components of equivalent stiffness matrix Mij can be obtained by minimizing the cost function F1 (M) as follows:
F 1 ( M ) M i j = 0 .
We can solve the linear system of six equations yielded by Equation (18) to obtain the equivalent stiffness tensor Mij; see Appendix A for more details.

2.2. Anisotropic Effective Elastic Properties for 3D Fractured Models

2.2.1. 3D Oscillatory Relaxation Tests

In this part, we apply Rubino’s 2D method to a 3D cracked medium [30]. The governing Equations (1)–(11) would remain the same. As illustrated in Figure 3, we need to conduct six simulation oscillatory relaxation tests for calculating the elements of an equivalent stiffness matrix for a 3D probed model according to the 3D stress–strain correlation.
Set Ω = ( 0 , L x ) × ( 0 , L y ) × ( 0 , L z ) as a domain that represents a 3D interest model. Figure 3a displays the first oscillatory relaxation test in which a uniform vertical displacement along the x-direction (blue arrow) is applied to the model’s red side, with the other sides employing periodic boundary conditions. The opposing side (gray side) is confined, and no shear stress is applied. Thus, the boundary conditions of the x-, y-, and z- directions are imposed as follows:
{ U · ν = Δ U ,       ( x , y , z ) A r , U · ν = 0 ,                     ( x , y , z ) A l , ( σ ν ) · χ = 0 ,         ( x , y , z ) A , U d s t = U s r c ,         ( x , y , z ) A f A k , U d s t = U s r c ,         ( x , y , z ) A t A b ,
where σ represents the total stress tensor; U denotes the solid displacement; Δ U denotes the displacement shift; U s r c denotes the source displacement field; U d s t denotes the displacement field of the target boundary; A is its side boundaries; Af, Ak, Al, Ar, At, and Ab are the left, right, front, back, top, and bottom boundaries, respectively; χ is a unit tangent; and ν denotes the unit outer normal on A, so that { ν , χ } is an orthonormal system.
Figure 3b illustrates the second test that is analogous to the first, but with a uniform horizontal displacement applied to the red side from the y-direction and the opposing side fixed and periodical boundary conditions applied to the other surfaces. So, we have
U · χ = Δ U ,       ( x , y , z ) A f , U · χ = 0 ,                     ( x , y , z ) A k , ( σ ν ) · ν = 0 ,         ( x , y , z ) A , U d s t = U s r c ,         ( x , y , z ) A l A r , U d s t = U s r c ,         ( x , y , z ) A t A b ,
Figure 3c–f shows the third to sixth tests whose loaded direction of solid displacement and the relevant boundary conditions are given by
U · ν = Δ U ,       ( x , y , z ) A t , U · ν = 0 ,                     ( x , y , z ) A b , ( σ ν ) · χ = 0 ,         ( x , y , z ) A , U d s t = U s r c ,         ( x , y , z ) A f A k , U d s t = U s r c ,         ( x , y , z ) A r A l ,
for the third test,
U · χ = Δ U ,       ( x , y , z ) A f , U · χ = 0 ,                     ( x , y , z ) A k , ( σ ν ) · ν = 0 ,         ( x , y , z ) A , U d s t = U s r c ,         ( x , y , z ) A l A r , U d s t = U s r c ,         ( x , y , z ) A t A b ,
for the fourth test,
U · χ = Δ U ,       ( x , y , z ) A t , U · χ = 0 ,                     ( x , y , z ) A b , ( σ ν ) · ν = 0 ,         ( x , y , z ) A , U d s t = U s r c ,         ( x , y , z ) A f A k , U d s t = U s r c ,         ( x , y , z ) A r A l ,
for the fifth test, and
U · χ = Δ U ,             ( x , y , z ) A r , U · χ = 0 ,                     ( x , y , z ) A l , ( σ ν ) · ν = 0 ,         ( x , y , z ) A , U d s t = U s r c ,         ( x , y , z ) A f A k , U d s t = U s r c ,         ( x , y , z ) A t A b .
for the sixth test.

2.2.2. Calculation of 3D Equivalent Stiffness Matrix

We solve Equations (1) and (11) by applying the FEM to obtain the corresponding displacement field, which promotes the stress and strain components for every element to be obtained using Equations (3) and (4). Then, parameters allow us to estimate the corresponding average over the sample’s volume:
e i j r ¯ = 1 V Ω e i j r d V ,
And
σ i j r ¯ = 1 V Ω σ i j r d V ,
where V is the volume of Ω , with r = 1–6 denoting the rth simulation oscillatory relaxation test described in Figure 3, respectively.
As in the case of 2D, we use a second-order equivalent Voigt stiffness matrix MMij to relate the averaging stress and strain components of 3D models. In the case of general anisotropy and under plane strain conditions, this relation is given by
( σ x x r ¯ σ y y r ¯ σ z z r ¯ σ y z r ¯ σ z x r ¯ σ x y r ¯ ) ( M 11 M 12 M 13 M 14 M 15 M 16 M 12 M 22 M 23 M 24 M 25 M 26 M 13 M 23 M 33 M 34 M 35 M 36 M 14 M 24 M 34 M 44 M 45 M 46 M 15 M 25 M 35 M 45 M 55 M 56 M 16 M 26 M 36 M 46 M 56 M 66 ) ( e x x r ¯ e y y r ¯ e z z r ¯ 2 e y z r ¯ 2 e z x r ¯ 2 e x y r ¯ )
where Mij is the equivalent stiffness components in the 3D media, and σ i j r ¯ and e i j r ¯ are cells of the average stress and strain vector, respectively. In general, the matrix M takes into account the highest degree of anisotropy for a 3D case. Like in the 2D case, we can obtain 21 unknown stiffness tensors by applying the least-squares procedure to minimize the following cost function F2 (M),
F 2 ( M ) M i j = 0 .
We can solve the linear system of twenty-one equations yielded by Equation (28) to obtain the equivalent stiffness matrix Mij, seeing Appendix B for more niceties.

3. Results

3.1. Stiffness Tensors for 2D Fractured Model

As a 2D case, a fractured model with sides of 0.1 m contains small, isolated, and penny-shaped elliptical cracks along the x-axis. Every elliptical fracture has an x-semiaxis of 1.5 cm and y-semiaxis of 0.08 cm. The following formula is used to define the fracture density [13].
  ζ = N a 2 s
where N and a denote the fracture numbers and major semiaxis, respectively; s is the area of the model. Five group parameters are designed with different fracture densities of 0.056, 0.11, 0.149, 0.23, and 0.49 in the horizontal direction. Table 1 shows the physical parameters of fluid parameters, fractures with saturated brine, and the embedding background matrix.
We solve Equations (1) and (11) using the relaxation experiments with FEM. Once the stress and strain fields are obtained, the stiffness tensor is calculated by the least-squares method. Figure 4 shows the resulting equivalent stiffness tensors M22 and M12. With increasing fracture density, the equivalent stiffness tensors M22 and M12 have a similar decreasing trend.

3.2. Stiffness Tensors for 3D Fractured Model

In the 3D case, as shown in Figure 5, we build four models, each being a cube of 0.1 m on all sides for validating the 3D numerical oscillatory relaxation tests. These models include penny-shaped ellipsoidal cracks along the horizontal direction, and x-, y-, and z-semiaxes of 0.02 m, 0.02 m, and 0.003 m, respectively, with different fracture densities.
As with 2D, the four fracture densities are set to 0.04, 0.112, 0.192, and 0.416, respectively, according to the basic definition,
ζ = N a 3 v
where v is the volume of the cube. The same physical parameters in Table 1 are applied to the 3D models, and the relaxation tests are conducted at a frequency of 20 Hz.
Equations (1) and (11) are also solved by the relaxation experiments with FEM. Once the stress and strain fields are obtained, the stiffness tensor is calculated by the least-squares method. Figure 6 shows the resulting equivalent stiffness tensors M12 and M44. With increasing fracture density, the equivalent stiffness tensors M12 and M44 tend to decrease.

3.3. Validation of Numerical Method

3.3.1. Validation of 2D Numerical Method

To validate the finite-element numerical simulation, the classical second-order Hudson’s model and the Eshelby–Cheng model are used. Specific formulas are included in Appendix C.
Figure 7 compares the resulting equivalent stiffness tensors M22 and M12 calculated from the 2D numerical oscillatory relaxation tests for a frequency of 20 Hz with the Hudson model and the Eshelby–Cheng model. As expected, the solutions obtained using the different methods are almost comparable for low fracture densities but significantly different for high crack densities. We see that the second-order Hudson model for pair-wise interactions between cracks produces an unreasonable result at a high crack density that is over the dilution to make the assumption invalid. Both the equivalent stiffness matrix M11 and M12 by the Eshelby–Cheng model keep a good trend close to the numerical results, but their difference becomes large with increasing crack densities.

3.3.2. Numerical Analysis of 3D Fractured Model

As illustrated in Figure 8, the resulting effective stiffness coefficients M12 and M44 are compared to the theoretical solutions obtained using the Hudson and Eshelby–Cheng models. As expected, these predictions exhibit a similar trend as facture densities increase. As illustrated in Figure 8a,b, there is a substantially better agreement between the numerical M12 and M44 and their theoretical counterpart solutions, especially at low crack densities. We also see that the numerical results closely match the Eshelby–Cheng model, with some minor departures between them at higher crack densities due to limited accuracies in the Eshelby–Cheng model. The simulation method provides a physical prediction at high fracture densities that control the overall elastic properties of rocks. We can analyze the fracture-induced anisotropy using these equivalent stiffness tensors.

4. Discussion

4.1. Stress Distribution Induced on Crack Interaction

Stress interactions (stress amplification and dilution) produced by fractures have a substantial effect on the anisotropic characteristics of cracks. To better illustrate the stress interactions across cracks at different fracture densities, we did numerical tests with three 2D simple fracture models (Figure 9). As shown in Figure 9a, each model (0.01 m2) contains two, six, and ten elliptic inclusions (fractures). Allow each fracture to have a main semiaxis and an aspect ratio of 0.015, and 0.053, respectively. The fractures are completely saturated with water. Table 1 contains the simulation parameters.
We repeated the numerical test in Figure 2a with the corresponding boundaries, this time using the model in Figure 9a. As illustrated in Figure 9b, the stress distributions of the models vary. The fractures alter the stress distribution in their neighborhood. The stress dilution and concentration occurring near the face (central part) and tip of a fracture in Figure 9b. The stress distribution with a high fracture density contributes a broader range of effects to the background than the low fracture density. Additionally, when fracture density increases, the stress concentrations around the tips decrease. On the basis of the foregoing, it is possible to explain in Figure 7 and Figure 8 that the elastic matrix with a high fracture density is lower than the theoretical method.

4.2. Anisotropy Properties for Real Fractures

Combined with Figure 7 and Figure 8, we see that different fracture densities can induce elastic anisotropy, which can be quantitatively characterized using the equivalent stiffness tensors. We apply 2D anisotropic finite-element modeling to a real crack for anisotropy properties at this part. Figure 10a illustrates an outcrop from the Ordovician carbonate formation in the Tarim Basin, northwest China. Multistage tectonic movements result in a complicated fracture network in the formation [33]. From the Cambrian to the Middle Ordovician, strike-slip faults with numerous fractures were formed by the joint action of north–south and north–east stresses [34]. These secondary fractures are considered as both migration channels and fracture reservoir spaces for hydrocarbon. The outcrop is selected because its fracture characteristics are similar to those of subsurface strata over 7 km, and a complicated fracture network satisfies the simulation scale. Due to the complexity of natural fractures, modeling of irregular geometries and anisotropic features are challenging. The outcrop picture is simplified into a model consisting of background media and superimposed fractures with different shapes, sizes, and orientations to facilitate research on the anisotropy properties of fractures and their stress interaction, as illustrated in Figure 10b.
We investigate the cracked model’s density and dip-angle. For calculating fracture density, we use the method developed by reference [35,36], which employs a circular scanning window to calculate the number of total fractures by counting the number of endpoints of complete visible fractures within the window. An equal-area rose diagram instead of the histogram of fracture length is used since it provides an indirect orientation distribution (see reference [37] for more details). Figure 10c,d shows the geometric properties of the real model, which clearly exhibits a range of densities and orientations in different areas.
We use a numerical approach combined with finite-element modeling to study the anisotropic properties of the real fracture model at different frequencies. The same physical parameters in Table 1 are used, and oscillatory relaxation tests at frequencies of 0.01, 1, 10, 100, and 1000 Hz are applied.
The resultant real parts of the effective stiffness coefficients M11 are illustrated in Figure 11. As we see in Figure 11, the real parts of the stiffness matrix of M11 (red circle and red dotted line) show an increasing tendency with increasing frequency, which is not as regular as the simple models in reference [26]. The pictures indicated by the shallow green arrow represent the stress ( σ x x ) distributions at different frequencies. Additionally, we observe that they change slowly at the beginning, then rapidly, and tend to be flat at high frequency, as the difference between protolith and faults affecting the stiffening of the fluid becomes smaller [38]. The stress ( σ x x ) occurs an obvious difference between high frequency (100 and 1000 Hz) and low frequency, and the interaction becomes more intense as the frequency increases. The resultant anisotropies of effective elastic properties under the different frequencies can be used to characterize fracture interactions and dynamic stress disturbances. We can characterize the velocity anisotropy and seismic attenuation using these equivalent stiffness tensors.
The existence of fractures surely affects the overall elasticity, as shown in Figure 4 and Figure 6. In the geophysical community, conventional analytical methods, such as the Hudson model and Eshelby–Cheng model, have been widely used to study the effective elastic properties of rocks, but they are limited to their assumption that fracture geometries are idealized [19,22] and hardly characterize the elastic properties of complex and irregular natural fracture networks. We see that the numerical results keep a good trend with the analytical methods at low crack density, while the difference between them becomes larger with the increase of fracture density (Figure 7 and Figure 8) in that the numerical method explicitly simulates the influence of fractures distribution and their interaction (Figure 9). It proves that our method can accurately simulate the equivalent stiffness tensors with different frequencies and low fracture density.
When the fracture density is very high, Figure 7 and Figure 8 show that the decreasing trend may change because the solution of the simulation equation may be unstable. Some non-dilute crack theories should be used; for example, non-dilute crack theories established successfully based on the inclusion model DEM theory [39]. Additionally, we need to consider that fractures in nature can be too large to satisfy the simulation conditions. Therefore, further investigation on effective stiffness tensors of arbitrary complex multi-dimensional fracture should also consider these elements and laboratory measurements.

5. Conclusions

Numerical methods must be used to determine the effective elasticity of genuine fractures. We modify Rubino’s numeric FEM for real fractures. For 2D simple fracture models with different fracture densities, the 2D technique is compared to theoretical models (Hudson’s model and Eshelby–Cheng model). The results closely match the theoretical predictions at low fracture densities. The simulation method provides a physical prediction at high fracture densities that controls the overall elastic properties of rocks.
When applied to a sample of natural fractures extracted from an outcrop at different frequencies, the stress interactions between cracks become highly complex. The stiffness matrix of M11 shows a trend as the frequency increases. With increasing frequency, the stress concentration spreads quickly to cover the network’s high fracture density, intensifying the interactions and becoming intense. The effective elastic properties of the resulting real fractured model can be altered to understand fracture interactions and dynamic stress disturbances.
The 2D finite-element numerical method is extended to 3D fractured media using the 3D stress–strain relationship defined by the six oscillatory relaxation simulation tests. A least-squares approach can be used to solvethe relevant equations for unknown stiffness tensors. The 3D fractured media are compared to effective medium theories in terms of their elastic responses to different fracture densities. The fracture density is found to have an obvious effect on the anisotropic effective properties, and the proposed method gives a physical prediction for high fracture densities. This work has significance calculating the effective elastic properties of multi-dimensional fracture models and fracture interaction mechanisms.

Author Contributions

Conceptualization, S.J. and L.F.; methodology, S.J.; software, S.J.; formal analysis, S.J.; investigation, S.J. and Y.C.; resources, S.J. and L.F.; data curation, S.J.; writing—original draft preparation, S.J.; writing—review and editing, S.J.; visualization, S.J.; supervision, L.F.; funding acquisition, L.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDA14010303) and the National Natural Science Foundation of China (Grant Nos. 41821002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available from the corresponding and the first author.

Acknowledgments

Thanks to the teachers and students of the FLY Team at the China University of Petroleum (East China). Shikai Jian wants to particularly thank the patience, care, and support from Xinxin Cao. Thanks to the anonymous reviewers for their constructive comments, and thanks to the editors for their hard work.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Least-Squares Procedure for 2D Equivalent Stiffness Tensors

The least-squares procedure to minimize the following cost function F1(M) for computing the equivalent stiffness matrix Mij is described according to [26].
F 1 ( M ) = m = 1 3 ( σ x x m ¯ M 11 e x x m ¯ M 12 e y y m ¯ 2 M 16 e x y m ¯ 2 + σ y y m ¯ M 12 e x x m ¯ M 22 e y y m ¯ 2 M 26 e x y m ¯ 2   + σ x y m ¯ M 16 e x x m ¯ M 26 e y y m ¯ 2 M 66 e x y m ¯ 2 ) ,  
where m = 1–3 denotes the mth simulation test, and σ i j m ¯ and e i j m ¯ are the average stress and average strain, respectively. With the goal, the partial derivatives of the cost function related to the elements Mij are set to be zero, resulting in the following linear equations:
F ( M ) M 11 = 0 m = 1 3 σ x x m ¯ e x x m ¯ = M 11 m = 1 3 e x x m ¯ e x x m ¯ + M 12 m = 1 3 e y y m ¯ e x x m ¯ + 2 M 16 m = 1 3 e x y m ¯ e x x m ¯ ,
F ( M ) M 12 = 0 m = 1 3 ( σ x x m ¯ e y y m ¯ + σ y y m ¯ e x x m ¯ ) = M 11 m = 1 3 e x x m ¯ e y y m ¯ + M 12 m = 1 3 ( e x x m ¯ e x x m ¯ + e y y m ¯ e y y m ¯ ) + 2 M 16 m = 1 3 e x y m ¯ e x x m ¯ + 2 M 26 m = 1 3 e x y m ¯ e x x m ¯ + M 22 m = 1 3 e x x m ¯ e y y m ¯ ,
F ( M ) M 16 = 0 m = 1 3 ( 2 σ x x m ¯ e x y m ¯ + σ x y m ¯ e x x m ¯ ) = 2 M 11 m = 1 3 e x x m ¯ e y y m ¯ + 2 M 12 m = 1 3 ( e y y m ¯ e x y m ¯ ) + M 16 ( m = 1 3 4 e x y m ¯ e x y m ¯ + e x x m ¯ e x x m ¯ ) + M 26 m = 1 3 e y y s ¯ e x x s ¯ + 2 M 66 m = 1 3 e x x m ¯ e x y m ¯ ,
F ( M ) M 22 = 0 m = 1 3 σ y y m ¯ e y y m ¯ = M 12 m = 1 3 e x x m ¯ e y y m ¯ + M 22 m = 1 3 e y y m ¯ e y y m ¯ + 2 M 26 m = 1 3 e x y m ¯ e y y m ¯ ,
F ( M ) M 26 = 0 m = 1 3 ( 2 σ y y m ¯ e x y m ¯ + σ x y m ¯ e y y m ¯ ) = 2 M 12 m = 1 3 e x x m ¯ e x y m ¯ + 2 M 22 m = 1 3 ( e y y m ¯ e x y m ¯ ) + 2 M 26 m = 1 3 ( 4 e x y m ¯ e x y m ¯ + e y y m ¯ e y y m ¯ ) + M 16 m = 1 3 e y y m ¯ e x x m ¯ + 2 M 66 m = 1 3 e x y m ¯ e y y m ¯ ,
and
F ( M ) M 66 = 0 m = 1 3 σ x y m ¯ e x y m ¯ = M 16 m = 1 3 e x x m ¯ e x y m ¯ + M 26 m = 1 3 e y y m ¯ e x y m ¯ + 2 M 66 m = 1 3 e x y m ¯ e x y m ¯ .

Appendix B. Least-Squares Procedure for 3D Equivalent Stiffness Tensors

The least-squares procedure to minimize the following cost function F2(M) for computing the equivalent stiffness matrix Mij is described as follows.
F 2 ( M ) = r = 1 6 ( σ x x r ¯ M 11 e x x r ¯ M 12 e y y r ¯ M 13 e z z r ¯ 2 M 14 e y z r ¯ 2 M 15 e z x r ¯ 2 M 16 e x y r ¯ 2 + σ y y r ¯ M 12 e x x r ¯ M 22 e y y r ¯ M 23 e z z r ¯ 2 M 24 e y z r ¯ 2 M 25 e z x r ¯ 2 M 26 e x y r ¯ 2   + σ z z r ¯ M 13 e x x r ¯ M 23 e y y r ¯ M 33 e z z r ¯ 2 M 34 e y z r ¯ 2 M 35 e z x r ¯ 2 M 36 e x y r ¯ 2 + σ y z r ¯ M 14 e x x r ¯ M 24 e y y r ¯ M 34 e z z r ¯ 2 M 44 e y z r ¯ 2 M 45 e z x r ¯ 2 M 46 e x y r ¯ 2   + σ z x r ¯ M 15 e x x r ¯ M 25 e y y r ¯ M 35 e z z r ¯ 2 M 45 e y z r ¯ 2 M 55 e z x r ¯ 2 M 56 e x y r ¯ 2 + σ x y r ¯ M 16 e x x r ¯ M 26 e y y r ¯ M 36 e z z r ¯ 2 M 46 e y z r ¯ 2 M 56 e z x r ¯ 2 M 66 e x y r ¯ 2 ) ,
where r = 1–6 denotes the rth simulation test, and σ i j r ¯ and e i j r ¯ represents the averaging stress and strain, respectively.
With the goal, the partial derivatives of the cost function related to the elements Mij are set to be zero, resulting in the following linear equations:
F ( M ) M 11 = 0 r = 1 6 σ x x r ¯ e x x r ¯ = M 11 r = 1 6 e x x r ¯ e x x r ¯ + M 12 r = 1 6 e y y r ¯ e x x r ¯ + M 13 r = 1 6 e z z r ¯ e x x r ¯ + 2 M 14 r = 1 6 e y z r ¯ e x x r ¯ + 2 M 15 r = 1 6 e z x r ¯ e x x r ¯ + 2 M 16 r = 1 6 e x y r ¯ e x x r ¯ ,
F ( M ) M 12 = 0 r = 1 6 ( σ x x r ¯ e y y r ¯ + σ y y r ¯ e x x r ¯ ) = M 11 r = 1 6 e x x r ¯ e y y r ¯ + M 12 r = 1 6 ( e x x r ¯ e x x r ¯ + e y y r ¯ e y y r ¯ ) + M 13 r = 1 6 e z z r ¯ e y y r ¯ + 2 M 14 r = 1 6 e y z r ¯ e y y r ¯ + 2 M 15 r = 1 6 e z x r ¯ e y y r ¯ + 2 M 16 r = 1 6 e x y r ¯ e y y r ¯ + M 22 r = 1 6 e y y r ¯ e x x r ¯ + M 23 r = 1 6 e z z r ¯ e x x r ¯ + 2 M 24 r = 1 6 e y z r ¯ e x x r ¯ + 2 M 25 r = 1 6 e z x r ¯ e x x r ¯ + 2 M 26 r = 1 6 e x y r ¯ e x x r ¯ ,
F ( M ) M 13 = 0 r = 1 6 ( σ x x r ¯ e z z r ¯ + σ z z r ¯ e x x r ¯ ) = M 11 r = 1 6 e z z r ¯ e x x r ¯ + M 12 r = 1 6 e y y r ¯ e z z r ¯ + M 13 r = 1 6 ( e z z r ¯ e z z r ¯ + e x x r ¯ e x x r ¯ ) + 2 M 14 r = 1 6 e y z r ¯ e z z r ¯ + 2 M 15 r = 1 6 e z x r ¯ e z z r ¯ + 2 M 16 r = 1 6 e x y r ¯ e z z r ¯ + M 23 r = 1 6 e y y r ¯ e x x r ¯ + M 33 r = 1 6 e z z r ¯ e x x r ¯ + 2 M 34 r = 1 6 e y z r ¯ e x x r ¯ + 2 M 35 r = 1 6 e z x r ¯ e x x r ¯ + 2 M 36 r = 1 6 e x y r ¯ e x x r ¯ ,
F ( M ) M 14 = 0 r = 1 6 ( 2 σ x x r ¯ e y z r ¯ + σ y z r ¯ e x x r ¯ ) = 2 M 11 r = 1 6 e x x r ¯ e y z r ¯ + 2 M 12 r = 1 6 e y y r ¯ e y z r ¯ + 2 M 13 r = 1 6 e z z r ¯ e y z r ¯ + M 14 r = 1 6 ( 4 e y z r ¯ e y z r ¯ + e x x r ¯ e x x r ¯ ) + 4 M 15 r = 1 6 e z x r ¯ e y z r ¯ + 4 M 16 r = 1 6 e x y r ¯ e y z r ¯ + M 24 r = 1 6 e y y r ¯ e x x r ¯ + M 34 r = 1 6 e z z r ¯ e x x r ¯ + 2 M 44 r = 1 6 e y z r ¯ e x x r ¯ + 2 M 45 r = 1 6 e z x r ¯ e x x r ¯ + 2 M 46 r = 1 6 e x y r ¯ e x x r ¯ ,
F ( M ) M 15 = 0 r = 1 6 ( 2 σ x x r ¯ e z x r ¯ + σ z x r ¯ e x x r ¯ ) = 2 M 11 r = 1 6 e x x r ¯ e z x r ¯ + 2 M 12 r = 1 6 e y y r ¯ e z x r ¯ + 2 M 13 r = 1 6 e z z r ¯ e z x r ¯ + 4 M 14 r = 1 6 e y z r ¯ e z x r ¯ + M 15 r = 1 6 ( 4 e z x r ¯ e z x r ¯ + e x x r ¯ e x x r ¯ ) + 4 M 16 k = 1 6 e x y r ¯ e z x r ¯ + M 25 r = 1 6 e y y r ¯ e x x r ¯ + M 35 r = 1 6 e z z r ¯ e x x r ¯ + 2 M 45 r = 1 6 e y z r ¯ e x x r ¯ + 2 M 55 r = 1 6 e z x r ¯ e x x r ¯ + 2 M 56 r = 1 6 e x y r ¯ e x x r ¯ ,
F ( M ) M 16 = 0 r = 1 6 ( 2 σ x x r ¯ e x y r ¯ + σ x y r ¯ e x x r ¯ ) = 2 M 11 r = 1 6 e x x r ¯ e x y r ¯ + 2 M 12 r = 1 6 e y y r ¯ e x y r ¯ + 2 M 13 r = 1 6 e z z r ¯ e x y r ¯ + 4 M 14 r = 1 6 e y z r ¯ e x y r ¯ + 4 M 15 r = 1 6 e z x r ¯ e x y r ¯ + M 16 r = 1 6 ( 4 e x y r ¯ e x y r ¯ + e x x r ¯ e x x r ¯ ) + M 26 r = 1 6 e y y r ¯ e x x r ¯ + M 36 r = 1 6 e z z r ¯ e x x r ¯ + 2 M 46 r = 1 6 e y z r ¯ e x x r ¯ + 2 M 56 r = 1 6 e z x r ¯ e x x r ¯ + 2 M 66 r = 1 6 e x y r ¯ e x x r ¯ ,
F ( M ) M 22 = 0 r = 1 6 σ y y r ¯ e y y r ¯ = M 12 r = 1 6 e x x r ¯ e y y r ¯ + M 22 r = 1 6 e y y r ¯ e y y r ¯ + M 23 r = 1 6 e z z r ¯ e y y r ¯ + 2 M 24 r = 1 6 e y z r ¯ e y y r ¯ + 2 M 25 r = 1 6 e z x r ¯ e y y r ¯ + 2 M 26 r = 1 6 e x y r ¯ e y y r ¯ ,
F ( M ) M 23 = 0 r = 1 6 ( σ y y r ¯ e z z r ¯ + σ z z r ¯ e y y r ¯ ) = M 12 r = 1 6 e x x r ¯ e z z r ¯ + M 22 r = 1 6 e y y r ¯ e z z r ¯ + M 23 r = 1 6 ( e z z r ¯ e z z r ¯ + e y y r ¯ e y y r ¯ ) + 2 M 24 r = 1 6 e y z r ¯ e z z r ¯ + 2 M 25 r = 1 6 e z x r ¯ e z z r ¯ + 2 M 26 r = 1 6 e x y r ¯ e z z r ¯ + M 13 r = 1 6 e x x r ¯ e y y r ¯ + M 33 r = 1 6 e z z r ¯ e y y r ¯ + 2 M 34 r = 1 6 e y z r ¯ e y y r ¯ + 2 M 35 r = 1 6 e z x r ¯ e y y r ¯ + 2 M 36 r = 1 6 e x y r ¯ e y y r ¯ ,
F ( M ) M 24 = 0 r = 1 6 ( 2 σ y y r ¯ e y z r ¯ + σ y z r ¯ e y y r ¯ ) = 2 M 12 r = 1 6 e x x r ¯ e y z r ¯ + 2 M 22 r = 1 6 e y y r ¯ e y z r ¯ + 2 M 23 r = 1 6 e z z r ¯ e y z r ¯ + M 24 r = 1 6 ( 4 e y z r ¯ e y z r ¯ + e y y r ¯ e y y r ¯ ) + 4 M 25 r = 1 6 e z x r ¯ e y z r ¯ + 4 M 26 r = 1 6 e x y r ¯ e y z r ¯ + M 14 r = 1 6 e x x r ¯ e y y r ¯ + M 34 r = 1 6 e z z r ¯ e y y r ¯ + 2 M 44 r = 1 6 e y z r ¯ e y y r ¯ + 2 M 45 r = 1 6 e z x r ¯ e y y r ¯ + 2 M 46 r = 1 6 e x y r ¯ e y y r ¯ ,
F ( M ) M 25 = 0 r = 1 6 ( 2 σ y y r ¯ e z x r ¯ + σ z x r ¯ e y y r ¯ ) = 2 M 12 r = 1 6 e x x r ¯ e z x r ¯ + 2 M 22 r = 1 6 e y y r ¯ e z x r ¯ + 2 M 23 r = 1 6 e z z r ¯ e z x r ¯ + 4 M 24 r = 1 6 e y z r ¯ e z x r ¯ + M 25 r = 1 6 ( 4 e z x r ¯ e z x r ¯ + e y y r ¯ e y y r ¯ ) + 4 M 26 r = 1 6 e x y r ¯ e z x r ¯ + M 15 r = 1 6 e x x r ¯ e y y r ¯ + M 35 r = 1 6 e z z r ¯ e y y r ¯ + 2 M 45 r = 1 6 e y z r ¯ e y y r ¯ + 2 M 55 r = 1 6 e z x r ¯ e y y r ¯ + 2 M 56 r = 1 6 e x y r ¯ e y y r ¯ ,
F ( M ) M 26 = 0 r = 1 6 ( 2 σ y y r ¯ e x y r ¯ + σ x y r ¯ e y y r ¯ ) = 2 M 12 r = 1 6 e x x r ¯ e x y r ¯ + 2 M 22 r = 1 6 e y y r ¯ e x y r ¯ + 2 M 23 r = 1 6 e z z r ¯ e x y r ¯ + 4 M 24 r = 1 6 e y z r ¯ e x y r ¯ + 4 M 25 r = 1 6 e z x r ¯ e x y r ¯ + M 26 r = 1 6 ( 4 e x y r ¯ e x y r ¯ + e y y r ¯ e y y r ¯ ) + M 16 r = 1 6 e x x r ¯ e y y r ¯ + M 36 r = 1 6 e z z r ¯ e y y r ¯ + 2 M 46 r = 1 6 e y z r ¯ e y y r ¯ + 2 M 56 r = 1 6 e z x r ¯ e y y r ¯ + 2 M 66 r = 1 6 e x y r ¯ e y y r ¯ ,
F ( M ) M 33 = 0 r = 1 6 σ z z r ¯ e z z r ¯ = M 13 r = 1 6 e x x r ¯ e z z r ¯ + M 23 r = 1 6 e y y r ¯ e z z r ¯ + M 33 r = 1 6 e z z r ¯ e z z r ¯ + 2 M 34 r = 1 6 e y z r ¯ e z z r ¯ + 2 M 35 r = 1 6 e z x r ¯ e z z r ¯ + 2 M 36 r = 1 6 e x y r ¯ e z z r ¯ ,
F ( M ) M 34 = 0 r = 1 6 ( 2 σ z z r ¯ e y z r ¯ + σ y z r ¯ e z z r ¯ ) = 2 M 13 r = 1 6 e x x r ¯ e y z r ¯ + 2 M 23 r = 1 6 e y y r ¯ e y z r ¯ + 2 M 33 r = 1 6 e z z r ¯ e y z r ¯ + M 34 r = 1 6 ( 4 e y z r ¯ e y z r ¯ + e z z r ¯ e z z r ¯ ) + 4 M 35 r = 1 6 e z x r ¯ e y z r ¯ + 4 M 36 r = 1 6 e x y r ¯ e y z r ¯ + M 14 r = 1 6 e x x r ¯ e z z r ¯ + M 24 r = 1 6 e y y r ¯ e z z r ¯ + 2 M 44 r = 1 6 e y z r ¯ e z z r ¯ + 2 M 45 r = 1 6 e z x r ¯ e z z r ¯ + 2 M 46 r = 1 6 e x y r ¯ e z z r ¯ ,
F ( M ) M 35 = 0 r = 1 6 ( 2 σ z z r ¯ e z x r ¯ + σ z x r ¯ e z z r ¯ ) = 2 M 13 r = 1 6 e x x r ¯ e z x r ¯ + 2 M 23 r = 1 6 e y y r ¯ e z x r ¯ + 2 M 33 r = 1 6 e z z r ¯ e z x r ¯ + 4 M 34 r = 1 6 e y z r ¯ e z x r ¯ + M 35 r = 1 6 ( 4 e z x r ¯ e z x r ¯ + e z z r ¯ e z z r ¯ ) + 4 M 36 r = 1 6 e x y r ¯ e z x r ¯ + M 15 r = 1 6 e x x r ¯ e z z r ¯ + M 25 r = 1 6 e y y r ¯ e z z r ¯ + 2 M 45 r = 1 6 e y z r ¯ e z z r ¯ + 2 M 55 r = 1 6 e z x r ¯ e z z r ¯ + 2 M 56 r = 1 6 e x y r ¯ e z z r ¯ ,
F ( M ) M 36 = 0 r = 1 6 ( 2 σ z z r ¯ e x y r ¯ + σ x y r ¯ e z z r ¯ ) = 2 M 13 r = 1 6 e x x r ¯ e x y r ¯ + 2 M 23 r = 1 6 e y y r ¯ e x y r ¯ + 2 M 33 r = 1 6 e z z r ¯ e x y r ¯ + 4 M 34 r = 1 6 e y z r ¯ e x y r ¯ + 4 M 35 r = 1 6 e z x r ¯ e x y r ¯ + M 36 r = 1 6 ( 4 e x y r ¯ e x y r ¯ + e z z r ¯ e z z r ¯ ) + M 16 r = 1 6 e x x r ¯ e z z r ¯ + M 26 r = 1 6 e y y r ¯ e z z r ¯ + 2 M 46 r = 1 6 e y z r ¯ e z z r ¯ + 2 M 56 r = 1 6 e z x r ¯ e z z r ¯ + 2 M 66 r = 1 6 e x y r ¯ e z z r ¯ ,
F ( M ) M 44 = 0 r = 1 6 σ y z r ¯ e y z r ¯ = M 14 r = 1 6 e x x r ¯ e y z r ¯ + M 24 r = 1 6 e y y r ¯ e y z r ¯ + M 34 r = 1 6 e z z r ¯ e y z r ¯ + 2 M 44 r = 1 6 e y z r ¯ e y z r ¯ + 2 M 45 r = 1 6 e z x r ¯ e y z r ¯ + 2 M 46 r = 1 6 e x y r ¯ e y z r ¯ ,
F ( M ) M 45 = 0 r = 1 6 ( σ y z r ¯ e z x r ¯ + σ z x r ¯ e y z r ¯ ) = M 14 r = 1 6 e x x r ¯ e z x r ¯ + M 24 r = 1 6 e y y r ¯ e z x r ¯ + M 34 r = 1 6 e z z r ¯ e z x r ¯ + 2 M 44 r = 1 6 e y z r ¯ e z x r ¯ + 2 M 45 r = 1 6 ( e z x r ¯ e z x r ¯ + e y z r ¯ e y z r ¯ ) + 2 M 46 r = 1 6 e x y r ¯ e z x r ¯ + M 15 r = 1 6 e x x r ¯ e y z r ¯ + M 25 r = 1 6 e y y r ¯ e y z r ¯ + M 35 r = 1 6 e z z r ¯ e y z r ¯ + 2 M 55 r = 1 6 e z x r ¯ e y z r ¯ + 2 M 56 r = 1 6 e x y r ¯ e y z r ¯ ,
F ( M ) M 46 = 0 r = 1 6 ( σ y z r ¯ e x y r ¯ + σ x y r ¯ e y z r ¯ ) = M 14 r = 1 6 e x x r ¯ e x y r ¯ + M 24 r = 1 6 e y y r ¯ e x y r ¯ + M 34 r = 1 6 e z z r ¯ e x y r ¯ + 2 M 44 r = 1 6 e y z r ¯ e x y r ¯ + 2 M 45 r = 1 6 e z x r ¯ e x y r ¯ + 2 M 46 r = 1 6 ( e x y r ¯ e x y r ¯ + e y z r ¯ e y z r ¯ ) + M 16 r = 1 6 e x x r ¯ e y z r ¯ + M 26 r = 1 6 e y y r ¯ e y z r ¯ + M 36 r = 1 6 e z z r ¯ e y z r ¯ + 2 M 56 r = 1 6 e z x r ¯ e y z r ¯ + 2 M 66 r = 1 6 e x y r ¯ e y z r ¯ ,
F ( M ) M 55 = 0 r = 1 6 σ z x r ¯ e z x r ¯ = M 15 r = 1 6 σ x x r ¯ e z x r ¯ + M 25 r = 1 6 σ y y r ¯ e z x r ¯ + M 35 r = 1 6 σ z z r ¯ e z x r ¯ + 2 M 45 r = 1 6 σ y z r ¯ e z x r ¯ + 2 M 55 r = 1 6 σ z x r ¯ e z x r ¯ + 2 M 56 r = 1 6 σ x y r ¯ e z x r ¯ ,
F ( M ) M 56 = 0 r = 1 6 ( σ z x r ¯ e x y r ¯ + σ x y r ¯ e z x r ¯ ) = M 15 r = 1 6 e x x r ¯ e x y r ¯ + M 25 r = 1 6 e y y r ¯ e x y r ¯ + M 35 r = 1 6 e z z r ¯ e x y r ¯ + 2 M 45 r = 1 6 e y z r ¯ e x y r ¯ + 2 M 55 r = 1 6 e z x r ¯ e x y r ¯ + 2 M 56 r = 1 6 ( e x y r ¯ e x y r ¯ + e z x r ¯ e z x r ¯ ) + M 16 r = 1 6 e x x r ¯ e z x r ¯ + M 26 r = 1 6 e y y r ¯ e z x r ¯ + M 36 r = 1 6 e z z r ¯ e z x r ¯ + 2 M 46 r = 1 6 e y z r ¯ e z x r ¯ + 2 M 66 r = 1 6 e x y r ¯ e z x r ¯ ,
F ( M ) M 66 = 0 r = 1 6 σ x y r ¯ e x y r ¯ = M 16 r = 1 6 e x x r ¯ e x y r ¯ + M 26 r = 1 6 e y y r ¯ e x y r ¯ + M 36 r = 1 6 e z z r ¯ e x y r ¯ + 2 M 46 r = 1 6 e y z r ¯ e x y r ¯ + 2 M 56 r = 1 6 e z x r ¯ e x y r ¯ + 2 M 66 r = 1 6 e x y r ¯ e x y r ¯ .

Appendix C. The Hudson’s Model and Eshelby–Cheng Model

The second-order Hudson [6,7] equivalent stiffness matrix is given as
M i j e f f = M i j 0 + M i j 1 + M i j 2
where M i j 0 is the elastic parameters of isotropy background, and M i j e f f is the total stiffness tensor. Additionally, M i j 1 is the first-order correction of effects of the isolate crack, and M i j 2 are the second-order correction of fractures interaction. If the z-axis is perpendicular to the normal of cracks, such as the VTI (Vertical Transverse Isotropy) media, the Hudson model’s first-order correction M i j 1 and second-order correction M i j 2 are given as follows:
{ M 11 1 = λ μ ξ H 3 , M 13 1 = λ ( λ + 2 μ ) μ ξ H 3 , M 33 1 = λ ( λ + 2 μ ) 2 μ ξ H 3 , M 44 1 = μ ξ H 1 M 66 1 = 0 ,
and
{ M 11 2 = q 15 λ 2 λ + 2 μ ( ξ H 3 ) 2 , M 13 2 = q 15 λ ( ξ H 3 ) 2 , M 33 2 = q 15 ( λ + 2 μ ) ( ξ H 3 ) 2 , M 44 2 = 2 15 μ ( 3 λ + 8 μ ) λ 2 λ + 2 μ ( ξ H 3 ) 2 M 66 2 = 0 ,
where q = q 15 λ 2 μ 2 + 28 λ μ + 28 . λ and μ represents the lame coefficients, and ξ denotes the fracture density. In terms of dry rock, we have
  H 3 = 4 ( λ + 2 μ ) 3 ( λ + μ ) 1 1 + B
and
H 1 = 16 ( λ + 2 μ ) 3 ( 3 λ + 4 μ ) 1 1 + k
where B = λ + 2 μ 3 ( λ + μ ) 4 μ π α μ , k = λ + 2 μ [ K + ( 4 / 3 ) μ ] π α μ ( λ + μ ) 4 μ π α μ , and K and μ denote the bulk and shear moduli of inclusions, respectively, and the aspect ratio of inclusions is α .
The first-order correction is calculated through exchanging M 11 1 and M 33 1 , M 44 1 and M 66 1 , as well as the second-order correction, if the x-axis is perpendicular to the normal of cracks, such as the HTI (Horizontal Transverse Isotropy) media.
For high fracture densities but small aspect ratios, the Eshelby–Cheng model can avoid some of the problems [4] given by
M i j e f f = M i j 0 1 a i j ξ 1 b i j ξ
where a i j and b i j represent the Padé coefficients, which are related to first- and second-order Hudson’s models.
b i j = M i j 2 M i j 1 ξ ,   a i j = M i j 1 M i j 0 ξ b i j

References

  1. Schultz, R. Introduction to Geologic Structural Discontinuities. In Geologic Fracture Mechanics; Cambridge University Press: Cambridge, UK, 2019; Chapter 1; pp. 1–26. [Google Scholar]
  2. Brandes, C.; Tanner, D. Fault mechanics and earthquakes. In Understanding Faults: Detecting, Dating, and Modeling; Elsevier: Amsterdam, The Netherlands, 2020; Chapter 2; pp. 11–80. [Google Scholar] [CrossRef]
  3. Grechka, V.; Kachanov, M. Effective elasticity of fractured rocks: A snapshot of the work in progress. Geophysics 2006, 71, W45–W58. [Google Scholar] [CrossRef]
  4. Cheng, C.H. Crack models for a transversely isotropic medium. J. Geophys. Res. 1993, 98, 675–684. [Google Scholar] [CrossRef]
  5. Hudson, J.A. Overall properties of a cracked solid. Math. Proc. Camb. Phil. Soc. 1980, 88, 371–384. [Google Scholar] [CrossRef]
  6. Hudson, J.A. A higher order approximation to the wave propagation constants for a cracked solid. Geophys. J. R. Astr. Soc. 1986, 87, 265–274. [Google Scholar] [CrossRef] [Green Version]
  7. Hudson, J.A. Overall elastic properties of isotropic materials with arbitrary distribution of circular cracks. Geophys. J. Intern. 1990, 102, 465–469. [Google Scholar] [CrossRef] [Green Version]
  8. Eshelby, J.D. The determination of the elastic field of an ellipsoidal inclusion and related problems. Proc. R. Soc. A Math. Phys. Sci. 1957, 241, 376–396. [Google Scholar] [CrossRef]
  9. Hudson, J.A. Overall properties of a material with inclusions or cavities. Geophys. J. Intern. 1994, 117, 555–561. [Google Scholar] [CrossRef] [Green Version]
  10. Chapman, M. Frequency-dependent anisotropy due to mesoscale fractures in the presence of equant porosity. Geophys. Prospect. 2003, 51, 369–379. [Google Scholar] [CrossRef] [Green Version]
  11. Guo, J.X.; Rubino, J.G.; Glubokovskikh, S.; Gurevich, B. Effects of fracture intersections on seismic dispersion: Theoretical predictions versus numerical simulations. Geophys. Prospect. 2016, 65, 1264–1276. [Google Scholar] [CrossRef]
  12. Berryman, J.G. Single-scattering approximations for coefficients in Biot’s equations of poroelasticity. J. Acoust. Soc. Am. 1992, 91, 551–571. [Google Scholar] [CrossRef]
  13. Mavko, G.; Mukerji, T.; Dvorkin, J. The Rock Physics Handbook, 2nd ed.; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  14. Ghosh, R.; Sen, M.K. Predicting subsurface CO2 movement: From laboratory to field scale. Geophysics 2012, 77, M27–M37. [Google Scholar] [CrossRef]
  15. Jakobsen, M.; Hudson, J.A.; Johansen, T.A. T-matrix approach to shale acoustics. Geophys. J. Intern. 2003, 154, 533–558. [Google Scholar] [CrossRef] [Green Version]
  16. Jing, L.; Hudson, J.A. Numerical methods in rock mechanics. Intern. J. Rock Mech. Min. Sci. 2002, 39, 409–427. [Google Scholar] [CrossRef]
  17. Vlastos, S.; Liu, E.; Main, I.G.; Li, X.Y. Numerical simulation of wave propagation in media with discrete distributions of fractures: Effects of fracture sizes and spatial distributions. Geophys. J. Intern. 2003, 152, 649–668. [Google Scholar] [CrossRef] [Green Version]
  18. Grechka, V.; Kachanov, M. Effective elasticity of rocks with closely spaced and intersecting cracks. Geophysics 2006, 71, D85–D91. [Google Scholar] [CrossRef]
  19. Quintal, B.; Steeb, H.; Frehner, M.; Schmalholz, S.M. Quasi-static finite element modeling of seismic attenuation and dispersion due to wave-induced fluid flow in poroelastic media. J. Geophys. Res. 2011, 116, B01201. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, L.X.; Yao, Q.L.; Han, D.H.; Yan, F.Y.; Nasser, M. Characterizing the effect of elastic interactions on the effective elastic properties of porous, cracked rocks. Geophys. Prospect. 2015, 64, 157–169. [Google Scholar] [CrossRef]
  21. Zhang, W.H.; Fu, L.Y.; Zhang, Y.; Jin, W.J. Computation of elastic properties of 3D digital cores from the Longmaxi shale. Appl. Geophys. 2016, 13, 364–374. [Google Scholar] [CrossRef]
  22. Lissa, S.; Barbosa, N.D.; Rubino, J.; Quintal, B. Seismic attenuation and dispersion in poroelastic media with fractures of variable aperture distributions. Solid Earth 2019, 10, 1321–1336. [Google Scholar] [CrossRef] [Green Version]
  23. Song, Y.J.; Hu, H.S.; Han, B. Effective properties of a porous medium with aligned cracks containing compressible fluid. Geophys. J. Intern. 2020, 221, 60–76. [Google Scholar] [CrossRef]
  24. Masson, Y.J.; Pride, S.R. On the correlation between material structure and seismic attenuation anisotropy in porous media. J. Geophys. Res. Solid Earth 2014, 119, 60–76. [Google Scholar] [CrossRef] [Green Version]
  25. Rubino, J.G.; Ravazzoli, C.L.; Santos, J.E. Equivalent viscoelastic solids for heterogeneous fluid-saturated porous rocks. Geophysics 2009, 74, N1–N13. [Google Scholar] [CrossRef]
  26. Rubino, J.G.; Caspari, E.; Müller, T.M.; Milani, M.; Barbosa, N.D.; Holliger, K. Numerical upscaling in 2-D heterogeneous poroelastic rocks: Anisotropic attenuation and dispersion of seismic waves. J. Geophys. Res. Solid Earth 2016, 121, 6698–6721. [Google Scholar] [CrossRef] [Green Version]
  27. Guo, J.X.; Rubino, J.G.; Barbosa, N.D.; Glubokovskikh, S.; Gurevich, B. Seismic dispersion and attenuation in saturated porous rocks with aligned fractures of finite thickness: Theory and numerical simulations—Part I: P-wave perpendicular to the fracture plane. Geophysics 2018, 83, WA49–WA62. [Google Scholar] [CrossRef]
  28. Favino, M.; Hunziker, J.; Caspari, E.; Quintal, B.; Krause, R. Fully-automated adaptive mesh refinement for media embedding complex heterogeneities: Application to poroelastic fluid pressure diffusion. Computa. Geosci. 2020, 24, 1101–1120. [Google Scholar] [CrossRef]
  29. Cao, C.H.; Fu, L.Y.; Ba, J.; Zhang, Y. Frequency- and incident-angle-dependent P-wave properties influenced by dynamic stress interactions in fractured porous media. Geophysics 2019, 84, MR173–MR184. [Google Scholar] [CrossRef]
  30. Cao, C.H.; Chen, F.Y.; Fu, L.Y.; Ba, J.; Han, T.C. Effect of stress interactions on anisotropic P-SV-wave dispersion and attenuation for closely spaced cracks in saturated porous media. Geophys. Prospect. 2020, 68, 2536–2556. [Google Scholar] [CrossRef]
  31. Jian, S.K.; Fu, L.Y.; Cao, C.H.; Han, T.C.; Du, Q.Z. 3D finite-element modeling of effective elastic properties for fracture density and multiscale natural fractures. J. Geophys. Eng. 2021, 18, 567–582. [Google Scholar] [CrossRef]
  32. Biot, M.A. Mechanics of deformation and acoustic propagation in porous media. J. Appl. Phys. 1962, 33, 1482–1498. [Google Scholar] [CrossRef]
  33. Deng, S.; Li, H.; Zhang, Z.P.; Zhang, J.B.; Yang, X. Structural characterization of intracratonic strike-slip faults in the central Tarim Basin. AAPG Bull. 2019, 103, 109–137. [Google Scholar] [CrossRef]
  34. Qi, L. Structural characteristics and storage control function of the shun I fault zone in the Shunbei region, Tarim Basin. J. Petrol. Sci. Eng. 2021, 203, 108653. [Google Scholar] [CrossRef]
  35. Healy, D.; Rizzo, R.E.; Cornwell, D.G.; Farrell, N.; Smith, M. FracPaQ: A MATLAB toolbox for the quantification of fracture patterns. J. Struct. Geol. 2017, 95, 1–16. [Google Scholar] [CrossRef] [Green Version]
  36. Mauldon, M.; Dunne, W.M.; Rohrbaugh, M.B. Circular scanlines and circular windows: New tools for characterizing the geometry of fracture traces. J. Struct. Geol. 2001, 23, 247–258. [Google Scholar] [CrossRef]
  37. Nemec, W. The shape of the rose. Sedimen. Geol. 1988, 59, 149–152. [Google Scholar] [CrossRef]
  38. Milani, M.; Rubino, J.G.; Müller, T.M.; Quintal, B.; Caspari, E.; Holliger, K. Representative elementary volumes for evaluating effective seismic properties of heterogeneous poroelastic media. Geophysics 2016, 81, D169–D181. [Google Scholar] [CrossRef]
  39. Ghosh, R. Effective medium modeling of gas hydrate-filled fractures using the sonic log in the Krishna-Godavari basin, offshore eastern India. J. Geophys. Res. Solid Earth 2010, 115, B06101. [Google Scholar] [CrossRef]
Figure 1. Flowchart of the modeling methodology.
Figure 1. Flowchart of the modeling methodology.
Applsci 12 01873 g001
Figure 2. Sketch map of oscillatory relaxation simulation tests to determine 2D stiffness tensors. The back arrows denote the different orientations of solid displacements. Let the solid displacement vertical to the boundary on the black edge lines be zero. (a) the first oscillatory relaxation test, a uniform vertical displacement is imposed on the bottom and top boundaries, with two lateral boundaries being confined; (b) the second oscillatory relaxation test with a uniform horizontal displacement to the right and left boundaries and the two other boundaries confined; (c) the third test, where the top and bottom boundaries are loaded with displacement.
Figure 2. Sketch map of oscillatory relaxation simulation tests to determine 2D stiffness tensors. The back arrows denote the different orientations of solid displacements. Let the solid displacement vertical to the boundary on the black edge lines be zero. (a) the first oscillatory relaxation test, a uniform vertical displacement is imposed on the bottom and top boundaries, with two lateral boundaries being confined; (b) the second oscillatory relaxation test with a uniform horizontal displacement to the right and left boundaries and the two other boundaries confined; (c) the third test, where the top and bottom boundaries are loaded with displacement.
Applsci 12 01873 g002
Figure 3. Schematic diagram of six oscillatory relaxation simulation tests to determine the equivalent stiffness matrix for a 3D model, including penny-shaped spheroid fractures. Giving every test, let a uniform solid displacement (blue arrow) be applied to the red side, the gray side be confined, and the periodical boundary conditions be imposed on the other sides. (a) the first oscillatory relaxation test; (b) the second oscillatory relaxation test; (c) the third oscillatory relaxation test; (d) the fourth oscillatory relaxation test; (e) the fifth oscillatory relaxation test; (f) the sixth oscillatory relaxation test.
Figure 3. Schematic diagram of six oscillatory relaxation simulation tests to determine the equivalent stiffness matrix for a 3D model, including penny-shaped spheroid fractures. Giving every test, let a uniform solid displacement (blue arrow) be applied to the red side, the gray side be confined, and the periodical boundary conditions be imposed on the other sides. (a) the first oscillatory relaxation test; (b) the second oscillatory relaxation test; (c) the third oscillatory relaxation test; (d) the fourth oscillatory relaxation test; (e) the fifth oscillatory relaxation test; (f) the sixth oscillatory relaxation test.
Applsci 12 01873 g003
Figure 4. Effective stiffness tensors M22 (a) and M12 (b) for 2D fracture model.
Figure 4. Effective stiffness tensors M22 (a) and M12 (b) for 2D fracture model.
Applsci 12 01873 g004
Figure 5. Four 3D models, each being a cube with 0.1 m on its sides, contain penny-shaped spheroid cracks with different fracture densities: 0.04 (a), 0.112 (b), 0.192 (c), and 0.416 (d).
Figure 5. Four 3D models, each being a cube with 0.1 m on its sides, contain penny-shaped spheroid cracks with different fracture densities: 0.04 (a), 0.112 (b), 0.192 (c), and 0.416 (d).
Applsci 12 01873 g005
Figure 6. Effective stiffness tensors M12 (a) and M44 (b) for 3D fracture model.
Figure 6. Effective stiffness tensors M12 (a) and M44 (b) for 3D fracture model.
Applsci 12 01873 g006
Figure 7. Comparison of numerical simulations and theoretical predictions for effective stiffness constants M22 (a) and M12 (b) for an isotropic background medium containing fractures saturated with brine.
Figure 7. Comparison of numerical simulations and theoretical predictions for effective stiffness constants M22 (a) and M12 (b) for an isotropic background medium containing fractures saturated with brine.
Applsci 12 01873 g007
Figure 8. The effective stiffness coefficients M12 (a) and M44 (b) of the 3D models (Figure 5), including water-saturated penny-shaped cracks, set different fracture densities.
Figure 8. The effective stiffness coefficients M12 (a) and M44 (b) of the 3D models (Figure 5), including water-saturated penny-shaped cracks, set different fracture densities.
Applsci 12 01873 g008
Figure 9. Stress distributions of 2D fractured models: (a) three 2D cracked models and (b) the stress ( σ x x ) distributions of the models. The stress dilution and concentration occur near the faces and tips (central parts) of fractures, respectively.
Figure 9. Stress distributions of 2D fractured models: (a) three 2D cracked models and (b) the stress ( σ x x ) distributions of the models. The stress dilution and concentration occur near the faces and tips (central parts) of fractures, respectively.
Applsci 12 01873 g009
Figure 10. 2D real-fracture model and its geometric characteristics: (a) rock outcrop of Ordovician carbonates in Tarim Basin, northwest China; (b) the model can be composed of the background media and extracted fractures; (c) the model’s fracture density distributions where the red and yellow areas in the maps indicate more fractures than others; (d) the equal-area rose diagrams of fracture length, which provide dip distribution.
Figure 10. 2D real-fracture model and its geometric characteristics: (a) rock outcrop of Ordovician carbonates in Tarim Basin, northwest China; (b) the model can be composed of the background media and extracted fractures; (c) the model’s fracture density distributions where the red and yellow areas in the maps indicate more fractures than others; (d) the equal-area rose diagrams of fracture length, which provide dip distribution.
Applsci 12 01873 g010
Figure 11. Real parts of the effective stiffness coefficients M11 (red circle) of the model with different frequencies indirectly describing the anisotropy characteristics of the model. The shallow green arrows are the stress ( σ x x ) distributions under different frequencies. The stress ( σ x x ) between high frequency (100 and 1000 Hz) and low frequency (0.01, 1 and 10 Hz) shows an obvious difference (black dotted rectangle). The magnitude of stress is denoted by the color bar.
Figure 11. Real parts of the effective stiffness coefficients M11 (red circle) of the model with different frequencies indirectly describing the anisotropy characteristics of the model. The shallow green arrows are the stress ( σ x x ) distributions under different frequencies. The stress ( σ x x ) between high frequency (100 and 1000 Hz) and low frequency (0.01, 1 and 10 Hz) shows an obvious difference (black dotted rectangle). The magnitude of stress is denoted by the color bar.
Applsci 12 01873 g011
Table 1. Simulation parameters for multi-dimensional cracked models.
Table 1. Simulation parameters for multi-dimensional cracked models.
ParametersPhase
Matrix
Grain Bulk moduli, ks (GPa)37.0
Grain   Density ,   ρ s (g/cm3)2.65
Porosity ,   ϕ 0.1
Bulk modulus, km (GPa)26
Shear   modulus ,   μ or μ m (GPa)31.0
Permeability , κ (m2)1 × 10−15
Fractures
Grain Bulk moduli, ks (GPa)37.0
Grain   Density ,   ρ s (g/cm3)2.65
Porosity ,  ϕ 0.8
Bulk modulus, km (GPa)0.04
Shear   modulus ,   μ or μ m (GPa)0.02
Permeability ,  κ (m2)1 × 10−10
Brine
Bulk modulus, kf (GPa)2.25
Density ,   ρ s (g/cm3)1.09
Viscosity ,   η   (Pa.s)0.01
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jian, S.; Fu, L.; Cheng, Y. Anisotropic Effective Elastic Properties for Multi-Dimensional Fractured Models. Appl. Sci. 2022, 12, 1873. https://doi.org/10.3390/app12041873

AMA Style

Jian S, Fu L, Cheng Y. Anisotropic Effective Elastic Properties for Multi-Dimensional Fractured Models. Applied Sciences. 2022; 12(4):1873. https://doi.org/10.3390/app12041873

Chicago/Turabian Style

Jian, Shikai, Liyun Fu, and Yifan Cheng. 2022. "Anisotropic Effective Elastic Properties for Multi-Dimensional Fractured Models" Applied Sciences 12, no. 4: 1873. https://doi.org/10.3390/app12041873

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