SciELO - Scientific Electronic Library Online

 
vol.68 issue3Amplitude and phase measument using reflection polarization mode of a prism-based surface plasmon resonanceCharacterizing errors for quantum Fourier transform on IBM quantum author indexsubject indexsearch form
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • Have no similar articlesSimilars in SciELO

Share


Revista mexicana de física

Print version ISSN 0035-001X

Rev. mex. fis. vol.68 n.3 México May./Jun. 2022  Epub Apr 14, 2022

https://doi.org/10.31349/revmexfis.68.031401 

Research

Other areas in Physics

Effect of the orientation distribution of thin highly conductive inhomogeneities on the overall electrical conductivity of heterogeneous material

V. Levina 

M. Markova  * 

G. Ronquillo Jarilloa 

a Instituto Mexicano del Petroleo, Eje Central Lazaro Cárdenas, 152, col. San Bartolo Atepehuacán, 07730, Ciudad de México, México.


Abstract

Many natural composite materials contain systems of partially oriented thin low-resistivity inclusions (for example, water-saturated microcracks in a double porosity sedimentary formation). We have calculated the components of the electrical conductivity tensor of such materials as a function of crack density. The results were obtained for thin ellipsoidal inclusions with conductivity (electrical or thermal) much larger than the matrix conductivity. To calculate the effective conductivity, we have used the effective field method (EFM). We have obtained the explicit expressions for the effective parameters of inhomogeneous materials. The application of the EFM allows one to describe the influence of the peculiarities in the spatial distribution of inclusions on the effective properties of the medium. General explicit expressions, obtained in this work, are illustrated by calculating examples for inclusions, homogeneously distributed in the sector [-β,β], where β is the disorientation angle, and some continuous angle distribution functions. The calculations have shown that the spatial distribution of the crack-like inclusions strongly affects the conductive properties of the effective medium and the symmetry of their tensor.

Keywords: Heterogeneous medium; highly conductive inhomogeneities; homogenization problem; effective field method; influence of the inclusion orientations

1. Introduction

Many microinhomogeneous materials, for instance, sedimentary rocks, contain a system of oriented or partially oriented inhomogeneities (inclusions). Thus, carbonate hydrocarbon reservoirs may be represented as a system of fluid-filled microcracks that are either randomly oriented or oriented along certain directions [1-18]. The presence of such systems of cracks leads to a considerable anisotropy of the physical characteristics of the medium, such as effective electrical or thermal conductivity or elastic moduli. Identification of oriented systems of cracks and the evaluation of their parameters is an important problem that is of interest for many areas of physics of composite materials and petrophysics. The materials, containing high-conductivity inclusions, may be considered as homogeneous with certain effective properties, when their physical properties are determined in a larger scale compared to the characteristic size of inclusions.

During the past one hundred years, starting from the pioneering works by [5,6,19] a number of methodologies of calculation of effective conductivity of a medium with inclusions of different shapes have been developed [25-29]. A review of such methods that includes the works, published before 1990, was given in the papers by [3,26,30]. An overview of the more recent works is given in Refs. [4,29]. As a rule, in the majority of the works, that are based on the effective medium approach, the authors consider the media that contain either random or parallel inclusions. In the case of real media, such models may be frequently considered only as a rough approximation [18]. In the current work, we consider the model of a microinhomogeneous medium with inclusions that are not strictly oriented. Their spatial distribution is described by some angular function. Such models have long been in the interest of researchers. For elastic cracked solids, for example, such models are considered in Refs. [15,28,31]. The solid review about the effect of orientation distribution on the effective properties of the fiber reinforced materials can be found in Ref. [20]. In the works by [8,9,17] the influence of the inhomogeneity orientation on the elastic and conductive properties of microinhomogeneous materials are studied. The results obtained in the last works were based on the so-called non-interaction approximation that slightly reduced the area of possible application. Our calculations are based on the so-named effective field method that allows considering the interaction between the inclusions. According to this method, every inclusion in the inhomogeneous medium is considered as an isolated one, embedded in the homogeneous background medium (matrix). The field that acts on this inclusion (effective field) does not coincide with the “external” field applied to the medium, but it is the sum of this “external” field and disturbances induced by all surrounding inclusions. This method has a long history, and it was mainly used in nuclear physics and the theory of phase transitions for the description of various types of many particle interactions. In application to the mechanics of composite materials, this method was developed by [10,13]. This method has gained popularity for the calculation of effective elastic properties due to its simplicity. The most complete explanation of the method is given in the monographs by [13].

However, this method is not frequently used in application to the problems of thermal or electric conductivity. In our previous work by [16], we have applied this method to calculate the coefficients of the conductivity tensor of randomly oriented and parallel inclusions. In the present work, we apply the effective field method for the calculation of conductivity tensor of microheterogeneous media containing thin (crack-like) high-conductivity inclusions that are distributed in the space, and this distribution is characterized by some distribution density. In the second section of the paper, we discuss the homogenization problem, while in the third one we present the solution of the so-called one-particle problem that is the building block of any homogenization scheme. Then, the effective field method is shortly presented, and the effective conductivity of the material containing a random set of thin high-conductivity inclusions are calculated. After that we study the changes of the overall conductive material symmetry depending on the changes in the parameters of the distribution function over the orientation of the inclusions.

2. Homogenization problem

Let us consider a set of dispersed isotropic particles (inclusions), having the conductivity coefficient C, randomly distributed in an infinitely isotropic homogeneous medium, having the conductivity coefficient C 0. The vector of the local flux q i (x) and the field ei (x) in such a medium satisfy the system of equations:

iqix=0,        qix=Cijxejx,      rotijej(x)=0,        ixi. (2.1)

where x is an arbitrary point in 3D-space. These equations can describe various physical processes in solids, including stationary thermal conductivity and filtration, static electroconductivity in conductors and dielectrics, as well as, magneto- and electrostriction. All functions in Eqs. (3) are random functions of coordinates. Determination of the relation between the mathematical expectations of the fields qix and eix that generally has the form:

qix=Cij*ejx, (2.2)

where Cij* is the tensor of effective (or overall) conductivity. The determination of this tensor is the central problem of micromechanics (the homogenization problem). For the random set of inclusions, the exact solution to this problem is impossible, and only approximate methods are available. There are several such approximate homogenization schemes that are named self-consistent methods [13]. The main distinguishing feature of these methods is the reduction of the problem for many randomly placed particles to the problem for only one separate particle (one-particle problem), that is the building block of these methods. In what follow, we consider a special kind of particle: the thin inclusion with high conductivity. Such inclusion can be used for example as a model of crack, filled with saltwater in a rock formation.

3. One-particle problem for a thin high-conductivity inclusion in a homogeneous medium

This problem has been solved by [16]. Here, we present this solution only briefly, the details can be seen in the mentioned publication. Let us examine a homogeneous isotropic medium having the conductivity coefficient C 0, containing a single inclusion with the conductivity coefficient C, occupying the region v. The fields This problem has been solved by [16]. Here, we present this solution only briefly, the details can be seen in the mentioned publication. Let us examine a homogeneous isotropic medium having the conductivity coefficient C 0, containing a single inclusion with the conductivity coefficient C, occupying the region v. The fields q i (x) and e i (x) satisfy the integral equations [16]:

ei(x)=ei0(x)-(C-C0)vPij(x-x')eJ(x')dx',      qi(x)=qi0(x)+1C-1C0vQij(x-x')qj(x')dx', (3.1)

where it is denoted

Pij(x)=ij14πC0|x|,        Qij(x)=C0C0Pij(x)-δijδ(x), (3.2)

and qi0x are the “ external” fields that would be in the medium without inclusion; δx is the 3D-Dirac-delta function.

We assume that one characteristic length h of the region v is smaller than two others of order l. Thus the ratio δ 1 = h/l is small. A thin inclusion with conductivity greater than that of the surrounding medium is of prime interest in application. In this case, the ratio δ 2 = C0/C is also small. The most valuable information about the fields ei (x) and q i (x) in the vicinity of the inclusion is contained in the principal terms of the asymptotic expansion of these fields over the parameters δ 1 and δ 2. In order to construct these terms, it is necessary to find the limiting solution of the conductivity problem, when δ 1 2 → 0, and the ratio δ 1 2 is of unit order and remains constant.

Let us assume that the middle surface of the inclusion Ω is a smooth enough surface with a given continuous field of its normal vector n i (x). The surface Ω bounded by the closed contour Γ. We take a point x on Ω and put it in the origin of the local coordinate system with We assume that one characteristic length h of the region v is smaller than two others of order l. Thus the ratio δ 1 = h/l is small. A thin inclusion with conductivity greater than that of the surrounding medium is of prime interest in application. In this case, the ratio δ 2 = C 0 /C is also small. The most valuable information about the fields e i (x) and q i (x) in the vicinity of the inclusion is contained in the principal terms of the asymptotic expansion of these fields over the parameters δ 1 and δ 2. In order to construct these terms, it is necessary to find the limiting solution of the conductivity problem, when δ 1 2 → 0, and the ratio δ 1 2 is of unit order and remains constant. Let us assume that the middle surface of the inclusion Ω is a smooth enough surface with a given continuous field of its normal vector We assume that one characteristic length h of the region v is smaller than two others of order l. Thus the ratio δ 1 = h/l is small. A thin inclusion with conductivity greater than that of the surrounding medium is of prime interest in application. In this case, the ratio δ 2 = C 0 /C is also small. The most valuable information about the fields e i (x) and q i (x) in the vicinity of the inclusion is contained in the principal terms of the asymptotic expansion of these fields over the parameters δ 1 and δ 2. In order to construct these terms, it is necessary to find the limiting solution of the conductivity problem, when δ 1 2 → 0, and the ratio δ 1 2 is of unit order and remains constant.

Let us assume that the middle surface of the inclusion Ω is a smooth enough surface with a given continuous field of its normal vector n i (x). The surface Ω bounded by the closed contour Γ. We take a point x on Ω and put it in the origin of the local coordinate system with z-axis directed along the normal n i (x). We denote by h(x) the transverse size of the inclusion along the normal n i (x). Then, we denote h(x) = δ 1 l and take into account that δ 1 2 is O(1). It follows from Eqs. (3.1) that the main terms of the field expansion of e i (x) and q i (x) over δ 1 and δ 2 in the medium with thin high-conductivity inclusion can be expressed as:

eix=ei0x-ΩPijx-x'ηjx'dx',        qi(x)=qi0(x)+1C0ΩQij(x-x')ηj(x')dx', (3.3)

where

ηk(x)=(C-C0-hx/2hx/2ek(x+n(x)z)dz, (3.4)

and it is necessary to consider the main terms of expansion over δ 1 and δ 2 in the expansion for η i (x). To construct these terms (for which we will preserve the same notation η i (x)), we will use the method of matched asymptotic expansion. Using the results of [12], where one can find the details of the proof, it is possible to show that η i (x) is the vector of the surface Ω satisfying the equality:

θij(x)ηj(x)=ηi(x),         θij(x)=δij-ni(x)nj(x),        (xΩ). (3.5)

This vector is the solution of the following integral equation:

μijxηjx-ΩUij(x-x')ηj(x')dΩ'=θij(x)ej0(x), (3.6)

where it is denoted

μij(x)=1Ch(x)θij(x),                    Uij(x)=θik(x)Pkl(x-x')θlj(x') (3.7)

and the action of the operator with kernel Uij (x,x0) on a smooth enough function (regularization of this operator) can be obtained in Ref. [12].

Solving (3.6) for the function ηix and substituting the result in the right-hand side of (3.3), we obtain expressions for the field eix and qix. These fields approximate the real fields in the medium with inhomogeneity except for the small vicinity of the contour Γ. For an inclusion of arbitrary shape, (3.6) can be solved only numerically. But for ellipsoidal thin inclusion and constant “external” field ei0 this equation can be solved in the closed analytical form.

Let the inclusion be a thin ellipsoid with semi-axes a1,a2,h,(h/a1,h/a21. Then Ω becomes a plane elliptical surface. Let ni denotes its normal. In the system of coordinates with the axes that coincide with the major ellipsoid axes, function hx can be written as:

hx=2hzx,          z(x)=1-x1a12-x2a22. (3.8)

In the results of calculations, the details of which are presented in Appendix A, we obtain that the fields eix and qix outside of the thin inclusion can be expressed as:

eix=ei0-ΩUij(x-x')ΛjkZ(x')dx'ek0,              qi(x)=qi0+1C0Qij(x-x,)ΛjkZ(x')dx'ek0 (3.9)

Here, Zx is denoted

Zx=2a12a2zx, (3.10)

and tensor Λ ij in same system of coordinates with unit vectors n 1 i ,n 2 i ,n 3 i (n 3 i is coincided with the normal n i of the surface Ω) has the form:

Λij=Λ1a1,a2ni1nj1+Λ2a1,a2ni2nj2, (3.11)

where

Λ1a1,a2=C0a1δ2a2δ1+Kk-Ekk2,            δ1=ha1,        δ2=C0C,             Λ2a1,a2=C0a1δ2a2δ1+Ek-1-k2Kk1-k2k2,            k2=1-a2a12,        a1a2, (3.12)

K(k) and E(k) are the complete elliptic integrals of the first and the second kind, respectively:

Kk=0π/2dt1-k2sin2t,            Ek=0π/21-k2sin2tdt. (3.13)

4. Random set of thin low-resistivity inclusions in an isotropic homogeneous medium

We consider a microinhomogeneous medium consisting of a homogeneous isotropic host material (matrix) and random set of thin low-resistivity inclusions. The “one-particle” problem, considered in the previous Section, is basic for self-consistent homogenization schemes [11,13]. Here, we use one of these schemes that is named the effective field method (EFM). According to this method, we introduce a local “external” field ei*x that acts on each inclusion. This field is composed of an “external” field ei0 and the fields induced by surrounding inclusions. The main hypothesis of the effective field method is as follows: every inclusion in the composite material can be considered as an isolated one in the homogeneous matrix in a local uniform “external” field ei*, which depends on the orientation of the inclusion (vector n). Using this hypothesis, the expressions for the fields eix and qix can be represented in a form like Eqs. (3.3), in which:

ηix=Λijxej*nxZxΩx,        Ω(x)=kΩk(x). (4.1)

Here, Ωkx is a generalized function concentrated on the surface of the k-th inclusion, the function nx coincides with the normal n to the surface Ωk, when xΩk. The function Λijx is equal to the constant value Λija1k,a2k determined in (3.11), when xΩk and:

Zx=2a1k2a2k1-x1a1k2-x2a2k2. (4.2)

The procedure of the homogenization by the effective field method is presented in Appendix B. The final result is:

Cij*=C0δij+n0δik-n0vΛim(n)Amk(n)-1vΛkj(n),         v=43πa13, (4.3)

where means the averaging over ensemble distribution of the sizes and orientation of the inclusions; n0 is the number concentration of the inclusions. Tensor Aijn in (4.3) is determined by the relations:

Aij(n)=A1ni1nj1+A2ni2nj2+A3ni3nj3, (4.4)

 Ak=α1α2α32C0×0dσ(αk2+σ)(α12+σ)(α22+σ)(α32+σ),                 (k=1,2,3), (4.5)

where α1,α2,α3 are the semi-axes of the ellipsoidal “correlation hole” (see Appendix B).

The cases of parallel and completely disoriented thin high-conductivity inclusions were considered by [16]. In what follows, our attention is concentrated on the effects of partly oriented inclusions.

5. Description of the inclusion orientation distribution

For the description of the thin inclusion orientation, we introduce a global Cartesian basis eii=1,2,3 of the axes x1,x2,x3, where x3 is the vertical axis (Fig. 1). The orientation of the basis nkk=1,2,3 that defines the symmetry axes of the elliptical thin inclusion with respect to the global basis ei is described by the three Euler angles ψ,θ,φ, and the connection between these bases is given by the relations:

n1=(cosφcosψ-sinφsinψcosθ)e1+(cosφsinψ+sinφcosψcosθ)e2+sinφsinθe3, n2=-(sinφcosψ+cosφsinψcosθ)e1+(-sinφsinψ+cosφcosψcosθ)e2+cosφsinθe3,n3=sinψsinθe1-cosψsinθe2+cosθe3. (5.1)

FIGURE 1 The global basis is ei , and the basis ni defines the symmetry axes of the elliptical crack-like inclusion, (ψ,θ,ϕ) are the Euler angles.  

Substituting these formulas into (4.3), we obtain the presentation of the tensor Cij* in the global basis ei.

Let us consider some special cases.

Let the thin inclusions be the same size and shape, but randomly oriented with respect to the global coordinate system. In this case, the formula (4.3) takes the form:

Cij*=C0δij+τδik-τΛim(n)Amk(n)-1Λkj(n),              τ=43πa13n0, (5.2)

and the Euler angles in Eqs. (5.1) become random variables. The parameter τ in Eq. (5.2) is the so-called crack density.

We introduce the function of distribution fψ,θ,φ of the inclusion orientation over the angles ψ,θ,φ. This function must satisfy the normalization conditions:

18π202πdψ02πdφ02πf(ψ,φ,θ)sinθdθ=1 (5.3)

Suppose that the orientations of the inclusions, described by the angles ψ,θ,φ, are statistically independent. Therefore, the function fψ,θ,φ can be represented as:

fψ,θ,φ=fψψfφφfθθ. (5.4)

The normalization conditions take the form:

12π02πfψ(ψ)dψ=1,        12π02πfφφdφ=1,        12π0πfθ(θ)sinθdθ=1 (5.5)

We consider the case, when fψψ=1,fφφ=1,fθθ1. The medium has the macroscopically transversely isotropic conductive properties with symmetry axis x3:

Cij*=C1*ei1ej1+ei2ej2+C3*ei3ej3, (5.6)

where it is denoted:

C1*=C0+τ(Λ1+Λ2)(1+Sθ)4-τ(Λ1A1+Λ2A2)(1+Sθ),         C3*=C0+τ(Λ1+Λ2)(1-Sθ)2-τ(Λ1A1+Λ2A2)(1-Sθ), (5.7)

Sθ=120πfθ(θ)cos2θsinθdθ. (5.8)

The case fθθ=1 (the uniform distribution over the angle θ) leads to Sθ=1/3, that corresponds to the full isotropy of the material:

Cij*=C*ei1ej1+ei2ej2+ei3ej3=C*δij,            C*=C0+τ3Λ1+Λ21-τ3Λ1A1+Λ2A2-1 (5.9)

5.1. System of vertical inclusions

We consider the system of vertical inclusions (θ=π/2,φ=0). In this case

n1=cosψe1+sinψe2,        n2=e3,           n3=sinψe1+cosψe2. (5.10)

We choose one of the simplest distribution functions over the angle ψ:

fψ(ψ)=1    when    ψ-β,β]0    when    ψ-β,β] (5.11)

It means that the horizontal crack-like inclusions homogeneously distributed in the sector -β,β, where β is the disorientation angle. For 0<β<π/2 the conductive properties of such material have the orthorhombic symmetry with the following tensor of effective conductivity coefficients:

Cij*=C1*ei1ej1+C2*ei2ej2+C3*ei3eJ3, (5.12)

where

C1*=C0+τΛ1Fβ1-τΛ1A1Fβ-1,  C2*=C0+τΛ11-Fβ1-τΛ1A11-Fβ-1, C3*=C0+τΛ21-τΛ2A2-1, (5.13)

F(β)=12β(β+sinβcosβ). (5.14)

When β0 (completely aligned vertical inclusions), Fβ1, and the medium is still orthotropic. The tensor of effective conductivity coefficients is determined by the same formula (5.12), in which:

C1*=C0+τΛ11-τΛ1A1-1,        C2*=C0,          C3*=C0+τΛ21-τΛ2A2-1 (5.15)

If βπ/2,Fβ1/2, and the medium becomes transversely isotropic with the symmetry axis x3. Expression for the C3* in (5.12) remains the same, but:

C1*=C2*=C0+τ2Λ11-τ2Λ1A1-1. (5.16)

Suppose that the inclusions have the same spheroidal shape: a1=a2=a,a3. For such inclusions:

Λ1=Λ2=Λ=C0δ2δ1+π4-1 (5.17)

If the shape of the correlation hole is also spheroidal (α1=α2=α,α3) and coaxial to the inclusion, then:

A1=A2=A=g(γ)C0,      A3=12C01-gγ,      γ=aa3>1,       g(γ)=γ2γ2-11-1γ2-1arctanγ2-1. (5.18)

For spheroidal inclusions and β=0 (parallel inclusions), the medium is transversely isotropic with the symmetry axis x2:

Cij*=C*ei1ej1+ei3ej3+C0ei2ej2, (5.19)

where

C*=C0+τΛ1-τΛA-1. (5.20)

It has to be noted, that the angle distribution function can be chosen as a continuous one [18]:

fψψ=12πexpcosψ/σ2, (5.21)

where σ is the parameter, which characterizes the inclusion disorientation. With this distribution function the tensor of effective conductivity coefficients is determined by the same formula (5.12), in which:

C1*=C0+τΛ1F1(σ)1-τΛ1A1F1(σ)-1,C2*=C0+τΛ1F2(σ)1-τΛ1A1F2(σ)-1,C3*=C0+τΛ21-τΛ2A2-1. (5.22)

Here, it is denoted

F1(σ)=σ2I1(1/σ2)+I2(1/σ2)I0(1/σ2),                  F2(σ)=σ2I1(1/σ2)I0(1/σ2). (5.23)

where Inz is the modified Bessel function.

Note, that the limit σ0 corresponds to the distribution function in the form:

fψψ=δψ, (5.24)

i.e. the inclusions are parallel to the plane x2x3. In this case, F1σ=1,F2σ=0 , and Eqs. (5.22) are transformed to Eqs. (5.15). The other limit σ corresponds to the homogeneous distribution of inclusions with respect to the angle ψ. For this limit F1σ=F2σ=1/2, the formula (5.16) is recovered from (5.22), i.e. the system becomes transversely isotropic with the symmetry axis x3.

In a more general case, the distribution function can be represented in the form of series of spherical harmonics. The coefficients of such series can be calculated from the analysis of measuring data.

5.2. System of horizontal thin inclusions with high conductivity.

Now we consider the case, when θ=0 (horizontal inclusions). It is well known that in this case the Euler angles degenerate (the line of nudes coincides with x1- axis, and the angles φ and ψ become uncertain). At the same time, it is possible to introduce the only one angle φ between the vectors n1 and e1 that determines the orientation of inclusions in the x1x2-plane. For this situation we have:

n1=cosφe1+sinφe2,     n2=-sinφe1+cosφe2,       n3=e3. (5.25)

With the same distribution function (5.11) the tensor of the effective conductivity coefficients Cij* has the form:

Cij*=C1*ei1ej1+C2*ei2eJ2+C0ei3ej3. (5.26)

One of these independent coefficients of conductivity coincides with those of the matrix in the direction of x3-axis, and two other coefficients are determined by the expressions:

C1*=C0+τΛ1F(β)+Λ2(1-F(β)×1-τΛ1A1F(β)+Λ2A2(1-F(β)-1,C2*=C0+τΛ1(1-F(β))+Λ2F(β)×1-τΛA1(1-F(β))+Λ2A2F(β)-1 (5.27)

where the function Fβis determined in (5.14).

When β0 (the symmetry axes of the horizontal ellipsoidal inclusions are completely aligned), the effective medium remains orthorhombic with the effective conductivities (see formula (5.19)):

C1*=C0+τΛ11-τΛ1A1-1,         C2*=C0+τΛ21-τΛ2A2-1. (5.28)

The other limiting case βπ/2 corresponds to the transversely isotropic material with the following tensor of effective coefficients of conductivity:

Cij*=C*ei1ej1+ei2ej2+C0ei3ej3, (5.29)

where

C*=C0+τ2Λ1+Λ21-τ2Λ1A1+Λ2A2-1. (5.30)

If we choose the distribution function in the form (5.21), the formula for the tensor of the effective conductivity coefficients remains the same (5.26), but in this case:

C1*=C0+τΛ1F1(σ)+Λ2F2(σ)×1-τΛ1A1F1(σ)+Λ2A2F2(σ)-1,C2*=C0+τΛ1F2(σ)+Λ2F1(σ)×1-τΛ1A1F2(σ)+Λ2A2F1(σ)-1 (5.31)

In both limit cases σ0 and σ, the (5.28)-(5.30) are recovered from Eqs. (5.31).

If the inclusions are identical spheroids, the medium with such inclusions is transversely isotropic (5.29). In this formula:

C*=C0+τΛ1-τΛA-1, (5.32)

independently on the choice of distribution functions (Λ and A are determined in (5.17) and (5.18)).

6. Numerical examples

To validate our model, we calculated the effective conductivity of the medium, which contains parallel spheroidal inclusions, using the effective field method and the well-known Effective Medium Approximation (EMA) by [3,4,5]. The semi-axes of the inclusions were chosen as a 1 = a 2 = 1, a 3 = 0.001. The inclusions conductivity is equal to 1, and the matrix conductivity is equal to 0.01. To compare the results obtained with the EMA model and the effective field method, we use the inclusion concentration Φ instead of the crack density τ. For spheroidal inclusions τ = 3Φ/(4πα), where α is the aspect ratio of the spheroid. The comparison of the results obtained shows that both methods give close results for parallel inclusions [Fig. 2], but to calculate the effective conductivity of the material with partially oriented inclusions by the EMA method is a complicated mathematical problem,while the (EFM) gives the explicit expressions for the effective conductivity coefficients.

FIGURE 2 Effective conductivity as a function of the inclusion concentration. The calculations are presented for parallel spheroidal inclusions.  

Further, we present examples of the effective conductivity calculation for partially oriented inclusions. We assume that the inclusion conductivity is equal to 1, and the matrix conductivity is equal to 0.01. The semi-axes of the inclusions were chosen as a 1 = 1, a2 = 0.5, a3 = 0.01. Such parameters are typical, for example, for sedimentary carbonate rocks containing cracks, filled with conductive formation water [1]. We assume that the correlation hole has a spherical shape: A 1 = A2 = A3 = 1. Figures 3 and 4 show the dependences of the conductivity tensor components of the horizontal and vertical crack-like inclusions, homogeneously distributed in the sector [−β,β], where β is the disorientation angle. The angle distribution function is described by (5.11). The calculations were fulfilled for the disorientation angle θ = π/2 and θ = 0(ϕ = 0) for vertical and horizontal inclusions, respectively. The aspect ratio of the correlation hole is close to 1. As expected, the effective medium is orthorhombic C1*C2*C3*, but with increasing of the value of the parameter β the difference between the conductivities C1* and C2* decreases. When the β value is close to π/2, the difference between C1* and C2* tends to zero, and the medium becomes transversely isotropic (Figs. 3 and 4).

FIGURE 3 Normalized components of the conductivity tensor as a function of the disorientation angle β. The crack density τ = 1. The results are presented for the system of vertical inclusions (θ = π/2, ϕ = 0) and the first model of the angular distribution (inclusions homogeneously distributed in the sector [−β,β]). 

FIGURE 4 Normalized components of the conductivity tensor as a function of the disorientation angle β. The crack density τ = 1. The results are presented for the system of horizontal inclusions (θ = 0, ϕ = 0) and the first model of the angular distribution (inclusions homogeneously distributed in the sector [−β,β]).  

Figures 5-6 show the dependences of the normalized components of the conductivity tensor as a function of the disorientation parameter σ (the distribution function (5.21). The results are presented for vertical and horizontal cracklike inclusions. In the general case, the effective medium is orthorhombic, but in the case, when σ, the effective medium becomes transversely isotropic.

FIGURE 5 Normalized components of the conductivity tensor as a function of the disorientation parameter σ. The crack density τ = 1. The results are presented for the system of vertical inclusions (θ = π/2, ϕ = 0) and the second model of the angular distribution of the inclusions (Eq.(5.21)).  

FIGURE 6 Normalized components of the conductivity tensor as a function of the disorientation parameter σ. The crack density τ = 1. The results are presented for the system of vertical inclusions (θ = 0, ϕ = 0) and the second model of the angular distribution of the inclusions (Eq.(5.21)).  

Figure 7 illustrates the dependences of the normalized components of the conductivity tensor as a function of the disorientation angle β for different crack densities. The results are presented for vertical inclusions. In the general case, the effective medium is orthorhombic, but in the case, when β=π/2, the effective medium becomes transversely isotropic.

FIGURE 7 Normalized components of the conductivity tensor as a function of the disorientation angle β. Different curves correspond to the different values of crack densities. Figure 7A corresponds to the conductivity in the direction x 1, and Figure 7B corresponds to the conductivity in the direction x 2. The results are presented for the first model of angular distribution (inclusions homogeneously distributed in the sector [−β,β]).  

7. Conclusion

We have presented an approach for calculating the effective conductivity tensor of material containing a system of cracks that are not strictly oriented. Their spatial distribution is described by some angular function. The approach is based on the effective field method. This method is sufficiently general and contains the well-known method of Mori-Tanaka and the Maxwell method as particular cases. The advantage of the effective field method lies in its simple numerical realization compared with other methods. The application of this method permits us to take into account the texture of a microinhomogeneous medium. The general theory was illustrated by numerical results obtained for crack-like inclusions homogeneously distributed in the sector. We have shown that in the case of circular penny-shaped inclusions with the same aspect ratio, the effective medium can be orthorhombic, transversely isotropic, or isotropic depending on the choice of the distribution function of the normal to the crack surface from Euler angles. From our point of view, the dependences, obtained in the framework of this work, may be interesting for many areas of applications, including the rock physics and physics of micro-inhomogeneous (cracked) materials.

Acknowledgements

The authors are grateful to the Mexican Petroleum Institute, where this study was fulfilled. We thank the anonym Reviewer for useful comments and discussion.

References

[1] R. Aguilera Naturally Fractured Reservoirs, PennWell Books, (1995) p. 521. [ Links ]

[2] Ya. Benveniste, On the effective thermal conductivity of multiphase composites, ZAMP, 37 (1986) 696, DOI: 10.1007/BF00947917. [ Links ]

[3] D.J. Bergman, D. Stroud, The physical properties of macroscopically inhomogeneous media, Solid State Phys. 46 (1992), 148, [ Links ]

[4] C. Brosseau, Modeling and simulation of dielectric heterostructures: a physical survey from an historical perspective. J. of Physics. D: Appl. Phys. 39 (2006) 1277, [ Links ]

[5] D.A. Bruggeman, Berechnung Verschidener Physikalischer Konstanten von Heterogenen Substanzen. Ann. Phys. Lpz 24 (1935) 636, [ Links ]

[6] H. Fricke, A mathematical treatment of the electric conductivity and capacity of disperse systems, Phys. Rev. 24 (1924) 575. [ Links ]

[7] S. Giordano, Effective medium theories of dielectric ellipsoids. J. of Electrostatics 58 (2003) 59, [ Links ]

[8] S. Giordano, Order and disorder in heterogeneous material microstructure: electric and elastic characterization of dispersions of pseudo oriented spheroids, International Journal of Engineering Science 43 (2005) 1033. [ Links ]

[9] S. Giordano, and L. Colombo, Effects of the orientation distribution of cracks in isotropic solids, Engineering Fracture Mechanics 74 (2007) 1983, [ Links ]

[10] S.K. Kanaun, Elastic medium with random fields of inhomogeneities, in: I.A. Kunin (Ed.), Elastic Media with Microstructure II, Three-dimensional Models, Springer, Berlin, (1983) 165. [ Links ]

[11] M. Kachanov, and I. Sevostianov, Effective Properties of Heterogeneous Materials. Solid Mechanics and Its Applications 193 (2013). [ Links ]

[12] S.K. Kanaun, The thin defect in a homogeneous elastic medium. Mech. Solids, 3 (1984) 74, [ Links ]

[13] S. K. Kanaun, and V. M. Levin, Self-Consistent Methods for Composites V.I Static Problems, Springer. Dordrecht (2008) p. 386. [ Links ]

[14] S. Kanaun, and V. Levin, Effective field method in the theory of heterogeneous media, in M. Kachanov M. , Sevostianov I. (Eds.), Effective Properties of Heterogeneous Materials. Solid Mechanics and Its Applications , 193 (2013) 199. ISBN 978-94-007-5715-8. [ Links ]

[15] V. Kushch, I. Sevostianov , and Jr. Mishnaevsky, Effect of crack orientation statistics on effective stiffness of microcracked solid. Int. J. of Solids and Structures 46 (2008) 2574. [ Links ]

[16] V. Levin , and M. Markov, Electroconductivity of a medium with thin low-resistivity inclusions, Journal of Electrostatics 61 (2004) 129, [ Links ]

[17] M. Markov, A. Mousatov, and E. Kazatchenko, Conductivity of carbonate formations with microfracture systems. J. of Petr. Sci. and Eng. 69 (2009) 247. [ Links ]

[18] G. Mavko, T. Mukerji, and J. Dvorkin, The rock physics handbook. Tools for seismic analysis in porous media. Cambridge university press, (2009) p. 330. [ Links ]

[19] J. C. Maxwell, A treatise on electricity and magnetism, 3rd ed.: Clarendon Press, (1892). (Reprinted by Dover, Mineola, N.Y., 1954). [ Links ]

[20] T. Mishurova et al., Evaluation of the probability density of orientations of inhomogeneities by computed tomography and its application to the calculation of the effective properties of a fiber reinforced concrete. International Journal of Engineering Science 122 (2018) 14, [ Links ]

[21] T. Mori, and K. Tanaka, Average stress in matrix and average energy of materials with misfitting inclusions, Acta Metallurgia 21 (1973) 571, [ Links ]

[22] N. Phan-Thien, and D. C. Pham, Differential multiphase models for polydispersed spheroidal inclusions: thermal conductivity and effective viscosity. International Journal of Engineering Science 38 (2000) 73, [ Links ]

[23] D. A. Robinson, and S. P. Friedman, Observations of the effects of particle shape and particle size distribution on avalanching of granular media. Physica A: Statistical Mechanics and its Applications 311 (2002) 97, http://doi.org/10.1016/S0378-4371(02)00815-4. [ Links ]

[24] B. Shafiro, andM. Kachanov , Anisotropic effective conductivity of materials with nonrandomly oriented inclusions of diverse ellipsoidal shapes. J. Appl. Phys., 87 (2000) 8561. [ Links ]

[25] P. Sen, C. Scala, and M.H. Cohen, A self-similar model for sedimentary rocks with application to the dielectric constant of fused glass beard. Geophysics 46 (1981) 781. [ Links ]

[26] M.I. Shvidler, Statistical hydrodynamics of porous media. Nedra, Moscow, (1985) p. 228 (In Russian). [ Links ]

[27] I. Sevostianov , and A. Giraud, On the compliance contribution tensor for a concave superspherical pore. International Journal of Fracture, 177 (2012) 199. [ Links ]

[28] I. Sevostianov , andM. Kachanov , Modeling of the anisotropic elastic properties of plasma-sprayed coatings in relation to their micro-structure. Acta Mater., 48 (2000) 1361. [ Links ]

[29] A. Sihvola, Electromagnetic Mixing Formulae and Applications, IEEE Electromagnetic Waves Series, 47 (1999) 296 [ Links ]

[30] L.K.H. Van Beek, Dielectric behavior of heterogeneous systems, Prog. Dielectr 7 (1976) 71. [ Links ]

[31] X. Zeng, and Y. Wei, The influence of crack-orientation distribution on the mechanical properties of pre-cracked brittle media. Int. J. of Solids and Structures 90 (2016) 64. [ Links ]

Appendix A.

The kernel of the integral operator in Eq. (3.6) has the form:

Uijx,x'=Uijx-x'=θiknPklx-x'θljn. (A.1)

It can be shown [13] that integral operator with such a kernel transforms the function zx into a constant on Ω. It allows us to find the solution of Eq. (3.5) for the elliptic domain Ω with ei0=const in the form of the constant vector multiplying on zx:

ηix=ηizx. (A.2)

Substituting (A.2) into Eq. (3.5) and taking into account (A.1), we obtain:

ηi=2a12a2μik0+Uik0-1ek0,μik0=12hCθik, (A.3)

and the constant tensor Uik0 is expressed in term of absolutely converging integral:

Uij0=θik(n)Pkl0θlj,               Pkl0=Pkl(x)[z(x)-1]dΩ. (A.4)

Here, the integration is over the plane x1x2, and the function zx vanishes outside Ω.

The explicit expression for Pijx is:

Pklx=-14πC0x3δkl-3xkxlx2. (A.5)

Substituting this expression into (A.4) and introducing the coordinates r and φ in the plane x1x2: x1=ra1cosφ,x2=ra2sinφ (a1,a2 are the ellipse Ω semi-axes), we obtain

Pkl0=-a1a24πC00z(r)-1r2×02π3mkl(φ)t2(φ)-δkldφt3(φ),         (k,l=1,2) (A.7)

t2(φ)=a12cos2φ+a22sin2φ,       z(r)=1-r2,(r1),       z(r)=0,    (r>1)m11=a12cos2φ,        m22=a22sin2φ,             m12=m21=a1a2sinφcosφ.

Calculation of the integral in (A.7) gives

U110=a22a12C0K(k)-E(k)k2,        k2=1-a2a1,       a1a2,       U12=0,      U220=a22a12C0E(k)-(1-k2)K(k)k2(1-k2). (A.8)

After the vector ηix has been determined, the function eix and qix outside the inclusion can be expressed by (3.8)-(3.9) of the main text, and tensor Λij is determined by (3.11)-(3.12).

Appendix B

If we introduce the function:

Ωx;x'=ijΩix',whenxΩj. (B.1)

The equation for the local “external” field at the point x located on the middle surface of an arbitrary inclusion can be presented in the following form:

ei*(x)=ei0-Pij(x-x')Λjk(x')×ek*(x')Z(x')Ω(x;x')dx',     xΩ. (B.2)

Let us average this equation under the condition that the point x is located on the middle surface of the inclusion with normal n. This averaging is denoted as |x,n. If the mean ei*x|x,n is identified with an effective field acting on the inclusion of orientation n:

ei*x|x,n=ei*n, (B.3)

then we obtain from (B.2)

ei*(x)=ei0-Pij(x-x')Λjk(x')×ek*(x')Z(x')Ω(x;x')|x,ndx'. (B.4)

Assuming that the conductivity properties of inclusions are statistically independent on their spatial position, one can find the expression for the mean under the integral in (B.4):

Λjkx'ek*x'Zx'Ωx;x'x,n=Λijnej*nΨnx-x',Λij(n)=Z(x)Ω(x)Λij(x),Ψn=Ω(x;x')x,nΩ(x) (B.5)

The mean Λijnej*n is calculated over the ensemble of inclusion distribution by orientation. The function Ψnx characterizes the spatial correlation of the random set of thin inclusions. It follows from definition of the function Ωx;x' that

Ψn0=0,Ψnx1,when|x|. (B.6)

This function defines the shape of the so-called “correlation hole” - the region in the vicinity of each inclusion inside which the existence of the center of some other inclusion is improbable. Let us assume that there exists a linear transformation of x-space rearrange the function Ψnx into a spatially symmetric one:

yi=αijn=Ψy. (B.7)

In this case, the ellipsoid A with semi-axes α1,α2,α3, defined by the equation:

αijnxj1, (B.8)

describes the shape of the correlation hole.

After the substitution (B.5) in Eq. (B.4), one can obtain an expression for ei*n in the form:

ei*n=ei0+AiknΛkjnej*n, (B.9)

where it is denoted:

Aijn=Pijx1-Ψnxdx. (B.10)

If the correlation hole is an ellipsoid, coaxial with the inclusion having the orientation n, then Aijn is defined by the formulas (4.4) of the main text.

Let us multiply both sides of Eq. (B.9) by the tensor Λijn and average the result over the ensemble of random sizes and orientation of the inclusions. Solving the obtained equation for the vector Λkjnej*n, we have:

Λkj(n)ej*(n)=δil-Λik(n)Akl(n)-1×Λlj(n)ej0. (B.11)

The expression for the effective field ei*n can be found, if we substitute Λijnej*n from (B.11) into the right-hand side of Eq. (B.9).

Let us average the Eqs. (4.1) over the ensemble of the random set of the inclusions. Taking into account the relation:

Λijxei*xZxΩx=Λijnej*n, (B.12)

we obtain:

ei=ei0-Pij(x-x')Λjk(n)ek*(n)dx',qi=qi0+1C0Qij(x-x')Λjk(n)ek*(n)dx'. (B.13)

Because the “external” field ei0 is fixed in the homogenization problem [13], Eqs. (B.13) yield:

ei=ei0,      qi=Cij*ej, (B.14)

where

Cij*=C0δik+δik-ΛilnAlkn-1Λkjn, (B.15)

is the tensor of the effective conductivity coefficients of the composite material with a random set of thin high conductive inclusions.

Note that the Mori-Tanaka approach [2,21] gives the same (B.15) expression for the effective conductive coefficients only if the shape of the correlation hole coincides with the shape of the typical inclusion. In the general case, this shape can be different [16].

The ergodic properties of the functions considered allow replacing the ensemble averaging over random set of inclusions by the volume averages over the fixed realization of this set, so that (a1a2):

Λijn=limW1WWΛijxZxΩxdx=n0vΛija1,a2, (B.16)

with analogous determination of the average ΛiknAkjn.

Received: July 26, 2021; Accepted: September 26, 2021

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License