Asymptotics of the stress concentration in high-contrast elastic composites
Abstract.
A long-standing area of materials science research has been the study of electrostatic, magnetic, and elastic fields in composite with densely packed inclusions whose material properties differ from that of the background. For a general elliptic system, when the coefficients are piecewise Hölder continuous and uniformly bounded, an -independent bound of the gradient was obtained by Li and Nirenberg [41], where represents the distance between the interfacial surfaces. However, in high-contrast composites, when tends to zero, the stress always concentrates in the narrow regions. As a contrast to the uniform boundedness result of Li and Nirenberg, in order to investigate the role of played in such kind of concentration phenomenon, in this paper we establish the blow-up asymptotic expressions of the gradients of solutions to the Lamé system with partially infinite coefficients in dimensions two and three. We discover the relationship between the blow-up rate of the stress and the relative convexity of adjacent surfaces, and find a family of blow-up factor matrices with respect to the boundary data. Therefore, this work completely solves the Babus̆ka problem on blow-up analysis of stress concentration in high-contrast composite media. Moreover, as a byproduct of these local analysis, we establish an extended Flaherty-Keller formula on the global effective elastic property of a periodic composite with densely packed fibers, which is related to the “Vigdergauz microstructure” in the shape optimization of fibers.
1. Introduction
In this paper we are concerned with the blow-up behavior of the gradients of solutions to a class of elliptic systems, stimulated by the study of composite media with closely spaced interfacial boundaries. It is a long-standing area of material science research to study the high concentration of electrostatics, magnetic, and elastic fields in high-contrast composites with densely packed inclusions since the time of Maxwell and Reyleigh. This requires an understanding of micro-structural effects, especially from the distances (say, ) between inclusions, because when the inclusions are close to touching, the charge density becomes nearly singular. To evaluate the electrostatic fields (where the potential function is scalar-valued), the potential theory, Fourier analysis, and numerical method have been fully developed. While, for the elastic field (where the deformation displacement is vector-valued), in order to predict damage initiation and growth in carbon-fiber epoxy composites at the fiber scale level, Babus̆ka, et al. [8] assumed the systems of linear elasticity
in unidirectional composites to numerically analyze the residual stresses and stresses due to mechanical loads, where expresses the displacement. Obviously, this multiscale problem need more regorous mathematical treatment and numerical analysis to control the errors of the analysis. On the other hand, we emphasize that there is a significant difficulty in applying the method developed for scalar equations to systems of equations. For instance, the maximum principle does not hold for the Lamé system. Due to these difficulties on PDE theory and numerical analysis as well as the importance in practical applications, it arouses great interest of many applied mathematicians and engineers. In the last two decades, there has been an extensive study on the gradient estimates of solutions to elliptic equations and systems with discontinuous coefficients, to show whether the stress remains bounded or blows up when inclusions touch or nearly touch.
Bonnetier and Vogelius [15] considered the elliptic equation with piecewise constant coefficients in dimension two
where the scalar is the out of plane displacement, represents the cross-section of a fiber-reinforced composite taken perpendicular to the fibers, containing a finite number of inhomohenuities, which are very closely spaced and may possible touch. The coefficients take two different constant values, after rescaling,
Despite the discontinuity of the coefficient along the interfaces, they proved that any variational solution is in , which actually improves a classical regularity result due to De Giorgi and Nash, which asserts that solution is in some Hölder class. A general result was established by Li and Vogelius [43] for a class of divergence form elliptic equations with piecewise Hölder continuous coefficients. They obtained a uniform bound of regardless of in any dimension . Li and Nirenberg [41] extended the results in [43] to general elliptic systems including systems of elasticity. This, in particular, answered in the affirmative the question that is naturally led to by the above mentioned numerical indication in [8] for the boundedness of the stress as tends to zero. Dong and Xu [21] further showed that a weak solution is Lipschitz and piecewise . Recently, Dong and Li [20] used an image charge method to construct a Green’s function for two adjacent circular inclusions and obtained more interesting higher-order derivative estimates for non-homogeneous equations making clear their specific dependence on and exactly. But for more general elliptic equations and systems, and more general shape of inclusions, it is still an open problem to estimate higher-order derivatives in any dimension. We draw the attention of readers to the open problem on page 894 of [41].
As mentioned above, the concentration of the stresses is greatly influenced by the thickness of the ligament between inclusions. To figure out the influence from this thickness , one assumes that the material parameters of the inclusions degenerate to infinity. However, this makes the situation become quite different. As a matter of fact, in 1960’s, in the context of electrostatics, Keller [33] computed the effective electrical conductivity for a composite medium consisting of a dense cubic array of identical perfectly conducting spheres (that is, degenerates to ) imbedded in a matrix medium and first discovered that it becomes infinite when sphere inclusions touch each other. Keller found that this singularity is not contained in the expressions given by Maxwell, and by Meredith and Tobias [52]. See also Budiansky and Carrier [16], and Markenscoff [49]. Rigorous proofs were later carried out by Ammari et al. [6, 7] for the case of circular inclusions by using layer potential method, together with the maximum principle. Since then, there is a long list of literature in this direction of research, for example, see [1, 3, 12, 13, 14, 19, 24, 25, 26, 29, 34, 38, 37, 39, 44, 45, 48, 47, 55, 56] and the references therein. It is proved that the blow-up rate of is in dimension two and in dimension three. From the perspective of practical application in engineering and the requirement of numerical algorithm design, it is more interesting and important to characterize the singular behavior of , see [2, 30, 31, 36, 42, 46, 40].
In the context of linear elasticity, for Lamé system with partially infinite coefficients, by building an iteration technique with respect to the energy, the first author and his collaborators overcame the difficulty caused by the lack of the maximum principle, obtained the upper and lower pointwise bounds of , and showed that may blow up on the shortest line between two adjacent inclusions, see [9, 10, 11, 35]. By using the polynomial function , , as a local expression of inclusion’s boundary to measure its order of convexity, Li and Hou [27] revealed the relationship between the blow-up rate of and the convexity order . However, under the same logic as in the electrostatics problem, what one cares more about in practical applications is how to obtain an asymptotic formula to characterize the singular behaviour of in the whole narrow region between two adjacent inclusions. The main contribution of the paper is that we completely solve this problem in two physically relevant dimensions and , and for all . For , the result is similar. Our asymptotic expressions of the gradients of solutions not only show the optimality of the blow-up rates, which depend only on the dimension and the convexity order of the inclusions, but also provide a family of blow-up factor matrices, which are linear functionals of boundary value data, determining whether or not blow-up occurs. Notice that when , the curvature of the inclusions vanishes at the two nearly touch points, so in general we can not use a spherical inclusion to approximate an -convex inclusion.
The asymptotic formulas obtained above clearly reflect the local property of . Beyond this, they can further influence the global property of a composite. For the effective elastic moduli of a composite, Flaherty and Keller [22] obtained an symptotic formula for a retangular array of cylinders in the nearly touching, when the cylinders are hard inclusions and showed their validity numerically. As an application of the above local asymptotic formulas, we give an extended Flaherty-Keller formula for -convex inclusions, which is also related to the “Vigdergauz microstructure” [54], having a large volume fraction in the theory of structure optimization, see Grabovsky and Kohn [23].
To end this introduction, we make some comments on the corresponding numerical problem. Accurate numerical computation of the gradient in the present of closely spaced inclusions is also a well-known challenging problem in computational mathematics and sciences. Here it should be noted that Lord Rayleigh, in his classic paper [53], use Fourier approach to determine the effective conductivity of a composite material consisting of a periodic array of disks in a uniform background. In the case that inclusions are reasonably well separated or have conductivities close to that of the background, Rayleigh’s method gives excellent result. Unfortunately, if the inclusions are close to touching and their conductivities differ greatly from that of the background, the charge density becomes nearly singular and the number of computation degrees of freedom required extremely large. Recently, a hybrid numerical method was developed by Cheng and Greengard [18], and Cheng [17]. Related works can be referred to Kang, Lim, and Yun [30], and McPhedran, Poladian, and Milton [51]. For high-contrast elastic composite, a serious difficulty arises in applying the methods for scalar equations to systems of equations. We expect our asymptotic formulas of in the narrow regions, the most difficult areas to deal with, can open up a way to do some computation for inclusions of arbitrary shape.
This paper consists of eight sections including introduction. In Section 2, we first fix our domain and formulate the problem with partially infinite coefficients, and then introduce a family of vector-valued auxiliary functions with several preliminary estimates including the main ingredient Proposition 2.3 for the asymptotics of , . In Section 3, a family of the blow-up factors is defined. Then our main results are stated. Theorem 3.2 and Theorem 3.5 are for 2-convex inclusions in 2D and 3D, respectively, Theorem 3.8 and Theorem 3.11 are for -convex inclusions. Finally we give an example to show the dependence on the precise geometry feature of and . In Section 4, we prove the two important ingredients Propositions 2.3 and 3.1, where two improved estimates, Theorem 2.2 and Theorem 2.6, are used . The proofs of Theorems 3.2 and 3.5 are given in Section 5. We prove Theorems 3.8 and 3.11 in Section 6. Finally, by applying the local asymptotic formulas established in the previous sections, we give an extended Flaherty-Keller formula in Section 7. The proof of Theorem 2.2 and Theorem 2.6 is given in the Appendix.
2. Problem formulation, Decomposition and Some preliminary results
In this section we first fix our notations and formulate the problems, then identify the key difficulties and present the strategy to solve them, and finally introduce our auxiliary functions, involving the parameters of Lamé system, and give some preliminary results.
2.1. Problem formulation
Because the aim of this paper is to study the asymptotic behavior of in the narrow region between two adjacent inclusions, we may without loss of generality restrict our attention to a situation with only two adjacent inclusions. The basic notations used in this paper follow from [10].
We use to denote a point in , where and . Let be a bounded open set in with boundary. and are two disjoint convex open subsets in with boundaries, -apart, and far away from . That is,
where is a constant independent of . We also assume that the norms of , , and are bounded by some positive constant independent of . Set
Assume that and are occupied, respectively, by two different isotropic and homogeneous materials with different Lamé constants and . Then the elasticity tensors for the background and the inclusion can be written, respectively, as and , with
and
where and is the kronecker symbol: for , for . Let denote the displacement field. For a given vector-valued function , we consider the following Dirichlet problem for the Lamé system with pieceiwise constant coefficients:
(2.1) |
where is the characteristic function of , and
is the strain tensor.
Assume that the standard ellipticity condition holds for (2.1), that is,
For , it is well known that there exists a unique solution to the Dirichlet problem (2.1), which is also the minimizer of the energy functional
on
As mentioned previously, Li and Nirenberg [41] proved that is uniformly bounded with respect to . But, in high-contrast composite media, the concentration of is a very usual phenomenon when the distance is sufficiently small. In order to investigate the role of in such concentration phenomenon, let us assume that the Lamé constant in degenerates to infinite and consider this extreme case. To this end, we first introduce the linear space of rigid displacement in :
with a basis , namely,
where denote the standard basis of . For fixed and , denoting as the solution of (2.1), then we have [10]
where is the unique solution of
(2.2) |
where
and is the unit outer normal of , . Here and throughout this paper the subscript indicates the limit from outside and inside the domain, respectively. The existence, uniqueness and regularity of weak solutions to (2.2) can be referred to the Appendix of [10]. We note that it suffices to consider the problem (2.2) with replaced by . Indeed, it follows from the maximum principle [50] that . Taking a slightly small domain , then in view of the interior derivative estimates for Lamé system, we find that satisfies
Without loss of generality, we assume that by considering if . If , then .
2.2. Main difficulties and decomposition
We first point out that problem (2.2) has free boundary value feature. Although implies in is linear combination of ,
(2.3) |
these constants are free, which will be uniquely determined by . We would like to emphasize that this is exactly the biggest difference with the conductivity model [12], where only two free constants need us to handle in any dimension. It is the increase of the number of free contants that makes elastic problem quite difficult to deal with. Therefore, how to determine such many constants is one of main difficulties we need to solve.
Our strategy in spirit follows from [10, 11]. First, by continuity of across , we can decompose the solution of (2.2),
(2.4) |
where , respectively, satisfying
(2.5) |
and
(2.6) |
So
(2.7) |
To investigate the symptotic behavior of , we need both the asymptotic formulas of and the exact value of . To solve , from the fourth line in (2.2) and the decomposition (2.4), we have the following linear system of these free constants ,
(2.8) |
where , . But these coefficients are all boundary integrals. By integration by parts,
(2.9) |
Therefore, in order to solve from (2.8), we have to calculate the energy integral on the right hand side of (2.9). This in turn needs to estimate . In fact, even if we can have the asymptotic formulas of , see Theorem 2.2 and Theorem 2.6 below, it is still hard to solve every . It seems to be a mission impossible. To avoid this difficulty, we rewrite (2.7) as
(2.10) |
where
(2.11) |
This is because the following boundedness estimates for and , together with the boundedness of , , makes is a “good” term, which has no singularity in the narrow region.
Thus, we reduce the establishment of the asymptotics of to that of the asymptotics of , , and to solving , . These are two main difficulties that we need to solve in this paper. For the former, we separate all singular terms of , up to constant terms, by using a family of improved auxiliary functions, which depend on the parameters of Lamé system and the geometry informations of and . This essentially improves the results in [10, 11], where only estimates of are obtained. For the latter, we need to characterize the coefficients to solve the big linear system generated by (2.10), and the convergence of , ; see Proposition 3.1. Finally, we remark that one crucial step in proving the asymptotics of is to derive the asymptotic expression of , which is of independent interest, see Proposition 2.3 and Remark 2.4 below. To state it precisely, we further fix our domain and notations.
2.3. Further assumptions on domain and construction of finer auxiliary functions
Recalling the assumptions about and , there exist two points and , respectively, such that
By a translation and rotation of coordinates, if necessary, we suppose that
Now we further assume that there exits a constant , independent of , such that the portions of and near and , respectively, can be represented by
Suppose that the convexity of and is of order () near the origin,
(2.12) |
and
(2.13) |
where , , and are constants independent of . We call these inclusions -convex inclusions. For simplicity, we assume that . For , set the narrow region between and as
By the standard theory for elliptic systems, we have
(2.14) |
Therefore, in the following we only need to deal with the problems in . To this end, we denote
and introduce a scalar auxiliary function such that
(2.15) |
on , on , and . Next we use the function to generate a family of vector-valued auxiliary functions , , which is much finer than before used in [10, 11].
For , recalling that
we define such that on , and, in
(2.16) |
where , on , and .
For , noting that
we define such that, in
(2.17) |
Notice that the corrected terms depend on the Lamé parameters and , which can help us capture all singular terms of , .
2.4. Asymptotic expression of , .
We now use these auxiliary functions to obtain the asymptotics of in the narrow region , .
Theorem 2.2.
For , let be the weak solutions of (2.5). Then for sufficiently small , we have
(2.18) |
Here we would like to emphasize the importance of the introduction of . Although can be used to obtain the upper bound estimates of by derived in [27, Corollary 5.2], it as well shows to us it is possible that there remains more other singular terms in . As it turns out, the appearance of can find all the singular terms of and make be bounded. The proof is an adaption of the iteration technique with respect to the energy which was first used in [39] and further developed in [10, 11] to obtain the gradient estimates. We leave it to the Appendix.
The asymptotic expression (2.18) is an essential improvement of the estimate . More importantly, it also allows us to obtain the asymptotics of , , defined in (2.9). This kind of formulas is also the key to establish the asymptotics of , . We here give the results for . For the general cases , see Section 6 below.
Proposition 2.3.
[The asymptotics of ] Under the assumptions (2.12) and (2.13) with , we have, for sufficiently small ,
(i) for ,
(ii) for , there exist constants and , independent of , such that for ,
Remark 2.4.
Here we would point out that if we only use as the auxiliary function like in [10], it is also not possible to obtain these asymptotic formula for . The details can be found in the proof of Proposition 2.3, see Section 4 below. So the introduction of is an essential improvement. Its advantage will be also shown in the calculation of other integrals in later sections, for instance, to calculate the global effective elastic property of a periodic composite containing -convex inclusions. This relates the “Vigdergauz microstructure” in the shape optimation of fibers. For more details, see Section 7.
Remark 2.5.
As we know, in electrostatics, the condenser capacity of relative to is given by
where satisfies
Henthforth, in this sense is an “elasticity capacity” of relative to .
Theorem 2.6.
We remark that Theorems 2.2 and 2.6 also hold true for by replacing with , where is a scalar function satisfying on , on , in , and . In this case we denote the auxiliary functions by .
Throughout the paper, unless otherwise stated, we use to denote some positive constant, whose values may vary from line to line, depending only on , and an upper bound of the norms of , and , but not on . We call a constant having such dependence a universal constant.
3. Main results
In this section, we state our main theorems. First, in Subsection 3.1, for the 2-convex inclusion case we introduce a family of blow-up factors, which is a linear functional of boundary data and then give the asymptotic formulas of in Theorem 3.2 for 2D and Theorem 3.5 for 3D under the assumption that the domains satisfy some proper symmetry condition. The results for the general -convex inclusions are presented in Subsection 3.2. We find that the blow-up rates depend on the dimension and the convexity order , and the blow-up points shift away from the origin with the increasing of . Finally, in Subsection 3.3, we give an example that and are assumed to be , respectively, for , to show how the geometric parameters and influence the blow-up of . It turns out that the mean curvature of inclusions plays the role.
3.1. For the -convex inclusions
As mentioned in Subsection 2.2, is uniformly bounded with respect to . In order to derive the asymptotics of the gradient of solution with respect to the sufficiently small parameter , we consider the case when two inclusions touch each other. Let be the solution of
(3.1) |
where
and the constants , , are uniquely determined by minimizing the energy
in an admissible function space
We emphasize that we use the third line of (3.1), which implies that total flux of along the boundaries of both two inclusions is zero, to make a distinction with the forth line of (2.2). This kind of limit function for conductivity problem was also used in [24, 26, 29, 36]. We define the functionals of , for ,
(3.2) |
and denote
We will prove that are convergent to , with rates .
Here we need some symmetric assumptions on the domain and the boundary data. First, we assume that
For boundary data , we assume that in dimension two, is odd with respect to and is even with respect to ; and in dimension three, is odd for , that is, for ,
We shall use to denote those quantities satisfying , for some constant independent of . We assume that for some ,
Then
Proposition 3.1.
The proof will be given in Section 4. We would like to point out that the functionals will determine whether or not the blow-up occurs, we call them blow-up factors. For more details, see the proof of Theorem 3.2 below.
The first asymptotic expression of in dimension and for follows.
Theorem 3.2.
Remark 3.3.
If and are disks, for example,
then the result in Theorem 3.2 holds for . More generally, we can choose for some constants and .
Remark 3.4.
For this moment, to get the exact asymptotic expression of near the origin, we only need to evaluate these boundary integrals in , which can be computed by numerical method in practical problems. We would like to point out that they no longer depend on and there is no singularity near the origin. It is completely a computation problem. We leave it to interested readers.
Theorem 3.5.
Remark 3.6.
(1) If for , and are spheres satisfying , then Theorem 3.5 holds true.
3.2. For the -convex inclusions,
In the following, we shall reveal the role of the order of the relatively convexity between and , , playing in the asymptotics of . Define
and
Then
Theorem 3.8.
Remark 3.9.
We would like to remark that the pointwise upper bound estimates of in [27] imply that when , the maximum of the upper bounds obtain at and for some positive constant . Therefore, for , recalling the definition of , we have
where
Remark 3.10.
If for some , then the concentration mechanism of the stress is determined by . Thus, Theorem 3.8, combining with the upper bounds in [27], implies that the blow-up occurs at the segments , with blow-up rate when and when . Consequently, the gradient will not blow up any more and will be more and more away from the shortest segment as goes to infinity. From (2.12) with , we can see that when , the boundaries of and parallel. However, in fact, it was showed in [27] that there is no blow-up in this situation. The result of Theorem 3.8 describes this diffuse process of the stress concentration phenomenon when changes from to .
Finally, when , we define
Then
Theorem 3.11.
(i) When , we assume that and satisfies . Then if satisfies , then for sufficiently small , and ,
(3.10) |
where are specific functions, constructed in (2.17).
Remark 3.12.
The above results, together with Theorem 3.8, imply that the blow-up rate of depends on the space dimension , the convexity order , and the first term’s coefficient . Furthermore, when , we do not need the symmetric assumption about domains and the boundary data, since in this case, tends to zero as ; see the proof of Theorem 3.8 and Theorem 3.11 in Section 6. For more generalized cases, we refer readers to Example 3.13 below. Our method can also be applied to study the cases in dimensions , which is left to the interested readers.
3.3. An example
Finally, we give an example in dimension three to show the dependence of the constants in above asymptotics upon the mean curvature of the inclusions more precisely.
Example 3.13.
For , we denote the top and bottom boundaries of as
and
where and are two positive constants, may different. For , denote
and
Then the results in Theorems 3.5 and 3.11 hold true, except that
(i) if , in (3.5) is replaced by , the square root of the relative Gauss curvature of and ; if , in (3.11) becomes ;
(ii) if , the terms and in (3.11) become and , respectively.
4. Proof of Proposition 2.3 and Proposition 3.1
This section is devoted to proving Proposition 2.3 and Proposition 3.1, for , which are two main ingredients to establish our asymptotic formulas of . Actually, our argument also holds in higher dimensions with a slight modification. We first need some preliminary estimates on , .
4.1. Auxiliary estimates for
Suppose satisfies
(4.1) |
We will prove that , for each , with proper convergence rates.
Lemma 4.1.
Proof.
For , we define similarly the auxiliary functions, and , as limits of and , where satisfies on , on and
where , . A direct computation yields
(4.4) |
and
(4.5) |
Recalling the construction of in (2.16), (2.17), and (2.19), we can construct in the same way.

Notice that satisfies
By (2.14), we have
(4.7) |
For (see Figure 1), by using mean value theorem and (4.7),
(4.8) |
For , by mean value theorem again and Theorem 2.2, we have
(4.9) |
where is some constant to be determined later. Similarly, for , we have
(4.10) |
and for , we have
(4.11) |
For with , it follows from Theorem 2.2 and (4.1) that
(4.12) |
Thus, for with , by using the triangle inequality, (4.11), the mean value theorem, and (4.1), we have
(4.13) |
By taking , we get . Substituting it into (4.1), (4.11), and (4.1), and using (4.1), (4.10), and on , we obtain
(4.14) |
Applying the maximum principle for Lamé systems in , see [50], we get
4.2. Convergence of ,
We use the notation in [35] in the following.
It follows from (2.10) and the forth line of (2.2) that
(4.16) |
where is defined in (2.9), and
(4.17) |
For the first equation of (4.16), we have
and
Adding these two equations together leads to
(4.18) |
Similarly, for the second equation of (4.16), we have
(4.19) |
Then from (4.18) and (4.19), we obtain
(4.20) |
where
and satisfies
(4.21) |
Let satisfy
(4.22) |
and satisfy
(4.23) |
By the definition of in (3.1), we find that
From the third line of (3.1), we have
(4.24) |
where
(4.25) |
We next use the symmetry properties of the domain and the boundary data to prove the convergence of , Proposition 4.4 below. Before this, we first prove the following two lemmas.
Proof.
Similar to Lemma 4.2, we can get
Proof.
Proposition 4.4.
Proof.
(1) We first prove the case when . First, on one hand, using the symmetry of the domain with respect to the origin and the boundary conditions of , we have, for ,
The using the equation , one can see that
Therefore,
(4.30) |
Similarly, by taking
we have are the solutions of (2.5). Thus,
(4.31) |
On the other hand, by using the symmetry of the domain with respect to and the boundary conditions of , we find that
and
are also the solutions of (2.5), respectively. Then we have
(4.32) |
Thus,
(4.33) |
Similarly,
admits (2.5). Then
Hence,
(4.34) |
Similarly, we have
(4.35) |
Combining (4.30)–(4.35), we obtain
Therefore, we obtain
(4.36) |
By , a direct calculation which is similar to (4.2) yields
(4.37) |
By (4.16), we have
where
By Cramer’s rule and a direct calculation, we get
and similarly,
Hence,
(4.38) |
Thus, by using (4.36)–(4.38), (4.20) becomes
(4.39) |
Similar to (4.36) and (4.37), by replicating the computation used in (4.2), we get
and
Then (4.24) becomes
(4.40) |
4.3. Proof of Proposition 3.1
In this section, we aim to prove Proposition 3.1. Recalling (2.11), (3.1), and the definitions of and , (3.2), we have
(4.42) |
where , , , and are, respectively, defined by (2.6), (4.23), (4.21), and (4.22). From the proof of Proposition 4.4, for , and the estimates of (see for instance, [10, Proposition 4.2] and[11, Proposition 4.1]), we have
(4.43) |
By using (4.3), (4.43), Lemmas 4.2 and 4.3, we have
(4.44) |
here we used the fact that in , see Theorem 2.1. Therefore, Proposition 3.1 is proved.
4.4. Proof of Proposition 2.3 (The symptotics of )
We will use Theorems 2.2 and 2.6, and Lemma 4.1 to prove Proposition 2.3. Let us first prove the case in dimention two.
Proof of Proposition 2.3 for ..
(1) First consider . We divide it into three parts:
(4.45) |
In the follow we estimate , , and one by one.
Step 1. Claim: There exists a constant , independent of , such that
(4.46) |
Notice that
and
Then since , and are , we have
(4.47) |
Moreover, we obtain from (4.2) that
(4.48) |
By using the interpolation inequality, (4.47), and (4.48), we obtain
(4.49) |
Denote
Then
It follows from , , and the boundedness of and in and , respectively, that
(4.50) |
So by using (4.49) and (4.50), we have
We henceforth get (4.46).
Step 2. Proof of
(4.51) |
where
We further divide into three terms:
(4.52) |
For , we rescale into a nearly cube in unit size, and into by using the following change of variables:
After rescaling, let
By the same reason that leads to (4.47), we get
Using the interpolation inequality, we obtain
Rescaling back to , we get
(4.53) |
Similarly, we have
(4.54) |
and
(4.55) |
Now, by (4.54) and , we have
Also, by using (4.53) and (4.55), we obtain
and by (4.53), we get
Substituting the estimates above into (4.4), we obtain (4.51).
Step 3. We next further approximate by some specific functions. Note that
By using Theorem 2.2, for the second term, we have
and for the third term,
Hence,
(4.56) |
where
is a constant independent of . Coming back to (4.4), and using (4.46), (4.51), and (4.56), so far we obtain
(4.57) |
Step 4. Now we are in a position to complete the rest of the proof by direct computations. First, similar to (4.56), we obtain
(4.58) |
Second, by a direct computation, we obtain in ,
(4.59) |
For any , recalling the definition of , a direct calculation yields
(4.60) |
Substituting and into (4.4) and using (4.59), we have
Notice that for the first term,
and for the second term,
Let two right hand sides subtract,
Then we obtain
(4.61) |
(2) For , we apply the auxiliary function defined in (2.16). The rest of the proof is the same as in (1), except that we substitute and into (4.4). A direct calculation gives
(4.63) |
By using (4.63) and the same argument as that in (4.4), we obtain
So we complete the proof of Proposition 2.3 in dimension two. ∎
Proof of Proposition 2.3 for ..
The proof is similar to the above until (4.4). For , (4.4) becomes
(4.64) |
Substituting the specific function , defined by (2.17), into (4.4) and recalling that in ,
we find that (4.62) becomes
where is a constant independent of and . Similarly,
where and are constants independent of and . Hence, Proposition 2.3 in dimension three is proved. ∎
5. The proof of Theorem 3.2 and Theorem 3.5
In this section, we prove our main results, Theorem 3.2 and Theorem 3.5. Because the estimates of have been proved in Theorems 2.1, 2.2, and 2.6, recalling the proof of Proposition 4.4, we have , , and (2.10) becomes
(5.1) |
Thus, it remains to derive the asymptotics of , . To achieve this, we use (5.1) and the forth line of (2.2) to obtain
(5.2) |
where is defined in (2.9), and is defined in (3.2). Let us first consider .
5.1. Proof of Theorem 3.2
5.2. Proof of Theorem 3.5
6. The proof of Theorem 3.8 and Theorem 3.11
6.1. The proof of Theorem 3.8
Proposition 6.1.
Proposition 6.2.
We are ready to finish the proof of Theorem 3.8.
Completion of the proof of Theorem 3.8..
6.2. The proof of Theorem 3.11
(i) The proof of the case is very similar to that in the proof of Theorem 3.2 when and . We omit it here.
(ii) For , we have
Proposition 6.3.
Proposition 6.4.
Under the above assumptions, we have for sufficiently small , if ,
and ,
where and are defined in (3.2), .
Finally, we close this section by giving the proof of Theorem 3.11 (ii).
7. Application: an extended Flaherty-Keller formula
As an application of the asymptotic expressions in Propositions 2.3 and 6.1, we prove an extended Flaherty-Keller formula on the effective elastic property of a periodic composite with densely packed inclusions. We are going to follow the setting in [22, 32, 37] other than the symmetry conditions. Specifically we denote the period cell by , where and are two positive numbers. Let be a -convex domain containing the origin with boundary. Assume that is close to the horizontal boundary of and away from the vertical boundary. Let be the distance between and the the horizontal boundary of , so that becomes the distance between two adjacent inclusions, see Figure 2.

As in [37], after translation, we denote
and set
Then we obtain the effective shear modulus and the effective extensional modulus defined by
where , and
is the solution to
Note that the definition of is similar to that of in (2.9). Then by using Proposition 2.3 and Proposition 6.1, we have
Theorem 7.1.
Given . As ,
where is the curvature of near the origin, and
8. Appendix: the proof of Theorem 2.2 and Theorem 2.6
We here give the proof of Theorem 2.2 and Theorem 2.6. The key point is that is improved to be controled by . This is due to the introduction of . Then we adapt the iteration technique first used in [39] and further developed in [10, 11], to capture all singular terms of and to obtain the asymptotic formulas.
Proof of Theorems 2.2 and 2.6..
Step 1. Claim:
(8.1) |
We will prove the claim in the light of the following two cases.
Case 1. . A direct calculation yields in ,
(8.2) |
and
(8.3) |
By using (8.2), we have
(8.4) |
By using (8.3), we get
which means that the “bad” terms in (8.3) are eliminated. Combining this and (8.2), we obtain
(8.5) |
We henceforth obtain from (8) and (8) that
Similarly, we have
Furthermore, we have
Then we obtain
Case 2. . We have in ,
(8.6) |
and
(8.7) |
By (8.6), we have
(8.8) |
and
(8.9) |
By (8.7), we obtain
(8.10) |
Then (8.6) and (8.10) imply that
(8.11) |
Similarly, we can get
For the corresponding estimates for , , we note that
We thus obtain
Therefore, (8.1) is proved.
Step 2. The proof of the boundedness of the global energy. We obtain from [10] that the global energy of are bounded, . Moreover, are also bounded because of . So it suffices to prove the bundedness of , , since , .
When . We obtain from (2.17) that
and
Thus,
Therefore, the boundedness of the global energy of is established.
Step 3. Proof of
(8.12) |
where
and , , satisfying
(8.13) |
We will use the iteration scheme developed in [10, 11, 39] to prove (8.12). For , let be a smooth cutoff function satisfying if , if , if , and . Multiplying the equation in (8.13) by and applying integration by parts, Hölder’s inequality, and Cauchy inequality, we get
(8.14) |
On one hand, we obtain from Hölder’s inequality that
(8.15) |
On the other hand, we estimate the second term on the right hand side of (8.14) according to the following two cases.
Case 1. . By using (8.1), we have for ,
(8.16) |
This is an improvement of [10, (3.32),(3.35)]. Denote
Then substituting (8.15) and (8.16) into (8.14), we have
(8.17) |
where is a universal canstant.
Let and . Then by (8.17) with and , we have
After iterations, using the global boundedness of , we have
Acknowledgements. H.G. Li was partially supported by NSFC (11631002, 11971061) and BJNSF (1202013). The authors are grateful to professor Yanyan Li for his encouragement and very helpful suggestions.
References
- [1] H. Ammari; E. Bonnetier; F. Triki; M. Vogelius. Elliptic estimates in composite media with smooth inclusions: an integral equation approach. Ann. Sci. éc. Norm. Supér. (4) 48 (2015), no. 2, 453-495.
- [2] H. Ammari; G. Ciraolo; H. Kang; H. Lee; K. Yun. Spectral analysis of the Neumann-Poincaré operator and characterization of the stress concentration in antiplane elasticity. Arch. Ration. Mech. Anal. 208 (2013), 275-304.
- [3] H. Ammari; H. Dassios; H. Kang; M. Lim. Estimates for the electric field in the presence of adjacent perfectly conducting spheres. Quat. Appl. Math. 65 (2007), 339-355.
- [4] S. Agmon; A. Douglis; L. Nirenberg. Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I. Comm. Pure Appl. Math. 12 (1959), 623-727.
- [5] S. Agmon; A. Douglis; L. Nirenberg. Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. II. Comm. Pure Appl. Math. 17 (1964), 35-92.
- [6] H. Ammari; H. Kang; M. Lim. Gradient estimates to the conductivity problem. Math. Ann. 332 (2005), 277-286.
- [7] H. Ammari; H. Kang; H. Lee; J. Lee; M. Lim. Optimal estimates for the electrical field in two dimensions. J. Math. Pures Appl. 88 (2007), 307-324.
- [8] I. Babuška; B. Andersson; P. Smith; K. Levin. Damage analysis of fiber composites. I. Statistical analysis on fiber scale. Comput. Methods Appl. Mech. Engrg. 172 (1999), 27-77.
- [9] J.G. Bao; H.J. Ju; H.G. Li. Optimal boundary gradient estimates for Lamé systems with partially infinite coefficients. Adv. Math. 314 (2017), 583-629.
- [10] J.G. Bao; H.G. Li; Y.Y. Li. Gradient estimates for solutions of the Lamé system with partially infinite coefficients. Arch. Ration. Mech. Anal. 215 (2015), no. 1, 307-351.
- [11] J.G. Bao; H.G. Li; Y.Y. Li. Gradient estimates for solutions of the Lamé system with partially infinite coefficients in dimensions greater than two. Adv. Math. 305 (2017), 298-338.
- [12] E.S. Bao; Y.Y. Li; B. Yin. Gradient estimates for the perfect conductivity problem. Arch. Ration. Mech. Anal. 193 (2009), 195-226.
- [13] E. Bao; Y.Y. Li; B. Yin. Gradient estimates for the perfect and insulated conductivity problems with multiple inclusions. Comm. Partial Differential Equations 35 (2010), 1982-2006.
- [14] E. Bonnetier; F. Triki. On the spectrum of the Poincaré variational problem for two close-to-touching inclusions in . Arch. Ration. Mech. Anal. 209 (2013), no. 2, 541-567.
- [15] E. Bonnetier; M. Vogelius. An elliptic regularity result for a composite medium with “touching” fibers of circular cross-section. SIAM J. Math. Anal. 31 (2000) 651-677.
- [16] B. Budiansky; G.F. Carrier. High shear stresses in stiff fiber composites. J. App. Mech. 51 (1984), 733-735.
- [17] H.W. Cheng. On the method of images for systems of closely spaced conducting spheres. SIAM J. Appl. Math. 61 (2000), no. 4, 1324-1337.
- [18] H.W. Cheng; L. Greengard. A method of images for the evaluation of electrostatic fileds in systems of closely spaced conducting cylinders. SIAM J. Appl. Math. 58 (2006), no. 1, 122-141.
- [19] H.J. Dong. Gradient estimates for parabolic and elliptic systems from linear laminates. Arch. Ration. Mech. Anal. 205 (2012), no. 1, 119-149.
- [20] H.J. Dong; H.G. Li. Optimal Estimates for the Conductivity Problem by Green’s Function Method. Arch. Ration. Mech. Anal. 231 (2019), no. 3, 1427-1453.
- [21] H.J. Dong; L.J. Xu. Gradient estimates for divergence form elliptic systems arising from composite material. SIAM J. Math. Anal. 51 (2019), no. 3, 2444-2478.
- [22] J.E. Flaherty; J.B. Keller. Elastic behavior of composite media. Comm. Pure. Appl. Math. 26 (1973), 565-580.
- [23] Y. Grabovsky; R.V. Kohn. Microstructure minimizing the energy of a two phase elastic composite in two space dimensions. II: the Vigdergauz microstructure. J. Mech. Phys. Solids. 43 (1995), 949-972.
- [24] Y. Gorb. Singular behavior of electric field of high-contrast concentrated composites. Multiscale Model. Simul. 13 (2015), no. 4, 1312-1326.
- [25] Y. Gorb and L. Berlyand. Asymptotics of the effective conductivity of composites with closely spaced inclusions of optimal shape. Quart. J. Mech. Appl. Math., 58 (2005), pp. 83-106.
- [26] Y. Gorb and A. Novikov. Blow-up of solutions to a -Laplace equation. Multiscale Model. Simul., 10 (2012), pp. 727-743.
- [27] Y.Y. Hou; H.G. Li. The convexity of inclusions and gradient’s concentration for the Lamé systems with partially infinite coefficients. arXiv: 1802.01412v1.
- [28] H.J. Ju; H.G. Li; L.J. Xu. Estimates for elliptic systems in a narrow region arising from composite materials. Quart. Appl. Math. 77 (2019), 177-199.
- [29] H. Kang; H. Lee; K. Yun. Optimal estimates and asymptotics for the stress concentration between closely located stiff inclusions. Math. Ann. 363 (2015), no. 3-4, 1281-1306.
- [30] H. Kang; M. Lim; K. Yun. Asymptotics and computation of the solution to the conductivity equation in the presence of adjacent inclusions with extreme conductivities. J. Math. Pures Appl. (9) 99 (2013), 234-249.
- [31] H. Kang; M. Lim; K. Yun. Characterization of the electric field concentration between two adjacent spherical perfect conductors. SIAM J. Appl. Math. 74 (2014), 125-146.
- [32] H. Kang; S. Yu. A proof of the Flaherty-Keller formula on the effective property of densely packed elastic composites. Calc. Var. Partial Differential Equations. 59 (2020), no. 1, 13 pp.
- [33] J.B. Keller. Conductivity of a medium containing a dense array of perfectly conducting spheres or cylinders or nonconducting cylinders. J. Appl. Phys. 34 (1963), 991-993.
- [34] J.B. Keller. Stresses in narrow regions. Trans. ASME J. Appl. Mech. 60 (1993), 1054-1056.
- [35] H.G. Li. Lower bounds of gradient’s blow-up for the Lamé system with partially infinite coefficients. to appear in J. Math. Pures Appl. (2020). DOI: https://doi.org/10.1016/j.matpur.2020.09.007.
- [36] H.G. Li. Asymptotics for the electric field concentration in the perfect conductivity problem. SIAM J. Math. Anal. 52 (2020), no.4, 3350-3375.
- [37] H.G. Li; Y. Li. An extension of flaherty-keller formula for density packed m-convex inclusion. arXiv: 1912.13261v1.
- [38] H.G. Li; Y.Y. Li. Gradient estimates for parabolic systems from composite material. Sci. China Math. 60 (2017), no. 11, 2011-2052.
- [39] H.G. Li; Y.Y. Li; E.S. Bao; B. Yin. Derivative estimates of solutions of elliptic systems in narrow regions. Quart. Appl. Math. 72 (2014), no. 3, 589-596.
- [40] H.G. Li; Y.Y. Li; Z.L. Yang. Asymptotics of the gradient of solutions to the perfect conductivity problem. Multiscale Model. Simul. 17 (2019), no. 3, 899-925.
- [41] Y.Y. Li; L. Nirenberg. Estimates for elliptic system from composite material. Comm. Pure Appl. Math. 56 (2003), 892-925.
- [42] H.G. Li; F. Wang; L.J. Xu. Characterization of Electric Fields between two Spherical Perfect Conductors with general radii in 3D. J. Differential Equations (2019), 267(11), 6644-6690.
- [43] Y.Y. Li; M. Vogelius. Gradient stimates for solutions to divergence form elliptic equations with discontinuous coefficients. Arch. Rational Mech. Anal. 153 (2000), 91-151.
- [44] H.G. Li; L.J. Xu. Optimal estimates for the perfect conductivity problem with inclusions close to the boundary. SIAM J. Math. Anal. 49 (2017), no. 4, 3125-3142.
- [45] H.G. Li; Z.W. Zhao. Boundary blow-up analysis of gradient estimates for Lamé systems in the presence of -convex hard inclusions. SIAM J. Math. Anal. 52 (2020), no. 4, 3777-3817.
- [46] M. Lim; S. Yu. Asymptotics of the solution to the conductivity equation in the presence of adjacent circular inclusions with finite conductivities. J. Math. Anal. Appl. 421 (2015), 131-156.
- [47] M. Lim; S. Yu. Stress concentration for two nearly touching circular holes. arXiv: 1705.10400v1. (2017)
- [48] M. Lim; K. Yun. Blow-up of electric fields between closely spaced spherical perfect conductors. Comm. Partial Differential Equations, 34 (2009), 1287-1315.
- [49] X. Markenscoff, Stress amplification in vanishingly small geometries. Computational Mechanics 19 (1996), 77-83.
- [50] V.G. Maz’ya; A.B. Movchan; M.J. Nieves. Uniform asymptotic formular for Green’s tensors in elastic singularly perturbed domains. Asym. Anal. 52 (2007), 173-206.
- [51] R. McPhedran; L. Poladian; G.W. Milton. Asymptotic studies of closely spaced, highly conducting cylinders. Proc. Roy. Soc. London Ser. A, 415, (1988), 185-196.
- [52] R. Meredith, C. Tobias. Resistance to potential flow through a cubical array of spheres. J. Applied Physics, 31 (1960), no. 7, 1270-1273.
- [53] L. Rayleigh. On the influence of obstacles arranged in rectangular order upon the properties of a medium. Phil. Mag. 34 (1892), 481-502.
- [54] S.B. Vigdergauz. Integral equation of the inverse problem of the plane theory of elasticity. PMM. 40 (1976), 518-521.
- [55] K. Yun. Estimates for electric fields blown up between closely adjacent conductors with arbitrary shape. SIAM J. Appl. Math. 67 (2007), 714-730.
- [56] K. Yun. Optimal bound on high stresses occurring between stiff fibers with arbitrary shaped cross-sections. J. Math. Anal. Appl. 350 (2009), 306-312.