This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Asymptotics of the stress concentration in high-contrast elastic composites

Haigang Li School of Mathematical Sciences, Beijing Normal University, Laboratory of MathematiCs and Complex Systems, Ministry of Education, Beijing 100875, China. [email protected]  and  Longjuan Xu Department of Mathematics, National University of Singapore, 10 Lower Kent Ridge Road, Singapore 119076. [email protected]
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 ε\varepsilon-independent bound of the gradient was obtained by Li and Nirenberg [41], where ε\varepsilon represents the distance between the interfacial surfaces. However, in high-contrast composites, when ε\varepsilon 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 ε\varepsilon 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, ε\varepsilon) 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

λ,μu=μΔu+(λ+μ)(u)\mathcal{L}_{\lambda,\mu}u=\mu\Delta u+(\lambda+\mu)\nabla(\nabla\cdot u)

in unidirectional composites to numerically analyze the residual stresses and stresses due to mechanical loads, where u=(u1,u2,u3)Tu=(u^{1},u^{2},u^{3})^{\mathrm{T}} 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

(ak(x)u)=0inD,\nabla(a_{k}(x)\nabla u)=0\quad\mbox{in}~{}D,

where the scalar uu is the out of plane displacement, DD 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 0<ak(x)<0<a_{k}(x)<\infty take two different constant values, after rescaling,

ak(x)\displaystyle a_{k}(x) =kfor xinside the cross-sections of the fibers,\displaystyle=k\quad\mbox{for~{}}x~{}\mbox{inside~{}the~{}cross-sections~{}of~{}the~{}fibers},
ak(x)\displaystyle a_{k}(x) =1elsewhere in D.\displaystyle=1\quad\mbox{elsewhere~{}in~{}}D.

Despite the discontinuity of the coefficient along the interfaces, they proved that any variational solution uu is in W1,W^{1,\infty}, which actually improves a classical regularity result due to De Giorgi and Nash, which asserts that H1H^{1} 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 |u||\nabla u| regardless of ε\varepsilon in any dimension d2d\geq 2. 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 ε\varepsilon tends to zero. Dong and Xu [21] further showed that a W1,1W^{1,1} weak solution is Lipschitz and piecewise C1C^{1}. 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 kk and ε\varepsilon 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 ε\varepsilon, 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, kk degenerates to \infty) 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 |u||\nabla u| is ε1/2\varepsilon^{-1/2} in dimension two and |εlogε|1|\varepsilon\log\varepsilon|^{-1} 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 u\nabla u, 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 |u||\nabla u|, and showed that |u||\nabla u| may blow up on the shortest line between two adjacent inclusions, see [9, 10, 11, 35]. By using the polynomial function xd=|x|mx_{d}=|x^{\prime}|^{m}, m2m\geq 2, 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 |u||\nabla u| and the convexity order mm. 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 u\nabla u 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 d=2d=2 and 33, and for all m2m\geq 2. For d4d\geq 4, 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 dd and the convexity order mm 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 m>2m>2, 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 mm-convex inclusion.

The asymptotic formulas obtained above clearly reflect the local property of u\nabla u. 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 (m=2)(m=2) 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 mm-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 u\nabla u 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 a11ααa_{11}^{\alpha\alpha}, α=1,,d\alpha=1,\cdots,d. 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 mm-convex inclusions. Finally we give an example to show the dependence on the precise geometry feature of D1D_{1} and D2D_{2}. 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 u\nabla u 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 x=(x,xd)x=(x^{\prime},x_{d}) to denote a point in d\mathbb{R}^{d}, where x=(x1,,xd1)x^{\prime}=(x_{1},\cdots,x_{d-1}) and d=2,3d=2,3. Let DD be a bounded open set in d\mathbb{R}^{d} with C2C^{2} boundary. D1D_{1} and D2D_{2} are two disjoint convex open subsets in DD with C2,γC^{2,\gamma} (0<γ<1)(0<\gamma<1) boundaries, ε\varepsilon-apart, and far away from D\partial D. That is,

D¯1,D¯2D,ε:=dist(D1,D2)>0,dist(D1D2,D)>κ0>0,\begin{split}\overline{D}_{1},\overline{D}_{2}\subset D,\quad\varepsilon:=\mbox{dist}(D_{1},D_{2})>0,\quad\mbox{dist}(D_{1}\cup D_{2},\partial D)>\kappa_{0}>0,\end{split}

where κ0\kappa_{0} is a constant independent of ε\varepsilon. We also assume that the C2,γC^{2,\gamma} norms of D1\partial{D}_{1}, D2\partial{D}_{2}, and D\partial{D} are bounded by some positive constant independent of ε\varepsilon. Set

Ω:=DD1D2¯.\Omega:=D\setminus\overline{D_{1}\cup D_{2}}.

Assume that Ω\Omega and D1D2D_{1}\cup D_{2} are occupied, respectively, by two different isotropic and homogeneous materials with different Lamé constants (λ,μ)(\lambda,\mu) and (λ1,μ1)(\lambda_{1},\mu_{1}). Then the elasticity tensors for the background and the inclusion can be written, respectively, as 0\mathbb{C}^{0} and 1\mathbb{C}^{1}, with

Cijkl0=λδijδkl+μ(δikδjl+δilδjk),C_{ijkl}^{0}=\lambda\delta_{ij}\delta_{kl}+\mu(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}),

and

Cijkl1=λ1δijδkl+μ1(δikδjl+δilδjk),C_{ijkl}^{1}=\lambda_{1}\delta_{ij}\delta_{kl}+\mu_{1}(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}),

where i,j,k,l=1,2,,di,j,k,l=1,2,\cdots,d and δij\delta_{ij} is the kronecker symbol: δij=0\delta_{ij}=0 for iji\neq j, δij=1\delta_{ij}=1 for i=ji=j. Let u=(u1,u2,,ud)T:Ddu=(u^{1},u^{2},\cdots,u^{d})^{\mathrm{T}}:D\rightarrow\mathbb{R}^{d} denote the displacement field. For a given vector-valued function φ=(φ1,φ2,,φd)T\varphi=(\varphi^{1},\varphi^{2},\cdots,\varphi^{d})^{\mathrm{T}}, we consider the following Dirichlet problem for the Lamé system with pieceiwise constant coefficients:

{((χΩ0+χD1D21)e(u))=0,inD,u=φ,onD,\displaystyle\begin{cases}\nabla\cdot\left((\chi_{\Omega}\mathbb{C}^{0}+\chi_{D_{1}\cup{D}_{2}}\mathbb{C}^{1})e(u)\right)=0,&\hbox{in}~{}~{}D,\\ u=\varphi,&\hbox{on}~{}~{}\partial{D},\end{cases} (2.1)

where χΩ\chi_{\Omega} is the characteristic function of Ωd\Omega\subset\mathbb{R}^{d}, and

e(u)=12(u+(u)T)e(u)=\frac{1}{2}(\nabla u+(\nabla u)^{\mathrm{T}})

is the strain tensor.

Assume that the standard ellipticity condition holds for (2.1), that is,

μ>0,dλ+2μ>0,μ1>0,dλ1+2μ1>0.\displaystyle\mu>0,\quad d\lambda+2\mu>0,\quad\mu_{1}>0,\quad d\lambda_{1}+2\mu_{1}>0.

For φC1,γ(D;d)\varphi\in C^{1,\gamma}(\partial D;\mathbb{R}^{d}), it is well known that there exists a unique solution uH1(D;d)u\in H^{1}(D;\mathbb{R}^{d}) to the Dirichlet problem (2.1), which is also the minimizer of the energy functional

J1[u]:=12Ω((χΩ0+χD1D21)e(u),e(u))𝑑xJ_{1}[u]:=\frac{1}{2}\int_{\Omega}\left((\chi_{\Omega}\mathbb{C}^{0}+\chi_{D_{1}\cup{D}_{2}}\mathbb{C}^{1})e(u),e(u)\right)dx

on

Hφ1(D;d):={uH1(D;d)|uφH01(D;d)}.\displaystyle H^{1}_{\varphi}(D;\mathbb{R}^{d}):=\left\{u\in H^{1}(D;\mathbb{R}^{d})~{}\big{|}~{}u-\varphi\in H^{1}_{0}(D;\mathbb{R}^{d})\right\}.

As mentioned previously, Li and Nirenberg [41] proved that u\nabla u is uniformly bounded with respect to ε\varepsilon. But, in high-contrast composite media, the concentration of u\nabla u is a very usual phenomenon when the distance ε\varepsilon is sufficiently small. In order to investigate the role of ε\varepsilon in such concentration phenomenon, let us assume that the Lamé constant in D1D2D_{1}\cup D_{2} degenerates to infinite and consider this extreme case. To this end, we first introduce the linear space of rigid displacement in d\mathbb{R}^{d}:

Ψ:={ψC1(d;d)|ψ+(ψ)T=0},\Psi:=\{\psi\in C^{1}(\mathbb{R}^{d};\mathbb{R}^{d})\ |\ \nabla\psi+(\nabla\psi)^{\mathrm{T}}=0\},

with a basis {ψα|α=1,2,,d(d+1)2}\left\{\psi_{\alpha}~{}|~{}\alpha=1,2,\cdots,\frac{d(d+1)}{2}\right\}, namely,

{ei,xjekxkej|1id,1j<kd},\left\{~{}e_{i},~{}x_{j}e_{k}-x_{k}e_{j}~{}\big{|}~{}1\leq\,i\leq\,d,~{}1\leq\,j<k\leq\,d~{}\right\},

where e1,,ede_{1},\cdots,e_{d} denote the standard basis of d\mathbb{R}^{d}. For fixed λ\lambda and μ\mu, denoting uλ1,μ1u_{\lambda_{1},\mu_{1}} as the solution of (2.1), then we have [10]

uλ1,μ1uinH1(D;d),asmin{μ1,dλ1+2μ1},\displaystyle u_{\lambda_{1},\mu_{1}}\rightarrow u\quad\hbox{in}\ H^{1}(D;\mathbb{R}^{d}),\quad\hbox{as}\ \min\{\mu_{1},d\lambda_{1}+2\mu_{1}\}\rightarrow\infty,

where uu is the unique H1(D;d)H^{1}(D;\mathbb{R}^{d}) solution of

{λ,μu:=(0e(u))=0,inΩ,u|+=u|,onDi,i=1,2,e(u)=0,inDi,i=1,2,Diuν|+ψα=0,i=1,2,α=1,2,,d(d+1)2,u=φ,onD,\displaystyle\begin{cases}\mathcal{L}_{\lambda,\mu}u:=\nabla\cdot(\mathbb{C}^{0}e(u))=0,\quad&\hbox{in}\ \Omega,\\ u|_{+}=u|_{-},&\hbox{on}\ \partial{D}_{i},i=1,2,\\ e(u)=0,&\hbox{in}~{}~{}D_{i},i=1,2,\\ \int_{\partial{D}_{i}}\frac{\partial u}{\partial\nu}\Big{|}_{+}\cdot\psi_{\alpha}=0,&i=1,2,\alpha=1,2,\cdots,\frac{d(d+1)}{2},\\ u=\varphi,&\hbox{on}\ \partial{D},\end{cases} (2.2)

where

uν|+\displaystyle\frac{\partial u}{\partial\nu}\Big{|}_{+} :=(0e(u))n=λ(u)n+μ(u+(u)T)n,\displaystyle:=(\mathbb{C}^{0}e(u))\vec{n}=\lambda(\nabla\cdot u)\vec{n}+\mu(\nabla u+(\nabla u)^{\mathrm{T}})\vec{n},

and n\vec{n} is the unit outer normal of DiD_{i}, i=1,2i=1,2. Here and throughout this paper the subscript ±\pm 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 φC0(D;d)\varphi\in C^{0}(\partial D;\mathbb{R}^{d}) replaced by φC1,γ(D;d)\varphi\in C^{1,\gamma}(\partial D;\mathbb{R}^{d}). Indeed, it follows from the maximum principle [50] that uL(D)CφC0(D)\|u\|_{L^{\infty}(D)}\leq C\|\varphi\|_{C^{0}(\partial D)}. Taking a slightly small domain D~D\widetilde{D}\subset\subset D, then in view of the interior derivative estimates for Lamé system, we find that φ~:=u|D~\tilde{\varphi}:=u\big{|}_{\partial\widetilde{D}} satisfies

φ~C1,γ(D~)CuL(D)CφC0(D).\|\tilde{\varphi}\|_{C^{1,\gamma}(\partial\widetilde{D})}\leq C\|u\|_{L^{\infty}(D)}\leq C\|\varphi\|_{C^{0}(\partial D)}.

Without loss of generality, we assume that φC0(D)=1\|\varphi\|_{C^{0}(\partial D)}=1 by considering u/φC0(D)u/\|\varphi\|_{C^{0}(\partial D)} if φC0(D)>0\|\varphi\|_{C^{0}(\partial D)}>0. If φ|D=0\varphi\big{|}_{\partial D}=0, then u0u\equiv 0.

2.2. Main difficulties and decomposition

We first point out that problem (2.2) has free boundary value feature. Although e(u)=0e(u)=0 implies uu in D¯i\overline{D}_{i} is linear combination of ψα\psi_{\alpha},

u=α=1d(d+1)/2CiαψαinD¯i,u=\sum_{\alpha=1}^{d(d+1)/2}C_{i}^{\alpha}\psi_{\alpha}\quad\text{in}~{}~{}\overline{D}_{i}, (2.3)

these d(d+1)d(d+1) constants CiαC_{i}^{\alpha} are free, which will be uniquely determined by uu. 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 uu across Di\partial D_{i}, we can decompose the solution of (2.2),

u(x)=i=12α=1d(d+1)/2Ciαviα(x)+v0(x),xΩ,u(x)=\sum_{i=1}^{2}\sum_{\alpha=1}^{d(d+1)/2}C_{i}^{\alpha}v_{i}^{\alpha}(x)+v_{0}(x),\quad x\in\,\Omega, (2.4)

where viα,v0C2(Ω;d)v_{i}^{\alpha},v_{0}\in{C}^{2}(\Omega;\mathbb{R}^{d}), respectively, satisfying

{λ,μviα=0,inΩ,viα=ψα,onDi,viα=0,onDjD,ji,i=1,2,α=1,,d(d+1)/2,\begin{cases}\mathcal{L}_{\lambda,\mu}v_{i}^{\alpha}=0,&\mathrm{in}~{}\Omega,\\ v_{i}^{\alpha}=\psi_{\alpha},&\mathrm{on}~{}\partial{D}_{i},\\ v_{i}^{\alpha}=0,&\mathrm{on}~{}\partial{D_{j}}\cup\partial{D},~{}j\neq i,\end{cases}\quad i=1,2,~{}\alpha=1,\cdots,d(d+1)/2, (2.5)

and

{λ,μv0=0,inΩ,v0=0,onD1D2,v0=φ,onD.\begin{cases}\mathcal{L}_{\lambda,\mu}v_{0}=0,&\mathrm{in}~{}\Omega,\\ v_{0}=0,&\mathrm{on}~{}\partial{D}_{1}\cup\partial{D_{2}},\\ v_{0}=\varphi,&\mathrm{on}~{}\partial{D}.\end{cases} (2.6)

So

u(x)=i=12α=1d(d+1)/2Ciαviα(x)+v0(x),xΩ.\nabla u(x)=\sum_{i=1}^{2}\sum_{\alpha=1}^{d(d+1)/2}C_{i}^{\alpha}\nabla v_{i}^{\alpha}(x)+\nabla v_{0}(x),\quad x\in\,\Omega. (2.7)

To investigate the symptotic behavior of u\nabla u, we need both the asymptotic formulas of viα\nabla v_{i}^{\alpha} and the exact value of CiαC_{i}^{\alpha}. To solve CiαC_{i}^{\alpha}, from the fourth line in (2.2) and the decomposition (2.4), we have the following linear system of these free constants CiαC_{i}^{\alpha},

i=12α=1d(d+1)2CiαDjviαν|+ψβ+Djv0ν|+ψβ=0,\sum_{i=1}^{2}\sum\limits_{\alpha=1}^{\frac{d(d+1)}{2}}C_{i}^{\alpha}\int_{\partial D_{j}}\frac{\partial v_{i}^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}+\int_{\partial D_{j}}\frac{\partial v_{0}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}=0, (2.8)

where j=1, 2j=1,\,2, β=1,,d(d+1)2\beta=1,\,\cdots,\frac{d(d+1)}{2}. But these coefficients are all boundary integrals. By integration by parts,

aijαβ:=Djviαν|+ψβ=Ω(0e(viα),e(vjβ))𝑑x.\displaystyle a_{ij}^{\alpha\beta}:=-\int_{\partial{D}_{j}}\frac{\partial{v}_{i}^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}=\int_{\Omega}\big{(}\mathbb{C}^{0}e(v_{i}^{\alpha}),e(v_{j}^{\beta})\big{)}\ dx. (2.9)

Therefore, in order to solve CiαC_{i}^{\alpha} from (2.8), we have to calculate the energy integral on the right hand side of (2.9). This in turn needs to estimate viα\nabla v_{i}^{\alpha}. In fact, even if we can have the asymptotic formulas of viα\nabla v_{i}^{\alpha}, see Theorem 2.2 and Theorem 2.6 below, it is still hard to solve every CiαC_{i}^{\alpha}. It seems to be a mission impossible. To avoid this difficulty, we rewrite (2.7) as

u=α=1d(d+1)/2(C1αC2α)v1α+ub,inΩ,\nabla{u}=\sum_{\alpha=1}^{d(d+1)/2}\left(C_{1}^{\alpha}-C_{2}^{\alpha}\right)\nabla{v}_{1}^{\alpha}+\nabla u_{b},\quad\mbox{in}~{}\Omega, (2.10)

where

ub:=α=1d(d+1)/2C2α(v1α+v2α)+v0.u_{b}:=\sum_{\alpha=1}^{d(d+1)/2}C_{2}^{\alpha}(v_{1}^{\alpha}+v_{2}^{\alpha})+v_{0}. (2.11)

This is because the following boundedness estimates for |(v1α+v2α)||\nabla(v_{1}^{\alpha}+v_{2}^{\alpha})| and |v0||\nabla v_{0}|, together with the boundedness of C2αC_{2}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2, makes ub\nabla u_{b} is a “good” term, which has no singularity in the narrow region.

Theorem 2.1 ([28, 39] ).

Let viαv_{i}^{\alpha} and v0v_{0} be defined in (2.5) and (2.6), respectively, i=1,2i=1,2. Then we have

(v1α+v2α)L(Ω)C,α=1,,d(d+1)/2,andv0L(Ω)C.\|\nabla(v_{1}^{\alpha}+v_{2}^{\alpha})\|_{L^{\infty}(\Omega)}\leq C,\quad\alpha=1,\cdots,d(d+1)/2,\quad\mbox{and}\quad\|\nabla v_{0}\|_{L^{\infty}(\Omega)}\leq C.

Thus, we reduce the establishment of the asymptotics of u\nabla u to that of the asymptotics of v1α\nabla v_{1}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2, and to solving C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2. These are two main difficulties that we need to solve in this paper. For the former, we separate all singular terms of v1α\nabla v_{1}^{\alpha}, 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 D1\partial D_{1} and D2\partial D_{2}. This essentially improves the results in [10, 11], where only estimates of |viα||\nabla v_{i}^{\alpha}| are obtained. For the latter, we need to characterize the coefficients aijαβa_{ij}^{\alpha\beta} to solve the big linear system generated by (2.10), and the convergence of D1ubν|+ψβ\int_{\partial D_{1}}\frac{\partial u_{b}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}, β=1,,d(d+1)/2\beta=1,\cdots,d(d+1)/2; see Proposition 3.1. Finally, we remark that one crucial step in proving the asymptotics of C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha} is to derive the asymptotic expression of a11ααa_{11}^{\alpha\alpha}, 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 D1D_{1} and D2D_{2}, there exist two points P1D1P_{1}\in\partial D_{1} and P2D2P_{2}\in\partial D_{2}, respectively, such that

dist(P1,P2)=dist(D1,D2)=ε.\mbox{dist}(P_{1},P_{2})=\mbox{dist}(\partial D_{1},\partial D_{2})=\varepsilon.

By a translation and rotation of coordinates, if necessary, we suppose that

P1=(0,ε2)D1,P2=(0,ε2)D2.P_{1}=(0^{\prime},\frac{\varepsilon}{2})\in\partial D_{1},\quad P_{2}=(0^{\prime},-\frac{\varepsilon}{2})\in\partial D_{2}.

Now we further assume that there exits a constant RR, independent of ε\varepsilon, such that the portions of D1\partial D_{1} and D2\partial D_{2} near P1P_{1} and P2P_{2}, respectively, can be represented by

xd=ε2+h1(x)andxd=ε2+h2(x),for|x|<2R.\displaystyle x_{d}=\frac{\varepsilon}{2}+h_{1}(x^{\prime})\quad\mbox{and}\quad x_{d}=-\frac{\varepsilon}{2}+h_{2}(x^{\prime}),\quad\mbox{for}~{}|x^{\prime}|<2R.

Suppose that the convexity of D1\partial D_{1} and D2\partial D_{2} is of order m2m\geq 2 (m+m\in\mathbb{N}^{+}) near the origin,

hi(x)={(1)i+1κi|x|2+O(|x|2+γ),m=2,(1)i+1κi|x|m+O(|x|m+1),m3,fori=1,2,\displaystyle h_{i}(x^{\prime})=\begin{cases}(-1)^{i+1}\kappa_{i}|x^{\prime}|^{2}+O(|x^{\prime}|^{2+\gamma}),&\quad m=2,\\ (-1)^{i+1}\kappa_{i}|x^{\prime}|^{m}+O(|x^{\prime}|^{m+1}),&\quad m\geq 3,\end{cases}\quad\mbox{for}~{}i=1,2, (2.12)

and

h1C2,γ(B2R)+h2C2,γ(B2R)C,\|h_{1}\|_{C^{2,\gamma}(B^{\prime}_{2R})}+\|h_{2}\|_{C^{2,\gamma}(B^{\prime}_{2R})}\leq C, (2.13)

where κ1\kappa_{1}, κ2\kappa_{2}, and CC are constants independent of ε\varepsilon. We call these inclusions mm-convex inclusions. For simplicity, we assume that κ1=κ2=κ2\kappa_{1}=\kappa_{2}=\dfrac{\kappa}{2}. For 0<r 2R0<r\leq\,2R, set the narrow region between D1\partial{D}_{1} and D2\partial{D}_{2} as

Ωr:={(x,xd)d|ε2+h2(x)<xd<ε2+h1(x),|x|<r}.\Omega_{r}:=\left\{(x^{\prime},x_{d})\in\mathbb{R}^{d}~{}\big{|}~{}-\frac{\varepsilon}{2}+h_{2}(x^{\prime})<x_{d}<\frac{\varepsilon}{2}+h_{1}(x^{\prime}),~{}|x^{\prime}|<r\right\}.

By the standard theory for elliptic systems, we have

viα(x)L(ΩΩR)C,α=1,,d(d+1)2,i=1,2.\|\nabla v_{i}^{\alpha}(x)\|_{L^{\infty}(\Omega\setminus\Omega_{R})}\leq\,C,\quad\alpha=1,\cdots,\frac{d(d+1)}{2},~{}i=1,2. (2.14)

Therefore, in the following we only need to deal with the problems in ΩR\Omega_{R}. To this end, we denote

δ(x):=ε+h1(x)h2(x),\delta(x^{\prime}):=\varepsilon+h_{1}(x^{\prime})-h_{2}(x^{\prime}),

and introduce a scalar auxiliary function u¯C2(d)\bar{u}\in C^{2}(\mathbb{R}^{d}) such that

u¯(x)=xd+ε2h2(x)δ(x),inΩ2R,\bar{u}(x)=\frac{x_{d}+\frac{\varepsilon}{2}-h_{2}(x^{\prime})}{\delta(x^{\prime})},\quad\hbox{in}\ \Omega_{2R}, (2.15)

u¯=1\bar{u}=1 on D1\partial{D}_{1}, u¯=0\bar{u}=0 on D2D\partial{D}_{2}\cup\partial{D}, and u¯C2(ΩΩR)C\|\bar{u}\|_{C^{2}(\Omega\setminus\Omega_{R})}\leq\,C. Next we use the function u¯\bar{u} to generate a family of vector-valued auxiliary functions u1αu_{1}^{\alpha}, α=1,,d\alpha=1,\cdots,d, which is much finer than before used in [10, 11].

For d=2d=2, recalling that

ψ1=(10)andψ2=(01),\psi_{1}=\begin{pmatrix}1\\ 0\end{pmatrix}\quad\mbox{and}\quad\psi_{2}=\begin{pmatrix}0\\ 1\end{pmatrix},

we define u1αC2(Ω)u_{1}^{\alpha}\in C^{2}(\Omega) such that u1α=v1αu_{1}^{\alpha}=v_{1}^{\alpha} on Ω\partial\Omega, and, in Ω2R\Omega_{2R}

u11:=u¯11+u~11:=u¯ψ1+λ+μλ+2μf(u¯)δ(x1)ψ2,u12:=u¯12+u~12:=u¯ψ2+λ+μμf(u¯)δ(x1)ψ1,\begin{split}u_{1}^{1}&:=\bar{u}_{1}^{1}+\tilde{u}_{1}^{1}:=\bar{u}\psi_{1}+\frac{\lambda+\mu}{\lambda+2\mu}f(\bar{u})\,\delta^{\prime}(x_{1})\,\psi_{2},\\ u_{1}^{2}&:=\bar{u}_{1}^{2}+\tilde{u}_{1}^{2}:=\bar{u}\psi_{2}+\frac{\lambda+\mu}{\mu}f(\bar{u})\,\delta^{\prime}(x_{1})\,\psi_{1},\end{split} (2.16)

where f(u¯):=12(u¯12)218f(\bar{u}):=\dfrac{1}{2}\Big{(}\bar{u}-\frac{1}{2}\Big{)}^{2}-\dfrac{1}{8}, f(u¯)=0f(\bar{u})=0 on (D1D2){|x1|<R}(\partial D_{1}\cup\partial D_{2})\cap\{|x_{1}|<R\}, and u1αC2(ΩΩR)C\|u_{1}^{\alpha}\|_{C^{2}(\Omega\setminus\Omega_{R})}\leq C.

For d=3d=3, noting that

ψ1=(100),ψ2=(010),andψ3=(001),\psi_{1}=\begin{pmatrix}1\\ 0\\ 0\end{pmatrix},\quad\psi_{2}=\begin{pmatrix}0\\ 1\\ 0\end{pmatrix},\quad\mbox{and}\quad\psi_{3}=\begin{pmatrix}0\\ 0\\ 1\end{pmatrix},

we define u1αC2(Ω)u_{1}^{\alpha}\in C^{2}(\Omega) such that, in Ω2R\Omega_{2R}

u1α:=u¯1α+u~1α:=u¯ψα+λ+μλ+2μf(u¯)xαδψ3,α=1,2,u13:=u¯13+u~13:=u¯ψ3+λ+μμf(u¯)(x1δψ1+x2δψ2).\begin{split}u_{1}^{\alpha}&:=\bar{u}_{1}^{\alpha}+\tilde{u}_{1}^{\alpha}:=\bar{u}\psi_{\alpha}+\frac{\lambda+\mu}{\lambda+2\mu}f(\bar{u})\,\partial_{x_{\alpha}}\delta\,\psi_{3},\qquad\qquad\alpha=1,2,\\ u_{1}^{3}&:=\bar{u}_{1}^{3}+\tilde{u}_{1}^{3}:=\bar{u}\psi_{3}+\frac{\lambda+\mu}{\mu}f(\bar{u})\,(\partial_{x_{1}}\delta\,\psi_{1}+\partial_{x_{2}}\delta\,\psi_{2}).\end{split} (2.17)

Notice that the corrected terms u~1α\tilde{u}_{1}^{\alpha} depend on the Lamé parameters λ\lambda and μ\mu, which can help us capture all singular terms of viα\nabla v_{i}^{\alpha}, α=1,,d\alpha=1,\cdots,d.

2.4. Asymptotic expression of v1α\nabla v_{1}^{\alpha}, α=1,,d\alpha=1,\cdots,d.

We now use these auxiliary functions u1αu_{1}^{\alpha} to obtain the asymptotics of v1α\nabla v_{1}^{\alpha} in the narrow region ΩR\Omega_{R}, α=1,,d\alpha=1,\cdots,d.

Theorem 2.2.

For d=2,3d=2,3, let v1αH1(Ω;d)v_{1}^{\alpha}\in{H}^{1}(\Omega;\mathbb{R}^{d}) be the weak solutions of (2.5). Then for sufficiently small 0<ε<1/20<\varepsilon<1/2, we have

v1α(x)=u1α+O(1),α=1,,d,xΩR.\displaystyle\nabla v_{1}^{\alpha}(x)=\nabla u_{1}^{\alpha}+O(1),\quad\alpha=1,\cdots,d,\quad x\in\Omega_{R}. (2.18)

Here we would like to emphasize the importance of the introduction of u~1α\tilde{u}_{1}^{\alpha}. Although u¯1α\bar{u}_{1}^{\alpha} can be used to obtain the upper bound estimates of |v1α||\nabla v_{1}^{\alpha}| by |(v1αu¯1α)|Cε|\nabla(v_{1}^{\alpha}-\bar{u}_{1}^{\alpha})|\leq\frac{C}{\sqrt{\varepsilon}} derived in [27, Corollary 5.2], it as well shows to us it is possible that there remains more other singular terms in (v1αu¯1α)\nabla(v_{1}^{\alpha}-\bar{u}_{1}^{\alpha}). As it turns out, the appearance of u~1α\nabla\tilde{u}_{1}^{\alpha} can find all the singular terms of v1α\nabla v_{1}^{\alpha} and make |(v1αu¯1αu~1α)||\nabla(v_{1}^{\alpha}-\bar{u}_{1}^{\alpha}-\tilde{u}_{1}^{\alpha})| 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 |(v1αu¯1α)|Cε|\nabla(v_{1}^{\alpha}-\bar{u}_{1}^{\alpha})|\leq\frac{C}{\sqrt{\varepsilon}}. More importantly, it also allows us to obtain the asymptotics of a11ααa_{11}^{\alpha\alpha}, α=1,,d\alpha=1,\cdots,d, defined in (2.9). This kind of formulas is also the key to establish the asymptotics of C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha}, α=1,,d\alpha=1,\cdots,d. We here give the results for m=2m=2. For the general cases m3m\geq 3, see Section 6 below.

Proposition 2.3.

[The asymptotics of a11ααa_{11}^{\alpha\alpha}] Under the assumptions (2.12) and (2.13) with m=2m=2, we have, for sufficiently small 0<ε<1/20<\varepsilon<1/2,

(i) for d=2d=2,

a1111=πμκ1ε(1+O(εγ+38))anda1122=π(λ+2μ)κ1ε(1+O(εγ+38));a_{11}^{11}=\frac{\pi\mu}{\sqrt{\kappa}}\cdot\frac{1}{\sqrt{\varepsilon}}\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right)\quad\mbox{and}\quad a_{11}^{22}=\frac{\pi(\lambda+2\mu)}{\sqrt{\kappa}}\cdot\frac{1}{\sqrt{\varepsilon}}\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right);

(ii) for d=3d=3, there exist constants 𝒞3α\mathcal{C}_{3}^{*\alpha} and 𝒞33\mathcal{C}_{3}^{*3}, independent of ε\varepsilon, such that for α=1,2\alpha=1,2,

a11αα=πμκ|logε|+𝒞3α+O(ε1/8)anda1133=π(λ+2μ)κ|logε|+𝒞33+O(ε1/8).a_{11}^{\alpha\alpha}=\frac{\pi\mu}{\kappa}|\log\varepsilon|+\mathcal{C}_{3}^{*\alpha}+O(\varepsilon^{1/8})\quad\mbox{and}\quad a_{11}^{33}=\frac{\pi(\lambda+2\mu)}{\kappa}|\log\varepsilon|+\mathcal{C}_{3}^{*3}+O(\varepsilon^{1/8}).
Remark 2.4.

Here we would point out that if we only use u¯1α\bar{u}_{1}^{\alpha} as the auxiliary function like in [10], it is also not possible to obtain these asymptotic formula for a11ααa_{11}^{\alpha\alpha}. The details can be found in the proof of Proposition 2.3, see Section 4 below. So the introduction of u~1β\tilde{u}_{1}^{\beta} 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 mm-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 D1\partial D_{1} relative to (DD2)\partial(D\setminus D_{2}) is given by

Cap(D1):=D1v1ν=Ω|v1|2𝑑x,\mbox{Cap}(D_{1}):=-\int_{\partial D_{1}}\frac{\partial v_{1}}{\partial\nu}=\int_{\Omega}|\nabla v_{1}|^{2}\ dx,

where v1C2(Ω)v_{1}\in{C}^{2}(\Omega) satisfies

{Δv1=0,inΩ,v1=1,onD1,v1=0,onD2D.\begin{cases}\Delta v_{1}=0,&\mathrm{in}~{}\Omega,\\ v_{1}=1,&\mathrm{on}~{}\partial{D}_{1},\\ v_{1}=0,&\mathrm{on}~{}\partial{D_{2}}\cup\partial{D}.\end{cases}

Henthforth, in this sense a11ααa_{11}^{\alpha\alpha} is an “elasticity capacity” of D1\partial D_{1} relative to (DD2)\partial(D\setminus D_{2}).

For α=d+1,,d(d+1)/2\alpha=d+1,\cdots,d(d+1)/2, we use the auxiliary functions as in [10, 11], that is,

u1α:=u¯ψα.\displaystyle u_{1}^{\alpha}:=\bar{u}\psi_{\alpha}. (2.19)
Theorem 2.6.

([10, 11]) For d=2,3d=2,3, let v1αH1(Ω;d)v_{1}^{\alpha}\in{H}^{1}(\Omega;\mathbb{R}^{d}) be the weak solutions of (2.5). Then for sufficiently small 0<ε<1/20<\varepsilon<1/2, we have

v1α(x)=u1α+O(1),α=d+1,,d(d+1)/2,xΩR.\displaystyle\nabla v_{1}^{\alpha}(x)=\nabla u_{1}^{\alpha}+O(1),\quad\alpha=d+1,\cdots,d(d+1)/2,\quad x\in\Omega_{R}.

We remark that Theorems 2.2 and 2.6 also hold true for v2αv_{2}^{\alpha} by replacing u¯\bar{u} with u¯\underline{u}, where u¯C2(d)\underline{u}\in C^{2}(\mathbb{R}^{d}) is a scalar function satisfying u¯=1\underline{u}=1 on D2\partial{D}_{2}, u¯=0\underline{u}=0 on D1D\partial{D}_{1}\cup\partial D, u¯=1u¯\underline{u}=1-\bar{u} in Ω2R\Omega_{2R}, and u¯C2(ΩΩR)C\|\underline{u}\|_{C^{2}(\Omega\setminus\Omega_{R})}\leq\,C. In this case we denote the auxiliary functions by u2αu_{2}^{\alpha}.

Throughout the paper, unless otherwise stated, we use CC to denote some positive constant, whose values may vary from line to line, depending only on d,κ0,Rd,\kappa_{0},R, and an upper bound of the C2,γC^{2,\gamma} norms of D1,D2\partial D_{1},\partial D_{2}, and D\partial D, but not on ε\varepsilon. 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 φ\varphi and then give the asymptotic formulas of u\nabla u 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 mm-convex inclusions are presented in Subsection 3.2. We find that the blow-up rates depend on the dimension dd and the convexity order mm, and the blow-up points shift away from the origin with the increasing of mm. Finally, in Subsection 3.3, we give an example that h1(x)h_{1}(x^{\prime}) and h2(x)h_{2}(x^{\prime}) are assumed to be x3=±κ2|x1|m±κ2|x2|mx_{3}=\pm\frac{\kappa}{2}|x_{1}|^{m}\pm\frac{\kappa^{\prime}}{2}|x_{2}|^{m}, respectively, for |x|<2R|x^{\prime}|<2R, to show how the geometric parameters κ\kappa and κ\kappa^{\prime} influence the blow-up of u\nabla u. It turns out that the mean curvature of inclusions plays the role.

3.1. For the 22-convex inclusions

As mentioned in Subsection 2.2, |ub||\nabla u_{b}| is uniformly bounded with respect to ε\varepsilon. In order to derive the asymptotics of the gradient of solution with respect to the sufficiently small parameter ε>0\varepsilon>0, we consider the case when two inclusions touch each other. Let uu^{*} be the solution of

{λ,μu=0,inΩ,u=α=1d(d+1)/2Cαψα,onD1D2,D1uν|+ψβ+D2uν|+ψβ=0,β=1,,d(d+1)/2,u=φ,onD,\displaystyle\begin{cases}\mathcal{L}_{\lambda,\mu}u^{*}=0,\quad&\hbox{in}\ \Omega^{*},\\ u^{*}=\sum_{\alpha=1}^{d(d+1)/2}C_{*}^{\alpha}\psi_{\alpha},&\hbox{on}\ \partial D_{1}^{*}\cup\partial D_{2}^{*},\\ \int_{\partial{D}_{1}^{*}}\frac{\partial{u}^{*}}{\partial\nu}\big{|}_{+}\cdot\psi_{\beta}+\int_{\partial{D}_{2}^{*}}\frac{\partial{u}^{*}}{\partial\nu}\big{|}_{+}\cdot\psi_{\beta}=0,&\beta=1,\cdots,d(d+1)/2,\\ u^{*}=\varphi,&\hbox{on}\ \partial{D},\end{cases} (3.1)

where

D1:={xd|x+P1D1},D2:=D2,andΩ:=DD1D2¯,D_{1}^{*}:=\{x\in\mathbb{R}^{d}\big{|}x+P_{1}\in D_{1}\},\quad D_{2}^{*}:=D_{2},\quad\mbox{and}\quad\Omega^{*}:=D\setminus\overline{D_{1}^{*}\cup D_{2}^{*}},

and the constants CαC_{*}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2, are uniquely determined by minimizing the energy

Ω((0)e(v),e(v))𝑑x\int_{\Omega^{*}}\left(\mathbb{C}^{(0)}e(v),e(v)\right)dx

in an admissible function space

𝒜:={vH1(D;d)|e(v)=0inD1D2,andv=φonD}.\mathcal{A}^{*}:=\left\{v\in{H}^{1}(D;\mathbb{R}^{d})~{}\big{|}~{}e(v)=0~{}~{}\mbox{in}~{}{D}^{*}_{1}\cup{D}^{*}_{2},~{}\mbox{and}~{}v=\varphi~{}\mbox{on}~{}\partial{D}\right\}.

We emphasize that we use the third line of (3.1), which implies that total flux of uu^{*} 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 φ\varphi, for β=1,,d(d+1)/2\beta=1,\cdots,d(d+1)/2,

b1β[φ]:=D1ubν|+ψβandb1β[φ]:=D1uν|+ψβ,\displaystyle b_{1}^{\beta}[\varphi]:=\int_{\partial{D}_{1}}\frac{\partial{u}_{b}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}\quad\mbox{and}\quad b_{1}^{*\beta}[\varphi]:=\int_{\partial{D}_{1}^{*}}\frac{\partial u^{*}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}, (3.2)

and denote

ρd(ε)={ε,d=2,|logε|1,d=3.\rho_{d}(\varepsilon)=\begin{cases}\sqrt{\varepsilon},&\quad d=2,\\ |\log\varepsilon|^{-1},&\quad d=3.\end{cases}

We will prove that b1β[φ]b_{1}^{*\beta}[\varphi] are convergent to b1β[φ]b_{1}^{\beta}[\varphi], with rates ρd(ε)\rho_{d}(\varepsilon).

Here we need some symmetric assumptions on the domain and the boundary data. First, we assume that

(S1):\displaystyle({\rm S_{1}}):~{} D1D2 and Dare  symmetric  with  repect  to  each  axis xi, and\displaystyle~{}D_{1}\cup D_{2}~{}\mbox{ and~{}}D~{}\mbox{are~{} symmetric~{} with~{} repect~{} to~{} each~{} axis~{}}x_{i},\mbox{~{}and~{}}
 the superplane {xd=0};\displaystyle~{}\mbox{ the superplane~{}}\{x_{d}=0\};

For boundary data φ\varphi, we assume that in dimension two, φ1(x1,x2)\varphi^{1}(x_{1},x_{2}) is odd with respect to x2x_{2} and φ2(x1,x2)\varphi^{2}(x_{1},x_{2}) is even with respect to x2x_{2}; and in dimension three, φi(x)\varphi^{i}(x) is odd for i=1,2,3i=1,2,3, that is, for xDx\in\partial D,

(S2):φ1(x1,x2)=φ1(x1,x2),φ2(x1,x2)=φ2(x1,x2),ford=2;\displaystyle({\rm S_{2}}):~{}~{}\varphi^{1}(x_{1},x_{2})=-\varphi^{1}(x_{1},-x_{2}),~{}\varphi^{2}(x_{1},x_{2})=\varphi^{2}(x_{1},-x_{2}),\quad\mbox{for}~{}d=2;
(S3):φi(x)=φi(x),i=1,2,3,ford=3.\displaystyle({\rm S_{3}}):~{}~{}\varphi^{i}(x)=-\varphi^{i}(-x),\quad\,i=1,2,3,~{}\quad\mbox{for}~{}d=3.

We shall use O(1)O(1) to denote those quantities satisfying |O(1)|C|O(1)|\leq C, for some constant independent of ε\varepsilon. We assume that for some δ0>0\delta_{0}>0,

δ0μ,dλ+2μ1δ0.\delta_{0}\leq\mu,d\lambda+2\mu\leq\frac{1}{\delta_{0}}.

Then

Proposition 3.1.

For d=2,3d=2,3, assume that the domain satisfies (2.12), (2.13) and (S1)({\rm S_{1}}). If φ\varphi satisfies (S2)({\rm S_{2}}) or (S3)({\rm S_{3}}), then, for sufficiently small ε>0\varepsilon>0,

b1β[φ]b1β[φ]=O(ρd(ε)),β=1,,d(d+1)/2.b_{1}^{\beta}[\varphi]-b_{1}^{*\beta}[\varphi]=O(\rho_{d}(\varepsilon)),\quad\beta=1,\cdots,d(d+1)/2. (3.3)

The proof will be given in Section 4. We would like to point out that the functionals b1β[φ]b_{1}^{*\beta}[\varphi] 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 u\nabla u in dimension d=2d=2 and for m=2m=2 follows.

Theorem 3.2.

For d=2d=2, let D1,D2DD_{1},D_{2}\subset D be defined as above and satisfy (2.12), (2.13) with m=2m=2, and (S1)({\rm S_{1}}). Let uH1(D;2)C1(Ω¯;2)u\in{H}^{1}(D;\mathbb{R}^{2})\cap{C}^{1}(\overline{\Omega};\mathbb{R}^{2}) be the solution to (2.2). Then, if φ\varphi satisfies (S2)({\rm S_{2}}), then for sufficiently small 0<ε<1/20<\varepsilon<1/2 and for xΩRx\in\Omega_{R},

u(x)=κπε(b11[φ]μu11+b12[φ]λ+2μu12)(1+O(εγ+38))+O(1)φC0(D),\displaystyle\nabla u(x)=\frac{\sqrt{\kappa}}{\pi}\cdot\sqrt{\varepsilon}\left(\frac{b_{1}^{*1}[\varphi]}{\mu}\nabla u_{1}^{1}+\frac{b_{1}^{*2}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{2}\right)(1+O(\varepsilon^{\frac{\gamma+3}{8}}))+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.4)

where u1αu_{1}^{\alpha} are specific functions, constructed in (2.16), α=1,2\alpha=1,2.

Remark 3.3.

If D1D2D_{1}\cup D_{2} and DD are disks, for example,

D1=B1(0,1+ε/2),D1=B1(0,1ε/2),andD=B3(0),D_{1}=B_{1}(0,1+\varepsilon/2),\quad D_{1}=B_{1}(0,-1-\varepsilon/2),\quad\mbox{and}~{}~{}D=B_{3}(0),

then the result in Theorem 3.2 holds for φ=(x2,0)T\varphi=(x_{2},0)^{\mathrm{T}}. More generally, we can choose φ=(cx2,c~x1)T\varphi=(cx_{2},\tilde{c}x_{1})^{\mathrm{T}} for some constants cc and c~\tilde{c}.

Remark 3.4.

Recalling the definition of u1αu_{1}^{\alpha}, (2.16), a direct calculation yields

u1α(x1,x2)=1δ(x1)Eα2+O(1),xΩε,\displaystyle\nabla u_{1}^{\alpha}(x_{1},x_{2})=\frac{1}{\delta(x_{1})}E_{\alpha 2}+O(1),\quad x\in\Omega_{\varepsilon},

where “EαβE_{\alpha\beta}” denotes the basic matrix with only one non-zero entry 1 in the αth\alpha^{th} row and the βth\beta^{th} column. So in a neighborhood of the origin, say Ωε\Omega_{\varepsilon}, we find that (3.4) becomes

u(x1,x2)=κπ𝔹2[φ]εδ(x1)(1+O(ε))+O(1)φC0(D),\displaystyle\nabla u(x_{1},x_{2})=\frac{\sqrt{\kappa}}{\pi}\mathbb{B}_{2}[\varphi]\cdot\frac{\sqrt{\varepsilon}}{\delta(x_{1})}(1+O(\sqrt{\varepsilon}))+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.5)

where

𝔹2[φ]:=1μb11[φ]E12+1λ+2μb12[φ]E22.\mathbb{B}_{2}[\varphi]:=\frac{1}{\mu}b_{1}^{*1}[\varphi]E_{12}+\frac{1}{\lambda+2\mu}b_{1}^{*2}[\varphi]E_{22}.

For this moment, to get the exact asymptotic expression of u\nabla u near the origin, we only need to evaluate these boundary integrals b1βb_{1}^{*\beta} in 𝔹2[φ]\mathbb{B}_{2}[\varphi], which can be computed by numerical method in practical problems. We would like to point out that they no longer depend on ε\varepsilon and there is no singularity near the origin. It is completely a computation problem. We leave it to interested readers.

Theorem 3.5.

For d=3d=3, let D1,D2DD_{1},D_{2}\subset D be defined as above and satisfy (2.12), (2.13) with m=2m=2, and (S1)({\rm S_{1}}). Let uH1(D;3)C1(Ω¯;3)u\in{H}^{1}(D;\mathbb{R}^{3})\cap{C}^{1}(\overline{\Omega};\mathbb{R}^{3}) is the solution to (2.2). Then if φ\varphi satisfies (S3)({\rm S_{3}}), we have, for sufficiently small 0<ε<1/20<\varepsilon<1/2, and for xΩRx\in\Omega_{R},

u\displaystyle\nabla u =κπ1|logε|(α=12b1α[φ]μu1α+b13[φ]λ+2μu13)(1+O(|logε|1))\displaystyle=\frac{\kappa}{\pi}\cdot\frac{1}{|\log\varepsilon|}\left(\sum_{\alpha=1}^{2}\frac{b_{1}^{*\alpha}[\varphi]}{\mu}\nabla u_{1}^{\alpha}+\frac{b_{1}^{*3}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{3}\right)\left(1+O(|\log\varepsilon|^{-1})\right)
+O(1)φC0(D),\displaystyle\quad\quad\quad+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.6)

where u1αu_{1}^{\alpha} are specific functions, constructed in (2.17).

Remark 3.6.

(1) If φi=xi\varphi^{i}=x_{i} for i=1,2,3i=1,2,3, D1D2D_{1}\cup D_{2} and DD are spheres satisfying (S1)({\rm S_{1}}), then Theorem 3.5 holds true.

(2) Similarly as in Remark 3.4, for xΩεx\in\Omega_{\sqrt{\varepsilon}}, from (3.5),

u(x,x3)=κπ|logε|𝔹3[φ]1δ(x)(1+O(|logε|1))+O(1)φC0(D),\displaystyle\nabla u(x^{\prime},x_{3})=\frac{\kappa}{\pi|\log\varepsilon|}\mathbb{B}_{3}[\varphi]\cdot\frac{1}{\delta(x^{\prime})}\left(1+O(|\log\varepsilon|^{-1})\right)+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.7)

where

𝔹3[φ]:=1μ(b11E13+b12E23)+1λ+2μb13[φ]E33.\displaystyle\mathbb{B}_{3}[\varphi]:=\frac{1}{\mu}\left(b_{1}^{*1}E_{13}+b_{1}^{*2}E_{23}\right)+\frac{1}{\lambda+2\mu}b_{1}^{*3}[\varphi]E_{33}. (3.8)

(3) It is worth mentioning that the asymptotic formulas (3.5) and (3.7) immediately imply that the blow-up rates of |u||\nabla u|, ε1/2\varepsilon^{-1/2} for d=2d=2, and (ε|logε|)1(\varepsilon|\log\varepsilon|)^{-1} for d=3d=3, obtained in [10, 11, 35], are optimal.

Remark 3.7.

If φ=(φ1,φ2,φ3)T\varphi=(\varphi^{1},\varphi^{2},\varphi^{3})^{\mathrm{T}} satisfies another symmetric condition

{φ1(x1,x2,x3)=φ1(x1,x2,x3)=φ1(x1,x2,x3),φ2(x1,x2,x3)=φ2(x1,x2,x3)=φ2(x1,x2,x3),φ3(x1,x2,x3)=φ3(x1,x2,x3)=φ3(x1,x2,x3),\begin{cases}\varphi^{1}(x_{1},x_{2},x_{3})=-\varphi^{1}(x_{1},x_{2},-x_{3})=-\varphi^{1}(x_{1},-x_{2},-x_{3}),\\ \varphi^{2}(x_{1},x_{2},x_{3})=-\varphi^{2}(x_{1},x_{2},-x_{3})=\varphi^{2}(x_{1},-x_{2},-x_{3}),\\ \varphi^{3}(x_{1},x_{2},x_{3})=\varphi^{3}(x_{1},x_{2},-x_{3})=\varphi^{3}(x_{1},-x_{2},-x_{3}),\end{cases}

then by replicating the proof of Proposition 4.4 below, we obtain C1α=C2αC_{1}^{\alpha}=C_{2}^{\alpha}, α=2,,6\alpha=2,\cdots,6. Thus, (3.5) becomes more simpler as follows

u=κπ1|logε|1μb11[φ]u11(1+O(|logε|1))+O(1)φC0(D).\displaystyle\nabla u=\frac{\kappa}{\pi}\cdot\frac{1}{|\log\varepsilon|}\frac{1}{\mu}b_{1}^{*1}[\varphi]\nabla u_{1}^{1}\left(1+O(|\log\varepsilon|^{-1})\right)+O(1)\|\varphi\|_{C^{0}(\partial D)}.

3.2. For the mm-convex inclusions, m3m\geq 3

In the following, we shall reveal the role of the order of the relatively convexity between D1\partial D_{1} and D2\partial D_{2}, mm, playing in the asymptotics of u\nabla u. Define

Qd,m=20td21+tm𝑑t,Q~d,m=20td1+tm𝑑t;Q_{d,m}=2\int_{0}^{\infty}\frac{t^{d-2}}{1+t^{m}}\ dt,\quad\quad\widetilde{Q}_{d,m}=2\int_{0}^{\infty}\frac{t^{d}}{1+t^{m}}\ dt;

and

ρm,2(ε)={|logε|1,m=3,ε1/4,m=4,0,m5,andE(κ,ε,m)={3κ21|logε|,m=3,κ3/mε13/mQ~2,m,m4.\displaystyle\rho_{m,2}(\varepsilon)=\begin{cases}|\log\varepsilon|^{-1},&~{}m=3,\\ \varepsilon^{1/4},&~{}m=4,\\ 0,&~{}m\geq 5,\end{cases}~{}\quad\quad\mbox{and}~{}\quad~{}E(\kappa,\varepsilon,m)=\begin{cases}\frac{3\kappa}{2}\frac{1}{|\log\varepsilon|},&~{}m=3,\\ \frac{\kappa^{3/m}\varepsilon^{1-3/m}}{\widetilde{Q}_{2,m}},&~{}m\geq 4.\end{cases}

Then

Theorem 3.8.

For d=2d=2, let D1,D2DD_{1},D_{2}\subset D be defined as above and satisfy (2.12) and (2.13) with m3m\geq 3. Let uH1(D;2)C1(Ω¯;2)u\in{H}^{1}(D;\mathbb{R}^{2})\cap{C}^{1}(\overline{\Omega};\mathbb{R}^{2}) be the solution to (2.2). Then for sufficiently small 0<ε<1/20<\varepsilon<1/2 and xΩRx\in\Omega_{R}, we have

u(x)\displaystyle\nabla u(x) =κ1/mε11/mQ2,m(b11[φ]μu11+b12[φ]λ+2μu12)(1+O(ρm,2(ε)))\displaystyle=\frac{\kappa^{1/m}\varepsilon^{1-1/m}}{Q_{2,m}}\left(\frac{b_{1}^{*1}[\varphi]}{\mu}\nabla u_{1}^{1}+\frac{b_{1}^{*2}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{2}\right)\left(1+O(\rho_{m,2}(\varepsilon))\right)
+E(κ,ε,m)b13[φ]λ+2μu13(1+O(ρm,2(ε)))+O(1)φC0(D),\displaystyle\quad\quad+E(\kappa,\varepsilon,m)\frac{b_{1}^{*3}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{3}\left(1+O(\rho_{m,2}(\varepsilon))\right)+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.9)

where u1αu_{1}^{\alpha} are defined in (2.16), α=1,2\alpha=1,2, and u13u_{1}^{3} is defined in (2.19).

Remark 3.9.

We would like to remark that the pointwise upper bound estimates of |u||\nabla u| in [27] imply that when md+1m\geq d+1, the maximum of the upper bounds obtain at xΩRx\in\Omega_{R} and |x|=cε1/m|x^{\prime}|=c\varepsilon^{1/m} for some positive constant c>0c>0. Therefore, for xΩε1/(m(m1))x\in\Omega_{\varepsilon^{1/(m(m-1))}}, recalling the definition of u1αu_{1}^{\alpha}, we have

u(x1,x2)\displaystyle\nabla u(x_{1},x_{2}) =κ1/mε11/mQ2,m𝔹2,I[φ]1δ(x1)(1+O(ρm,2(ε)))\displaystyle=\frac{\kappa^{1/m}\varepsilon^{1-1/m}}{Q_{2,m}}\mathbb{B}_{2,\rm{I}}[\varphi]\cdot\frac{1}{\delta(x_{1})}\left(1+O(\rho_{m,2}(\varepsilon))\right)
+E(κ,ε,m)𝔹2,II[φ]x1δ(x1)(1+O(ρm,2(ε)))+O(1)φC0(D),\displaystyle\quad+E(\kappa,\varepsilon,m)\mathbb{B}_{2,\rm{II}}[\varphi]\cdot\frac{x_{1}}{\delta(x_{1})}\left(1+O(\rho_{m,2}(\varepsilon))\right)+O(1)\|\varphi\|_{C^{0}(\partial D)},

where

𝔹2,I[φ]:=1μb11[φ]E12+1λ+2μb12[φ]E22,𝔹2,II[φ]:=1λ+2μb13[φ]E22.\displaystyle\mathbb{B}_{2,\text{I}}[\varphi]:=\frac{1}{\mu}b_{1}^{*1}[\varphi]E_{12}+\frac{1}{\lambda+2\mu}b_{1}^{*2}[\varphi]E_{22},\quad\mathbb{B}_{2,\text{II}}[\varphi]:=\frac{1}{\lambda+2\mu}b_{1}^{*3}[\varphi]E_{22}.
Remark 3.10.

If 𝔹2,I[φ]=0\mathbb{B}_{2,\text{I}}[\varphi]=0 for some φ\varphi, then the concentration mechanism of the stress is determined by 𝔹2,II[φ]\mathbb{B}_{2,\text{II}}[\varphi]. Thus, Theorem 3.8, combining with the upper bounds in [27], implies that the blow-up occurs at the segments 𝒮:={(x¯1,x¯2)ΩR||x¯1|=ε1/m}\mathcal{S}:=\{(\bar{x}_{1},\bar{x}_{2})\in\Omega_{R}\big{|}~{}~{}|\bar{x}_{1}|=\varepsilon^{1/m}\}, with blow-up rate 1|logε|ε2/3\frac{1}{|\log\varepsilon|\varepsilon^{2/3}} when m=3m=3 and ε2m\varepsilon^{-\frac{2}{m}} when m4m\geq 4. Consequently, the gradient will not blow up any more and 𝒮\mathcal{S} will be more and more away from the shortest segment P1P2¯\overline{P_{1}P_{2}} as mm goes to infinity. From (2.12) with R<1R<1, we can see that when mm\rightarrow\infty, the boundaries of D1\partial D_{1} and D2\partial D_{2} 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 mm changes from 22 to ++\infty.

Finally, when d=3d=3, we define

ρm,3(ε)={|logε|1,m=4,ε14/m,5m7,0,m8;andF(κ,ε,m)={2κπ|logε|,m=4,κ4/mε14/mπQ~3,m,m5.\displaystyle\rho_{m,3}(\varepsilon)=\begin{cases}|\log\varepsilon|^{-1},&~{}m=4,\\ \varepsilon^{1-4/m},&~{}5\leq m\leq 7,\\ 0,&~{}m\geq 8;\end{cases}\quad\mbox{and}~{}F(\kappa,\varepsilon,m)=\begin{cases}\frac{2\kappa}{\pi|\log\varepsilon|},&~{}m=4,\\ \frac{\kappa^{4/m}\varepsilon^{1-4/m}}{\pi\widetilde{Q}_{3,m}},&~{}m\geq 5.\end{cases}

Then

Theorem 3.11.

For d=3d=3, let D1,D2DD_{1},D_{2}\subset D be defined as above and satisfy (2.12) and (2.13) with m3m\geq 3. Let uH1(D;3)C1(Ω¯;3)u\in{H}^{1}(D;\mathbb{R}^{3})\cap{C}^{1}(\overline{\Omega};\mathbb{R}^{3}) be the solution to (2.2).

(i) When m=3m=3, we assume that D1D2D_{1}\cup D_{2} and DD satisfies (S1)({\rm S_{1}}). Then if φ\varphi satisfies (S3)({\rm S_{3}}), then for sufficiently small 0<ε<1/20<\varepsilon<1/2, and xΩRx\in\Omega_{R},

u(x)\displaystyle\nabla u(x) =κ2/3πε1/3Q3,3(α=12b1α[φ]μu1α+b13[φ]λ+2μu13)(1+O(ε1/3))\displaystyle=\frac{\kappa^{2/3}}{\pi}\cdot\frac{\varepsilon^{1/3}}{Q_{3,3}}\left(\sum_{\alpha=1}^{2}\frac{b_{1}^{*\alpha}[\varphi]}{\mu}\nabla u_{1}^{\alpha}+\frac{b_{1}^{*3}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{3}\right)\Big{(}1+O(\varepsilon^{1/3})\Big{)}
+O(1)φC0(D),\displaystyle\quad\quad\quad+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.10)

where u1αu_{1}^{\alpha} are specific functions, constructed in (2.17).

(ii) When m4m\geq 4, for sufficiently small 0<ε<1/20<\varepsilon<1/2 and xΩRx\in\Omega_{R},

u(x)\displaystyle\nabla u(x) =κ2/mε12/mπQ3,m(α=12b1α[φ]μu1α+b13[φ]λ+2μu13)(1+O(ρm,3(ε)))\displaystyle=\frac{\kappa^{2/m}\varepsilon^{1-2/m}}{\pi Q_{3,m}}\left(\sum_{\alpha=1}^{2}\frac{b_{1}^{*\alpha}[\varphi]}{\mu}\nabla u_{1}^{\alpha}+\frac{b_{1}^{*3}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{3}\right)\left(1+O(\rho_{m,3}(\varepsilon))\right)
+F(κ,ε,m)(b14[φ]μu14+α=56b1α[φ]λ+2μu1α)(1+O(ρm,3(ε)))\displaystyle\quad\quad+F(\kappa,\varepsilon,m)\left(\frac{b_{1}^{*4}[\varphi]}{\mu}\nabla u_{1}^{4}+\sum_{\alpha=5}^{6}\frac{b_{1}^{*\alpha}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{\alpha}\right)\left(1+O(\rho_{m,3}(\varepsilon))\right)
+O(1)φC0(D),\displaystyle\quad\quad+O(1)\|\varphi\|_{C^{0}(\partial D)}, (3.11)

where u1αu_{1}^{\alpha} are specific functions, constructed in (2.17) and (2.19), respectively.

Remark 3.12.

The above results, together with Theorem 3.8, imply that the blow-up rate of |u||\nabla u| depends on the space dimension dd, the convexity order mm, and the first term’s coefficient κ\kappa. Furthermore, when md+1m\geq d+1, we do not need the symmetric assumption about domains and the boundary data, since in this case, C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha} tends to zero as ε0\varepsilon\rightarrow 0; 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 d4d\geq 4, 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 m2m\geq 2, we denote the top and bottom boundaries of ΩR\Omega_{R} as

ΓR+={x3|x3=ε2+κ2|x1|m+κ2|x2|m}\displaystyle\Gamma_{R}^{+}=\left\{x\in\mathbb{R}^{3}~{}\big{|}~{}x_{3}=\frac{\varepsilon}{2}+\frac{\kappa}{2}|x_{1}|^{m}+\frac{\kappa^{\prime}}{2}|x_{2}|^{m}\right\}

and

ΓR={x3|x3=ε2κ2|x1|mκ2|x2|m},\displaystyle\Gamma_{R}^{-}=\left\{x\in\mathbb{R}^{3}~{}\big{|}~{}x_{3}=-\frac{\varepsilon}{2}-\frac{\kappa}{2}|x_{1}|^{m}-\frac{\kappa^{\prime}}{2}|x_{2}|^{m}\right\},

where κ\kappa and κ\kappa^{\prime} are two positive constants, may different. For θ[0,2π]\theta\in[0,2\pi], denote

Em(θ):=sin2m1θcos2m+1θ+cos2m1θsin2m+1θ,E_{m}(\theta):=\sin^{\frac{2}{m}-1}\theta\cos^{\frac{2}{m}+1}\theta+\cos^{\frac{2}{m}-1}\theta\sin^{\frac{2}{m}+1}\theta,
Fm(θ):=(cos2θκ)2/m+(sin2θκ)2/m,F_{m}(\theta):=\left(\frac{\cos^{2}\theta}{\kappa}\right)^{2/m}+\left(\frac{\sin^{2}\theta}{\kappa^{\prime}}\right)^{2/m},

and

Gm:=02πEm(θ)𝑑θ,G~m:=02πEm(θ)Fm(θ)𝑑θ.\quad G_{m}:=\int_{0}^{2\pi}E_{m}(\theta)\ d\theta,\quad\widetilde{G}_{m}:=\int_{0}^{2\pi}E_{m}(\theta)F_{m}(\theta)\ d\theta.

Then the results in Theorems 3.5 and 3.11 hold true, except that

(i) if m=2m=2, κ\kappa in (3.5) is replaced by κκ\sqrt{\kappa\kappa^{\prime}}, the square root of the relative Gauss curvature of D1\partial D_{1} and D2\partial D_{2}; if m=3m=3, κ2/3π\frac{\kappa^{2/3}}{\pi} in (3.11) becomes 3(κκ)1/32G3\frac{3(\kappa\kappa^{\prime})^{1/3}}{2G_{3}};

(ii) if m4m\geq 4, the terms κ2/mπ\frac{\kappa^{2/m}}{\pi} and κ4/mπ\frac{\kappa^{4/m}}{\pi} in (3.11) become m(κκ)1/mGm\frac{m(\kappa\kappa^{\prime})^{1/m}}{G_{m}} and m(κκ)2/mG~m\frac{m(\kappa\kappa^{\prime})^{2/m}}{\widetilde{G}_{m}}, 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 d=2,3d=2,3, which are two main ingredients to establish our asymptotic formulas of u\nabla u. Actually, our argument also holds in higher dimensions d4d\geq 4 with a slight modification. We first need some preliminary estimates on v1αv_{1}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2.

4.1. Auxiliary estimates for v1αv_{1}^{\alpha}

Suppose v1αv_{1}^{*\alpha} satisfies

{λ,μv1α=0,inΩ,v1α=ψα,onD1{0},v1α=0,onD2D,α=1,,d(d+1)/2.\begin{cases}\mathcal{L}_{\lambda,\mu}v_{1}^{*\alpha}=0,&\mathrm{in}~{}\Omega^{*},\\ v_{1}^{*\alpha}=\psi_{\alpha},&\mathrm{on}~{}\partial{D}_{1}^{*}\setminus\{0\},\\ v_{1}^{*\alpha}=0,&\mathrm{on}~{}\partial{D_{2}^{*}}\cup\partial{D},\end{cases}\quad\alpha=1,\cdots,d(d+1)/2. (4.1)

We will prove that v1αv1αv_{1}^{\alpha}\rightarrow v_{1}^{*\alpha}, for each α\alpha, with proper convergence rates.

Define

V:=DD1D2D1D2¯,V:=D\setminus\overline{D_{1}\cup D_{2}\cup D_{1}^{*}\cup D_{2}^{*}},

see Figure 1, and

𝒞r:={xd||x|<r,ε2+2min|x|=rh2(x)xdε2+2max|x|=rh1(x)},r<R.\mathcal{C}_{r}:=\left\{x\in\mathbb{R}^{d}\big{|}|x^{\prime}|<r,~{}-\frac{\varepsilon}{2}+2\min_{|x^{\prime}|=r}h_{2}(x^{\prime})\leq x_{d}\leq\frac{\varepsilon}{2}+2\max_{|x^{\prime}|=r}h_{1}(x^{\prime})\right\},\quad r<R.

Then we have

Lemma 4.1.

Let v1αv_{1}^{\alpha} and v1αv_{1}^{*\alpha} satisfy (2.5) and (4.1), respectively. Then we have

|(v1αv1α)(x)|Cε1/2,α=1,,d,xV𝒞ε1/4,\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x)|\leq C\varepsilon^{1/2},\quad\alpha=1,\cdots,d,\quad x\in V\setminus\mathcal{C}_{\varepsilon^{1/4}}, (4.2)

and

|(v1αv1α)(x)|Cε2/3,α=d+1,,d(d+1)/2,xV𝒞ε1/3,\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x)|\leq C\varepsilon^{2/3},\quad\alpha=d+1,\cdots,d(d+1)/2,\quad x\in V\setminus\mathcal{C}_{\varepsilon^{1/3}}, (4.3)

where CC is a universal constant.

Proof.

For ε=0\varepsilon=0, we define similarly the auxiliary functions, u¯\bar{u}^{*} and u1αu_{1}^{*\alpha}, as limits of u¯\bar{u} and u1αu_{1}^{\alpha}, where u¯C2(d)\bar{u}^{*}\in C^{2}(\mathbb{R}^{d}) satisfies u¯=1\bar{u}^{*}=1 on D1\partial{D}_{1}^{*}, u¯=0\bar{u}^{*}=0 on D2D\partial{D}_{2}^{*}\cup\partial{D} and

u¯=xdh2(x)h1(x)h2(x),inΩ2R,u¯C2(ΩΩR)C,\bar{u}^{*}=\frac{x_{d}-h_{2}(x^{\prime})}{h_{1}(x^{\prime})-h_{2}(x^{\prime})},\quad\hbox{in}\ \Omega_{2R}^{*},\quad\|\bar{u}^{*}\|_{C^{2}(\Omega^{*}\setminus\Omega_{R}^{*})}\leq\,C,

where Ωr:={(x,xd)d|h2(x)<xd<h1(x),|x|<r}\Omega_{r}^{*}:=\left\{(x^{\prime},x_{d})\in\mathbb{R}^{d}~{}\big{|}~{}h_{2}(x^{\prime})<x_{d}<h_{1}(x^{\prime}),~{}|x^{\prime}|<r\right\}, r<Rr<R. A direct computation yields

|xu¯(x)|C|x|ε+|x|2,xdu¯(x)=1δ(x),xΩR,|\partial_{x^{\prime}}\bar{u}(x)|\leq\frac{C|x^{\prime}|}{\varepsilon+|x^{\prime}|^{2}},\qquad\partial_{x_{d}}\bar{u}(x)=\frac{1}{\delta(x^{\prime})},\quad x\in\Omega_{R}, (4.4)

and

|xu¯(x)|C|x|,1C|x|2xdu¯(x)C|x|2,xΩR.|\partial_{x^{\prime}}\bar{u}^{*}(x)|\leq\frac{C}{|x^{\prime}|},\quad\frac{1}{C|x^{\prime}|^{2}}\leq\partial_{x_{d}}\bar{u}^{*}(x)\leq\frac{C}{|x^{\prime}|^{2}},\quad x\in\Omega_{R}^{*}. (4.5)

Recalling the construction of u1αu_{1}^{\alpha} in (2.16), (2.17), and (2.19), we can construct u1αu_{1}^{*\alpha} in the same way.

Refer to caption
Figure 1. Region ΩR\Omega_{R}^{*}.

Case 1. α=1,,d\alpha=1,\cdots,d. By using (2.16), (2.17), (4.4) and (4.5), we have for xΩRx\in\Omega_{R}^{*},

|xd(u1αu1α)|\displaystyle|\partial_{x_{d}}(u_{1}^{\alpha}-u_{1}^{*\alpha})| |xd(u¯1αu¯1α)|+|xd(u~1αu~1α)|\displaystyle\leq|\partial_{x_{d}}(\bar{u}_{1}^{\alpha}-\bar{u}_{1}^{*\alpha})|+|\partial_{x_{d}}(\tilde{u}_{1}^{\alpha}-\tilde{u}_{1}^{*\alpha})|
Cε|x|2(ε+|x|2)+Cε|x|3Cε|x|2(ε+|x|2)(1+ε|x|).\displaystyle\leq\frac{C\varepsilon}{|x^{\prime}|^{2}(\varepsilon+|x^{\prime}|^{2})}+\frac{C\varepsilon}{|x^{\prime}|^{3}}\leq\frac{C\varepsilon}{|x^{\prime}|^{2}(\varepsilon+|x^{\prime}|^{2})}\left(1+\frac{\varepsilon}{|x^{\prime}|}\right). (4.6)

Notice that v1αv1αv_{1}^{\alpha}-v_{1}^{*\alpha} satisfies

{λ,μ(v1αv1α)=0,inV,v1αv1α=ψαv1α,onD1D1,v1αv1α=v1α,onD2D2,v1αv1α=v1αψα,onD1(D1{0}),v1αv1α=v1α,onD2D2,v1αv1α=0,onD.\begin{cases}\mathcal{L}_{\lambda,\mu}(v_{1}^{\alpha}-v_{1}^{*\alpha})=0,&\mathrm{in}~{}V,\\ v_{1}^{\alpha}-v_{1}^{*\alpha}=\psi_{\alpha}-v_{1}^{*\alpha},&\mathrm{on}~{}\partial{D}_{1}\setminus D_{1}^{*},\\ v_{1}^{\alpha}-v_{1}^{*\alpha}=-v_{1}^{*\alpha},&\mathrm{on}~{}\partial{D}_{2}\setminus D_{2}^{*},\\ v_{1}^{\alpha}-v_{1}^{*\alpha}=v_{1}^{\alpha}-\psi_{\alpha},&\mathrm{on}~{}\partial{D}_{1}^{*}\setminus(D_{1}\cup\{0\}),\\ v_{1}^{\alpha}-v_{1}^{*\alpha}=v_{1}^{\alpha},&\mathrm{on}~{}\partial{D}_{2}^{*}\setminus D_{2},\\ v_{1}^{\alpha}-v_{1}^{*\alpha}=0,&\mathrm{on}~{}\partial{D}.\end{cases}

By (2.14), we have

|xdv1α|C,inΩΩR.|\partial_{x_{d}}v_{1}^{*\alpha}|\leq C,\quad\mbox{in}~{}\Omega^{*}\setminus\Omega_{R}^{*}. (4.7)

For xD1D1ΩΩRx\in\partial{D}_{1}\setminus D_{1}^{*}\subset\Omega^{*}\setminus\Omega_{R}^{*} (see Figure 1), by using mean value theorem and (4.7),

|(v1αv1α)(x,xd)|\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})| =|(ψαv1α)(x,xd)|\displaystyle=|(\psi_{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})|
=|v1α(x,xdε/2)v1α(x,xd)|Cε.\displaystyle=|v_{1}^{*\alpha}(x^{\prime},x_{d}-\varepsilon/2)-v_{1}^{*\alpha}(x^{\prime},x_{d})|\leq C\varepsilon. (4.8)

For xD1(D1𝒞εθ)x\in\partial{D}_{1}^{*}\setminus(D_{1}\cup\mathcal{C}_{\varepsilon^{\theta}}), by mean value theorem again and Theorem 2.2, we have

|(v1αv1α)(x,xd)|\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})| =|(v1αψα)(x,xd)|\displaystyle=|(v_{1}^{\alpha}-\psi_{\alpha})(x^{\prime},x_{d})|
=|v1α(x,xd)v1α(x,xd+ε/2)|Cεε+|x|2Cε12θ,\displaystyle=|v_{1}^{\alpha}(x^{\prime},x_{d})-v_{1}^{\alpha}(x^{\prime},x_{d}+\varepsilon/2)|\leq\frac{C\varepsilon}{\varepsilon+|x^{\prime}|^{2}}\leq C\varepsilon^{1-2\theta}, (4.9)

where 0<θ<10<\theta<1 is some constant to be determined later. Similarly, for xD2D2x\in\partial{D}_{2}\setminus D_{2}^{*}, we have

|(v1αv1α)(x,xd)|Cε,\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})|\leq C\varepsilon, (4.10)

and for xD2(D2𝒞εθ)x\in\partial{D}_{2}^{*}\setminus(D_{2}\cup\mathcal{C}_{\varepsilon^{\theta}}), we have

|(v1αv1α)(x,xd)|Cε12θ.\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})|\leq C\varepsilon^{1-2\theta}. (4.11)

For xΩRx\in\Omega_{R}^{*} with |x|=εθ|x^{\prime}|=\varepsilon^{\theta}, it follows from Theorem 2.2 and (4.1) that

|xd(v1αv1α)(x,xd)|\displaystyle\left|\partial_{x_{d}}(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})\right| =|xd(v1αu1α)+xd(u1αu1α)+xd(u1αv1α)|(x,xd)\displaystyle=\left|\partial_{x_{d}}(v_{1}^{\alpha}-u_{1}^{\alpha})+\partial_{x_{d}}(u_{1}^{\alpha}-u_{1}^{*\alpha})+\partial_{x_{d}}(u_{1}^{*\alpha}-v_{1}^{*\alpha})\right|(x^{\prime},x_{d})
C(1+ε|x|2(ε+|x|2))C(1+1ε4θ1).\displaystyle\leq C\left(1+\frac{\varepsilon}{|x^{\prime}|^{2}(\varepsilon+|x^{\prime}|^{2})}\right)\leq C\left(1+\frac{1}{\varepsilon^{4\theta-1}}\right). (4.12)

Thus, for xΩRx\in\Omega_{R}^{*} with |x|=εθ|x^{\prime}|=\varepsilon^{\theta}, by using the triangle inequality, (4.11), the mean value theorem, and (4.1), we have

|(v1αv1α)(x,xd)|\displaystyle\left|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})\right| |(v1αv1α)(x,xd)(v1αv1α)(x,h2(x)|+Cε12θ\displaystyle\leq\left|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})-(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},h_{2}(x^{\prime})\right|+C\varepsilon^{1-2\theta}
|xd(v1αv1α)|(h1(x)h2(x))||x|=εθ+Cε12θ\displaystyle\leq\left|\partial_{x_{d}}(v_{1}^{\alpha}-v_{1}^{*\alpha})\right|\cdot(h_{1}(x^{\prime})-h_{2}(x^{\prime}))\Big{|}_{|x^{\prime}|=\varepsilon^{\theta}}+C\varepsilon^{1-2\theta}
C(ε2θ+ε12θ).\displaystyle\leq C\left(\varepsilon^{2\theta}+\varepsilon^{1-2\theta}\right). (4.13)

By taking 2θ=12θ2\theta=1-2\theta, we get θ=14\theta=\frac{1}{4}. Substituting it into (4.1), (4.11), and (4.1), and using (4.1), (4.10), and v1αv1α=0v_{1}^{\alpha}-v_{1}^{*\alpha}=0 on D\partial D, we obtain

|v1αv1α|Cε1/2,on(V𝒞ε1/4).\displaystyle|v_{1}^{\alpha}-v_{1}^{*\alpha}|\leq C\varepsilon^{1/2},\quad\mbox{on}~{}\partial{(V\setminus\mathcal{C}_{\varepsilon^{1/4}})}. (4.14)

Applying the maximum principle for Lamé systems in V𝒞ε1/4V\setminus\mathcal{C}_{\varepsilon^{1/4}}, see [50], we get

|(v1αv1α)(x)|Cε1/2,inV𝒞ε1/4.\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x)|\leq C\varepsilon^{1/2},\quad\mbox{in}~{}V\setminus\mathcal{C}_{\varepsilon^{1/4}}.

Case 2. α=d+1,,d(d+1)/2\alpha=d+1,\cdots,d(d+1)/2. By using (2.19), (4.4), and (4.5), we obtain for xΩRx\in\Omega_{R}^{*},

|xd(u1αu1α)|\displaystyle|\partial_{x_{d}}(u_{1}^{\alpha}-u_{1}^{*\alpha})| |xd(u¯u¯)ψα|+|(u¯u¯)xdψα|\displaystyle\leq|\partial_{x_{d}}(\bar{u}-\bar{u}^{*})\psi_{\alpha}|+|(\bar{u}-\bar{u}^{*})\partial_{x_{d}}\psi_{\alpha}|
C(1+ε|x|(ε+|x|2)).\displaystyle\leq C\left(1+\frac{\varepsilon}{|x^{\prime}|(\varepsilon+|x^{\prime}|^{2})}\right). (4.15)

By using Theorem 2.6, we find that (4.1) and (4.11) become

|(v1αv1α)(x,xd)|Cε1θ,x(D1(D1𝒞εθ))(D2(D2𝒞εθ)),\displaystyle|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})|\leq C\varepsilon^{1-\theta},\quad x\in\Big{(}\partial{D}_{1}^{*}\setminus(D_{1}\cup\mathcal{C}_{\varepsilon^{\theta}})\Big{)}\cup\Big{(}\partial{D}_{2}^{*}\setminus(D_{2}\cup\mathcal{C}_{\varepsilon^{\theta}})\Big{)},

where 0<θ<10<\theta<1 is some constant to be fixed later. For xΩRx\in\Omega_{R}^{*} with |x|=εθ|x^{\prime}|=\varepsilon^{\theta}, (4.1) becomes

|(v1αv1α)(x,xd)|C(ε2θ+ε1θ).\displaystyle\left|(v_{1}^{\alpha}-v_{1}^{*\alpha})(x^{\prime},x_{d})\right|\leq C\left(\varepsilon^{2\theta}+\varepsilon^{1-\theta}\right).

By taking 2θ=1θ2\theta=1-\theta, we get θ=13\theta=\frac{1}{3}. We henceforth obtain

|v1αv1α|Cε2/3,on(V𝒞ε1/3).\displaystyle|v_{1}^{\alpha}-v_{1}^{*\alpha}|\leq C\varepsilon^{2/3},\quad\mbox{on}~{}(V\setminus\mathcal{C}_{\varepsilon^{1/3}}).

The proof of Lemma 4.1 is finished. ∎

4.2. Convergence of C1α+C2α2Cα\frac{C_{1}^{\alpha}+C_{2}^{\alpha}}{2}-C_{*}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2

We use the notation in [35] in the following.

It follows from (2.10) and the forth line of (2.2) that

{α=1d(d+1)/2C1αa11αβ+α=1d(d+1)/2C2αa21αβb~1β=0,α=1d(d+1)/2C1αa12αβ+α=1d(d+1)/2C2αa22αβb~2β=0,β=1,,d(d+1)/2,\displaystyle\left\{\begin{aligned} \sum_{\alpha=1}^{d(d+1)/2}C_{1}^{\alpha}a_{11}^{\alpha\beta}+\sum_{\alpha=1}^{d(d+1)/2}C_{2}^{\alpha}a_{21}^{\alpha\beta}-\tilde{b}_{1}^{\beta}&=0,\\ \sum_{\alpha=1}^{d(d+1)/2}C_{1}^{\alpha}a_{12}^{\alpha\beta}+\sum_{\alpha=1}^{d(d+1)/2}C_{2}^{\alpha}a_{22}^{\alpha\beta}-\tilde{b}_{2}^{\beta}&=0,\end{aligned}\right.\quad\quad~{}~{}\beta=1,\cdots,d(d+1)/2, (4.16)

where aijαβa_{ij}^{\alpha\beta} is defined in (2.9), and

b~jβ=Djv0ν|+ψβ.\tilde{b}_{j}^{\beta}=\int_{\partial D_{j}}\frac{\partial v_{0}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}. (4.17)

For the first equation of (4.16), we have

α=13(C1α+C2α)a11αβ+α=13C2α(a21αβa11αβ)b~1β=0,\sum_{\alpha=1}^{3}(C_{1}^{\alpha}+C_{2}^{\alpha})a_{11}^{\alpha\beta}+\sum_{\alpha=1}^{3}C_{2}^{\alpha}(a_{21}^{\alpha\beta}-a_{11}^{\alpha\beta})-\tilde{b}_{1}^{\beta}=0,

and

α=13(C1α+C2α)a21αβ+α=13C1α(a11αβa21αβ)b~1β=0.\sum_{\alpha=1}^{3}(C_{1}^{\alpha}+C_{2}^{\alpha})a_{21}^{\alpha\beta}+\sum_{\alpha=1}^{3}C_{1}^{\alpha}(a_{11}^{\alpha\beta}-a_{21}^{\alpha\beta})-\tilde{b}_{1}^{\beta}=0.

Adding these two equations together leads to

α=13(C1α+C2α)(a11αβ+a21αβ)+α=13(C1αC2α)(a11αβa21αβ)2b~1β=0.\sum_{\alpha=1}^{3}(C_{1}^{\alpha}+C_{2}^{\alpha})(a_{11}^{\alpha\beta}+a_{21}^{\alpha\beta})+\sum_{\alpha=1}^{3}(C_{1}^{\alpha}-C_{2}^{\alpha})(a_{11}^{\alpha\beta}-a_{21}^{\alpha\beta})-2\tilde{b}_{1}^{\beta}=0. (4.18)

Similarly, for the second equation of (4.16), we have

α=13(C1α+C2α)(a12αβ+a22αβ)+α=13(C1αC2α)(a12αβa22αβ)2b~2β=0.\sum_{\alpha=1}^{3}(C_{1}^{\alpha}+C_{2}^{\alpha})(a_{12}^{\alpha\beta}+a_{22}^{\alpha\beta})+\sum_{\alpha=1}^{3}(C_{1}^{\alpha}-C_{2}^{\alpha})(a_{12}^{\alpha\beta}-a_{22}^{\alpha\beta})-2\tilde{b}_{2}^{\beta}=0. (4.19)

Then from (4.18) and (4.19), we obtain

α=13C1α+C2α2aαβ+α=13C1αC2α2(a11αβa22αβ+a12αβa21αβ)(b~1β+b~2β)=0.\sum_{\alpha=1}^{3}\frac{C_{1}^{\alpha}+C_{2}^{\alpha}}{2}a^{\alpha\beta}+\sum_{\alpha=1}^{3}\frac{C_{1}^{\alpha}-C_{2}^{\alpha}}{2}(a_{11}^{\alpha\beta}-a_{22}^{\alpha\beta}+a_{12}^{\alpha\beta}-a_{21}^{\alpha\beta})-(\tilde{b}_{1}^{\beta}+\tilde{b}_{2}^{\beta})=0. (4.20)

where

aαβ=i,j=12aijαβ=(D1vαν|+ψβ+D2vαν|+ψβ).a^{\alpha\beta}=\sum_{i,j=1}^{2}a_{ij}^{\alpha\beta}=-\left(\int_{\partial{D}_{1}}\frac{\partial{v}^{\alpha}}{\partial\nu}\big{|}_{+}\cdot\psi_{\beta}+\int_{\partial{D}_{2}}\frac{\partial{v}^{\alpha}}{\partial\nu}\big{|}_{+}\cdot\psi_{\beta}\right).

and vα:=v1α+v2αv^{\alpha}:=v_{1}^{\alpha}+v_{2}^{\alpha} satisfies

{λ,μvα=0,inΩ,vα=ψα,onD1D2,vα=0,onD.\begin{cases}\mathcal{L}_{\lambda,\mu}v^{\alpha}=0,&\mathrm{in}~{}\Omega,\\ v^{\alpha}=\psi_{\alpha},&\mathrm{on}~{}\partial{D}_{1}\cup\partial D_{2},\\ v^{\alpha}=0,&\mathrm{on}~{}\partial{D}.\end{cases} (4.21)

Let vα:=v1α+v2αv^{*\alpha}:=v_{1}^{*\alpha}+v_{2}^{*\alpha} satisfy

{λ,μvα=0,inΩ,vα=ψα,onD1D2,vα=0,onD.\begin{cases}\mathcal{L}_{\lambda,\mu}v^{*\alpha}=0,&\mathrm{in}~{}\Omega^{*},\\ v^{*\alpha}=\psi_{\alpha},&\mathrm{on}~{}\partial{D}_{1}^{*}\cup\partial{D}_{2}^{*},\\ v^{*\alpha}=0,&\mathrm{on}~{}\partial{D}.\end{cases} (4.22)

and v0v_{0}^{*} satisfy

{λ,μv0=0,inΩ,v0=0,onD1D2,v0=φ,onD,\begin{cases}\mathcal{L}_{\lambda,\mu}v_{0}^{*}=0,&\mathrm{in}~{}\Omega^{*},\\ v_{0}^{*}=0,&\mathrm{on}~{}\partial{D}_{1}^{*}\cup\partial{D_{2}^{*}},\\ v_{0}^{*}=\varphi,&\mathrm{on}~{}\partial{D},\end{cases} (4.23)

By the definition of uu^{*} in (3.1), we find that

u=α=13Cαvα+v0.u^{*}=\sum_{\alpha=1}^{3}C_{*}^{\alpha}v^{*\alpha}+v_{0}^{*}.

From the third line of (3.1), we have

α=13Cαaαβ(b~1β+b~2β)=0,β=1,2,3,\displaystyle\sum_{\alpha=1}^{3}C_{*}^{\alpha}a_{*}^{\alpha\beta}-(\tilde{b}_{1}^{*\beta}+\tilde{b}_{2}^{*\beta})=0,\quad\beta=1,2,3, (4.24)

where

aαβ=D1D2vαν|+ψβ,andb~jβ=Djv0ν|+ψβ,j=1,2.\displaystyle a_{*}^{\alpha\beta}=-\int_{\partial{D}_{1}^{*}\cup\partial{D}_{2}^{*}}\frac{\partial{v}^{*\alpha}}{\partial\nu}\big{|}_{+}\cdot\psi_{\beta},\quad\mbox{and}~{}~{}\tilde{b}_{j}^{*\beta}=\int_{\partial{D}_{j}^{*}}\frac{\partial{v}_{0}^{*}}{\partial\nu}\big{|}_{+}\cdot\psi_{\beta},\quad j=1,2. (4.25)

We next use the symmetry properties of the domain and the boundary data to prove the convergence of C1α+C2α2Cα\frac{C_{1}^{\alpha}+C_{2}^{\alpha}}{2}-C_{*}^{\alpha}, Proposition 4.4 below. Before this, we first prove the following two lemmas.

Lemma 4.2.

Let b~1β\tilde{b}_{1}^{\beta} and b~1β\tilde{b}_{1}^{*\beta} be defined in (4.17) and (4.25), respectively. Then

|b~1βb~1β|CφL1(D){ε1/2,β=1,,d,ε2/3,β=d+1,,d(d+1)2.\left|\tilde{b}_{1}^{\beta}-\tilde{b}_{1}^{*\beta}\right|\leq C\|\varphi\|_{L^{1}(\partial D)}\begin{cases}\varepsilon^{1/2},~{}\beta=1,\cdots,d,\\ \varepsilon^{2/3},~{}\beta=d+1,\cdots,\frac{d(d+1)}{2}.\end{cases} (4.26)
Proof.

It follows from (2.6), (4.23), and the integration by parts that

b~1β=D1v0ν|+ψβ\displaystyle\tilde{b}_{1}^{\beta}=\int_{\partial D_{1}}\frac{\partial v_{0}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta} =D1v0ν|+v1β=Dv1βν|+φ,\displaystyle=\int_{\partial D_{1}}\frac{\partial v_{0}}{\partial\nu}\Big{|}_{+}\cdot v_{1}^{\beta}=-\int_{\partial D}\frac{\partial v_{1}^{\beta}}{\partial\nu}\Big{|}_{+}\cdot\varphi,

and

b~1β=D1v0ν|+ψβ\displaystyle\tilde{b}_{1}^{*\beta}=\int_{\partial D_{1}^{*}}\frac{\partial v_{0}^{*}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta} =Dv1βν|+φ.\displaystyle=-\int_{\partial D}\frac{\partial v_{1}^{*\beta}}{\partial\nu}\Big{|}_{+}\cdot\varphi.

Thus, we have

b~1βb~1β=D(v1βv1β)ν|+φ.\tilde{b}_{1}^{\beta}-\tilde{b}_{1}^{*\beta}=-\int_{\partial D}\frac{\partial(v_{1}^{\beta}-v_{1}^{*\beta})}{\partial\nu}\Big{|}_{+}\cdot\varphi. (4.27)

By using Lemma 4.1 and the standard boundary gradient estimates for Lamé system (see [4]), we have (4.26). We hence finish the proof of Lemma 4.2. ∎

Similar to Lemma 4.2, we can get

Lemma 4.3.

Let vαv^{\alpha} and vαv^{*\alpha} be defined by (4.21) and (4.22), respectively, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2. Then

|D1vαν|+ψβD1vαν|+ψβ|CψαL(D){ε1/2,β=1,,d,ε2/3,β=d+1,,d(d+1)2.\displaystyle\left|\int_{\partial D_{1}}\frac{\partial v^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}-\int_{\partial D_{1}^{*}}\frac{\partial v^{*\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}\right|\leq C\|\psi_{\alpha}\|_{L^{\infty}(\partial D)}\begin{cases}\varepsilon^{1/2},~{}\beta=1,\cdots,d,\\ \varepsilon^{2/3},~{}\beta=d+1,\cdots,\frac{d(d+1)}{2}.\end{cases} (4.28)
Proof.

For α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2, it follows from (4.21) that

{λ,μ(vαψα)=0,inΩ,vαψα=0,onD1D2,vαψα=ψα,onD.\begin{cases}\mathcal{L}_{\lambda,\mu}(v^{\alpha}-\psi_{\alpha})=0,&\mbox{in}~{}\Omega,\\ v^{\alpha}-\psi_{\alpha}=0,&\mbox{on}~{}\partial{D}_{1}\cup\partial{D}_{2},\\ v^{\alpha}-\psi_{\alpha}=-\psi_{\alpha},&\mbox{on}~{}\partial D.\end{cases}

By using the integration by parts, we have for α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2,

D1vαν|+ψβ=D1(vαψα)ν|+v1β=Dv1βν|+ψα.\displaystyle\int_{\partial D_{1}}\frac{\partial v^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}=\int_{\partial D_{1}}\frac{\partial(v^{\alpha}-\psi_{\alpha})}{\partial\nu}\Big{|}_{+}\cdot v_{1}^{\beta}=\int_{\partial D}\frac{\partial v_{1}^{\beta}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\alpha}.

Similarly,

D1vαν|+ψβ=Dv1βν|+ψα.\displaystyle\int_{\partial D_{1}^{*}}\frac{\partial v^{*\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}=\int_{\partial D}\frac{\partial v_{1}^{*\beta}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\alpha}.

Hence,

D1vαν|+ψβD1vαν|+ψβ=D(v1βv1β)ν|+ψα.\int_{\partial D_{1}}\frac{\partial v^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}-\int_{\partial D_{1}^{*}}\frac{\partial v^{*\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}=\int_{\partial D}\frac{\partial(v_{1}^{\beta}-v_{1}^{*\beta})}{\partial\nu}\Big{|}_{+}\cdot\psi_{\alpha}.

Thus, similar to (4.26), we obtain (4.28). The proof of Lemma 4.3 is finished. ∎

Proposition 4.4.

For d=2,3d=2,3, assume that D1D2D_{1}\cup D_{2} and DD satisfies (S1)({\rm S_{1}}), and φ\varphi satisfies (S2)({\rm S_{2}}) or (S3)({\rm S_{3}}). Let C1α,C2αC_{1}^{\alpha},C_{2}^{\alpha}, and CαC_{*}^{\alpha} be defined in (2.3) and (3.1), respectively. Then

|C1α+C2α2Cα|Cρd(ε),α=1,,d(d+1)/2.\left|\frac{C_{1}^{\alpha}+C_{2}^{\alpha}}{2}-C_{*}^{\alpha}\right|\leq C\rho_{d}(\varepsilon),\quad\alpha=1,\cdots,d(d+1)/2. (4.29)
Proof.

(1) We first prove the case when d=2d=2. First, on one hand, using the symmetry of the domain with respect to the origin and the boundary conditions of viαv_{i}^{\alpha}, we have, for α=1,2\alpha=1,2,

v2α(x1,x2)|D1=v1α(x1,x2)|D2\displaystyle v_{2}^{\alpha}(x_{1},x_{2})\big{|}_{\partial D_{1}}=v_{1}^{\alpha}(-x_{1},-x_{2})\big{|}_{\partial D_{2}} =0,v2α(x1,x2)|D2=v1α(x1,x2)|D1=ψα,\displaystyle=0,~{}~{}v_{2}^{\alpha}(x_{1},x_{2})\big{|}_{\partial D_{2}}=v_{1}^{\alpha}(-x_{1},-x_{2})\big{|}_{\partial D_{1}}=\psi_{\alpha},
v2α(x1,x2)|D\displaystyle v_{2}^{\alpha}(x_{1},x_{2})\big{|}_{\partial D} =v1α(x1,x2)|D=0.\displaystyle=v_{1}^{\alpha}(-x_{1},-x_{2})\big{|}_{\partial D}=0.

The using the equation λ,μviα=0\mathcal{L}_{\lambda,\mu}v_{i}^{\alpha}=0, one can see that

v2α(x1,x2)=v1α(x1,x2),inΩ,α=1,2.v_{2}^{\alpha}(x_{1},x_{2})=v_{1}^{\alpha}(-x_{1},-x_{2}),\quad\mbox{in}~{}\Omega,\quad~{}\alpha=1,2.

Therefore,

a11αβ=a22αβ,a12αβ=a21αβ,α,β=1,2.a_{11}^{\alpha\beta}=a_{22}^{\alpha\beta},\quad a_{12}^{\alpha\beta}=a_{21}^{\alpha\beta},\quad~{}\alpha,\beta=1,2. (4.30)

Similarly, by taking

v23(x1,x2)|D1=v13(x1,x2)|D2\displaystyle v_{2}^{3}(x_{1},x_{2})\big{|}_{\partial D_{1}}=-v_{1}^{3}(-x_{1},-x_{2})\big{|}_{\partial D_{2}} =0,v23(x1,x2)|D2=v13(x1,x2)|D1=ψ3,\displaystyle=0,~{}~{}v_{2}^{3}(x_{1},x_{2})\big{|}_{\partial D_{2}}=-v_{1}^{3}(-x_{1},-x_{2})\big{|}_{\partial D_{1}}=\psi_{3},
v23(x1,x2)|D\displaystyle v_{2}^{3}(x_{1},x_{2})\big{|}_{\partial D} =v13(x1,x2)|D=0,\displaystyle=-v_{1}^{3}(-x_{1},-x_{2})\big{|}_{\partial D}=0,

we have v23(x1,x2)=v13(x1,x2)v_{2}^{3}(x_{1},x_{2})=-v_{1}^{3}(-x_{1},-x_{2}) are the solutions of (2.5). Thus,

a11α3=a22α3,a1133=a2233,a12α3=a21α3,a123α=a213α,α=1,2.\displaystyle a_{11}^{\alpha 3}=-a_{22}^{\alpha 3},\quad a_{11}^{33}=a_{22}^{33},\quad a_{12}^{\alpha 3}=-a_{21}^{\alpha 3},\quad a_{12}^{3\alpha}=-a_{21}^{3\alpha},\quad\quad~{}~{}\alpha=1,2. (4.31)

On the other hand, by using the symmetry of the domain with respect to {x2=0}\{x_{2}=0\} and the boundary conditions of viαv_{i}^{\alpha}, we find that

v22(x1,x2)=((v22)1(x1,x2)(v22)2(x1,x2))=((v12)1(x1,x2)(v12)2(x1,x2))\displaystyle v_{2}^{2}(x_{1},x_{2})=\begin{pmatrix}(v_{2}^{2})^{1}(x_{1},x_{2})\\ \\ (v_{2}^{2})^{2}(x_{1},x_{2})\end{pmatrix}=\begin{pmatrix}-(v_{1}^{2})^{1}(x_{1},-x_{2})\\ \\ (v_{1}^{2})^{2}(x_{1},-x_{2})\end{pmatrix}

and

v23(x1,x2)=((v23)1(x1,x2)(v23)2(x1,x2))=((v13)1(x1,x2)(v13)2(x1,x2))\displaystyle v_{2}^{3}(x_{1},x_{2})=\begin{pmatrix}(v_{2}^{3})^{1}(x_{1},x_{2})\\ \\ (v_{2}^{3})^{2}(x_{1},x_{2})\end{pmatrix}=\begin{pmatrix}-(v_{1}^{3})^{1}(x_{1},-x_{2})\\ \\ (v_{1}^{3})^{2}(x_{1},-x_{2})\end{pmatrix}

are also the solutions of (2.5), respectively. Then we have

((0)e(v22),e(v23))\displaystyle\left(\mathbb{C}^{(0)}e(v_{2}^{2}),e(v_{2}^{3})\right) =λ(x1(v12)1x2(v12)2)(x1(v13)1x2(v13)2)\displaystyle=\lambda\left(-\partial_{x_{1}}(v_{1}^{2})^{1}-\partial_{x_{2}}(v_{1}^{2})^{2}\right)\left(-\partial_{x_{1}}(v_{1}^{3})^{1}-\partial_{x_{2}}(v_{1}^{3})^{2}\right)
+μ(2x1(v12)1x1(v13)1+2x2(v12)2x2(v13)2\displaystyle\quad+\mu\Big{(}2\partial_{x_{1}}(v_{1}^{2})^{1}\cdot\partial_{x_{1}}(v_{1}^{3})^{1}+2\partial_{x_{2}}(v_{1}^{2})^{2}\cdot\partial_{x_{2}}(v_{1}^{3})^{2}
+(x2(v12)1+x1(v12)2)(x2(v13)1+x1(v13)2))\displaystyle\quad+(\partial_{x_{2}}(v_{1}^{2})^{1}+\partial_{x_{1}}(v_{1}^{2})^{2})(\partial_{x_{2}}(v_{1}^{3})^{1}+\partial_{x_{1}}(v_{1}^{3})^{2})\Big{)}
=((0)e(v12),e(v13)).\displaystyle=\left(\mathbb{C}^{(0)}e(v_{1}^{2}),e(v_{1}^{3})\right). (4.32)

Thus,

a2223=Ω((0)e(v22),e(v23))𝑑x=Ω((0)e(v12),e(v13))𝑑x=a1123.a_{22}^{23}=\int_{\Omega}\left(\mathbb{C}^{(0)}e(v_{2}^{2}),e(v_{2}^{3})\right)dx=\int_{\Omega}\left(\mathbb{C}^{(0)}e(v_{1}^{2}),e(v_{1}^{3})\right)dx=a_{11}^{23}. (4.33)

Similarly,

v21(x1,x2)=((v21)1(x1,x2)(v21)2(x1,x2))=((v11)1(x1,x2)(v11)2(x1,x2))\displaystyle v_{2}^{1}(x_{1},x_{2})=\begin{pmatrix}(v_{2}^{1})^{1}(x_{1},x_{2})\\ \\ (v_{2}^{1})^{2}(x_{1},x_{2})\end{pmatrix}=\begin{pmatrix}(v_{1}^{1})^{1}(x_{1},-x_{2})\\ \\ -(v_{1}^{1})^{2}(x_{1},-x_{2})\end{pmatrix}

admits (2.5). Then

((0)e(v21),e(v23))\displaystyle\left(\mathbb{C}^{(0)}e(v_{2}^{1}),e(v_{2}^{3})\right) =λ(x1(v11)1+x2(v11)2)(x1(v13)1x2(v13)2)\displaystyle=\lambda\left(\partial_{x_{1}}(v_{1}^{1})^{1}+\partial_{x_{2}}(v_{1}^{1})^{2}\right)\left(-\partial_{x_{1}}(v_{1}^{3})^{1}-\partial_{x_{2}}(v_{1}^{3})^{2}\right)
+μ(2x1(v11)1x1(v13)12x2(v11)2x2(v13)2\displaystyle\quad+\mu\Big{(}-2\partial_{x_{1}}(v_{1}^{1})^{1}\cdot\partial_{x_{1}}(v_{1}^{3})^{1}-2\partial_{x_{2}}(v_{1}^{1})^{2}\cdot\partial_{x_{2}}(v_{1}^{3})^{2}
+(x2(v11)1x1(v11)2)(x2(v13)1+x1(v13)2))\displaystyle\quad+(-\partial_{x_{2}}(v_{1}^{1})^{1}-\partial_{x_{1}}(v_{1}^{1})^{2})(\partial_{x_{2}}(v_{1}^{3})^{1}+\partial_{x_{1}}(v_{1}^{3})^{2})\Big{)}
=((0)e(v11),e(v13)).\displaystyle=-\left(\mathbb{C}^{(0)}e(v_{1}^{1}),e(v_{1}^{3})\right).

Hence,

a2213=a1113.a_{22}^{13}=-a_{11}^{13}. (4.34)

Similarly, we have

a2212=a1112,a1212=a2112,a1221=a2121,a1213=a2113,a1231=a2131,a1223=a2123,a1232=a2132.\begin{split}&a_{22}^{12}=-a_{11}^{12},\quad a_{12}^{12}=-a_{21}^{12},\quad a_{12}^{21}=-a_{21}^{21},\\ &a_{12}^{13}=-a_{21}^{13},\quad a_{12}^{31}=-a_{21}^{31},\quad a_{12}^{23}=a_{21}^{23},\quad a_{12}^{32}=a_{21}^{32}.\end{split} (4.35)

Combining (4.30)–(4.35), we obtain

a1112=a2212=a1212=a2112=a1221=a2121=a1123=a2223=a1223=a2123=a1232=a2132=0.a_{11}^{12}=a_{22}^{12}=a_{12}^{12}=a_{21}^{12}=a_{12}^{21}=a_{21}^{21}=a_{11}^{23}=a_{22}^{23}=a_{12}^{23}=a_{21}^{23}=a_{12}^{32}=a_{21}^{32}=0.

Therefore, we obtain

A:=(a11αβa21αβa21αβa22αβ)α,β=1,2,3=(a11110a1113a12110a12130a112200a12220a11130a1133a12130a1233a12110a1213a11110a11130a122200a11220a12130a1233a11130a1133).\displaystyle A:=\begin{pmatrix}~{}a_{11}^{\alpha\beta}&a_{21}^{\alpha\beta}~{}\\ \\ ~{}a_{21}^{\alpha\beta}&a_{22}^{\alpha\beta}~{}\end{pmatrix}_{\alpha,\beta=1,2,3}=\begin{pmatrix}~{}a_{11}^{11}&0&a_{11}^{13}&a_{12}^{11}&0&a_{12}^{13}~{}\\ \\ ~{}0&a_{11}^{22}&0&0&a_{12}^{22}&0~{}\\ \\ ~{}a_{11}^{13}&0&a_{11}^{33}&-a_{12}^{13}&0&a_{12}^{33}~{}\\ \\ ~{}a_{12}^{11}&0&-a_{12}^{13}&a_{11}^{11}&0&-a_{11}^{13}~{}\\ \\ ~{}0&a_{12}^{22}&0&0&a_{11}^{22}&0~{}\\ \\ ~{}a_{12}^{13}&0&a_{12}^{33}&-a_{11}^{13}&0&a_{11}^{33}\end{pmatrix}. (4.36)

By (S2)({\rm S_{2}}), a direct calculation which is similar to (4.2) yields

b~11=b~21,b~12=b~22,b~13=b~23.\tilde{b}_{1}^{1}=-\tilde{b}_{2}^{1},\quad\tilde{b}_{1}^{2}=\tilde{b}_{2}^{2},\quad\tilde{b}_{1}^{3}=\tilde{b}_{2}^{3}. (4.37)

By (4.16), we have

AX=b,AX=b,

where

X=(C11,C12,C13,C21,C22,C23)T, and b=(b~11,b~12,b~13,b~21,b~22,b~23)T.X=(C_{1}^{1},C_{1}^{2},C_{1}^{3},C_{2}^{1},C_{2}^{2},C_{2}^{3})^{\mathrm{T}},\quad\mbox{ and }~{}~{}b=(\tilde{b}_{1}^{1},\tilde{b}_{1}^{2},\tilde{b}_{1}^{3},\tilde{b}_{2}^{1},\tilde{b}_{2}^{2},\tilde{b}_{2}^{3})^{\mathrm{T}}.

By Cramer’s rule and a direct calculation, we get

C13\displaystyle C_{1}^{3} =(a1122)2(a1222)2detA|a1111b~11a1211a1213a1113b~13a1213a1233a1211b~11a1111a1113a1213b~13a1113a1133|\displaystyle=\frac{(a_{11}^{22})^{2}-(a_{12}^{22})^{2}}{\det{A}}\begin{vmatrix}~{}a_{11}^{11}&\tilde{b}_{1}^{1}&a_{12}^{11}&a_{12}^{13}~{}\\ \\ ~{}a_{11}^{13}&\tilde{b}_{1}^{3}&-a_{12}^{13}&a_{12}^{33}~{}\\ \\ ~{}a_{12}^{11}&-\tilde{b}_{1}^{1}&a_{11}^{11}&-a_{11}^{13}~{}\\ \\ ~{}a_{12}^{13}&\tilde{b}_{1}^{3}&-a_{11}^{13}&a_{11}^{33}\end{vmatrix}
=(a1122)2(a1222)2detA(|a1111a1211b~11a1113+a1213b~13||a1111+a1211a1213a1113a1213a1113a1133a1233|),\displaystyle=\frac{(a_{11}^{22})^{2}-(a_{12}^{22})^{2}}{\det{A}}\left(\begin{vmatrix}~{}a_{11}^{11}-a_{12}^{11}&\tilde{b}_{1}^{1}~{}\\ \\ ~{}a_{11}^{13}+a_{12}^{13}&\tilde{b}_{1}^{3}\end{vmatrix}\cdot\begin{vmatrix}~{}a_{11}^{11}+a_{12}^{11}&a_{12}^{13}-a_{11}^{13}~{}\\ \\ ~{}a_{12}^{13}-a_{11}^{13}&a_{11}^{33}-a_{12}^{33}\end{vmatrix}\right),

and similarly,

C23\displaystyle C_{2}^{3} =(a1122)2(a1222)2detA(|a1111a1211b~11a1113a1213b~13||a1111+a1211a1113a1213a1113a1213a1133a1233|).\displaystyle=\frac{(a_{11}^{22})^{2}-(a_{12}^{22})^{2}}{\det{A}}\left(\begin{vmatrix}~{}a_{11}^{11}-a_{12}^{11}&-\tilde{b}_{1}^{1}~{}\\ \\ ~{}-a_{11}^{13}-a_{12}^{13}&\tilde{b}_{1}^{3}\end{vmatrix}\cdot\begin{vmatrix}~{}a_{11}^{11}+a_{12}^{11}&a_{11}^{13}-a_{12}^{13}~{}\\ \\ ~{}a_{11}^{13}-a_{12}^{13}&a_{11}^{33}-a_{12}^{33}\end{vmatrix}\right).

Hence,

C13=C23.\displaystyle C_{1}^{3}=C_{2}^{3}. (4.38)

Thus, by using (4.36)–(4.38), (4.20) becomes

(a11000a22000a33)(C11+C212C12+C222C13+C232)=(02b~122b~13(C11C21)(a1113+a1213)).\displaystyle\begin{pmatrix}~{}a^{11}&0&0~{}\\ \\ ~{}0&a^{22}&0~{}\\ \\ ~{}0&0&a^{33}\end{pmatrix}\begin{pmatrix}~{}\frac{C_{1}^{1}+C_{2}^{1}}{2}~{}\\ \\ \frac{C_{1}^{2}+C_{2}^{2}}{2}\\ \\ \frac{C_{1}^{3}+C_{2}^{3}}{2}\end{pmatrix}=\begin{pmatrix}~{}0~{}\\ \\ 2\tilde{b}_{1}^{2}\\ \\ 2\tilde{b}_{1}^{3}-(C_{1}^{1}-C_{2}^{1})(a_{11}^{13}+a_{12}^{13})\end{pmatrix}. (4.39)

Similar to (4.36) and (4.37), by replicating the computation used in (4.2), we get

a12=a21=a13=a31=a23=a32=0a_{*}^{12}=a_{*}^{21}=a_{*}^{13}=a_{*}^{31}=a_{*}^{23}=a_{*}^{32}=0

and

b~11=b~21,b~12=b~22,b~13=b~23.\tilde{b}_{1}^{*1}=-\tilde{b}_{2}^{*1},\quad\tilde{b}_{1}^{*2}=\tilde{b}_{2}^{*2},\quad\tilde{b}_{1}^{*3}=\tilde{b}_{2}^{*3}.

Then (4.24) becomes

(a11000a22000a33)(C1C2C3)=(02b~122b~13).\displaystyle\begin{pmatrix}~{}a_{*}^{11}&0&0~{}\\ \\ ~{}0&a_{*}^{22}&0~{}\\ \\ ~{}0&0&a_{*}^{33}\end{pmatrix}\begin{pmatrix}~{}C_{*}^{1}~{}\\ \\ C_{*}^{2}\\ \\ C_{*}^{3}\end{pmatrix}=\begin{pmatrix}~{}0~{}\\ \\ 2\tilde{b}_{1}^{*2}\\ \\ 2\tilde{b}_{1}^{*3}\end{pmatrix}. (4.40)

Finally, we turn to the estimate of C1α+C2α2Cα\frac{C_{1}^{\alpha}+C_{2}^{\alpha}}{2}-C_{*}^{\alpha}, α=1,2,3\alpha=1,2,3. Combining (4.39) and (4.40), we have

(a11000a22000a33)(C11+C212C1C12+C222C2C13+C232C3)=(123),\displaystyle\begin{pmatrix}~{}a^{11}&0&0~{}\\ \\ ~{}0&a^{22}&0~{}\\ \\ ~{}0&0&a^{33}\end{pmatrix}\begin{pmatrix}~{}\frac{C_{1}^{1}+C_{2}^{1}}{2}-C_{*}^{1}~{}\\ \\ \frac{C_{1}^{2}+C_{2}^{2}}{2}-C_{*}^{2}\\ \\ \frac{C_{1}^{3}+C_{2}^{3}}{2}-C_{*}^{3}\end{pmatrix}=\begin{pmatrix}~{}\mathcal{B}^{1}~{}\\ \\ \mathcal{B}^{2}\\ \\ \mathcal{B}^{3}\end{pmatrix}, (4.41)

where

1\displaystyle\mathcal{B}^{1} =C1(a11a11)=O(ε),\displaystyle=-C_{*}^{1}(a^{11}-a_{*}^{11})=O(\sqrt{\varepsilon}),
2\displaystyle\mathcal{B}^{2} =2(b~12b~12)C2(a22a22)=O(ε),\displaystyle=2(\tilde{b}_{1}^{2}-\tilde{b}_{1}^{*2})-C_{*}^{2}(a^{22}-a_{*}^{22})=O(\sqrt{\varepsilon}),
3\displaystyle\mathcal{B}^{3} =2(b~13b~13)C3(a33a33)(C11C21)(a1113+a1213)=O(ε),\displaystyle=2(\tilde{b}_{1}^{3}-\tilde{b}_{1}^{*3})-C_{*}^{3}(a^{33}-a_{*}^{33})-(C_{1}^{1}-C_{2}^{1})(a_{11}^{13}+a_{12}^{13})=O(\sqrt{\varepsilon}),

here we used Lemmas 4.2 and 4.3. By Cramer’s rule, we obtain

C1α+C2α2Cα=O(ε),α=1,2,3.\displaystyle\frac{C_{1}^{\alpha}+C_{2}^{\alpha}}{2}-C_{*}^{\alpha}=O(\sqrt{\varepsilon}),\quad\alpha=1,2,3.

(2) For d=3d=3, we can similarly prove by using the symmetry property of domain that

C1α=C2α,α=4,5,6.C_{1}^{\alpha}=C_{2}^{\alpha},\quad\alpha=4,5,6.

We omit the details here. Similar to (4.41), we have

(a11000000a66)(C11+C212C1C16+C262C6)=(16),\displaystyle\begin{pmatrix}~{}a^{11}&0&0&0~{}\\ \\ ~{}\vdots&\vdots&\vdots&\vdots~{}\\ \\ ~{}0&0&0&a^{66}\end{pmatrix}\begin{pmatrix}~{}\frac{C_{1}^{1}+C_{2}^{1}}{2}-C_{*}^{1}~{}\\ \\ \vdots\\ \\ \frac{C_{1}^{6}+C_{2}^{6}}{2}-C_{*}^{6}\end{pmatrix}=\begin{pmatrix}~{}\mathcal{B}^{1}~{}\\ \\ \vdots\\ \\ \mathcal{B}^{6}\end{pmatrix},

where β=O(|logε|1)\mathcal{B}^{\beta}=O(|\log\varepsilon|^{-1}) as ε0\varepsilon\rightarrow 0, β=1,,6\beta=1,\cdots,6. Hence, by Cramer’s rule, we get (4.29) for d=3d=3. Proposition 4.4 is thus proved. ∎

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 b1β[φ]b_{1}^{\beta}[\varphi] and b1β[φ]b_{1}^{*\beta}[\varphi], (3.2), we have

b1β[φ]b1β[φ]=D1ubν|+ψβD1uν|+ψβ\displaystyle b_{1}^{\beta}[\varphi]-b_{1}^{*\beta}[\varphi]=\int_{\partial{D}_{1}}\frac{\partial{u}_{b}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}-\int_{\partial{D}_{1}^{*}}\frac{\partial u^{*}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}
=D1v0ν|+ψβD1v0ν|+ψβ+α=1d(d+1)/2C2αD1vαν|+ψβ\displaystyle=\int_{\partial D_{1}}\frac{\partial v_{0}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}-\int_{\partial D_{1}^{*}}\frac{\partial v_{0}^{*}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}+\sum_{\alpha=1}^{d(d+1)/2}C_{2}^{\alpha}\int_{\partial D_{1}}\frac{\partial v^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}
α=1d(d+1)/2CαD1vαν|+ψβ\displaystyle\quad-\sum_{\alpha=1}^{d(d+1)/2}C_{*}^{\alpha}\int_{\partial D_{1}^{*}}\frac{\partial v^{*\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}
=D1v0ν|+ψβD1v0ν|+ψβ+α=1d(d+1)/2C2α(D1vαν|+ψβ\displaystyle=\int_{\partial D_{1}}\frac{\partial v_{0}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}-\int_{\partial D_{1}^{*}}\frac{\partial v_{0}^{*}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}+\sum_{\alpha=1}^{d(d+1)/2}C_{2}^{\alpha}\Bigg{(}\int_{\partial D_{1}}\frac{\partial v^{\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}
D1vαν|+ψβ)+α=1d(d+1)/2(C2αCα)D1vαν|+ψβ,\displaystyle\quad-\int_{\partial D_{1}^{*}}\frac{\partial v^{*\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}\Bigg{)}+\sum_{\alpha=1}^{d(d+1)/2}\Big{(}C_{2}^{\alpha}-C_{*}^{\alpha}\Big{)}\int_{\partial D_{1}^{*}}\frac{\partial v^{*\alpha}}{\partial\nu}\Big{|}_{+}\cdot\psi_{\beta}, (4.42)

where v0v_{0}, v0v_{0}^{*}, vαv^{\alpha}, and vαv^{*\alpha} are, respectively, defined by (2.6), (4.23), (4.21), and (4.22). From the proof of Proposition 4.4, C1α=C2αC_{1}^{\alpha}=C_{2}^{\alpha} for α=d+1,,d(d+1)/2\alpha=d+1,\cdots,d(d+1)/2, and the estimates of C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha} (see for instance, [10, Proposition 4.2] and[11, Proposition 4.1]), we have

|CαC2α|Cρd(ε),α=1,,d(d+1)/2.|C_{*}^{\alpha}-C_{2}^{\alpha}|\leq C\rho_{d}(\varepsilon),\quad\alpha=1,\cdots,d(d+1)/2. (4.43)

By using (4.3), (4.43), Lemmas 4.2 and 4.3, we have

|b1β[φ]b1β[φ]|Cρd(ε),\displaystyle|b_{1}^{\beta}[\varphi]-b_{1}^{*\beta}[\varphi]|\leq C\rho_{d}(\varepsilon), (4.44)

here we used the fact that |vα|C|\nabla v^{*\alpha}|\leq C in Ω\Omega^{*}, see Theorem 2.1. Therefore, Proposition 3.1 is proved.

4.4. Proof of Proposition 2.3 (The symptotics of a11ααa_{11}^{\alpha\alpha})

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 d=2d=2..

(1) First consider a1111a_{11}^{11}. We divide it into three parts:

a1111\displaystyle a_{11}^{11} =Ω(0e(v11),e(v11))𝑑x\displaystyle=\int_{\Omega}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx
=ΩΩR(0e(v11),e(v11))𝑑x+ΩRΩε1/8(0e(v11),e(v11))𝑑x\displaystyle=\int_{\Omega\setminus\Omega_{R}}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx+\int_{\Omega_{R}\setminus\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx
+Ωε1/8(0e(v11),e(v11))dx=:I1+I2+I3.\displaystyle\quad+\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx=:\mbox{I}_{1}+\mbox{I}_{2}+\mbox{I}_{3}. (4.45)

In the follow we estimate I1\mbox{I}_{1}, I2\mbox{I}_{2}, and I3\mbox{I}_{3} one by one.

Step 1. Claim: There exists a constant M1M_{1}^{*}, independent of ε\varepsilon, such that

I1=M1+O(ε1/4).\mbox{I}_{1}=M_{1}^{*}+O(\varepsilon^{1/4}). (4.46)

Notice that

λ,μ(v11v11)=0,xDD1D2D1D2ΩR,\mathcal{L}_{\lambda,\mu}(v_{1}^{1}-v_{1}^{*1})=0,\quad x\in D\setminus D_{1}\cup D_{2}\cup D_{1}^{*}\cup D_{2}^{*}\cup\Omega_{R},

and

0|v11|,|v11|1,xDD1D2D1D2ΩR.0\leq|v_{1}^{1}|,|v_{1}^{*1}|\leq 1,\quad x\in D\setminus D_{1}\cup D_{2}\cup D_{1}^{*}\cup D_{2}^{*}\cup\Omega_{R}.

Then since D1,D1,D2\partial D_{1},\partial D_{1}^{*},\partial D_{2}, and D\partial D are C2,αC^{2,\alpha}, we have

|2(v11v11)||2v11|+|2v11|CinD(D1D1D2D2ΩR).|\nabla^{2}(v_{1}^{1}-v_{1}^{*1})|\leq|\nabla^{2}v_{1}^{1}|+|\nabla^{2}v_{1}^{*1}|\leq C\quad\mbox{in}~{}D\setminus(D_{1}\cup D_{1}^{*}\cup D_{2}\cup D_{2}^{*}\cup\Omega_{R}). (4.47)

Moreover, we obtain from (4.2) that

v11v11L(D(D1D1D2D2Ωε1/4))Cε1/2.\|v_{1}^{1}-v_{1}^{*1}\|_{L^{\infty}(D\setminus(D_{1}\cup D_{1}^{*}\cup D_{2}\cup D_{2}^{*}\cup\Omega_{\varepsilon^{1/4}}))}\leq C\varepsilon^{1/2}. (4.48)

By using the interpolation inequality, (4.47), and (4.48), we obtain

|(v11v11)|Cε1/2(112)=Cε1/4inD(D1D1D2D2ΩR).|\nabla(v_{1}^{1}-v_{1}^{*1})|\leq C\varepsilon^{1/2(1-\frac{1}{2})}=C\varepsilon^{1/4}\quad\mbox{in}~{}D\setminus(D_{1}\cup D_{1}^{*}\cup D_{2}\cup D_{2}^{*}\cup\Omega_{R}). (4.49)

Denote

M1:=ΩΩR(0e(v11),e(v11))𝑑x.M_{1}^{*}:=\int_{\Omega^{*}\setminus\Omega_{R}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{*1})\big{)}\ dx.

Then

I1M1\displaystyle\mbox{I}_{1}-M_{1}^{*} =Ω(D1ΩR)((0e(v11),e(v11))(0e(v11),e(v11)))𝑑x\displaystyle=\int_{\Omega\setminus(D_{1}^{*}\cup\Omega_{R})}\Big{(}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}-\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{*1})\big{)}\Big{)}\ dx
+D1(D1ΩR)(0e(v11),e(v11))𝑑xD1D1(0e(v11),e(v11))𝑑x.\displaystyle\quad+\int_{D_{1}^{*}\setminus(D_{1}\cup\Omega_{R})}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx-\int_{D_{1}\setminus D_{1}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{*1})\big{)}\ dx.

It follows from |D1(D1ΩR)|Cε|D_{1}^{*}\setminus(D_{1}\cup\Omega_{R})|\leq C\varepsilon, |D1D1|Cε|D_{1}\setminus D_{1}^{*}|\leq C\varepsilon, and the boundedness of |v11||\nabla v_{1}^{1}| and |v11||\nabla v_{1}^{*1}| in D1(D1ΩR)D_{1}^{*}\setminus(D_{1}\cup\Omega_{R}) and D1D1D_{1}\setminus D_{1}^{*}, respectively, that

|D1(D1ΩR)(0e(v11),e(v11))𝑑xD1D1(0e(v11),e(v11))𝑑x|Cε.\displaystyle\left|\int_{D_{1}^{*}\setminus(D_{1}\cup\Omega_{R})}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx-\int_{D_{1}\setminus D_{1}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{*1})\big{)}\ dx\right|\leq C\varepsilon. (4.50)

So by using (4.49) and (4.50), we have

I1M1\displaystyle\mbox{I}_{1}-M_{1}^{*} =Ω(D1ΩR)(0e(v11v11),e(v11v11))𝑑x\displaystyle=\int_{\Omega\setminus(D_{1}^{*}\cup\Omega_{R})}\big{(}\mathbb{C}^{0}e(v_{1}^{1}-v_{1}^{*1}),e(v_{1}^{1}-v_{1}^{*1})\big{)}\ dx
+2Ω(D1ΩR)(0e(v11),e(v11v11))𝑑x+O(ε)\displaystyle\quad+2\int_{\Omega\setminus(D_{1}^{*}\cup\Omega_{R})}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{1}-v_{1}^{*1})\big{)}\ dx+O(\varepsilon)
=O(ε1/4).\displaystyle=O(\varepsilon^{1/4}).

We henceforth get (4.46).

Step 2. Proof of

I2=I2+O(ε1/8),\mbox{I}_{2}=\mbox{I}_{2}^{*}+O(\varepsilon^{1/8}), (4.51)

where

I2=ΩRΩε1/8(0e(v11),e(v11))𝑑x.\mbox{I}_{2}^{*}=\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{*1})\big{)}\ dx.

We further divide I2I2\mbox{I}_{2}-\mbox{I}_{2}^{*} into three terms:

I2I2\displaystyle\mbox{I}_{2}-\mbox{I}_{2}^{*}
=(ΩRΩε1/8)(ΩRΩε1/8)(0e(v11),e(v11))𝑑x+2ΩRΩε1/8(0e(v11),e(v11v11))𝑑x\displaystyle=\int_{(\Omega_{R}\setminus\Omega_{\varepsilon^{1/8}})\setminus(\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*})}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx+2\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}),e(v_{1}^{1}-v_{1}^{*1})\big{)}\ dx
+ΩRΩε1/8(0e(v11v11),e(v11v11))𝑑x\displaystyle\quad+\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{1}-v_{1}^{*1}),e(v_{1}^{1}-v_{1}^{*1})\big{)}\ dx
=:I2,1+I2,2+I2,3.\displaystyle=:\mbox{I}_{2,1}+\mbox{I}_{2,2}+\mbox{I}_{2,3}. (4.52)

For ε1/8|z1|R\varepsilon^{1/8}\leq|z_{1}|\leq R, we rescale Ω|z1|+|z1|2Ω|z1|\Omega_{|z_{1}|+|z_{1}|^{2}}\setminus\Omega_{|z_{1}|} into a nearly cube Q1Q_{1} in unit size, and Ω|z1|+|z1|2Ω|z1|\Omega_{|z_{1}|+|z_{1}|^{2}}^{*}\setminus\Omega_{|z_{1}|}^{*} into Q1Q_{1}^{*} by using the following change of variables:

{x1z1=|z1|2y1,x2=|z1|2y2.\displaystyle\begin{cases}x_{1}-z_{1}=|z_{1}|^{2}y_{1},\\ x_{2}=|z_{1}|^{2}y_{2}.\end{cases}

After rescaling, let

V11=v11(z1+z12y1,|z1|2y2),inQ1,andV11=v11(z1+z12y1,|z1|2y2),inQ1.V_{1}^{1}=v_{1}^{1}(z_{1}+z_{1}^{2}y_{1},|z_{1}|^{2}y_{2}),\quad\mbox{in}~{}Q_{1},\quad\mbox{and}\quad V_{1}^{*1}=v_{1}^{*1}(z_{1}+z_{1}^{2}y_{1},|z_{1}|^{2}y_{2}),\quad\mbox{in}~{}Q_{1}^{*}.

By the same reason that leads to (4.47), we get

|2V11|CinQ1,and|2V11|CinQ1.|\nabla^{2}V_{1}^{1}|\leq C\quad\mbox{in}~{}Q_{1},\quad\mbox{and}~{}|\nabla^{2}V_{1}^{*1}|\leq C\quad\mbox{in}~{}Q_{1}^{*}.

Using the interpolation inequality, we obtain

|(V11V11)|Cε1/4.|\nabla(V_{1}^{1}-V_{1}^{*1})|\leq C\varepsilon^{1/4}.

Rescaling back to v11v11v_{1}^{1}-v_{1}^{*1}, we get

|(v11v11)|Cε1/4|x1|2inΩRΩε1/8.|\nabla(v_{1}^{1}-v_{1}^{*1})|\leq C\varepsilon^{1/4}|x_{1}|^{-2}\quad\mbox{in}~{}\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}. (4.53)

Similarly, we have

|v11|C|x1|2inΩRΩε1/8,|\nabla v_{1}^{1}|\leq C|x_{1}|^{-2}\quad\mbox{in}~{}\Omega_{R}\setminus\Omega_{\varepsilon^{1/8}}, (4.54)

and

|v11|C|x1|2inΩRΩε1/8.|\nabla v_{1}^{*1}|\leq C|x_{1}|^{-2}\quad\mbox{in}~{}\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}. (4.55)

Now, by (4.54) and |(ΩRΩε1/8)(ΩRΩε1/8)|Cε|(\Omega_{R}\setminus\Omega_{\varepsilon^{1/8}})\setminus(\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*})|\leq C\varepsilon, we have

|I2,1|Cεε1/8<|x1|Rdx1|x1|4Cε5/8.\displaystyle|\mbox{I}_{2,1}|\leq C\varepsilon\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\frac{dx_{1}}{|x_{1}|^{4}}\leq C\varepsilon^{5/8}.

Also, by using (4.53) and (4.55), we obtain

|I2,2|Cε1/4ε1/8<|x1|Rdx1|x1|2Cε1/8,\displaystyle|\mbox{I}_{2,2}|\leq C\varepsilon^{1/4}\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\frac{dx_{1}}{|x_{1}|^{2}}\leq C\varepsilon^{1/8},

and by (4.53), we get

|I2,3|Cε1/2ε1/8<|x1|Rdx1|x1|2Cε3/8.\displaystyle|\mbox{I}_{2,3}|\leq C\varepsilon^{1/2}\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\frac{dx_{1}}{|x_{1}|^{2}}\leq C\varepsilon^{3/8}.

Substituting the estimates above into (4.4), we obtain (4.51).

Step 3. We next further approximate I2\mbox{I}_{2}^{*} by some specific functions. Note that

I2\displaystyle\mbox{I}_{2}^{*} =ΩRΩε1/8(0e(u11),e(u11))𝑑x+2ΩRΩε1/8(0e(u11),e(v11u11))𝑑x\displaystyle=\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(u_{1}^{*1})\big{)}\ dx+2\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(v_{1}^{*1}-u_{1}^{*1})\big{)}\ dx
+ΩRΩε1/8(0e(v11u11),e(v11u11))𝑑x.\displaystyle\quad+\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}-u_{1}^{*1}),e(v_{1}^{*1}-u_{1}^{*1})\big{)}\ dx.

By using Theorem 2.2, for the second term, we have

2|Ωε1/8(0e(u11),e(v11u11))𝑑x|Cε1/8,\displaystyle 2\left|\int_{\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(v_{1}^{*1}-u_{1}^{*1})\big{)}\ dx\right|\leq C\varepsilon^{1/8},

and for the third term,

|Ωε1/8(0e(v11u11),e(v11u11))𝑑x|Cε3/8.\displaystyle\left|\int_{\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}-u_{1}^{*1}),e(v_{1}^{*1}-u_{1}^{*1})\big{)}\ dx\right|\leq C\varepsilon^{3/8}.

Hence,

I2=ΩRΩε1/8(0e(u11),e(u11))𝑑x+M2+O(ε1/8),\displaystyle\mbox{I}_{2}^{*}=\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(u_{1}^{*1})\big{)}\ dx+M_{2}^{*}+O(\varepsilon^{1/8}), (4.56)

where

M2:=2ΩR(0e(u11),e(v11u11))𝑑x+ΩR(0e(v11u11),e(v11u11))𝑑xM_{2}^{*}:=2\int_{\Omega_{R}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(v_{1}^{*1}-u_{1}^{*1})\big{)}\ dx+\int_{\Omega_{R}^{*}}\big{(}\mathbb{C}^{0}e(v_{1}^{*1}-u_{1}^{*1}),e(v_{1}^{*1}-u_{1}^{*1})\big{)}\ dx

is a constant independent of ε\varepsilon. Coming back to (4.4), and using (4.46), (4.51), and (4.56), so far we obtain

a1111\displaystyle a_{11}^{11} =I3+ΩRΩε1/8(0e(u11),e(u11))𝑑x+M1+M2+O(ε1/8).\displaystyle=\mbox{I}_{3}+\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(u_{1}^{*1})\big{)}\ dx+M_{1}^{*}+M_{2}^{*}+O(\varepsilon^{1/8}). (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

I3\displaystyle\mbox{I}_{3} =Ωε1/8(0e(v11),e(v11))𝑑x\displaystyle=\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(v_{1}^{1}),e(v_{1}^{1})\big{)}\ dx
=Ωε1/8(0e(u11),e(u11))𝑑x+2Ωε1/8(0e(u11),e(v11u11))𝑑x\displaystyle=\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(u_{1}^{1}),e(u_{1}^{1})\big{)}\ dx+2\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(u_{1}^{1}),e(v_{1}^{1}-u_{1}^{1})\big{)}\ dx
+Ωε1/8(0e(v11u11),e(v11u11))𝑑x\displaystyle\quad+\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(v_{1}^{1}-u_{1}^{1}),e(v_{1}^{1}-u_{1}^{1})\big{)}\ dx
=Ωε1/8(0e(u11),e(u11))𝑑x+O(ε1/8).\displaystyle=\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(u_{1}^{1}),e(u_{1}^{1})\big{)}\ dx+O(\varepsilon^{1/8}). (4.58)

Second, by a direct computation, we obtain in ΩR\Omega_{R},

|x1(u¯11)1|,|x2(u~11)2|C|x1|δ(x1),|x1(u~11)2|C,x2(u¯11)1=1δ(x1).\displaystyle|\partial_{x_{1}}(\bar{u}_{1}^{1})^{1}|,~{}|\partial_{x_{2}}(\tilde{u}_{1}^{1})^{2}|\leq\frac{C|x_{1}|}{\delta(x_{1})},\quad|\partial_{x_{1}}(\tilde{u}_{1}^{1})^{2}|\leq C,~{}\partial_{x_{2}}(\bar{u}_{1}^{1})^{1}=\frac{1}{\delta(x_{1})}. (4.59)

For any u=(u1,u2)Tu=(u^{1},u^{2})^{\mathrm{T}}, recalling the definition of 0\mathbb{C}^{0}, a direct calculation yields

(0e(u),e(u))\displaystyle\big{(}\mathbb{C}^{0}e(u),e(u)\big{)}
=λ(x1u1+x2u2)2+μ(2(x1u1)2+(x2u1+x1u2)2+2(x2u2)2).\displaystyle=\lambda(\partial_{x_{1}}u^{1}+\partial_{x_{2}}u^{2})^{2}+\mu\big{(}2(\partial_{x_{1}}u^{1})^{2}+(\partial_{x_{2}}u^{1}+\partial_{x_{1}}u^{2})^{2}+2(\partial_{x_{2}}u^{2})^{2}\big{)}. (4.60)

Substituting u11u_{1}^{1} and u11u_{1}^{*1} into (4.4) and using (4.59), we have

Ωε1/8(0e(u11),e(u11))𝑑x+ΩRΩε1/8(0e(u11),e(u11))𝑑x\displaystyle\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(u_{1}^{1}),e(u_{1}^{1})\big{)}\ dx+\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(u_{1}^{*1})\big{)}\ dx
=μ|x1|ε1/8dx1ε+κ|x1|2+o(|x1|2)+με1/8<|x1|Rdx1κ|x1|2+o(|x1|2)+O(ε1/8).\displaystyle=\mu\int_{|x_{1}|\leq\varepsilon^{1/8}}\frac{dx_{1}}{\varepsilon+\kappa|x_{1}|^{2}+o(|x_{1}|^{2})}+\mu\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\frac{dx_{1}}{\kappa|x_{1}|^{2}+o(|x_{1}|^{2})}+O(\varepsilon^{1/8}).

Notice that for the first term,

|x1|ε1/8dx1ε+κ|x1|2+o(|x1|2)=|x1|ε1/8dx1ε+κ|x1|2+O(εγ18),\displaystyle\int_{|x_{1}|\leq\varepsilon^{1/8}}\frac{dx_{1}}{\varepsilon+\kappa|x_{1}|^{2}+o(|x_{1}|^{2})}=\int_{|x_{1}|\leq\varepsilon^{1/8}}\frac{dx_{1}}{\varepsilon+\kappa|x_{1}|^{2}}+O(\varepsilon^{\frac{\gamma-1}{8}}),

and for the second term,

ε1/8<|x1|Rdx1κ|x1|2+o(|x1|2)=ε1/8<|x1|Rdx1κ|x1|2+O(εγ18).\displaystyle\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\frac{dx_{1}}{\kappa|x_{1}|^{2}+o(|x_{1}|^{2})}=\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\frac{dx_{1}}{\kappa|x_{1}|^{2}}+O(\varepsilon^{\frac{\gamma-1}{8}}).

Let two right hand sides subtract,

ε1/8<|x1|R(1κ|x1|21ε+κ|x1|2)𝑑x1=O(ε5/8).\displaystyle\int_{\varepsilon^{1/8}<|x_{1}|\leq R}\left(\frac{1}{\kappa|x_{1}|^{2}}-\frac{1}{\varepsilon+\kappa|x_{1}|^{2}}\right)\ dx_{1}=O(\varepsilon^{5/8}).

Then we obtain

Ωε1/8(0e(u11),e(u11))𝑑x+ΩRΩε1/8(0e(u11),e(u11))𝑑x\displaystyle\int_{\Omega_{\varepsilon^{1/8}}}\big{(}\mathbb{C}^{0}e(u_{1}^{1}),e(u_{1}^{1})\big{)}\ dx+\int_{\Omega_{R}^{*}\setminus\Omega_{\varepsilon^{1/8}}^{*}}\big{(}\mathbb{C}^{0}e(u_{1}^{*1}),e(u_{1}^{*1})\big{)}\ dx
=πμκε+O(|logε|).\displaystyle=\frac{\pi\mu}{\sqrt{\kappa}\sqrt{\varepsilon}}+O(|\log\varepsilon|). (4.61)

Substituting (4.4) into (4.57) and using (4.4), we have

a1111=πμκε+O(|logε|)=πμκε(1+O(εγ+38)).a_{11}^{11}=\frac{\pi\mu}{\sqrt{\kappa}\sqrt{\varepsilon}}+O(|\log\varepsilon|)=\frac{\pi\mu}{\sqrt{\kappa}\sqrt{\varepsilon}}\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right). (4.62)

(2) For a1122a_{11}^{22}, we apply the auxiliary function defined in (2.16). The rest of the proof is the same as in (1), except that we substitute u12u_{1}^{2} and u12u_{1}^{*2} into (4.4). A direct calculation gives

|x1(u¯12)2|,|x2(u~12)1|C|x1|δ(x1),|x1(u~12)1|C,x2(u¯12)2=1δ(x1).\displaystyle|\partial_{x_{1}}(\bar{u}_{1}^{2})^{2}|,~{}|\partial_{x_{2}}(\tilde{u}_{1}^{2})^{1}|\leq\frac{C|x_{1}|}{\delta(x_{1})},\quad|\partial_{x_{1}}(\tilde{u}_{1}^{2})^{1}|\leq C,~{}\partial_{x_{2}}(\bar{u}_{1}^{2})^{2}=\frac{1}{\delta(x_{1})}. (4.63)

By using (4.63) and the same argument as that in (4.4), we obtain

a1122=π(λ+2μ)κε(1+O(εγ+38)).a_{11}^{22}=\frac{\pi(\lambda+2\mu)}{\sqrt{\kappa}\sqrt{\varepsilon}}\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right).

So we complete the proof of Proposition 2.3 in dimension two. ∎

Proof of Proposition 2.3 for d=3d=3..

The proof is similar to the above until (4.4). For u=(u1,u2,u3)Tu=(u^{1},u^{2},u^{3})^{\mathrm{T}}, (4.4) becomes

(0e(u),e(u))\displaystyle\big{(}\mathbb{C}^{0}e(u),e(u)\big{)}
=λ(x1u1+x2u2+x3u3)2+μ(2(x1u1)2+2(x2u2)2+2(x3u3)2\displaystyle=\lambda\Big{(}\partial_{x_{1}}u^{1}+\partial_{x_{2}}u^{2}+\partial_{x_{3}}u^{3}\Big{)}^{2}+\mu\Big{(}2(\partial_{x_{1}}u^{1})^{2}+2(\partial_{x_{2}}u^{2})^{2}+2(\partial_{x_{3}}u^{3})^{2}
+(x2u1+x1u2)2+(x3u1+x1u3)2+(x2u3+x3u2)2).\displaystyle\quad+(\partial_{x_{2}}u^{1}+\partial_{x_{1}}u^{2})^{2}+(\partial_{x_{3}}u^{1}+\partial_{x_{1}}u^{3})^{2}+(\partial_{x_{2}}u^{3}+\partial_{x_{3}}u^{2})^{2}\Big{)}. (4.64)

Substituting the specific function u11u_{1}^{1}, defined by (2.17), into (4.4) and recalling that in ΩR\Omega_{R},

|xku11|C|x|δ(x)andx3u11=1δ(x),k=1,2,\displaystyle|\partial_{x_{k}}u_{1}^{1}|\leq\frac{C|x^{\prime}|}{\delta(x^{\prime})}\quad\mbox{and}\quad\partial_{x_{3}}u_{1}^{1}=\frac{1}{\delta(x^{\prime})},\quad k=1,2,

we find that (4.62) becomes

a1111\displaystyle a_{11}^{11} =μ|x|ε1/8dxε+κ|x|2+o(|x|2)+με1/8<|x|Rdxκ|x|2+o(|x|2)+O(ε1/8)\displaystyle=\mu\int_{|x^{\prime}|\leq\varepsilon^{1/8}}\frac{dx^{\prime}}{\varepsilon+\kappa|x^{\prime}|^{2}+o(|x^{\prime}|^{2})}+\mu\int_{\varepsilon^{1/8}<|x^{\prime}|\leq R}\frac{dx^{\prime}}{\kappa|x^{\prime}|^{2}+o(|x^{\prime}|^{2})}+O(\varepsilon^{1/8})
=μ|x|Rdxε+κ|x|2+O(ε1/8)\displaystyle=\mu\int_{|x^{\prime}|\leq R}\frac{dx^{\prime}}{\varepsilon+\kappa|x^{\prime}|^{2}}+O(\varepsilon^{1/8})
=πμκ|logε|+𝒞31+O(ε1/8),\displaystyle=\frac{\pi\mu}{\kappa}|\log\varepsilon|+\mathcal{C}_{3}^{*1}+O(\varepsilon^{1/8}),

where 𝒞31\mathcal{C}_{3}^{*1} is a constant independent of ε\varepsilon and RR. Similarly,

a1122=πμκ|logε|+𝒞32+O(ε1/8),anda1133=π(λ+2μ)κ|logε|+𝒞33+O(ε1/8),\displaystyle a_{11}^{22}=\frac{\pi\mu}{\kappa}|\log\varepsilon|+\mathcal{C}_{3}^{*2}+O(\varepsilon^{1/8}),\quad\mbox{and}\quad a_{11}^{33}=\frac{\pi(\lambda+2\mu)}{\kappa}|\log\varepsilon|+\mathcal{C}_{3}^{*3}+O(\varepsilon^{1/8}),

where 𝒞32\mathcal{C}_{3}^{*2} and 𝒞33\mathcal{C}_{3}^{*3} are constants independent of ε\varepsilon and RR. 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 viα\nabla v_{i}^{\alpha} have been proved in Theorems 2.1, 2.2, and 2.6, recalling the proof of Proposition 4.4, we have C1α=C2αC_{1}^{\alpha}=C_{2}^{\alpha}, α=d+1,,d(d+1)/2\alpha=d+1,\cdots,d(d+1)/2, and (2.10) becomes

u=α=1d(C1αC2α)v1α+ub,inΩ.\nabla{u}=\sum_{\alpha=1}^{d}\left(C_{1}^{\alpha}-C_{2}^{\alpha}\right)\nabla{v}_{1}^{\alpha}+\nabla u_{b},\quad\mbox{in}~{}\Omega. (5.1)

Thus, it remains to derive the asymptotics of C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha}, α=1,,d\alpha=1,\cdots,d. To achieve this, we use (5.1) and the forth line of (2.2) to obtain

α=1d(C1αC2α)a11αβ\displaystyle\sum_{\alpha=1}^{d}(C_{1}^{\alpha}-C_{2}^{\alpha})a_{11}^{\alpha\beta} =b1β,β=1,,d(d+1)/2,\displaystyle=b_{1}^{\beta},\quad~{}~{}\beta=1,\cdots,d(d+1)/2, (5.2)

where a11αβa_{11}^{\alpha\beta} is defined in (2.9), and b1βb_{1}^{\beta} is defined in (3.2). Let us first consider d=2d=2.

5.1. Proof of Theorem 3.2

Proof.

We denote (5.2) in block matrix as follows:

(a111100a1122)(C11C21C12C22)=(b11b12),\displaystyle\begin{pmatrix}~{}a_{11}^{11}&0~{}\\ \\ ~{}0&a_{11}^{22}\end{pmatrix}\begin{pmatrix}~{}C_{1}^{1}-C_{2}^{1}~{}\\ \\ C_{1}^{2}-C_{2}^{2}\end{pmatrix}=\begin{pmatrix}~{}b_{1}^{1}~{}\\ \\ b_{1}^{2}\end{pmatrix}, (5.3)

here we used a1112=a1121=0a_{11}^{12}=a_{11}^{21}=0 in (4.36).

By using Cramer’s rule, Proposition 2.3, and Proposition 3.1, we have

C11C21=b11a1111=κπμε(b11[φ]+O(ε))(1+O(εγ+38)).\displaystyle C_{1}^{1}-C_{2}^{1}=\frac{b_{1}^{1}}{a_{11}^{11}}=\frac{\sqrt{\kappa}}{\pi\mu}\cdot\sqrt{\varepsilon}\left(b_{1}^{*1}[\varphi]+O(\sqrt{\varepsilon})\right)\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right). (5.4)

Similarly,

C12C22=κπ(λ+2μ)ε(b12[φ]+O(ε))(1+O(εγ+38)).\displaystyle C_{1}^{2}-C_{2}^{2}=\frac{\sqrt{\kappa}}{\pi(\lambda+2\mu)}\cdot\sqrt{\varepsilon}\left(b_{1}^{*2}[\varphi]+O(\sqrt{\varepsilon})\right)\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right). (5.5)

Then for any xΩRx\in\Omega_{R}, by using (5.1), Theorems 2.1 and 2.2, (5.4), and (5.5), we obtain

u\displaystyle\nabla u =α=12(C1αC2α)u1α+O(1)\displaystyle=\sum_{\alpha=1}^{2}\left(C_{1}^{\alpha}-C_{2}^{\alpha}\right)\nabla u_{1}^{\alpha}+O(1)
=κεπμb11[φ]u11(1+O(εγ+38))+κεπ(λ+2μ)b12[φ]u12(1+O(εγ+38))\displaystyle=\frac{\sqrt{\kappa\varepsilon}}{\pi\mu}b_{1}^{*1}[\varphi]\nabla u_{1}^{1}\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right)+\frac{\sqrt{\kappa\varepsilon}}{\pi(\lambda+2\mu)}b_{1}^{*2}[\varphi]\nabla u_{1}^{2}\left(1+O(\varepsilon^{\frac{\gamma+3}{8}})\right)
+O(1)\displaystyle\quad+O(1)
=κπε(b11[φ]μu11+b12[φ]λ+2μu12)(1+O(εγ+38))+O(1).\displaystyle=\frac{\sqrt{\kappa}}{\pi}\cdot\sqrt{\varepsilon}\left(\frac{b_{1}^{*1}[\varphi]}{\mu}\nabla u_{1}^{1}+\frac{b_{1}^{*2}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{2}\right)(1+O(\varepsilon^{\frac{\gamma+3}{8}}))+O(1).

The proof of Theorem 3.2 is finished. ∎

5.2. Proof of Theorem 3.5

Proof.

In dimension three, in order to solve C1αC2αC_{1}^{\alpha}-C_{2}^{\alpha} from (5.2), we rewrite it as

(a1111000a1122000a1133)(C11C21C12C22C13C23)=(b11b12b13),\displaystyle\begin{pmatrix}~{}a_{11}^{11}&0&0~{}\\ \\ ~{}0&a_{11}^{22}&0\\ \\ ~{}0&0&a_{11}^{33}\end{pmatrix}\begin{pmatrix}~{}C_{1}^{1}-C_{2}^{1}~{}\\ \\ C_{1}^{2}-C_{2}^{2}~{}\\ \\ C_{1}^{3}-C_{2}^{3}\end{pmatrix}=\begin{pmatrix}~{}b_{1}^{1}~{}\\ \\ b_{1}^{2}~{}\\ \\ b_{1}^{3}\end{pmatrix}, (5.6)

here we also used the fact that a11αβ=0a_{11}^{\alpha\beta}=0 by the symmetry of domains, α,β=1,2,3\alpha,\beta=1,2,3, αβ\alpha\neq\beta. Then similar to (5.4) and (5.5), we have

C1αC2α\displaystyle C_{1}^{\alpha}-C_{2}^{\alpha} =κπμ1|logε|b1α[φ](1+O(|logε|1)),α=1,2,\displaystyle=\frac{\kappa}{\pi\mu}\cdot\frac{1}{|\log\varepsilon|}b_{1}^{*\alpha}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right),\quad\alpha=1,2,
C13C23\displaystyle C_{1}^{3}-C_{2}^{3} =κπ(λ+2μ)1|logε|b13[φ](1+O(|logε|1)).\displaystyle=\frac{\kappa}{\pi(\lambda+2\mu)}\cdot\frac{1}{|\log\varepsilon|}b_{1}^{*3}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right). (5.7)

For any xΩRx\in\Omega_{R}, by using (5.1), Theorems 2.2 and 2.6, and (5.2), we have

u\displaystyle\nabla u =α=13(C1αC2α)u1α+O(1)\displaystyle=\sum_{\alpha=1}^{3}(C_{1}^{\alpha}-C_{2}^{\alpha})\nabla u_{1}^{\alpha}+O(1)
=κπ1|logε|(1μα=12b1α[φ]u1α+1λ+2μb13[φ]u13)(1+O(|logε|1))+O(1).\displaystyle=\frac{\kappa}{\pi}\cdot\frac{1}{|\log\varepsilon|}\left(\frac{1}{\mu}\sum_{\alpha=1}^{2}b_{1}^{*\alpha}[\varphi]\nabla u_{1}^{\alpha}+\frac{1}{\lambda+2\mu}b_{1}^{*3}[\varphi]\nabla u_{1}^{3}\right)\left(1+O(|\log\varepsilon|^{-1})\right)+O(1).

The proof of Theorem 3.5 is completed. ∎

6. The proof of Theorem 3.8 and Theorem 3.11

6.1. The proof of Theorem 3.8

By using a similar argument that led to Propositions 2.3 and 3.1, we obtain

Proposition 6.1.

Under the assumptions (2.12) and (2.13) with m3m\geq 3, we have, for sufficiently small ε>0\varepsilon>0,

a1111=μQ2,mκ1/mε11/m+O(1),a1122=(λ+2μ)Q2,mκ1/mε11/m+O(1),m3,a_{11}^{11}=\frac{\mu Q_{2,m}}{\kappa^{1/m}\varepsilon^{1-1/m}}+O(1),\quad a_{11}^{22}=\frac{(\lambda+2\mu)Q_{2,m}}{\kappa^{1/m}\varepsilon^{1-1/m}}+O(1),\quad m\geq 3,

and

a1133={2(λ+2μ)3κ|logε|+O(1),m=3,(λ+2μ)Q~2,mκ3/mε13/m+O(1),m4,a_{11}^{33}=\begin{cases}\frac{2(\lambda+2\mu)}{3\kappa}|\log\varepsilon|+O(1),\quad m=3,\\ \frac{(\lambda+2\mu)\widetilde{Q}_{2,m}}{\kappa^{3/m}\varepsilon^{1-3/m}}+O(1),\quad m\geq 4,\end{cases}

where

Q2,m=2011+tm𝑑t,Q~2,m=20t21+tm𝑑t.Q_{2,m}=2\int_{0}^{\infty}\frac{1}{1+t^{m}}\ dt,\quad\widetilde{Q}_{2,m}=2\int_{0}^{\infty}\frac{t^{2}}{1+t^{m}}\ dt.
Proposition 6.2.

Under the above assumptions, we have for small enough ε>0\varepsilon>0, if α=1,2\alpha=1,2,

b1α[φ]b1α[φ]={O(|logε|1),m=3,O(ε1/4),m=4,O(ε1/3),m5,b13[φ]b13[φ]={O(|logε|1),m=3,O(ε13m),m=4,5,O(εm+23m),m6,\displaystyle b_{1}^{\alpha}[\varphi]-b_{1}^{*\alpha}[\varphi]=\begin{cases}O(|\log\varepsilon|^{-1}),~{}m=3,\\ O(\varepsilon^{1/4}),~{}m=4,\\ O(\varepsilon^{1/3}),~{}m\geq 5,\end{cases}b_{1}^{3}[\varphi]-b_{1}^{*3}[\varphi]=\begin{cases}O(|\log\varepsilon|^{-1}),~{}m=3,\\ O(\varepsilon^{1-\frac{3}{m}}),~{}m=4,5,\\ O(\varepsilon^{\frac{m+2}{3m}}),~{}m\geq 6,\end{cases}

where b1α[φ]b_{1}^{\alpha}[\varphi]and b1αb_{1}^{*\alpha} are defined in (3.2), α=1,2,3\alpha=1,2,3.

We are ready to finish the proof of Theorem 3.8.

Completion of the proof of Theorem 3.8..

For any xΩRx\in\Omega_{R}, by using (2.10), Theorems 2.1 and 2.2, we have

u=α=13(C1αC2α)u1α+O(1),m3.\displaystyle\nabla u=\sum_{\alpha=1}^{3}\left(C_{1}^{\alpha}-C_{2}^{\alpha}\right)\nabla u_{1}^{\alpha}+O(1),\quad m\geq 3. (6.1)

Recalling (5.3), and using Cramer’s rule, Propositions 6.1 and 6.2, if m=3m=3, we have

C11C21\displaystyle C_{1}^{1}-C_{2}^{1} =b11[φ]a1111=κ1/3ε2/3μQ2,3b11[φ](1+O(|logε|1)),\displaystyle=\frac{b_{1}^{1}[\varphi]}{a_{11}^{11}}=\frac{\kappa^{1/3}\varepsilon^{2/3}}{\mu Q_{2,3}}b_{1}^{*1}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right),
C12C22\displaystyle C_{1}^{2}-C_{2}^{2} =b12[φ]a1122=κ1/3ε2/3(λ+2μ)Q2,3b12[φ](1+O(|logε|1))\displaystyle=\frac{b_{1}^{2}[\varphi]}{a_{11}^{22}}=\frac{\kappa^{1/3}\varepsilon^{2/3}}{(\lambda+2\mu)Q_{2,3}}b_{1}^{*2}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right)
C13C23\displaystyle C_{1}^{3}-C_{2}^{3} =3κ2(λ+2μ)|logε|b13[φ](1+O(|logε|1)).\displaystyle=\frac{3\kappa}{2(\lambda+2\mu)|\log\varepsilon|}b_{1}^{*3}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right).

Hence, coming back to (6.1), for xΩRx\in\Omega_{R},

u\displaystyle\nabla u =κ1/3ε2/3Q2,3(b11[φ]μu11+b12[φ]λ+2μu12)(1+O(|logε|1))\displaystyle=\frac{\kappa^{1/3}\varepsilon^{2/3}}{Q_{2,3}}\left(\frac{b_{1}^{*1}[\varphi]}{\mu}\nabla u_{1}^{1}+\frac{b_{1}^{*2}[\varphi]}{\lambda+2\mu}\nabla u_{1}^{2}\right)\left(1+O(|\log\varepsilon|^{-1})\right)
+3κ2(λ+2μ)|logε|b13[φ]u13(1+O(|logε|1))+O(1).\displaystyle\quad+\frac{3\kappa}{2(\lambda+2\mu)|\log\varepsilon|}b_{1}^{*3}[\varphi]\nabla u_{1}^{3}\left(1+O(|\log\varepsilon|^{-1})\right)+O(1).

The cases m4m\geq 4 follow from the same argument as above, we omit them here. This completes the proof of Theorem 3.8. ∎

6.2. The proof of Theorem 3.11

(i) The proof of the case m=3m=3 is very similar to that in the proof of Theorem 3.2 when d=3d=3 and m=2m=2. We omit it here.

(ii) For m4m\geq 4, we have

Proposition 6.3.

Under the assumptions (2.12) and (2.13) with m4m\geq 4, we have, for sufficiently small ε>0\varepsilon>0 and m4m\geq 4,

a11αα=πμQ3,mκ2/mε12/m+O(1),α=1,2,a1133=π(λ+2μ)Q3,mκ2/mε12/m+O(1),a_{11}^{\alpha\alpha}=\frac{\pi\mu Q_{3,m}}{\kappa^{2/m}\varepsilon^{1-2/m}}+O(1),~{}\alpha=1,2,\quad a_{11}^{33}=\frac{\pi(\lambda+2\mu)Q_{3,m}}{\kappa^{2/m}\varepsilon^{1-2/m}}+O(1),
a1144={πμ2κ|logε|+O(1),m=4,πμQ~3,mκ4/mε14/m+O(1),m5,a_{11}^{44}=\begin{cases}\frac{\pi\mu}{2\kappa}|\log\varepsilon|+O(1),~{}\quad m=4,\\ \frac{\pi\mu\widetilde{Q}_{3,m}}{\kappa^{4/m}\varepsilon^{1-4/m}}+O(1),~{}\quad m\geq 5,\end{cases}

and for α=5,6\alpha=5,6,

a11αα={π(λ+2μ)4κ|logε|+O(1),m=4,π(λ+2μ)Q~3,m2κ4/mε14/m+O(1),m5,a_{11}^{\alpha\alpha}=\begin{cases}\frac{\pi(\lambda+2\mu)}{4\kappa}|\log\varepsilon|+O(1),~{}\quad m=4,\\ \frac{\pi(\lambda+2\mu)\widetilde{Q}_{3,m}}{2\kappa^{4/m}\varepsilon^{1-4/m}}+O(1),~{}\quad m\geq 5,\end{cases}

where

Q3,m=20t1+tm𝑑t,Q~3,m=20t31+tm𝑑t.Q_{3,m}=2\int_{0}^{\infty}\frac{t}{1+t^{m}}\ dt,\quad\widetilde{Q}_{3,m}=2\int_{0}^{\infty}\frac{t^{3}}{1+t^{m}}\ dt.
Proposition 6.4.

Under the above assumptions, we have for sufficiently small ε>0\varepsilon>0, if α=1,2,3\alpha=1,2,3,

b1α[φ]b1α[φ]\displaystyle b_{1}^{\alpha}[\varphi]-b_{1}^{*\alpha}[\varphi] ={O(|logε|1),m=4,O(ε1/5),m=5,O(ε1/3),m6,\displaystyle=\begin{cases}O(|\log\varepsilon|^{-1}),&\quad m=4,\\ O(\varepsilon^{1/5}),&\quad m=5,\\ O(\varepsilon^{1/3}),&\quad m\geq 6,\end{cases}

and α=4,5,6\alpha=4,5,6,

b1α[φ]b1α[φ]\displaystyle b_{1}^{\alpha}[\varphi]-b_{1}^{*\alpha}[\varphi] ={O(|logε|1),m=4,O(ε14m),m=5,6,O(εm+23m),m7,\displaystyle=\begin{cases}O(|\log\varepsilon|^{-1}),&\quad m=4,\\ O(\varepsilon^{1-\frac{4}{m}}),&\quad m=5,6,\\ O(\varepsilon^{\frac{m+2}{3m}}),&\quad m\geq 7,\end{cases}

where b1α[φ]b_{1}^{\alpha}[\varphi] and b1α[φ]b_{1}^{*\alpha}[\varphi] are defined in (3.2), α=1,2,3,4,5,6\alpha=1,2,3,4,5,6.

Finally, we close this section by giving the proof of Theorem 3.11 (ii).

Proof of Theorem 3.11 (ii).

For any xΩRx\in\Omega_{R}, by using (2.10), Theorems 2.1 and 2.2, we have

u=α=16(C1αC2α)u1α+O(1),m3.\displaystyle\nabla u=\sum_{\alpha=1}^{6}\left(C_{1}^{\alpha}-C_{2}^{\alpha}\right)\nabla u_{1}^{\alpha}+O(1),\quad m\geq 3. (6.2)

By (2.10), Propostions 6.3 and 6.4, we give the proof only for m=4m=4. The other cases are similar. When m=4m=4,

C1αC2α\displaystyle C_{1}^{\alpha}-C_{2}^{\alpha} =κεπμQ3,4b1α[φ](1+O(|logε|1)),α=1,2,\displaystyle=\frac{\sqrt{\kappa}\sqrt{\varepsilon}}{\pi\mu Q_{3,4}}b_{1}^{*\alpha}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right),\quad\alpha=1,2,
C13C23\displaystyle C_{1}^{3}-C_{2}^{3} =κεπ(λ+2μ)Q3,4b13[φ](1+O(|logε|1)),\displaystyle=\frac{\sqrt{\kappa}\sqrt{\varepsilon}}{\pi(\lambda+2\mu)Q_{3,4}}b_{1}^{*3}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right),
C14C24\displaystyle C_{1}^{4}-C_{2}^{4} =2κπμ|logε|b14[φ](1+O(|logε|1)),\displaystyle=\frac{2\kappa}{\pi\mu|\log\varepsilon|}b_{1}^{*4}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right),

and for α=5,6\alpha=5,6,

C1αC2α\displaystyle C_{1}^{\alpha}-C_{2}^{\alpha} =4κπ(λ+2μ)|logε|b1α[φ](1+O(|logε|1)).\displaystyle=\frac{4\kappa}{\pi(\lambda+2\mu)|\log\varepsilon|}b_{1}^{*\alpha}[\varphi]\left(1+O(|\log\varepsilon|^{-1})\right).

Substituting the above terms into (6.2), for xΩRx\in\Omega_{R},

u\displaystyle\nabla u =α=16(C1αC2α)u1α+O(1)\displaystyle=\sum_{\alpha=1}^{6}(C_{1}^{\alpha}-C_{2}^{\alpha})\nabla u_{1}^{\alpha}+O(1)
=κεπQ3,4(α=121μb1α[φ]u1α+1λ+2μb13[φ]u13)(1+O(|logε|1))\displaystyle=\frac{\sqrt{\kappa}\sqrt{\varepsilon}}{\pi Q_{3,4}}\left(\sum_{\alpha=1}^{2}\frac{1}{\mu}b_{1}^{*\alpha}[\varphi]\nabla u_{1}^{\alpha}+\frac{1}{\lambda+2\mu}b_{1}^{*3}[\varphi]\nabla u_{1}^{3}\right)\left(1+O(|\log\varepsilon|^{-1})\right)
+2κπ|logε|(1μb14[φ]u14+α=561λ+2μb1α[φ]u1α)(1+O(|logε|1))+O(1).\displaystyle\quad+\frac{2\kappa}{\pi|\log\varepsilon|}\left(\frac{1}{\mu}b_{1}^{*4}[\varphi]\nabla u_{1}^{4}+\sum_{\alpha=5}^{6}\frac{1}{\lambda+2\mu}b_{1}^{*\alpha}[\varphi]\nabla u_{1}^{\alpha}\right)\left(1+O(|\log\varepsilon|^{-1})\right)+O(1).

The proof is finished. ∎

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 Y:=(L1,L1)×(L2,L2)Y:=(-L_{1},L_{1})\times(-L_{2},L_{2}), where L1L_{1} and L2L_{2} are two positive numbers. Let DYD\subset Y be a mm-convex domain containing the origin with C2C^{2} boundary. Assume that DD is close to the horizontal boundary of YY and away from the vertical boundary. Let ε/2\varepsilon/2 be the distance between DD and the the horizontal boundary of YY, so that ε\varepsilon becomes the distance between two adjacent inclusions, see Figure 2.

Refer to caption
Figure 2. mm-convex inclusions (say, x4+y4=1x^{4}+y^{4}=1).

As in [37], after translation, we denote

Yt:=Y+(0,L2)=(L1,L1)×(0,2L2),\displaystyle Y_{t}:=Y+(0,L_{2})=(-L_{1},L_{1})\times(0,2L_{2}),
D1:=D+(0,2L2+ε/2),D2:=D+(0,ε/2),Y:=YtD1D2¯,\displaystyle D_{1}:=D+(0,2L_{2}+\varepsilon/2),\quad D_{2}:=D+(0,\varepsilon/2),\quad Y^{\prime}:=Y_{t}\setminus\overline{D_{1}\cup D_{2}},

and set

Γ+:=(D1{x2=2L2+ε/2})Y,Γ:=(D2{x2=ε/2})Y.\displaystyle\Gamma_{+}:=(\partial D_{1}\cup\{x_{2}=2L_{2}+\varepsilon/2\})\cap\partial Y^{\prime},\quad\Gamma_{-}:=(\partial D_{2}\cup\{x_{2}=\varepsilon/2\})\cap\partial Y^{\prime}.

Then we obtain the effective shear modulus μm\mu_{m}^{*} and the effective extensional modulus EmE_{m}^{*} defined by

μm=L2L111,Em=Eλ+2μL2L112,\displaystyle\mu_{m}^{*}=\frac{L_{2}}{L_{1}}\mathcal{E}_{1}^{1},\quad E_{m}^{*}=\frac{E}{\lambda+2\mu}\frac{L_{2}}{L_{1}}\mathcal{E}_{1}^{2},

where E=μ(3λ+2μ)λ+μE=\frac{\mu(3\lambda+2\mu)}{\lambda+\mu}, and

1α=Y(0e(v1α),e(v1α)dx,α=1,2,\displaystyle\mathcal{E}_{1}^{\alpha}=\int_{Y^{\prime}}(\mathbb{C}^{0}e(v_{1}^{\alpha}),e(v_{1}^{\alpha})\ dx,\quad\alpha=1,2,

v1αH1(Y)v_{1}^{\alpha}\in H^{1}(Y^{\prime}) is the solution to

{λ,μv1α:=(0e(v1α))=0,inY,v1α=ψα,onΓ+,v1α=0,onΓ,v1αν|+=0,onx1=±L1.\displaystyle\begin{cases}\mathcal{L}_{\lambda,\mu}v_{1}^{\alpha}:=\nabla\cdot(\mathbb{C}^{0}e(v_{1}^{\alpha}))=0,\quad&\hbox{in}\ Y^{\prime},\\ v_{1}^{\alpha}=\psi_{\alpha},&\hbox{on}~{}\Gamma_{+},\\ v_{1}^{\alpha}=0,&\hbox{on}~{}\Gamma_{-},\\ \frac{\partial v_{1}^{\alpha}}{\partial\nu}\Big{|}_{+}=0,&\hbox{on}~{}x_{1}=\pm L_{1}.\end{cases}

Note that the definition of 1α\mathcal{E}_{1}^{\alpha} is similar to that of a11ααa_{11}^{\alpha\alpha} in (2.9). Then by using Proposition 2.3 and Proposition 6.1, we have

Theorem 7.1.

Given m2m\geq 2. As ε0\varepsilon\rightarrow 0,

μm=μL2L1Q2,mκ1/mε11/m+O(1),andEm=EL2L1Q2,mκ1/mε11/m+O(1),\displaystyle\mu_{m}^{*}=\mu\frac{L_{2}}{L_{1}}\frac{Q_{2,m}}{\kappa^{1/m}\varepsilon^{1-1/m}}+O(1),\quad\mbox{and}\quad E_{m}^{*}=E\frac{L_{2}}{L_{1}}\frac{Q_{2,m}}{\kappa^{1/m}\varepsilon^{1-1/m}}+O(1),

where κ\kappa is the curvature of D\partial D near the origin, and

Q2,m=2011+tm𝑑t.Q_{2,m}=2\int_{0}^{\infty}\frac{1}{1+t^{m}}\ dt.

Clearly, by a direct calculation, we find that Theorem 7.1 for m=2m=2 is actually the result in [32]. Furthermore, we would like to remark that compared with [32], our method can do not need to assume that D1D_{1} and D2D_{2} are symmetric.

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 |λ,μu1α||\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}| is improved to be controled by Cδ(x)\frac{C}{\delta(x^{\prime})}. This is due to the introduction of u~1α\tilde{u}_{1}^{\alpha}. Then we adapt the iteration technique first used in [39] and further developed in [10, 11], to capture all singular terms of v1α\nabla v_{1}^{\alpha} and to obtain the asymptotic formulas.

Proof of Theorems 2.2 and 2.6..

Step 1. Claim:

|λ,μu1α|C{1δ(x)+|x1|γ1,m=2,|x|m2δ(x)(1+ε|x|),m3,α=1,,d;|λ,μu1α|Cδ(x),α=d+1,,d(d+1)2.\begin{split}|\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}|&\leq C\begin{cases}\frac{1}{\delta(x^{\prime})}+|x_{1}|^{\gamma-1},&~{}m=2,\\ \frac{|x^{\prime}|^{m-2}}{\delta(x^{\prime})}\left(1+\frac{\varepsilon}{|x^{\prime}|}\right),&~{}m\geq 3,\end{cases}\quad\alpha=1,\cdots,d;\\ |\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}|&\leq\frac{C}{\delta(x^{\prime})},\quad\alpha=d+1,\cdots,\frac{d(d+1)}{2}.\end{split} (8.1)

We will prove the claim in the light of the following two cases.

Case 1. d=2d=2. A direct calculation yields in ΩR\Omega_{R},

|x1x1(u¯11)1|,|x1x2(u~11)2|C|x1|m2δ(x1),|x1x1(u~11)2|C{|x1|δ(x1)+|x1|γ1,m=2,|x1|m3,m3;\Big{|}\partial_{x_{1}x_{1}}(\bar{u}_{1}^{1})^{1}\Big{|},~{}\Big{|}\partial_{x_{1}x_{2}}(\tilde{u}_{1}^{1})^{2}\Big{|}\leq\frac{C|x_{1}|^{m-2}}{\delta(x_{1})},\quad\Big{|}\partial_{x_{1}x_{1}}(\tilde{u}_{1}^{1})^{2}\Big{|}\leq C\begin{cases}\frac{|x_{1}|}{\delta(x_{1})}+|x_{1}|^{\gamma-1},&~{}m=2,\\ |x_{1}|^{m-3},&~{}m\geq 3;\end{cases} (8.2)

and

x1x2(u¯11)1=x1(h1h2)δ2(x1),x2x2(u~11)2=λ+μλ+2μx1(h1h2)δ2(x1).\displaystyle\partial_{x_{1}x_{2}}(\bar{u}_{1}^{1})^{1}=-\frac{\partial_{x_{1}}(h_{1}-h_{2})}{\delta^{2}(x_{1})},\quad\partial_{x_{2}x_{2}}(\tilde{u}_{1}^{1})^{2}=\frac{\lambda+\mu}{\lambda+2\mu}\frac{\partial_{x_{1}}(h_{1}-h_{2})}{\delta^{2}(x_{1})}. (8.3)

By using (8.2), we have

|(λ,μu11)1|\displaystyle\Big{|}(\mathcal{L}_{\lambda,\mu}u_{1}^{1})^{1}\Big{|}
=λ(x1x1(u11)1+x2x1(u11)2)+μ(2x1x1(u11)1+x2x2(u11)1+x1x2(u11)2)\displaystyle=\lambda\Big{(}\partial_{x_{1}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{2}x_{1}}(u_{1}^{1})^{2}\Big{)}+\mu\Big{(}2\partial_{x_{1}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{2}x_{2}}(u_{1}^{1})^{1}+\partial_{x_{1}x_{2}}(u_{1}^{1})^{2}\Big{)}
=|μΔ(u11)1+(λ+μ)(x1x1(u11)1+x1x2(u11)2)|\displaystyle=\Big{|}\mu\Delta(u_{1}^{1})^{1}+(\lambda+\mu)\big{(}\partial_{x_{1}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{1}x_{2}}(u_{1}^{1})^{2}\big{)}\Big{|}
=|μΔ(u¯11)1+(λ+μ)(x1x1(u¯11)1+x1x2(u~11)2)|C|x1|m2δ(x1).\displaystyle=\Big{|}\mu\Delta(\bar{u}_{1}^{1})^{1}+(\lambda+\mu)\big{(}\partial_{x_{1}x_{1}}(\bar{u}_{1}^{1})^{1}+\partial_{x_{1}x_{2}}(\tilde{u}_{1}^{1})^{2}\big{)}\Big{|}\leq\frac{C|x_{1}|^{m-2}}{\delta(x_{1})}. (8.4)

By using (8.3), we get

(λ+μ)x1x2(u¯11)1+(λ+2μ)x2x2(u~11)2=0,\displaystyle(\lambda+\mu)\partial_{x_{1}x_{2}}(\bar{u}_{1}^{1})^{1}+(\lambda+2\mu)\partial_{x_{2}x_{2}}(\tilde{u}_{1}^{1})^{2}=0,

which means that the “bad” terms in (8.3) are eliminated. Combining this and (8.2), we obtain

|(λ,μu11)2|\displaystyle\Big{|}(\mathcal{L}_{\lambda,\mu}u_{1}^{1})^{2}\Big{|} =|(λ+μ)(x1x2(u11)1+x2x2(u11)2)+μΔu12|\displaystyle=\Big{|}(\lambda+\mu)\big{(}\partial_{x_{1}x_{2}}(u_{1}^{1})^{1}+\partial_{x_{2}x_{2}}(u_{1}^{1})^{2}\big{)}+\mu\Delta u_{1}^{2}\Big{|}
=|μx1x1(u~11)2|C{|x1|δ(x1)+|x1|γ1,m=2,|x1|m3,m3.\displaystyle=\Big{|}\mu\partial_{x_{1}x_{1}}(\tilde{u}_{1}^{1})^{2}\Big{|}\leq C\begin{cases}\frac{|x_{1}|}{\delta(x_{1})}+|x_{1}|^{\gamma-1},&~{}m=2,\\ |x_{1}|^{m-3},&~{}m\geq 3.\end{cases} (8.5)

We henceforth obtain from (8) and (8) that

|λ,μu11|C{1δ(x1)+|x1|γ1,m=2,|x1|m2δ(x1)(1+ε|x1|),m3.\displaystyle\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{1}\Big{|}\leq C\begin{cases}\frac{1}{\delta(x_{1})}+|x_{1}|^{\gamma-1},&~{}m=2,\\ \frac{|x_{1}|^{m-2}}{\delta(x_{1})}\left(1+\frac{\varepsilon}{|x_{1}|}\right),&~{}m\geq 3.\end{cases}

Similarly, we have

|λ,μu12|C{1δ(x1)+|x1|γ1,m=2,|x1|m2δ(x1)(1+ε|x1|),m3.\displaystyle\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{2}\Big{|}\leq C\begin{cases}\frac{1}{\delta(x_{1})}+|x_{1}|^{\gamma-1},&~{}m=2,\\ \frac{|x_{1}|^{m-2}}{\delta(x_{1})}\left(1+\frac{\varepsilon}{|x_{1}|}\right),&~{}m\geq 3.\end{cases}

Furthermore, we have

|x1x1u13|C|x1|m1δ(x1),|x1x2u13|,|x2x2u13|Cδ(x1).\displaystyle\Big{|}\partial_{x_{1}x_{1}}u_{1}^{3}\Big{|}\leq\frac{C|x_{1}|^{m-1}}{\delta(x_{1})},~{}\Big{|}\partial_{x_{1}x_{2}}u_{1}^{3}\Big{|},\Big{|}\partial_{x_{2}x_{2}}u_{1}^{3}\Big{|}\leq\frac{C}{\delta(x_{1})}.

Then we obtain

|λ,μu13|Cδ(x1).\displaystyle\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{3}\Big{|}\leq\frac{C}{\delta(x_{1})}.

Case 2. d=3d=3. We have in ΩR\Omega_{R},

|x1x1(u¯11)1|,|x2x2(u¯11)1|,|x1x2(u¯11)1|,|x1x3(u~11)3|,|x2x3(u~11)3|C|x|m2δ(x),|x1x1(u~11)3|,|x2x2(u~11)3|C{|x|δ(x)+1,m=2,|x|m3,m3;\begin{split}&\Big{|}\partial_{x_{1}x_{1}}(\bar{u}_{1}^{1})^{1}\Big{|},~{}\Big{|}\partial_{x_{2}x_{2}}(\bar{u}_{1}^{1})^{1}\Big{|},~{}\Big{|}\partial_{x_{1}x_{2}}(\bar{u}_{1}^{1})^{1}\Big{|},~{}\Big{|}\partial_{x_{1}x_{3}}(\tilde{u}_{1}^{1})^{3}\Big{|},~{}\Big{|}\partial_{x_{2}x_{3}}(\tilde{u}_{1}^{1})^{3}\Big{|}\leq\frac{C|x^{\prime}|^{m-2}}{\delta(x^{\prime})},\\ &\Big{|}\partial_{x_{1}x_{1}}(\tilde{u}_{1}^{1})^{3}\Big{|},~{}\Big{|}\partial_{x_{2}x_{2}}(\tilde{u}_{1}^{1})^{3}\Big{|}\leq C\begin{cases}\frac{|x^{\prime}|}{\delta(x^{\prime})}+1,&~{}m=2,\\ |x^{\prime}|^{m-3},&~{}m\geq 3;\end{cases}\end{split} (8.6)

and

x3x1(u¯11)1=x1(h1h2)δ(x)2,x3x3(u~11)3=λ+μλ+2μx1(h1h2)δ(x)2.\displaystyle\partial_{x_{3}x_{1}}(\bar{u}_{1}^{1})^{1}=-\frac{\partial_{x_{1}}(h_{1}-h_{2})}{\delta(x^{\prime})^{2}},\quad\partial_{x_{3}x_{3}}(\tilde{u}_{1}^{1})^{3}=\frac{\lambda+\mu}{\lambda+2\mu}\frac{\partial_{x_{1}}(h_{1}-h_{2})}{\delta(x^{\prime})^{2}}. (8.7)

By (8.6), we have

|(λ,μu11)1|\displaystyle\Big{|}(\mathcal{L}_{\lambda,\mu}u_{1}^{1})^{1}\Big{|}
=|λ(x1x1(u11)1+x3x1(u11)3)+μ(2x1x1(u11)1+x2x2(u11)1+x1x3(u11)3)|\displaystyle=\Big{|}\lambda\big{(}\partial_{x_{1}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{3}x_{1}}(u_{1}^{1})^{3}\big{)}+\mu\big{(}2\partial_{x_{1}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{2}x_{2}}(u_{1}^{1})^{1}+\partial_{x_{1}x_{3}}(u_{1}^{1})^{3}\big{)}\Big{|}
C|x|m2δ(x),\displaystyle\leq\frac{C|x^{\prime}|^{m-2}}{\delta(x^{\prime})}, (8.8)

and

|(λ,μu11)2|\displaystyle\Big{|}(\mathcal{L}_{\lambda,\mu}u_{1}^{1})^{2}\Big{|} =|λ(x1x2(u11)1+x3x2(u11)3)+μ(x2x1(u11)1+x2x3(u11)3)|\displaystyle=\Big{|}\lambda\big{(}\partial_{x_{1}x_{2}}(u_{1}^{1})^{1}+\partial_{x_{3}x_{2}}(u_{1}^{1})^{3}\big{)}+\mu\big{(}\partial_{x_{2}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{2}x_{3}}(u_{1}^{1})^{3}\big{)}\Big{|}
C|x|m2δ(x).\displaystyle\leq\frac{C|x^{\prime}|^{m-2}}{\delta(x^{\prime})}. (8.9)

By (8.7), we obtain

(λ+μ)x3x1(u¯11)1+(λ+2μ)x3x3(u~11)3=0.\displaystyle(\lambda+\mu)\partial_{x_{3}x_{1}}(\bar{u}_{1}^{1})^{1}+(\lambda+2\mu)\partial_{x_{3}x_{3}}(\tilde{u}_{1}^{1})^{3}=0. (8.10)

Then (8.6) and (8.10) imply that

|(λ,μu11)3|\displaystyle\Big{|}(\mathcal{L}_{\lambda,\mu}u_{1}^{1})^{3}\Big{|} =|λ(x1x3(u11)1+x3x3(u11)3)\displaystyle=\Big{|}\lambda\big{(}\partial_{x_{1}x_{3}}(u_{1}^{1})^{1}+\partial_{x_{3}x_{3}}(u_{1}^{1})^{3}\big{)}
+μ(x1x1(u11)3+x3x1(u11)1+x2x2(u11)3+2x3x3(u11)3)|\displaystyle\quad+\mu\big{(}\partial_{x_{1}x_{1}}(u_{1}^{1})^{3}+\partial_{x_{3}x_{1}}(u_{1}^{1})^{1}+\partial_{x_{2}x_{2}}(u_{1}^{1})^{3}+2\partial_{x_{3}x_{3}}(u_{1}^{1})^{3}\big{)}\Big{|}
=|μ(x1x1(u~11)3+x2x2(u~11)3)|C{|x|δ(x)+1,m=2,|x|m3,m3.\displaystyle=\Big{|}\mu\big{(}\partial_{x_{1}x_{1}}(\tilde{u}_{1}^{1})^{3}+\partial_{x_{2}x_{2}}(\tilde{u}_{1}^{1})^{3}\big{)}\Big{|}\leq C\begin{cases}\frac{|x^{\prime}|}{\delta(x^{\prime})}+1,&~{}m=2,\\ |x^{\prime}|^{m-3},&~{}m\geq 3.\end{cases} (8.11)

Hence, (8), (8), and (8) give

|λ,μu11|C{1δ(x),m=2,|x|m2δ(x)(1+ε|x|),m3.\displaystyle\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{1}\Big{|}\leq C\begin{cases}\frac{1}{\delta(x^{\prime})},&~{}m=2,\\ \frac{|x^{\prime}|^{m-2}}{\delta(x^{\prime})}\left(1+\frac{\varepsilon}{|x^{\prime}|}\right),&~{}m\geq 3.\end{cases}

Similarly, we can get

|λ,μu12|,|λ,μu13|C{1δ(x),m=2,|x|m2δ(x)(1+ε|x|),m3.\displaystyle\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{2}\Big{|},~{}\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{3}\Big{|}\leq C\begin{cases}\frac{1}{\delta(x^{\prime})},&~{}m=2,\\ \frac{|x^{\prime}|^{m-2}}{\delta(x^{\prime})}\left(1+\frac{\varepsilon}{|x^{\prime}|}\right),&~{}m\geq 3.\end{cases}

For the corresponding estimates for u1αu_{1}^{\alpha}, α=4,5,6\alpha=4,5,6, we note that

|xkxlu1α|C|x|m1δ(x),|xkx3u1α|,|x3x3u1α|Cδ(x),k,l=1,2.\displaystyle\Big{|}\partial_{x_{k}x_{l}}u_{1}^{\alpha}\Big{|}\leq\frac{C|x^{\prime}|^{m-1}}{\delta(x^{\prime})},~{}\Big{|}\partial_{x_{k}x_{3}}u_{1}^{\alpha}\Big{|},\Big{|}\partial_{x_{3}x_{3}}u_{1}^{\alpha}\Big{|}\leq\frac{C}{\delta(x^{\prime})},\quad k,l=1,2.

We thus obtain

|λ,μu1α|Cδ(x),α=4,5,6.\displaystyle\Big{|}\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}\Big{|}\leq\frac{C}{\delta(x^{\prime})},\quad\alpha=4,5,6.

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 (v1αu¯1α)\nabla(v_{1}^{\alpha}-\bar{u}_{1}^{\alpha}) are bounded, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2. Moreover, ΩΩR|u~1α|2\int_{\Omega\setminus\Omega_{R}}|\nabla\tilde{u}_{1}^{\alpha}|^{2} are also bounded because of u1αC2(ΩΩR)C\|u_{1}^{\alpha}\|_{C^{2}(\Omega\setminus\Omega_{R})}\leq C. So it suffices to prove the bundedness of ΩR|u~1α|2\int_{\Omega_{R}}|\nabla\tilde{u}_{1}^{\alpha}|^{2}, α=1,,d\alpha=1,\cdots,d, since u~1α=0\tilde{u}_{1}^{\alpha}=0, α=d+1,,d(d+1)/2\alpha=d+1,\cdots,d(d+1)/2.

When d=2d=2. Recalling the definition of u~1α\tilde{u}_{1}^{\alpha} in (2.16), we have

(u~11)1=0,|x1(u~11)2|C,|x2(u~11)2|C|x1|m1δ(x1);\displaystyle\nabla(\tilde{u}_{1}^{1})^{1}=0,\quad\Big{|}\partial_{x_{1}}(\tilde{u}_{1}^{1})^{2}\Big{|}\leq C,\quad\Big{|}\partial_{x_{2}}(\tilde{u}_{1}^{1})^{2}\Big{|}\leq\frac{C|x_{1}|^{m-1}}{\delta(x_{1})};

and

(u~12)2=0,|x1(u~12)1|C,|x2(u~12)1|C|x1|m1δ(x1).\displaystyle\nabla(\tilde{u}_{1}^{2})^{2}=0,\quad\Big{|}\partial_{x_{1}}(\tilde{u}_{1}^{2})^{1}\Big{|}\leq C,\quad\Big{|}\partial_{x_{2}}(\tilde{u}_{1}^{2})^{1}\Big{|}\leq\frac{C|x_{1}|^{m-1}}{\delta(x_{1})}.

Hence,

ΩR|u~1α|2𝑑xΩRC|x1|2(m1)δ2(x1)𝑑xC|x1|<R|x1|2(m1)δ(x1)𝑑x1C.\displaystyle\int_{\Omega_{R}}|\nabla\tilde{u}_{1}^{\alpha}|^{2}\ dx\leq\int_{\Omega_{R}}\frac{C|x_{1}|^{2(m-1)}}{\delta^{2}(x_{1})}\ dx\leq C\int_{|x_{1}|<R}\frac{|x_{1}|^{2(m-1)}}{\delta(x_{1})}\ dx_{1}\leq C.

When d=3d=3. We obtain from (2.17) that

(u~1α)β=0,|x(u~1α)3|C,|x3(u~1α)3|C|x|m1δ(x),α,β=1,2;\displaystyle\nabla(\tilde{u}_{1}^{\alpha})^{\beta}=0,\quad\Big{|}\nabla_{x^{\prime}}(\tilde{u}_{1}^{\alpha})^{3}\Big{|}\leq C,\quad\Big{|}\partial_{x_{3}}(\tilde{u}_{1}^{\alpha})^{3}\Big{|}\leq\frac{C|x^{\prime}|^{m-1}}{\delta(x^{\prime})},\quad\alpha,\beta=1,2;

and

|x(u~13)β|C,|x3(u~13)β|C|x|m1δ(x),(u~13)3=0,β=1,2.\displaystyle\Big{|}\nabla_{x^{\prime}}(\tilde{u}_{1}^{3})^{\beta}\Big{|}\leq C,\quad\Big{|}\partial_{x_{3}}(\tilde{u}_{1}^{3})^{\beta}\Big{|}\leq\frac{C|x^{\prime}|^{m-1}}{\delta(x^{\prime})},\quad\nabla(\tilde{u}_{1}^{3})^{3}=0,\quad\beta=1,2.

Thus,

ΩR|u~1α|2𝑑xΩRC|x|2(m1)δ2(x)𝑑xC|x|<R|x|2(m1)δ(x)𝑑xC.\displaystyle\int_{\Omega_{R}}|\nabla\tilde{u}_{1}^{\alpha}|^{2}\ dx\leq\int_{\Omega_{R}}\frac{C|x^{\prime}|^{2(m-1)}}{\delta^{2}(x^{\prime})}\ dx\leq C\int_{|x^{\prime}|<R}\frac{|x^{\prime}|^{2(m-1)}}{\delta(x^{\prime})}\ dx^{\prime}\leq C.

Therefore, the boundedness of the global energy of (v1αu1α)\nabla(v_{1}^{\alpha}-u_{1}^{\alpha}) is established.

Step 3. Proof of

Ωδ(z)|w1α|2𝑑xCδd(z){δ2(12m)(z),α=1,,d,1,α=(d+1),,d(d+1)/2.\displaystyle\int_{\Omega_{\delta}(z^{\prime})}|\nabla w_{1}^{\alpha}|^{2}\ dx\leq C\delta^{d}(z^{\prime})\begin{cases}\delta^{2(1-\frac{2}{m})}(z^{\prime}),&\quad\alpha=1,\cdots,d,\\ 1,&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases} (8.12)

where

Ωs(z):={(x,xd)d|h2(x)<xd<ε+h1(x),|xz|<r},s<R,\Omega_{s}(z^{\prime}):=\left\{(x^{\prime},x_{d})\in\mathbb{R}^{d}~{}\big{|}~{}h_{2}(x^{\prime})<x_{d}<\varepsilon+h_{1}(x^{\prime}),~{}|x^{\prime}-z^{\prime}|<r\right\},\quad s<R,

and w1α=v1αu1αw_{1}^{\alpha}=v_{1}^{\alpha}-u_{1}^{\alpha}, α=1,,d(d+1)/2\alpha=1,\cdots,d(d+1)/2, satisfying

{λ,μw1α=λ,μu1α,inΩ,w=0,onΩ.\displaystyle\begin{cases}\mathcal{L}_{\lambda,\mu}w_{1}^{\alpha}=-\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha},&\hbox{in}\ \Omega,\\ w=0,\quad&\hbox{on}\ \partial\Omega.\end{cases} (8.13)

We will use the iteration scheme developed in [10, 11, 39] to prove (8.12). For 0<t<s<R0<t<s<R, let η\eta be a smooth cutoff function satisfying η(x)=1\eta(x^{\prime})=1 if |xz|<t|x^{\prime}-z^{\prime}|<t, η(x)=0\eta(x^{\prime})=0 if |xz|>s|x^{\prime}-z^{\prime}|>s, 0η(x)10\leq\eta(x^{\prime})\leq 1 if t|xz|st\leq|x^{\prime}-z^{\prime}|\leq\,s, and |xη(x)|2st|\nabla_{x^{\prime}}\eta(x^{\prime})|\leq\frac{2}{s-t}. Multiplying the equation in (8.13) by w1αη2w_{1}^{\alpha}\eta^{2} and applying integration by parts, Hölder’s inequality, and Cauchy inequality, we get

Ωt(z)|w1α|2𝑑xC(st)2Ωs(z)|w1α|2𝑑x+C(st)2Ωs(z)|λ,μu1α|2𝑑x.\displaystyle\int_{\Omega_{t}(z^{\prime})}|\nabla w_{1}^{\alpha}|^{2}\ dx\leq\frac{C}{(s-t)^{2}}\int_{\Omega_{s}(z^{\prime})}|w_{1}^{\alpha}|^{2}\ dx+C(s-t)^{2}\int_{\Omega_{s}(z^{\prime})}|\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}|^{2}\ dx. (8.14)

On one hand, we obtain from Hölder’s inequality that

Ωs(z)|w1α|2𝑑x=Ωs(z)|h2(x)xdxdw1α(x,ξ)dξ|2𝑑xCδ2(z)Ωs(z)|w1α|2𝑑x.\displaystyle\int_{\Omega_{s}(z^{\prime})}|w_{1}^{\alpha}|^{2}\ dx=\int_{\Omega_{s}(z^{\prime})}\left|\int_{h_{2}(x^{\prime})}^{x_{d}}\partial_{x_{d}}w_{1}^{\alpha}(x^{\prime},\xi)\ d\xi\right|^{2}\ dx\leq\,C\delta^{2}(z^{\prime})\int_{\Omega_{s}(z^{\prime})}|\nabla w_{1}^{\alpha}|^{2}\ dx. (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. |z|ε1/m|z^{\prime}|\leq\varepsilon^{1/m}. By using (8.1), we have for 0<s<ε1/m0<s<\varepsilon^{1/m},

Ωs(z)|λ,μu1α|2𝑑xCsd1{ε2(m2)m1,α=1,,d,ε1,α=(d+1),,d(d+1)/2.\displaystyle\int_{\Omega_{s}(z^{\prime})}|\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}|^{2}\ dx\leq Cs^{d-1}\begin{cases}\varepsilon^{\frac{2(m-2)}{m}-1},&\quad\alpha=1,\cdots,d,\\ \varepsilon^{-1},&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases} (8.16)

This is an improvement of [10, (3.32),(3.35)]. Denote

F(t):=Ωt(z)|w1α|2.F(t):=\int_{\Omega_{t}(z^{\prime})}|\nabla w_{1}^{\alpha}|^{2}.

Then substituting (8.15) and (8.16) into (8.14), we have

F(t)(c1εst)2F(s)+C(st)2sd1{ε2(m2)m1,α=1,,d,ε1,α=(d+1),,d(d+1)/2,F(t)\leq\left(\frac{c_{1}\varepsilon}{s-t}\right)^{2}F(s)+C(s-t)^{2}s^{d-1}\begin{cases}\varepsilon^{\frac{2(m-2)}{m}-1},&~{}\alpha=1,\cdots,d,\\ \varepsilon^{-1},&~{}\alpha=(d+1),\cdots,d(d+1)/2,\end{cases} (8.17)

where c1c_{1} is a universal canstant.

Let k=[14c1ε1/m]k=\left[\frac{1}{4c_{1}\varepsilon^{1/m}}\right] and ti=2c1iε,i=1,,kt_{i}=2c_{1}i\varepsilon,i=1,\cdots,k. Then by (8.17) with s=ti+1s=t_{i+1} and t=tit=t_{i}, we have

F(ti)14F(ti+1)+C(i+1)d1εd{ε2(m2)m,α=1,,d,1,α=(d+1),,d(d+1)/2.F(t_{i})\leq\frac{1}{4}F(t_{i+1})+C(i+1)^{d-1}\varepsilon^{d}\begin{cases}\varepsilon^{\frac{2(m-2)}{m}},&\quad\alpha=1,\cdots,d,\\ 1,&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases}

After kk iterations, using the global boundedness of w1α\nabla w_{1}^{\alpha}, we have

F(t1)Cεd{ε2(m2)m,α=1,,d,1,α=(d+1),,d(d+1)/2.F(t_{1})\leq C\varepsilon^{d}\begin{cases}\varepsilon^{\frac{2(m-2)}{m}},&\quad\alpha=1,\cdots,d,\\ 1,&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases}

Case 2. ε1/m<|z|<R\varepsilon^{1/m}<|z^{\prime}|<R. For 0<s<23|z|0<s<\frac{2}{3}|z^{\prime}|, (8.16) becomes

Ωs(z)|λ,μu1α|2𝑑xCsd1{|z|m4,α=1,,d,|z|m,α=(d+1),,d(d+1)/2.\displaystyle\int_{\Omega_{s}(z^{\prime})}|\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}|^{2}\ dx\leq Cs^{d-1}\begin{cases}|z^{\prime}|^{m-4},&\quad\alpha=1,\cdots,d,\\ |z^{\prime}|^{-m},&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases}

(8.17) becomes

F(t)(c2|z|mst)2F(s)+C(st)2sd1{|z|m4,α=1,,d,|z|m,α=(d+1),,d(d+1)/2,F(t)\leq\left(\frac{c_{2}|z^{\prime}|^{m}}{s-t}\right)^{2}F(s)+C(s-t)^{2}s^{d-1}\begin{cases}|z^{\prime}|^{m-4},&\quad\alpha=1,\cdots,d,\\ |z^{\prime}|^{-m},&\quad\alpha=(d+1),\cdots,d(d+1)/2,\end{cases}

where c2c_{2} is another universal canstant. Let k=[14c2|z|]k=\left[\frac{1}{4c_{2}|z^{\prime}|}\right] and ti=2c2i|z|m,i=1,,kt_{i}=2c_{2}i|z^{\prime}|^{m},i=1,\cdots,k. Then by (8.17) with s=ti+1s=t_{i+1} and t=tit=t_{i}, we have

F(ti)14F(ti+1)+C(i+1)d1|z|md{|z|2(m2),α=1,,d,1,α=(d+1),,d(d+1)/2.F(t_{i})\leq\frac{1}{4}F(t_{i+1})+C(i+1)^{d-1}|z^{\prime}|^{md}\begin{cases}|z^{\prime}|^{2(m-2)},&\quad\alpha=1,\cdots,d,\\ 1,&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases}

After kk iterations, using the global boundedness of w1α\nabla w_{1}^{\alpha}, we have

F(t1)C|z|md{|z|2(m2),α=1,,d,1,α=(d+1),,d(d+1)/2.F(t_{1})\leq C|z^{\prime}|^{md}\begin{cases}|z^{\prime}|^{2(m-2)},&\quad\alpha=1,\cdots,d,\\ 1,&\quad\alpha=(d+1),\cdots,d(d+1)/2.\end{cases}

So (8.12) is proved.

Step 4. Scaling and LL^{\infty}-estimates. It follows from [10, (3.40)] that

w1αL(Ωδ/2(z))Cδ(δ1d2w1αL2(Ωδ(z))+δ2(z)λ,μu1αL(Ωδ(z))).\displaystyle\|\nabla w_{1}^{\alpha}\|_{L^{\infty}(\Omega_{\delta/2}(z^{\prime}))}\leq\frac{C}{\delta}\left(\delta^{1-\frac{d}{2}}\|\nabla w_{1}^{\alpha}\|_{L^{2}(\Omega_{\delta}(z^{\prime}))}+\delta^{2}(z^{\prime})\|\mathcal{L}_{\lambda,\mu}u_{1}^{\alpha}\|_{L^{\infty}(\Omega_{\delta}(z^{\prime}))}\right).

By using (8.12) and (8.1), we obtain

|w1α(z,xd)|C,h2(z)<xd<ε+h1(z).|\nabla w_{1}^{\alpha}(z^{\prime},x_{d})|\leq\,C,\quad h_{2}(z^{\prime})<x_{d}<\varepsilon+h_{1}(z^{\prime}).

Theorems 2.2 and 2.6 are proved. ∎

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 2D2D. 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 pp-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 MM-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.