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

Sharp estimate of electric field from a conductive rod and application in EIT

Xiaoping Fang School of Mathematics and Statistics, Key Laboratory of Hunan Province for Statistical Learning and Intelligent Computation, Hunan University of Technology and Business, Changsha 410205, China [email protected] Youjun Deng School of Mathematics and Statistics, Central South University, Changsha, Hunan, China. [email protected], [email protected]  and  Hongyu Liu Department of Mathematics, City University of Hong Kong, Kowloon, Hong Kong SAR, China. [email protected]; [email protected]
Abstract.

We are concerned with the quantitative study of the electric field perturbation due to the presence of an inhomogeneous conductive rod embedded in a homogenous conductivity. We sharply quantify the dependence of the perturbed electric field on the geometry of the conductive rod. In particular, we accurately characterise the localisation of the gradient field (i.e. the electric current) near the boundary of the rod where the curvature is sufficiently large. We develop layer-potential techniques in deriving the quantitative estimates and the major difficulty comes from the anisotropic geometry of the rod. The result complements and sharpens several existing studies in the literature. It also generates an interesting application in EIT (electrical impedance tomography) in determining the conductive rod by a single measurement, which is also known as the Calderón’s inverse inclusion problem in the literature.

Keywords:  conductivity equation, rod inclusion, Neumann-Poincaré operator, asymptotic analysis, electrical impedance tomography, single measurement.

2010 Mathematics Subject Classification:  35Q60, 35J05, 31B10, 78A40

1. Introduction

1.1. Mathematical setup

Initially focusing on the mathematics, but not the physics, we consider the following elliptic PDE system in 2\mathbb{R}^{2}:

{(σ(𝐱)u(𝐱))=0,𝐱=(x1,x2)2,u(𝐱)H(𝐱)=𝒪(|𝐱|1),|𝐱|,\left\{\begin{split}\nabla\cdot\Big{(}\sigma(\mathbf{x})\nabla u(\mathbf{x})\Big{)}=0,&\quad\mathbf{x}=(x_{1},x_{2})\in\mathbb{R}^{2},\\ u(\mathbf{x})-H(\mathbf{x})=\mathcal{O}(|\mathbf{x}|^{-1}),&\quad|\mathbf{x}|\rightarrow\infty,\end{split}\right. (1.1)

where u(𝐱)Hloc1(2)u(\mathbf{x})\in H_{loc}^{1}(\mathbb{R}^{2}) is a potential field, and σ(𝐱)\sigma(\mathbf{x}) is of the following form:

σ:=(σ01)χ(D)+1,σ0+andσ01,\sigma:=(\sigma_{0}-1)\chi(D)+1,\ \ \sigma_{0}\in\mathbb{R}_{+}\ \mbox{and}\ \sigma_{0}\neq 1, (1.2)

with DD being a bounded domain with a connected complement 2\D¯\mathbb{R}^{2}\backslash\overline{D}. H(𝐱)H(\mathbf{x}) in (1.1) is a (nontrivial) harmonic function in 2\mathbb{R}^{2}, which stands for a background potential. We are mainly concerned with the quantitative properties of the solution u(𝐱)u(\mathbf{x}) in (1.1) and in particular, its dependence on the geometry of DD. To that end, we next introduce the rod-geometry of DD for our subsequent study. Let Γ0\Gamma_{0} be a straight line of length L+L\in\mathbb{R}_{+} with Γ0=(x1,0)\Gamma_{0}=(x_{1},0), x1(L/2,L/2)x_{1}\in(-L/2,L/2). Let 𝐧:=(0,1)\mathbf{n}:=(0,1) and define the two points P:=(L/2,0)P:=(-L/2,0) and Q:=(L/2,0)Q:=(L/2,0). Then the rod DD is introduced as D¯=Da¯Df¯Db¯\overline{D}=\overline{D^{a}}\cup\overline{D^{f}}\cup\overline{D^{b}}, where DfD^{f} is defined by

Df:={𝐱;𝐱=Γ0±t𝐧,t(δ,δ)},δ+.D^{f}:=\{\mathbf{x};\ \mathbf{x}=\Gamma_{0}\pm t\mathbf{n},\ t\in(-\delta,\delta)\},\ \ \delta\in\mathbb{R}_{+}. (1.3)

The two end-caps DaD^{a} and DbD^{b} are two half disks with radius δ\delta and centering at PP and QQ, respectively. More precisely,

Da={𝐱;|𝐱P|<δ,x1<L/2},Db={𝐱;|𝐱Q|<δ,x1>L/2}.D^{a}=\{\mathbf{x};\ |\mathbf{x}-P|<\delta,\ x_{1}<-L/2\},\quad D^{b}=\{\mathbf{x};\ |\mathbf{x}-Q|<\delta,\ x_{1}>L/2\}.

It can be verified that DD is of class C1,αC^{1,\alpha} for a certain α+\alpha\in\mathbb{R}_{+}. In what follows, we define Sc:=Dc=(DaDb)S^{c}:=\partial D^{c}=\partial(D^{a}\cup D^{b}), and Sf:=DfS^{f}:=\partial D^{f}. Specially, Sf=Γ1Γ2S^{f}=\Gamma_{1}\cup\Gamma_{2}, where Γj\Gamma_{j}, j=1,2j=1,2 are defined by

Γ1={𝐱;𝐱=Γ0δ𝐧},Γ2={𝐱;𝐱=Γ0+δ𝐧}.\Gamma_{1}=\{\mathbf{x};\ \mathbf{x}=\Gamma_{0}-\delta\mathbf{n}\},\quad\Gamma_{2}=\{\mathbf{x};\ \mathbf{x}=\Gamma_{0}+\delta\mathbf{n}\}. (1.4)

Moreover, we shall always use 𝐳x\mathbf{z}_{x} and 𝐳y\mathbf{z}_{y} to signify the projections of 𝐱Sf\mathbf{x}\in S^{f} and 𝐲Sf\mathbf{y}\in S^{f} on Γ0\Gamma_{0}, respectively.

The elliptic PDE system (1.1) describes the perturbation of an electric field H(𝐱)H(\mathbf{x}) due to the presence of a conductive body DD. u(𝐱)u(\mathbf{x}) signifies the electric potential and σ(𝐱)\sigma(\mathbf{x}) signifies the conductivity of the space. The homogeneous background space possesses a conductivity being normalized to be 1, whereas the conductive rod DD possesses an inhomogeneous conductivity being σ0\sigma_{0}. The perturbed electric potential is uHu-H and the gradient field (uH)\nabla(u-H) is the corresponding electric current.

1.2. Background discussion and literature review

The conductivity equation (1.1) is a fundamental problem in many existing studies. It is the master equation for the electrical impedance tomography (EIT) which is an important medical imaging modality and it can also find important applications in materials science; see e.g. [5, 6, 15, 43] and the references cited therein. There are rich results in the literature devoted to the quantitative properties of the solution to (1.1) and its geometric relationship to the conductive inclusion DD. In this work, we shall derive an accurate characterisation of the perturbed electric field uHu-H and its dependence on the geometry of DD. There are mainly two motivations for our study as described in what follows.

First, in [4, 9, 14, 34], the authors studied the electric field perturbation from thin/slender structures, which are originated from the study of imaging crack defects in EIT [4, 9, 41]. This is closely related to the current study. Indeed, the geometric setup in the aforementioned works are more general than the one considered in the present article. However, with the specific rod-geometry, we can derive an accurate characterisation of the perturbed electric field and its geometric dependence on DD. In fact, in our asymptotic formula of the electric field uHu-H (with respect to δ1\delta\ll 1), the leading-order term is exact and can be used to fully decode DD. This is in a sharp contrast to the existing studies mentioned above, which inevitably involve some qualitative estimates due to the more general geometries. The rod-geometry, though special, also possesses several local features that are worth our investigation, which is the second motivation of our study as described below.

Clearly, the studies mentioned above are mainly concerned with extracting the global geometry of D\partial D from the perturbed electric field uHu-H. On the other hand, there are studies on the relationships between the local geometry of D\partial D and the perturbed field uHu-H. In fact, there are classical results concerning the singularities of the solution near a boundary corner point [28]. Roughly speaking, if D\partial D possesses a corner, then the solution uu locally around the corner point can be decomposed into the sum of a regular part and a singular part. Such a qualitative property has been used to establish novel uniqueness and stability results for the Calderón inverse inclusion problem in EIT by a single partial boundary measurement [17, 37, 38]. The Calderón inverse inclusion problem is a longstanding problem in the literature and we shall present more background discussion in Section 4. In [18], the corner singularities of a conductive inclusion have been characterized in terms of the generalised polarisation tensors associated with the electric potential uu, and the results are directly applied to EIT. Recently, in [39], the authors consider the case that D\partial D is smooth, but possesses high-curvature points. In two dimensions, a high-curvature point means that the extrinsic curvature of the boundary curve D\partial D at that point is sufficiently large. It is shown in [39] that the quantitative property of u\nabla u around the high-curvature point enables one to recover the local part of D\partial D around that high-curvature point. However, the sharp curvature-dependence of u\nabla u in [39] is established through numerically refining the upper-bound estimate in [24]. As mentioned before, the rod-geometry possesses a few interesting local features that consolidate the numerical study in [39]. First, it is geometrically anisotropic where the two dimensions are of different scales. In fact, the curvature at the two end-caps (i.e. Sc=DaDbS^{c}=\partial D^{a}\cup\partial D^{b}) is δ11\delta^{-1}\geq 1, whereas the curvature at the facade part (i.e. SfS^{f}) is 0. Hence, the rod-geometry, though special, provides rich insights on the curvature dependence of the electric field with respect to the shape of the conductive inclusion. In fact, we shall see that the perturbed electric energy is localized at the two end-caps of the rod. Similar to [39], the result enables us to rigorously justify that one can uniquely determine the conductive rod by a single measurement in EIT. It is emphasized that in three dimensions or in the case that the rod is curved, the situation would become much more complicated. Hence, we mainly consider the case with a straight rod in the two dimensions. Nevertheless, even in such a case, the mathematical analysis is technically involved and highly nontrivial. We shall develop layer potential techniques to tackle the problem. It turns out that the so-called Neumann-Poincaré (NP) operator shall play a critical role in our analysis. We would like to mention that the NP operator and its spectral properties have received considerable attentions recently in the literature due to its important applications in several intriguing fields of mathematical physics, including plasmon resonances and invisibility cloaking [7, 8, 10, 11, 19, 20, 23, 26, 29, 32, 31, 35]. Finally, we would also like to mention in passing that more general rod-geometries were also considered in the literature in different contexts of physical importance [21, 22, 36].

The rest of the paper is organized as follows. In Section 2, we derive several auxiliary results and in Section 3, we present the main results on the quantitative analysis of the solution uu to (1.1) with respect to the geometry of the inclusion DD. Finally, we consider in Section 4 the application of the quantitative result derived in Section 3 to Calderón’s inverse inclusion problem.

2. Some auxiliary results

In this section, we shall establish several auxiliary results for our subsequent use. We first present some preliminary knowledge on the layer potential operators for solving the conductivity problem (1.1), and we also refer to [6, 40] for more related results and discussions.

2.1. Layer potentials

Let GG be the radiating fundamental solution to the Laplacian Δ\Delta in 2\mathbb{R}^{2}, which is given by

G(𝐱)=12πln|𝐱|,\displaystyle G(\mathbf{x})=\frac{1}{2\pi}\ln|\mathbf{x}|, (2.1)

For any bounded Lipschitz domain B2B\subset\mathbb{R}^{2}, we denote by 𝒮B:L2(B)H1(2B)\mathcal{S}_{B}:L^{2}(\partial B)\rightarrow H^{1}(\mathbb{R}^{2}\setminus\partial B) the single-layer potential operator given by

𝒮B[ϕ](𝐱):=BG(𝐱𝐲)ϕ(𝐲)𝑑s𝐲,\mathcal{S}_{B}[\phi](\mathbf{x}):=\int_{\partial B}G(\mathbf{x}-\mathbf{y})\phi(\mathbf{y})\;ds_{\mathbf{y}}, (2.2)

and 𝒦B:H1/2(B)H1/2(B)\mathcal{K}_{B}^{*}:H^{-1/2}(\partial B)\rightarrow H^{-1/2}(\partial B) the Neumann-Poincaré (NP) operator:

𝒦B[ϕ](𝐱):=p.v.12πB𝐱𝐲,ν𝐱|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲,\mathcal{K}_{B}^{*}[\phi](\mathbf{x}):=\mbox{p.v.}\quad\frac{1}{2\pi}\int_{\partial B}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{\mathbf{x}}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})\;ds_{\mathbf{y}}, (2.3)

where p.v. stands for the Cauchy principle value. In (2.3) and also in what follows, unless otherwise specified, ν\nu signifies the exterior unit normal vector to the boundary of the domain concerned. It is known that the single-layer potential operator 𝒮B\mathcal{S}_{B} is continuous across B\partial B and satisfies the following trace formula

ν𝒮B[ϕ]|±=(±12I+𝒦B)[ϕ]onB,\frac{\partial}{\partial\nu}\mathcal{S}_{B}[\phi]\Big{|}_{\pm}=\Big{(}\pm\frac{1}{2}I+\mathcal{K}_{B}^{*}\Big{)}[\phi]\quad\mbox{on}\quad\partial B, (2.4)

where ν\frac{\partial}{\partial\nu} stands for the normal derivative and the subscripts ±\pm indicate the limits from outside and inside of a given inclusion BB, respectively.

By using the layer-potential techniques, one can readily find the integral solution to (1.1) by

u=H+𝒮D[φ],u=H+\mathcal{S}_{D}[\varphi], (2.5)

where the density function φH1/2(D)\varphi\in H^{-1/2}(\partial D) is determined by

φ=(λI𝒦D)1[Hν|D].\varphi=\Big{(}\lambda I-\mathcal{K}_{D}^{*}\Big{)}^{-1}\Big{[}\frac{\partial H}{\partial\nu}\Big{|}_{\partial D}\Big{]}. (2.6)

Here, the constant λ\lambda is defined by

λ:=σ0+12(σ01).\lambda:=\frac{\sigma_{0}+1}{2(\sigma_{0}-1)}.

2.2. Asymptotic expansion of the Neumann-Poincaré operator

In what follows, we always suppose that δ1\delta\ll 1. We shall present some asymptotic expansions of the Neumann-Poincaré operator with respect to δ\delta. Recalling that D=SaSf¯Sb\partial D=S^{a}\cup\overline{S^{f}}\cup S^{b}, we decompose the Neumann-Poincaré operator into several parts accordingly. To that end, we introduce the following boundary integral operator:

𝒦𝒮,𝒮[ϕ](𝐱):=χ(𝒮)12π𝒮𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲,for𝒮𝒮=.\mathcal{K}_{\mathcal{S},\mathcal{S}^{\prime}}[\phi](\mathbf{x}):=\chi(\mathcal{S}^{\prime})\frac{1}{2\pi}\int_{\mathcal{S}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}},\quad for\quad\mathcal{S}\cap\mathcal{S}^{\prime}=\emptyset. (2.7)

It is obvious that 𝒦𝒮,𝒮\mathcal{K}_{\mathcal{S},\mathcal{S}^{\prime}} is a bounded operator from L2(𝒮)L^{2}(\mathcal{S}) to L2(𝒮)L^{2}(\mathcal{S}^{\prime}). For the case 𝒮=Sc\mathcal{S}=S^{c}, we mean Sc=SaSbS^{c}=S^{a}\cup S^{b}. In what follows, we define S1aS_{1}^{a} and S1bS_{1}^{b} by

S1a={𝐱;|𝐱P|=1,x1<L/2},S1b={𝐱;|𝐱Q|=1,x1>L/2}.S_{1}^{a}=\{\mathbf{x};\ |\mathbf{x}-P|=1,\ x_{1}<-L/2\},\quad S_{1}^{b}=\{\mathbf{x};\ |\mathbf{x}-Q|=1,\ x_{1}>L/2\}. (2.8)

For the subsequent use, we also introduce the following regions:

ιδ(P):={𝐱;|P𝐳x|=𝒪(δ),𝐱Sf},\displaystyle\iota_{\delta}(P):=\{\mathbf{x};\ |P-\mathbf{z}_{x}|=\mathcal{O}(\delta),\ \mathbf{x}\in S^{f}\}, (2.9)
ιδ(Q):={𝐱;|Q𝐳x|=𝒪(δ),𝐱Sf}.\displaystyle\iota_{\delta}(Q):=\{\mathbf{x};\ |Q-\mathbf{z}_{x}|=\mathcal{O}(\delta),\ \mathbf{x}\in S^{f}\}. (2.10)

Define ϕ~(𝐱~):=ϕ(𝐱)\tilde{\phi}(\tilde{\mathbf{x}}):=\phi(\mathbf{x}), where 𝐱Sa,Sb\mathbf{x}\in S^{a},S^{b} and 𝐱~S1a,S1b\tilde{\mathbf{x}}\in S_{1}^{a},S_{1}^{b}.

We can prove the following result on the asymptotic expansion of the Neumann-Poincaré operator.

Lemma 2.1.

The Neumann-Poincaré operator 𝒦D\mathcal{K}_{D}^{*} admits the following asymptotic expansion:

𝒦D[ϕ](𝐱)=𝒦0[ϕ](𝐱)+δ𝒦1[ϕ](𝐱)+𝒪(δ2),\mathcal{K}_{D}^{*}[\phi](\mathbf{x})=\mathcal{K}_{0}[\phi](\mathbf{x})+\delta\mathcal{K}_{1}[\phi](\mathbf{x})+\mathcal{O}(\delta^{2}), (2.11)

where 𝒦0\mathcal{K}_{0} is defined by

𝒦0[ϕ](𝐱)=χ(Sa)(𝒦Sf,Sa[ϕ](𝐱)+14πS1aϕ~(𝐲))+χ(Sb)(𝒦Sf,Sb[ϕ](𝐱)+14πS1bϕ~)+𝒜Γ2,Γ1[ϕ]+𝒜Γ1,Γ2[ϕ]+χ(ιδ(P))𝒦Sa,Sf[ϕ](𝐱)+χ(ιδ(Q))𝒦Sb,Sf[ϕ](𝐱),\begin{split}\mathcal{K}_{0}[\phi](\mathbf{x})&=\chi(S^{a})\Big{(}\mathcal{K}_{S^{f},S^{a}}[\phi](\mathbf{x})+\frac{1}{4\pi}\int_{S_{1}^{a}}\tilde{\phi}(\mathbf{y})\Big{)}+\chi(S^{b})\Big{(}\mathcal{K}_{S^{f},S^{b}}[\phi](\mathbf{x})+\frac{1}{4\pi}\int_{S_{1}^{b}}\tilde{\phi}\Big{)}\\ &+\mathcal{A}_{\Gamma_{2},\Gamma_{1}}[\phi]+\mathcal{A}_{\Gamma_{1},\Gamma_{2}}[\phi]+\chi(\iota_{\delta}(P))\mathcal{K}_{S^{a},S^{f}}[\phi](\mathbf{x})+\chi(\iota_{\delta}(Q))\mathcal{K}_{S^{b},S^{f}}[\phi](\mathbf{x}),\end{split} (2.12)

and

𝒦1[ϕ]=χ(Sb)𝐱P,νx2π|𝐱P|2S1aϕ~+χ(Sa)𝐱Q,νx2π|𝐱Q|2S1bϕ~+χ(Sfιδ(P))(δ|𝐱P|2S1a(1𝐲~P,νx)ϕ~(𝐲~)𝑑s𝐲~+o(δ|𝐱P|2))+χ(Sfιδ(Q))(δ|𝐱Q|2S1b(1𝐲~Q,νx)ϕ~(𝐲~)𝑑s𝐲~+o(δ|𝐱P|2)).\begin{split}\mathcal{K}_{1}[\phi]=&\chi(S^{b})\frac{\langle\mathbf{x}-P,\nu_{x}\rangle}{2\pi|\mathbf{x}-P|^{2}}\int_{S_{1}^{a}}\tilde{\phi}+\chi(S^{a})\frac{\langle\mathbf{x}-Q,\nu_{x}\rangle}{2\pi|\mathbf{x}-Q|^{2}}\int_{S_{1}^{b}}\tilde{\phi}\\ &+\chi(S^{f}\setminus\iota_{\delta}(P))\left(\frac{\delta}{|\mathbf{x}-P|^{2}}\int_{S_{1}^{a}}(1-\langle\tilde{\mathbf{y}}-P,\nu_{x}\rangle)\tilde{\phi}(\tilde{\mathbf{y}})ds_{\tilde{\mathbf{y}}}+o\Big{(}\frac{\delta}{|\mathbf{x}-P|^{2}}\Big{)}\right)\\ &+\chi(S^{f}\setminus\iota_{\delta}(Q))\left(\frac{\delta}{|\mathbf{x}-Q|^{2}}\int_{S_{1}^{b}}(1-\langle\tilde{\mathbf{y}}-Q,\nu_{x}\rangle)\tilde{\phi}(\tilde{\mathbf{y}})ds_{\tilde{\mathbf{y}}}+o\Big{(}\frac{\delta}{|\mathbf{x}-P|^{2}}\Big{)}\right).\end{split} (2.13)

Here, the operators 𝒜Γ1,Γ2\mathcal{A}_{\Gamma_{1},\Gamma_{2}} and 𝒜Γ2,Γ1\mathcal{A}_{\Gamma_{2},\Gamma_{1}} are defined by

𝒜Γ1,Γ2[ϕ](𝐱)=1πχ(Γ2)Γ1δ|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲,𝒜Γ2,Γ1[ϕ](𝐱)=1πχ(Γ1)Γ2δ|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲.\begin{split}\mathcal{A}_{\Gamma_{1},\Gamma_{2}}[\phi](\mathbf{x})=&\frac{1}{\pi}\chi(\Gamma_{2})\int_{\Gamma_{1}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}},\\ \mathcal{A}_{\Gamma_{2},\Gamma_{1}}[\phi](\mathbf{x})=&\frac{1}{\pi}\chi(\Gamma_{1})\int_{\Gamma_{2}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}.\end{split} (2.14)
Proof.

First, one has the following separation:

𝒦D[ϕ](𝐱)=12πSf𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+12πSaSb𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲=:A1[ϕ](𝐱)+A2[ϕ](𝐱).\begin{split}\mathcal{K}_{D}^{*}[\phi](\mathbf{x})=&\frac{1}{2\pi}\int_{S^{f}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}+\frac{1}{2\pi}\int_{S^{a}\cup S^{b}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =:&A_{1}[\phi](\mathbf{x})+A_{2}[\phi](\mathbf{x}).\end{split} (2.15)

Note that for 𝐱,𝐲Γj\mathbf{x},\mathbf{y}\in\Gamma_{j}, j=2,3j=2,3, one can easily obtain that 𝐱𝐲,νx=0\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle=0. Thus one has

A1[ϕ](𝐱)=12πSf𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲=χ(SaSb)12πSf𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+χ(Γ1)12πΓ2(𝐱2δνx𝐲)+2δνx,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+χ(Γ2)12πΓ1(𝐱2δνx𝐲)+2δνx,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲=𝒦Sf,Sc[ϕ](𝐱)+δχ(Γ1)1πΓ21|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+δχ(Γ2)1πΓ11|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲.\begin{split}&A_{1}[\phi](\mathbf{x})=\frac{1}{2\pi}\int_{S^{f}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =&\chi(S^{a}\cup S^{b})\frac{1}{2\pi}\int_{S^{f}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ &+\chi(\Gamma_{1})\frac{1}{2\pi}\int_{\Gamma_{2}}\frac{\langle(\mathbf{x}-2\delta\nu_{x}-\mathbf{y})+2\delta\nu_{x},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ &+\chi(\Gamma_{2})\frac{1}{2\pi}\int_{\Gamma_{1}}\frac{\langle(\mathbf{x}-2\delta\nu_{x}-\mathbf{y})+2\delta\nu_{x},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =&\mathcal{K}_{S^{f},S^{c}}[\phi](\mathbf{x})+\delta\chi(\Gamma_{1})\frac{1}{\pi}\int_{\Gamma_{2}}\frac{1}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}+\delta\chi(\Gamma_{2})\frac{1}{\pi}\int_{\Gamma_{1}}\frac{1}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}.\end{split} (2.16)

On the other hand,

A2[ϕ](𝐱)=12πSaSb𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲=χ(Sa)12πSa𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+χ(Sb)12πSb𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+χ(Sa)12πSb𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+χ(Sb)12πSa𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲+χ(Sf)12πSaSb𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲=χ(Sa)14πS1aϕ~(𝐲~)𝑑s𝐲~+χ(Sb)14πS1bϕ~(𝐲~)𝑑s𝐲~+𝒦Sb,Sa[ϕ]+𝒦Sa,Sb[ϕ]+𝒦Sc,Sf[ϕ].\begin{split}&A_{2}[\phi](\mathbf{x})=\frac{1}{2\pi}\int_{S^{a}\cup S^{b}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =&\chi(S^{a})\frac{1}{2\pi}\int_{S^{a}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}+\chi(S^{b})\frac{1}{2\pi}\int_{S^{b}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ &+\chi(S^{a})\frac{1}{2\pi}\int_{S^{b}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}+\chi(S^{b})\frac{1}{2\pi}\int_{S^{a}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ &+\chi(S^{f})\frac{1}{2\pi}\int_{S^{a}\cup S^{b}}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =&\chi(S^{a})\frac{1}{4\pi}\int_{S_{1}^{a}}\tilde{\phi}(\tilde{\mathbf{y}})ds_{\tilde{\mathbf{y}}}+\chi(S^{b})\frac{1}{4\pi}\int_{S_{1}^{b}}\tilde{\phi}(\tilde{\mathbf{y}})ds_{\tilde{\mathbf{y}}}+\mathcal{K}_{S^{b},S^{a}}[\phi]+\mathcal{K}_{S^{a},S^{b}}[\phi]+\mathcal{K}_{S^{c},S^{f}}[\phi].\end{split} (2.17)

For 𝐲Sa\mathbf{y}\in S^{a} and 𝐱Sb\mathbf{x}\in S^{b}, by using Taylor’s expansions one has

|𝐱𝐲|=|𝐱P(𝐲P))|=|𝐱Pδ(𝐲~P)|=|𝐱P|+δ𝐱P,𝐲~P+𝒪(δ2).|\mathbf{x}-\mathbf{y}|=|\mathbf{x}-P-(\mathbf{y}-P))|=|\mathbf{x}-P-\delta(\tilde{\mathbf{y}}-P)|=|\mathbf{x}-P|+\delta\langle\mathbf{x}-P,\tilde{\mathbf{y}}-P\rangle+\mathcal{O}(\delta^{2}). (2.18)

Thus one has

𝒦Sa,Sb[ϕ](𝐱)=δ𝐱P,νx2π|𝐱P|2S1aϕ~+𝒪(δ2).\mathcal{K}_{S^{a},S^{b}}[\phi](\mathbf{x})=\delta\frac{\langle\mathbf{x}-P,\nu_{x}\rangle}{2\pi|\mathbf{x}-P|^{2}}\int_{S_{1}^{a}}\tilde{\phi}+\mathcal{O}(\delta^{2}). (2.19)

Similarly, one can obtain

𝒦Sb,Sa[ϕ](𝐱)=δ𝐱Q,νx2π|𝐱Q|2S1bϕ~+𝒪(δ2).\mathcal{K}_{S^{b},S^{a}}[\phi](\mathbf{x})=\delta\frac{\langle\mathbf{x}-Q,\nu_{x}\rangle}{2\pi|\mathbf{x}-Q|^{2}}\int_{S_{1}^{b}}\tilde{\phi}+\mathcal{O}(\delta^{2}). (2.20)

For 𝐱Sf\mathbf{x}\in S^{f} 𝐲Sa\mathbf{y}\in S^{a}, by direct computations, one can obtain

𝒦Sa,Sf[ϕ](𝐱)=δ2S1a1𝐲~P,νx|𝐱P|22δ𝐱P,𝐲~P+δ2ϕ~(𝐲~)𝑑s𝐲~.\mathcal{K}_{S^{a},S^{f}}[\phi](\mathbf{x})=\delta^{2}\int_{S_{1}^{a}}\frac{1-\langle\tilde{\mathbf{y}}-P,\nu_{x}\rangle}{|\mathbf{x}-P|^{2}-2\delta\langle\mathbf{x}-P,\tilde{\mathbf{y}}-P\rangle+\delta^{2}}\tilde{\phi}(\tilde{\mathbf{y}})ds_{\tilde{\mathbf{y}}}.

We decompose Sf=(Sfιδ(P))ιδ(P)S^{f}=(S^{f}\setminus\iota_{\delta}(P))\cup\iota_{\delta}(P), then one has

𝒦Sa,Sf[ϕ](𝐱)=δ2|𝐱P|2S1a(1𝐲~P,νx)ϕ~(𝐲~)𝑑s𝐲~+o(δ2|𝐱P|2),𝐱Sfιδ(P).\mathcal{K}_{S^{a},S^{f}}[\phi](\mathbf{x})=\frac{\delta^{2}}{|\mathbf{x}-P|^{2}}\int_{S_{1}^{a}}(1-\langle\tilde{\mathbf{y}}-P,\nu_{x}\rangle)\tilde{\phi}(\tilde{\mathbf{y}})\,ds_{\tilde{\mathbf{y}}}+o\Big{(}\frac{\delta^{2}}{|\mathbf{x}-P|^{2}}\Big{)},\quad\mathbf{x}\in S^{f}\setminus\iota_{\delta}(P). (2.21)

Similarly, one can derive the asymptotic expansion for 𝒦Sb,Sf\mathcal{K}_{S^{b},S^{f}}. By substituting (2.19)-(2.21) back into (2.17) and combining (2.16) one finally achieves (2.11), which completes the proof.

The proof is complete. ∎

Lemma 2.2.

The operators 𝒜Γj,Γk\mathcal{A}_{\Gamma_{j},\Gamma_{k}}, {j,k}={1,2},{2,1}\{j,k\}=\{1,2\},\{2,1\}, defined in (2.14) are bounded operators from L2(Γj)L^{2}(\Gamma_{j}) to L2(Γk)L^{2}(\Gamma_{k}). Furthermore, the operators χ(ιδ(P))𝒦Sa,Sf\chi(\iota_{\delta}(P))\mathcal{K}_{S^{a},S^{f}} and χ(ιδ(Q))𝒦Sb,Sf\chi(\iota_{\delta}(Q))\mathcal{K}_{S^{b},S^{f}} are bounded operators from L2(Sa)L^{2}(S^{a}) to L2(Sf)L^{2}(S^{f}), and from L2(Sb)L^{2}(S^{b}) to L2(Sf)L^{2}(S^{f}), respectively.

Proof.

We only prove that 𝒜Γ2,Γ1\mathcal{A}_{\Gamma_{2},\Gamma_{1}} is a bounded operator L2(Γ2)L^{2}(\Gamma_{2}) to L2(Γ1)L^{2}(\Gamma_{1}). First, for ϕ1L2(Γ1)\phi_{1}\in L^{2}(\Gamma_{1}) and ϕ2L2(Γ2)\phi_{2}\in L^{2}(\Gamma_{2}), one has

|𝒜Γ2,Γ1[ϕ2],ϕ1L2(Γ1)|=12π|Γ1Γ2δ|𝐱𝐲|2ϕ2(𝐲)𝑑s𝐲ϕ1(𝐱)𝑑s𝐱|14πΓ1Γ2δ|𝐱𝐲|2ϕ22(𝐲)𝑑s𝐲𝑑s𝐱+14πΓ1Γ2δ|𝐱𝐲|2ϕ12(𝐱)𝑑s𝐲𝑑s𝐱=14πL/2L/2L/2L/2δ|x1y1|2+4δ2𝑑x1ϕ22(𝐲)𝑑y1+14πL/2L/2L/2L/2δ|x1y1|2+4δ2𝑑y1ϕ12(𝐱)𝑑x1=18πL/2L/2(arctanL/2y12δarctanL/2y12δ)ϕ22(𝐲)𝑑y1+18πL/2L/2(arctanL/2x12δarctanL/2x12δ)ϕ12(𝐱)𝑑x1C(ϕ1L2(Γ1)2+ϕ2L2(Γ2)2),\begin{split}&|\langle\mathcal{A}_{\Gamma_{2},\Gamma_{1}}[\phi_{2}],\phi_{1}\rangle_{L^{2}(\Gamma_{1})}|\\ =&\frac{1}{2\pi}\Big{|}\int_{\Gamma_{1}}\int_{\Gamma_{2}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\phi_{2}(\mathbf{y})ds_{\mathbf{y}}\phi_{1}(\mathbf{x})ds_{\mathbf{x}}\Big{|}\\ \leq&\frac{1}{4\pi}\int_{\Gamma_{1}}\int_{\Gamma_{2}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\phi_{2}^{2}(\mathbf{y})ds_{\mathbf{y}}ds_{\mathbf{x}}+\frac{1}{4\pi}\int_{\Gamma_{1}}\int_{\Gamma_{2}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\phi_{1}^{2}(\mathbf{x})ds_{\mathbf{y}}ds_{\mathbf{x}}\\ =&\frac{1}{4\pi}\int_{-L/2}^{L/2}\int_{-L/2}^{L/2}\frac{\delta}{|x_{1}-y_{1}|^{2}+4\delta^{2}}dx_{1}\phi_{2}^{2}(\mathbf{y})dy_{1}+\frac{1}{4\pi}\int_{-L/2}^{L/2}\int_{-L/2}^{L/2}\frac{\delta}{|x_{1}-y_{1}|^{2}+4\delta^{2}}dy_{1}\phi_{1}^{2}(\mathbf{x})dx_{1}\\ =&\frac{1}{8\pi}\int_{-L/2}^{L/2}\Big{(}\arctan\frac{L/2-y_{1}}{2\delta}-\arctan\frac{-L/2-y_{1}}{2\delta}\Big{)}\phi_{2}^{2}(\mathbf{y})dy_{1}\\ &+\frac{1}{8\pi}\int_{-L/2}^{L/2}\Big{(}\arctan\frac{L/2-x_{1}}{2\delta}-\arctan\frac{-L/2-x_{1}}{2\delta}\Big{)}\phi_{1}^{2}(\mathbf{x})dx_{1}\\ \leq&C(\|\phi_{1}\|_{L^{2}(\Gamma_{1})}^{2}+\|\phi_{2}\|_{L^{2}(\Gamma_{2})}^{2}),\end{split}

where the constant CC is independent of δ\delta. By following a similar arguments as in [6] (pp. 18), one can show that 𝒜Γ2,Γ1\mathcal{A}_{\Gamma_{2},\Gamma_{1}} is a bounded operator L2(Γ2)L^{2}(\Gamma_{2}) to L2(Γ1)L^{2}(\Gamma_{1}). ∎

Lemma 2.3.

Suppose 𝐱Sc\mathbf{x}\in S^{c}, then for any function ϕL2(Sf)\phi\in L^{2}(S^{f}), which satisfies

ϕ(𝐲)=ϕ(𝐲+2δ𝐧),𝐲Γ1,\phi(\mathbf{y})=-\phi(\mathbf{y}+2\delta\mathbf{n}),\quad\mathbf{y}\in\Gamma_{1}, (2.22)

there holds

𝒦Sf(ιδ(P)ιδ(Q)),Sc[ϕ](𝐱)=o(1).\mathcal{K}_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q)),S^{c}}[\phi](\mathbf{x})=o(1). (2.23)
Proof.

Note that Sf=Γ1Γ2S^{f}=\Gamma_{1}\cup\Gamma_{2}. Straightforward computations show that

𝒦Sf(ιδ(P)ιδ(Q)),Sc[ϕ](𝐱)=12πSf(ιδ(P)ιδ(Q))𝐱𝐲,νx|𝐱𝐲|2ϕ(𝐲)𝑑s𝐲=12πSf(ιδ(P)ιδ(Q))𝐱𝐳yδνy,νx|𝐱𝐳yδνy|2ϕ(𝐲)𝑑s𝐲=12πSf(ιδ(P)ιδ(Q))𝐱𝐳y,νx|𝐱𝐳y|2ϕ(𝐲)𝑑s𝐲+o(1)=12πΓ1(ιδ(P)ιδ(Q))𝐱𝐳y,νx|𝐱𝐳y|2ϕ(𝐲)𝑑s𝐲+12πΓ1(ιδ(P)ιδ(Q))𝐱𝐳y,νx|𝐱𝐳y|2ϕ(𝐲+2δ𝐧)𝑑s𝐲+o(1)=o(1),\begin{split}&\mathcal{K}_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q)),S^{c}}[\phi](\mathbf{x})=\frac{1}{2\pi}\int_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =&\frac{1}{2\pi}\int_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}\frac{\langle\mathbf{x}-\mathbf{z}_{y}-\delta\nu_{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{z}_{y}-\delta\nu_{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ =&\frac{1}{2\pi}\int_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}\frac{\langle\mathbf{x}-\mathbf{z}_{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{z}_{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}+o(1)\\ =&\frac{1}{2\pi}\int_{\Gamma_{1}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}\frac{\langle\mathbf{x}-\mathbf{z}_{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{z}_{y}|^{2}}\phi(\mathbf{y})ds_{\mathbf{y}}\\ &+\frac{1}{2\pi}\int_{\Gamma_{1}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}\frac{\langle\mathbf{x}-\mathbf{z}_{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{z}_{y}|^{2}}\phi(\mathbf{y}+2\delta\mathbf{n})ds_{\mathbf{y}}+o(1)=o(1),\end{split}

which completes the proof. ∎

3. Quantitative analysis of the electric field

In this section, we present the quantitative analysis of the solution to the conductivity equation (1.1) as well as its geometric relationship to the inclusion DD.

3.1. Several auxiliary lemmas

Recall that uu is represented by (2.5). We first derive some asymptotic properties of the density function φ\varphi in (2.6). Let 𝐳xΓ0\mathbf{z}_{x}\in\Gamma_{0} be defined by

𝐳x={𝐱+δ𝐧,𝐱Γ1,𝐱δ𝐧,𝐱Γ2.\mathbf{z}_{x}=\left\{\begin{array}[]{ll}\mathbf{x}+\delta\mathbf{n},&\mathbf{x}\in\Gamma_{1},\\ \mathbf{x}-\delta\mathbf{n},&\mathbf{x}\in\Gamma_{2}.\end{array}\right. (3.1)

One has the following asymptotic expansion for HH around Γ0\Gamma_{0}:

H(𝐱)=H(𝐳x)+H(𝐳x)(𝐱𝐳x)+𝒪(|𝐱𝐳x|2)=H(𝐳x)+δH(𝐳x)(𝐱~𝐳x)+𝒪(δ2),H(\mathbf{x})=H(\mathbf{z}_{x})+\nabla H(\mathbf{z}_{x})\cdot(\mathbf{x}-\mathbf{z}_{x})+\mathcal{O}(|\mathbf{x}-\mathbf{z}_{x}|^{2})=H(\mathbf{z}_{x})+\delta\nabla H(\mathbf{z}_{x})\cdot(\tilde{\mathbf{x}}-\mathbf{z}_{x})+\mathcal{O}(\delta^{2}), (3.2)

for 𝐱Sf\mathbf{x}\in S^{f} and 𝐱~S1f\tilde{\mathbf{x}}\in S_{1}^{f}. Similarly, one has

H(𝐱)=H(P)+H(P)(𝐱P)+𝒪(|𝐱P|2)=H(P)+δH(P)νx+𝒪(δ2),H(\mathbf{x})=H(P)+\nabla H(P)\cdot(\mathbf{x}-P)+\mathcal{O}(|\mathbf{x}-P|^{2})=H(P)+\delta\nabla H(P)\cdot\nu_{x}+\mathcal{O}(\delta^{2}), (3.3)

for 𝐱Sa\mathbf{x}\in S^{a} and 𝐱~S1a\tilde{\mathbf{x}}\in S_{1}^{a}. Moreover,

H(𝐱)=H(Q)+H(P)(𝐱Q)+𝒪(|𝐱Q|2)=H(Q)+δH(Q)νx+𝒪(δ2),H(\mathbf{x})=H(Q)+\nabla H(P)\cdot(\mathbf{x}-Q)+\mathcal{O}(|\mathbf{x}-Q|^{2})=H(Q)+\delta\nabla H(Q)\cdot\nu_{x}+\mathcal{O}(\delta^{2}), (3.4)

for 𝐱Sb\mathbf{x}\in S^{b} and 𝐱~S1b\tilde{\mathbf{x}}\in S_{1}^{b}.

We now can show the following asymptotic result.

Lemma 3.1.

Suppose φ\varphi is defined in (2.6), then one has

φ(𝐱)={(λI+Aδ)1[(1)jx2H(,0)](x1)+δ(λIAδ)1[x22H(,0)](x1)+χ(ιδϵ(P)ιδϵ(Q))𝒪(δ2(1ϵ))+𝒪(δ2),𝐱Γj(ιδ(P)ιδ(Q)),(λI𝒦1)1[H(P)ν]+o(1),𝐱Saιδ(P),(λI𝒦2)1[H(Q)ν]+o(1),𝐱Sbιδ(Q),\varphi(\mathbf{x})=\left\{\begin{array}[]{l}(\lambda I+A_{\delta})^{-1}[(-1)^{j}\partial_{x_{2}}H(\cdot,0)](x_{1})+\delta(\lambda I-A_{\delta})^{-1}[\partial_{x_{2}}^{2}H(\cdot,0)](x_{1})\vskip 6.0pt plus 2.0pt minus 2.0pt\\ \quad+\chi(\iota_{\delta^{\epsilon}}(P)\cup\iota_{\delta^{\epsilon}}(Q))\mathcal{O}(\delta^{2(1-\epsilon)})+\mathcal{O}(\delta^{2}),\quad\mathbf{x}\in\Gamma_{j}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q)),\vskip 6.0pt plus 2.0pt minus 2.0pt\\ (\lambda I-\mathcal{K}_{1}^{*})^{-1}[\nabla H(P)\cdot\nu]+o(1),\quad\quad\quad\quad\quad\mathbf{x}\in S^{a}\cup\iota_{\delta}(P),\vskip 6.0pt plus 2.0pt minus 2.0pt\\ (\lambda I-\mathcal{K}_{2}^{*})^{-1}[\nabla H(Q)\cdot\nu]+o(1),\quad\quad\quad\quad\quad\mathbf{x}\in S^{b}\cup\iota_{\delta}(Q),\end{array}\right. (3.5)

where 0<ϵ<10<\epsilon<1 and the operator AδA_{\delta} is defined by

Aδ[ψ](x1):=12πL/2L/2δ(x1y1)2+4δ2ψ(y1)𝑑y1,ψL2(L/2,L/2).A_{\delta}[\psi](x_{1}):=\frac{1}{2\pi}\int_{-L/2}^{L/2}\frac{\delta}{(x_{1}-y_{1})^{2}+4\delta^{2}}\psi(y_{1})dy_{1},\quad\psi\in L^{2}(-L/2,L/2). (3.6)

The operators 𝒦1\mathcal{K}_{1}^{*} and 𝒦2\mathcal{K}_{2}^{*} are defined by

𝒦1[φ1](𝐱):=Saιδ(P)𝐱𝐲,νx|𝐱𝐲|2φ1(𝐲)𝑑s𝐲+χ(ιδ(P))𝒜Sfιδ(P)[φ1](𝐱)𝒦2[φ2](𝐱):=Sbιδ(Q)𝐱𝐲,νx|𝐱𝐲|2φ1(𝐲)𝑑s𝐲+χ(ιδ(Q))𝒜Sfιδ(Q)[φ2](𝐱),\begin{split}\mathcal{K}_{1}^{*}[\varphi_{1}](\mathbf{x}):=&\int_{S^{a}\cup\iota_{\delta}(P)}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\varphi_{1}(\mathbf{y})ds_{\mathbf{y}}+\chi(\iota_{\delta}(P))\mathcal{A}_{S^{f}\cap\iota_{\delta}(P)}[\varphi_{1}](\mathbf{x})\\ \mathcal{K}_{2}^{*}[\varphi_{2}](\mathbf{x}):=&\int_{S^{b}\cup\iota_{\delta}(Q)}\frac{\langle\mathbf{x}-\mathbf{y},\nu_{x}\rangle}{|\mathbf{x}-\mathbf{y}|^{2}}\varphi_{1}(\mathbf{y})ds_{\mathbf{y}}+\chi(\iota_{\delta}(Q))\mathcal{A}_{S^{f}\cap\iota_{\delta}(Q)}[\varphi_{2}](\mathbf{x}),\end{split} (3.7)

respectively.

Proof.

Since

(λI𝒦D)[φ]=[Hν|D].\Big{(}\lambda I-\mathcal{K}_{D}^{*}\Big{)}[\varphi]=\Big{[}\frac{\partial H}{\partial\nu}\Big{|}_{\partial D}\Big{]}.

By combining (2.11) and (3.2) one can readily verify that

(λI𝒦0)[φ](𝐱)=H(𝐳x)νx+o(1),𝐱Sf.\Big{(}\lambda I-\mathcal{K}_{0}\Big{)}[\varphi](\mathbf{x})=\nabla H(\mathbf{z}_{x})\cdot\nu_{x}+o(1),\quad\mathbf{x}\in S^{f}.

By using (2.12), one thus has

λφ(𝐱)1πΓ2δ|𝐱𝐲|2φ(𝐲)𝑑s𝐲=H(𝐳x)𝐧+o(1),𝐱Γ1(ιδ(P)ιδ(Q)),λφ(𝐱)1πΓ1δ|𝐱𝐲|2φ(𝐲)𝑑s𝐲=H(𝐳x)𝐧+o(1),𝐱Γ2(ιδ(P)ιδ(Q)).\begin{split}\lambda\varphi(\mathbf{x})-\frac{1}{\pi}\int_{\Gamma_{2}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\varphi(\mathbf{y})ds_{\mathbf{y}}=-\nabla H(\mathbf{z}_{x})\cdot\mathbf{n}+o(1),\quad&\mathbf{x}\in\Gamma_{1}\setminus\big{(}\iota_{\delta}(P)\cup\iota_{\delta}(Q)\big{)},\\ \lambda\varphi(\mathbf{x})-\frac{1}{\pi}\int_{\Gamma_{1}}\frac{\delta}{|\mathbf{x}-\mathbf{y}|^{2}}\varphi(\mathbf{y})ds_{\mathbf{y}}=\nabla H(\mathbf{z}_{x})\cdot\mathbf{n}+o(1),\quad&\mathbf{x}\in\Gamma_{2}\setminus\big{(}\iota_{\delta}(P)\cup\iota_{\delta}(Q)\big{)}.\end{split} (3.8)

By direct computations, one can show

λφ(x1,δ)1πL/2L/2δ(x1y1)2+4δ2φ(y1,δ)𝑑y1=x2H(x1,0)+o(1),|x1|L/2𝒪(δ),λφ(x1,δ)1πL/2L/2δ(x1y1)2+4δ2φ(y1,δ)𝑑y1=x2H(x1,0)+o(1),|x1|L/2𝒪(δ).\begin{split}&\lambda\varphi(x_{1},-\delta)-\frac{1}{\pi}\int_{-L/2}^{L/2}\frac{\delta}{(x_{1}-y_{1})^{2}+4\delta^{2}}\varphi(y_{1},\delta)dy_{1}\\ &\hskip 113.81102pt=-\partial_{x_{2}}H(x_{1},0)+o(1),\quad|x_{1}|\leq L/2-\mathcal{O}(\delta),\\ &\lambda\varphi(x_{1},\delta)-\frac{1}{\pi}\int_{-L/2}^{L/2}\frac{\delta}{(x_{1}-y_{1})^{2}+4\delta^{2}}\varphi(y_{1},-\delta)dy_{1}\\ &\hskip 113.81102pt=\partial_{x_{2}}H(x_{1},0)+o(1),\quad|x_{1}|\leq L/2-\mathcal{O}(\delta).\end{split} (3.9)

Thus one can derive that φ(x1,δ)=φ(x1,δ)+o(1)\varphi(x_{1},-\delta)=-\varphi(x_{1},\delta)+o(1), for |x1|L/2𝒪(δ)|x_{1}|\leq L/2-\mathcal{O}(\delta). Furthermore, for 𝐱S1a\mathbf{x}\in S_{1}^{a}, by making use of (2.12), (3.3) and Lemma 3.3, one has

(λI𝒦1)[φ](𝐱)=H(P)νx+o(1),inSaιδ(P).(\lambda I-\mathcal{K}_{1}^{*})[\varphi](\mathbf{x})=\nabla H(P)\cdot\nu_{x}+o(1),\quad\mbox{in}\quad S^{a}\cup\iota_{\delta}(P). (3.10)

In a similar manner, one can show that

(λI𝒦2)[φ](𝐱)=H(Q)νx+o(1),inSbιδ(Q).(\lambda I-\mathcal{K}_{2}^{*})[\varphi](\mathbf{x})=\nabla H(Q)\cdot\nu_{x}+o(1),\quad\mbox{in}\quad S^{b}\cup\iota_{\delta}(Q). (3.11)

and so the last equation in (3.5) follows.

Next, by combining (2.11), (2.12), and (2.13) again for 𝐱Γj(ιδ(P)ιδ(Q))\mathbf{x}\in\Gamma_{j}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q)), j=1,2j=1,2, and using the second and third equations in (3.5), one has

λφ(x1,(1)jδ)12πL/2L/2δ(x1y1)2+4δ2φ(y1,(1)j+1δ)𝑑y1=(1)jx2H(x1,0)+δx22H(x1,0)+χ(ιδϵ(P)ιδϵ(Q))𝒪(δ2(1ϵ))+𝒪(δ2),0<ϵ<1,\begin{split}&\lambda\varphi(x_{1},(-1)^{j}\delta)-\frac{1}{2\pi}\int_{-L/2}^{L/2}\frac{\delta}{(x_{1}-y_{1})^{2}+4\delta^{2}}\varphi(y_{1},(-1)^{j+1}\delta)dy_{1}\\ =&(-1)^{j}\partial_{x_{2}}H(x_{1},0)+\delta\partial_{x_{2}}^{2}H(x_{1},0)+\chi(\iota_{\delta^{\epsilon}}(P)\cup\iota_{\delta^{\epsilon}}(Q))\mathcal{O}(\delta^{2(1-\epsilon)})+\mathcal{O}(\delta^{2}),\quad 0<\epsilon<1,\end{split} (3.12)

which verifies the first equation in (3.5) and completes the proof. ∎

Before presenting our main result, we need to further analyze the operator AδA_{\delta} defined in (3.6)

Lemma 3.2.

Suppose AδA_{\delta} is defined in (3.6), then it holds that

Aδ[y1n](x1)=12x1n+o(1),𝐱Γj(ιδ(P)ιδ(Q)),n0.A_{\delta}[y_{1}^{n}](x_{1})=\frac{1}{2}x_{1}^{n}+o(1),\quad\mathbf{x}\in\Gamma_{j}\setminus\big{(}\iota_{\delta}(P)\cup\iota_{\delta}(Q)\big{)},\quad n\geq 0.\\ (3.13)
Proof.

We use deduction to prove the assertion. Since 𝐱Γj(ιδ(P)ιδ(Q))\mathbf{x}\in\Gamma_{j}\setminus\big{(}\iota_{\delta}(P)\cup\iota_{\delta}(Q)\big{)}, one has

|L/2x1|=𝒪(δϵ),and|L/2+x1|=𝒪(δϵ),0ϵ<1.|L/2-x_{1}|=\mathcal{O}(\delta^{\epsilon}),\quad\mbox{and}\quad|L/2+x_{1}|=\mathcal{O}(\delta^{\epsilon}),\quad 0\leq\epsilon<1.

Then for n=0n=0, it is straightforward to verify that

Aδ[1](x1)=1πL/2L/2δ(x1y1)2+4δ2𝑑y1=12π(arctanL/2x12δarctanL/2x12δ)=12+o(1).\begin{split}A_{\delta}[1](x_{1})=&\frac{1}{\pi}\int_{-L/2}^{L/2}\frac{\delta}{(x_{1}-y_{1})^{2}+4\delta^{2}}dy_{1}\\ =&\frac{1}{2\pi}\Big{(}\arctan\frac{L/2-x_{1}}{2\delta}-\arctan\frac{-L/2-x_{1}}{2\delta}\Big{)}=\frac{1}{2}+o(1).\end{split} (3.14)

Next, we suppose that (3.13) holds for nNn\leq N. Then by using change of variables, one can derive that

Aδ[y1N+1](x1)=1πL/2L/2δ(x1y1)2+4δ2y1Ny1𝑑y1=12πL/2x12δL/2x12δ11+t2y1N(2δt+x1)𝑑t=1πδL/2x12δL/2x12δ11+t2y1Nt𝑑t+x112x1N+o(1)=1πδ𝒪(ln(1+δ2(ϵ1)))+12x1N+1+o(1)=12x1N+1+o(1),\begin{split}A_{\delta}[y_{1}^{N+1}](x_{1})=&\frac{1}{\pi}\int_{-L/2}^{L/2}\frac{\delta}{(x_{1}-y_{1})^{2}+4\delta^{2}}y_{1}^{N}y_{1}dy_{1}\\ =&\frac{1}{2\pi}\int_{\frac{-L/2-x_{1}}{2\delta}}^{\frac{L/2-x_{1}}{2\delta}}\frac{1}{1+t^{2}}y_{1}^{N}(2\delta t+x_{1})dt\\ =&\frac{1}{\pi}\delta\int_{\frac{-L/2-x_{1}}{2\delta}}^{\frac{L/2-x_{1}}{2\delta}}\frac{1}{1+t^{2}}y_{1}^{N}tdt+x_{1}\frac{1}{2}x_{1}^{N}+o(1)\\ =&\frac{1}{\pi}\delta\mathcal{O}\left(\ln(1+\delta^{2(\epsilon-1)})\right)+\frac{1}{2}x_{1}^{N+1}+o(1)=\frac{1}{2}x_{1}^{N+1}+o(1),\end{split} (3.15)

which completes the proof. ∎

The following lemma is also of critical importance

Lemma 3.3.

There holds the following that

Saιδ(P)(λI𝒦1)1[H(P)ν]=2δ(λ12)1x1H(P)+o(δ),Sbιδ(Q)(λI𝒦2)1[H(Q)ν]=2δ(λ12)1x1H(Q)+o(δ).\begin{split}\int_{S^{a}\cup\iota_{\delta}(P)}(\lambda I-\mathcal{K}_{1}^{*})^{-1}[\nabla H(P)\cdot\nu]=&-2\delta\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\partial_{x_{1}}H(P)+o(\delta),\\ \int_{S^{b}\cup\iota_{\delta}(Q)}(\lambda I-\mathcal{K}_{2}^{*})^{-1}[\nabla H(Q)\cdot\nu]=&2\delta\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\partial_{x_{1}}H(Q)+o(\delta).\end{split} (3.16)
Proof.

For any fL2(D)f\in L^{2}(\partial D), we consider the following boundary integral equation

(λI𝒦D)[ϕ]=f.(\lambda I-\mathcal{K}_{D}^{*})[\phi]=f. (3.17)

By using the decomposition (2.12) (see also (3.10) and (3.11)), one has

χ(Saιδ(P))(λI𝒦1+o(1))[ϕ]+χ(Sbιδ(Q))(λI𝒦2+o(1))[ϕ]+𝒜Γ2,Γ1[ϕ]+𝒜Γ1,Γ2[ϕ]+χ(ιδϵ(P)ιδϵ(Q))𝒪(δ2(1ϵ))+𝒪(δ2)=f,0<ϵ<1.\begin{split}&\chi(S^{a}\cup\iota_{\delta}(P))(\lambda I-\mathcal{K}_{1}^{*}+o(1))[\phi]+\chi(S^{b}\cup\iota_{\delta}(Q))(\lambda I-\mathcal{K}_{2}^{*}+o(1))[\phi]\\ &+\mathcal{A}_{\Gamma_{2},\Gamma_{1}}[\phi]+\mathcal{A}_{\Gamma_{1},\Gamma_{2}}[\phi]+\chi(\iota_{\delta^{\epsilon}}(P)\cup\iota_{\delta^{\epsilon}}(Q))\mathcal{O}(\delta^{2(1-\epsilon)})+\mathcal{O}(\delta^{2})=f,\quad 0<\epsilon<1.\end{split} (3.18)

Note that D\partial D is of C1,αC^{1,\alpha}. By taking the boundary integral of both sides of (3.17) on D\partial D and making use of (3.18), one then has

(λ12)Dϕ=Saιδ(P)(λI𝒦1+o(1))[ϕ]+Sbιδ(Q)(λI𝒦2+o(1))[ϕ]+Γ1𝒜Γ2,Γ1[ϕ]+Γ2𝒜Γ1,Γ2[ϕ]+o(δ)=Df.\begin{split}\Big{(}\lambda-\frac{1}{2}\Big{)}\int_{\partial D}\phi=&\int_{S^{a}\cup\iota_{\delta}(P)}(\lambda I-\mathcal{K}_{1}^{*}+o(1))[\phi]+\int_{S^{b}\cup\iota_{\delta}(Q)}(\lambda I-\mathcal{K}_{2}^{*}+o(1))[\phi]\\ &+\int_{\Gamma_{1}}\mathcal{A}_{\Gamma_{2},\Gamma_{1}}[\phi]+\int_{\Gamma_{2}}\mathcal{A}_{\Gamma_{1},\Gamma_{2}}[\phi]+o(\delta)=\int_{\partial D}f.\end{split} (3.19)

By assuming f=χ(Saιδ(P))H(P)νf=\chi(S^{a}\cup\iota_{\delta}(P))\nabla H(P)\cdot\nu and plugging into (3.19), one thus has

(λ12)Saιδ(P)(λI𝒦1+o(1))1[H(P)ν]=Saιδ(P)H(P)ν=2δx1H(P),\Big{(}\lambda-\frac{1}{2}\Big{)}\int_{S^{a}\cup\iota_{\delta}(P)}(\lambda I-\mathcal{K}_{1}^{*}+o(1))^{-1}[\nabla H(P)\cdot\nu]=\int_{S^{a}\cup\iota_{\delta}(P)}\nabla H(P)\cdot\nu=-2\delta\partial_{x_{1}}H(P), (3.20)

which verifies the first equation in (3.16). Similarly, by assuming f=χ(Sbιδ(Q))H(q)νf=\chi(S^{b}\cup\iota_{\delta}(Q))\nabla H(q)\cdot\nu, one can prove the second equation in (3.16). The proof is complete. ∎

3.2. Sharp asymptotic approximation of the solution uu

With Lemmas 3.1, 3.2 and 3.3, we can now establish one of the main results of this paper as follows.

Theorem 3.1.

Let uu be the solution to (1.1) and (1.2), with DD of the rod-shape described in Section 1.1. Then for 𝐱2D¯\mathbf{x}\in\mathbb{R}^{2}\setminus\overline{D}, it holds that

u(𝐱)=H(𝐱)+δ12π(λ12)1L/2L/2ln(x1y1)2+x22(x1+L/2)2+x22y22H(y1,0)dy1+δ1π(λ12)1L/2L/2x2(x1y1)2+x22y2H(y1,0)dy1+δ12π(λ12)1ln(x1L/2)2+x22(x1+L/2)2+x22x1H(L/2,0)+o(δ).\begin{split}u(\mathbf{x})=&H(\mathbf{x})+\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\int_{-L/2}^{L/2}\ln\frac{(x_{1}-y_{1})^{2}+x_{2}^{2}}{(x_{1}+L/2)^{2}+x_{2}^{2}}\partial_{y_{2}}^{2}H(y_{1},0)\,dy_{1}\\ &\quad\quad\,+\delta\frac{1}{\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\int_{-L/2}^{L/2}\frac{x_{2}}{(x_{1}-y_{1})^{2}+x_{2}^{2}}\partial_{y_{2}}H(y_{1},0)\,dy_{1}\\ &\quad\quad\,+\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\ln\frac{(x_{1}-L/2)^{2}+x_{2}^{2}}{(x_{1}+L/2)^{2}+x_{2}^{2}}\partial_{x_{1}}H(L/2,0)+o(\delta).\end{split} (3.21)
Proof.

By using (2.5) and Taylor’s expansion along with Γ0\Gamma_{0}, one has

u(𝐱)=H(𝐱)+Sf(ιδ(P)ιδ(Q))G(𝐱𝐳y)φ(𝐲)𝑑s𝐲+δSf(ιδ(P)ιδ(Q))𝐲G(𝐱𝐳y)ν𝐲φ(𝐲)𝑑s𝐲+Saιδ(P)G(𝐱𝐳y)φ(𝐲)𝑑s𝐲~+Sbιδ(Q)G(𝐱𝐳y)φ~(𝐲~)𝑑s𝐲~+𝒪(δ2).\begin{split}u(\mathbf{x})=&H(\mathbf{x})+\int_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}G(\mathbf{x}-\mathbf{z}_{y})\varphi(\mathbf{y})\,ds_{\mathbf{y}}\\ &\quad\quad\ \,+\delta\int_{S^{f}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}\nabla_{\mathbf{y}}G(\mathbf{x}-\mathbf{z}_{y})\cdot\nu_{\mathbf{y}}\varphi(\mathbf{y})\,ds_{\mathbf{y}}\\ &\quad\quad\ \,+\int_{S^{a}\cup\iota_{\delta}(P)}G(\mathbf{x}-\mathbf{z}_{y})\varphi(\mathbf{y})\,ds_{\tilde{\mathbf{y}}}+\int_{S^{b}\cup\iota_{\delta}(Q)}G(\mathbf{x}-\mathbf{z}_{y})\tilde{\varphi}(\tilde{\mathbf{y}})\,ds_{\tilde{\mathbf{y}}}+\mathcal{O}(\delta^{2}).\end{split} (3.22)

First, by using (3.12), one can derive that

SfG(𝐱𝐳y)φ(𝐲)𝑑s𝐲=2δΓ1(ιδ(P)ιδ(Q))G(𝐱𝐳y)(λIAδ)1[x22H(,0)](y1)𝑑y1+o(δ)=δ12π(λ12)1L/2L/2ln((x1y1)2+x22)y22H(y1,0)dy1+o(δ).\begin{split}&\int_{S^{f}}G(\mathbf{x}-\mathbf{z}_{y})\varphi(\mathbf{y})\,ds_{\mathbf{y}}\\ =&2\delta\int_{\Gamma_{1}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}G(\mathbf{x}-\mathbf{z}_{y})(\lambda I-A_{\delta})^{-1}[\partial_{x_{2}}^{2}H(\cdot,0)](y_{1})dy_{1}+o(\delta)\\ =&\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\int_{-L/2}^{L/2}\ln((x_{1}-y_{1})^{2}+x_{2}^{2})\partial_{y_{2}}^{2}H(y_{1},0)\,dy_{1}+o(\delta).\end{split} (3.23)

Similarly, one has

Sf𝐲G(𝐱𝐳y)ν𝐲φ(𝐲)𝑑s𝐲=1π(λ12)1L/2L/2x2(x1y1)2+x22y2H(y1,0)dy1+o(1).\begin{split}&\int_{S^{f}}\nabla_{\mathbf{y}}G(\mathbf{x}-\mathbf{z}_{y})\cdot\nu_{\mathbf{y}}\varphi(\mathbf{y})\,ds_{\mathbf{y}}\\ =&\frac{1}{\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\int_{-L/2}^{L/2}\frac{x_{2}}{(x_{1}-y_{1})^{2}+x_{2}^{2}}\partial_{y_{2}}H(y_{1},0)\,dy_{1}+o(1).\end{split} (3.24)

By using Lemma 3.3, one then obtains that

Sbιδ(Q)G(𝐱𝐳y)φ(𝐲)𝑑s𝐲=Sbιδ(Q)G(𝐱Q)φ(𝐲)𝑑s𝐲+o(δ)=δ1πln|𝐱Q|(λ12)1x1H(Q)+o(δ)=δ12π(λ12)1ln((x1L/2)2+x22)x1H(L/2,0)+o(δ).\begin{split}&\int_{S^{b}\cup\iota_{\delta}(Q)}G(\mathbf{x}-\mathbf{z}_{y})\varphi(\mathbf{y})\,ds_{\mathbf{y}}=\int_{S^{b}\cup\iota_{\delta}(Q)}G(\mathbf{x}-Q)\varphi(\mathbf{y})\,ds_{\mathbf{y}}+o(\delta)\\ =&\delta\frac{1}{\pi}\ln|\mathbf{x}-Q|\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\partial_{x_{1}}H(Q)+o(\delta)\\ =&\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\ln((x_{1}-L/2)^{2}+x_{2}^{2})\partial_{x_{1}}H(L/2,0)+o(\delta).\end{split} (3.25)

Noting that

Dφ(𝐲)𝑑s𝐲=D(λI𝒦D)1[Hν](𝐲)𝑑s𝐲=0,\int_{\partial D}\varphi(\mathbf{y})ds_{\mathbf{y}}=\int_{\partial D}\Big{(}\lambda I-\mathcal{K}_{D}^{*}\Big{)}^{-1}\Big{[}\frac{\partial H}{\partial\nu}\Big{]}(\mathbf{y})ds_{\mathbf{y}}=0,

and by combining (3.23), one can readily show that

Saιδ(P)G(𝐱𝐳y)φ(𝐲)𝑑s𝐲=Saιδ(P)G(𝐱P)φ(𝐲)𝑑s𝐲+o(δ)=12πln|𝐱P|(Sbιδ(Q)φ(𝐲)ds𝐲+2Γ1(ιδ(P)ιδ(Q))(λIAδ)1[x22H(,0)](y1)dy1)+o(δ)=δ12π(λ12)1ln((x1+L/2)2+x22)x1H(L/2,0)δ12π(λ12)1ln((x1+L/2)2+x22)L/2L/2y22H(y1,0)dy1+o(δ).\begin{split}&\int_{S^{a}\cup\iota_{\delta}(P)}G(\mathbf{x}-\mathbf{z}_{y})\varphi(\mathbf{y})\,ds_{\mathbf{y}}=\int_{S^{a}\cup\iota_{\delta}(P)}G(\mathbf{x}-P)\varphi(\mathbf{y})\,ds_{\mathbf{y}}+o(\delta)\\ =&-\frac{1}{2\pi}\ln|\mathbf{x}-P|\Big{(}\int_{S^{b}\cup\iota_{\delta}(Q)}\varphi(\mathbf{y})\,ds_{\mathbf{y}}\\ &\quad\quad\quad\quad+2\int_{\Gamma_{1}\setminus(\iota_{\delta}(P)\cup\iota_{\delta}(Q))}(\lambda I-A_{\delta})^{-1}[\partial_{x_{2}}^{2}H(\cdot,0)](y_{1})dy_{1}\Big{)}+o(\delta)\\ =&-\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\ln((x_{1}+L/2)^{2}+x_{2}^{2})\partial_{x_{1}}H(L/2,0)\\ &-\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\ln((x_{1}+L/2)^{2}+x_{2}^{2})\int_{-L/2}^{L/2}\partial_{y_{2}}^{2}H(y_{1},0)\,dy_{1}+o(\delta).\end{split} (3.26)

Finally, by substituting (3.23)-(3.26) into (3.22) one has (3.21), which completes the proof. ∎

We finally can derive the sharp asymptotic expansion of the electric field as follows.

Theorem 3.2.

Suppose H(𝐱)=𝐚𝐱H(\mathbf{x})=\mathbf{a}\cdot\mathbf{x}, where 𝐚=(a1,a2)2\mathbf{a}=(a_{1},a_{2})\in\mathbb{R}^{2}. Then for 𝐱2D¯\mathbf{x}\in\mathbb{R}^{2}\setminus\overline{D}, the electric field uu satisfies

u(𝐱)=𝐚𝐱+δ1π(λ12)1a2(arctan(L/2x1x2)+arctan(L/2+x1x2))+δ12π(λ12)1a1ln(x1L/2)2+x22(x1+L/2)2+x22+o(δ).\begin{split}u(\mathbf{x})=&\mathbf{a}\cdot\mathbf{x}+\delta\frac{1}{\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}a_{2}\left(\arctan\Big{(}\frac{L/2-x_{1}}{x_{2}}\Big{)}+\arctan\Big{(}\frac{L/2+x_{1}}{x_{2}}\Big{)}\right)\\ &+\delta\frac{1}{2\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}a_{1}\ln\frac{(x_{1}-L/2)^{2}+x_{2}^{2}}{(x_{1}+L/2)^{2}+x_{2}^{2}}+o(\delta).\end{split} (3.27)

Furthermore, the perturbed gradient field admits the following asymptotic expansion:

u(𝐱)=𝐚+δ1π(λ12)1(f2(𝐱)a1f1(𝐱)a2f1(𝐱)a1+f2(𝐱)a2)+o(δ),\nabla u(\mathbf{x})=\mathbf{a}+\delta\frac{1}{\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\left(\begin{array}[]{c}f_{2}(\mathbf{x})a_{1}-f_{1}(\mathbf{x})a_{2}\\ f_{1}(\mathbf{x})a_{1}+f_{2}(\mathbf{x})a_{2}\end{array}\right)+o(\delta), (3.28)

where the functions fjf_{j}, j=1,2j=1,2 are defined by

f1(𝐱):=x2(x1L/2)2+x22x2(x1+L/2)2+x22f2(𝐱):=x1L/2(x1L/2)2+x22x1+L/2(x1+L/2)2+x22.\begin{split}&f_{1}(\mathbf{x}):=\frac{x_{2}}{(x_{1}-L/2)^{2}+x_{2}^{2}}-\frac{x_{2}}{(x_{1}+L/2)^{2}+x_{2}^{2}}\\ &f_{2}(\mathbf{x}):=\frac{x_{1}-L/2}{(x_{1}-L/2)^{2}+x_{2}^{2}}-\frac{x_{1}+L/2}{(x_{1}+L/2)^{2}+x_{2}^{2}}.\end{split} (3.29)
Proof.

The proof is given by using (3.21) together with direct computations. ∎

3.3. Quantitative analysis and numerical illustrations

Define the following vector field

𝐄s:=δ1π(λ12)1(f2(𝐱)a1f1(𝐱)a2f1(𝐱)a1+f2(𝐱)a2).\mathbf{E}^{s}:=\delta\frac{1}{\pi}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-1}\left(\begin{array}[]{c}f_{2}(\mathbf{x})a_{1}-f_{1}(\mathbf{x})a_{2}\\ f_{1}(\mathbf{x})a_{1}+f_{2}(\mathbf{x})a_{2}\end{array}\right). (3.30)

According to (3.28), 𝐄s\mathbf{E}^{s} is the leading order term of the perturbed gradient field. It is noted that the distribution of |𝐄s||\mathbf{E}^{s}| is independent of the uniform gradient potential 𝐚\mathbf{a}. In fact, one has

|𝐄s|2=δ21π2(λ12)2(a12+a22)(f1(𝐱)2+f2(𝐱)2).|\mathbf{E}^{s}|^{2}=\delta^{2}\frac{1}{\pi^{2}}\Big{(}\lambda-\frac{1}{2}\Big{)}^{-2}(a_{1}^{2}+a_{2}^{2})(f_{1}(\mathbf{x})^{2}+f_{2}(\mathbf{x})^{2}). (3.31)

Moreover, further computations show that

f1(𝐱)2+f2(𝐱)2=(1|𝐱Q|1|𝐱P|)2+2|𝐱P||𝐱Q|(1𝐱P,𝐱Q|𝐱P||𝐱Q|).f_{1}(\mathbf{x})^{2}+f_{2}(\mathbf{x})^{2}=\left(\frac{1}{|\mathbf{x}-Q|}-\frac{1}{|\mathbf{x}-P|}\right)^{2}+\frac{2}{|\mathbf{x}-P||\mathbf{x}-Q|}\left(1-\frac{\langle\mathbf{x}-P,\mathbf{x}-Q\rangle}{|\mathbf{x}-P||\mathbf{x}-Q|}\right). (3.32)

One can thus derive that |𝐄s||\mathbf{E}^{s}| is maximized near the two caps (high curvature parts) of the inclusion DD. In fact, near the caps one has

|𝐱P|=δ+o(δ),or|𝐱Q|=δ+o(δ).|\mathbf{x}-P|=\delta+o(\delta),\quad\mbox{or}\quad|\mathbf{x}-Q|=\delta+o(\delta).

By (3.32) one then has

f1(𝐱)2+f2(𝐱)2=δ2(1+o(1)),f_{1}(\mathbf{x})^{2}+f_{2}(\mathbf{x})^{2}=\delta^{-2}(1+o(1)), (3.33)

while near the centering parts of the rod,

f1(𝐱)2+f2(𝐱)2=𝒪(1).f_{1}(\mathbf{x})^{2}+f_{2}(\mathbf{x})^{2}=\mathcal{O}(1).

To better illustrate the result, we next present some numerical solutions with different background fields. The parameters of the rod-shape inclusion are selected as follows:

σ0=2,L=10,δ=5tan(π/36)0.4374.\sigma_{0}=2,\quad L=10,\quad\delta=5*\tan(\pi/36)\approx 0.4374. (3.34)

We choose three different uniform background fields, i.e., 𝐚=(1,0),(0,1),(1,1)\mathbf{a}=(1,0),(0,1),(1,1), respectively, and plot the absolute values of the perturbed fields as well as the corresponding gradient fields, which are scaled for better display. It is clearly shown from Figure 1 to Figure 3 that the gradient fields behave much stronger near the high curvature parts of the inclusion.

Refer to caption
Refer to caption
Figure 1. 𝐚=(1,0)\mathbf{a}=(1,0). Left: Perturbed field |u𝐚𝐱||u-\mathbf{a}\cdot\mathbf{x}| (scaled) Right: Perturbed gradient field |u𝐚||\nabla u-\mathbf{a}| (scaled).
Refer to caption
Refer to caption
Figure 2. 𝐚=(0,1)\mathbf{a}=(0,1). Left: Perturbed field |u𝐚𝐱||u-\mathbf{a}\cdot\mathbf{x}| (scaled) Right: Perturbed gradient field |u𝐚||\nabla u-\mathbf{a}| (scaled).
Refer to caption
Refer to caption
Figure 3. 𝐚=(1,1)\mathbf{a}=(1,1). Left: Perturbed field |u𝐚𝐱||u-\mathbf{a}\cdot\mathbf{x}| (scaled) Right: Perturbed gradient field |u𝐚||\nabla u-\mathbf{a}| (scaled).

4. Application to Calderón inverse inclusion problem

In this section, we consider the application of the quantitative results derived in the previous section to the Calderón inverse inclusion problem. To that end, we let DD denote a generic rod inclusion that is obtained through rigid motions performed on special case described in Section 1.1. We write D(L,δ,𝐳0,σ0)D(L,\delta,\mathbf{z}_{0},\sigma_{0}) to signify its dependence on the length LL, width δ\delta, position 𝐳0\mathbf{z}_{0} (which is the geometric centre of DD) as well as the conductivity parameter σ0\sigma_{0}. Consider the conductivity system (1.1) associated with a generic inclusion described above. The inverse inclusion problem is concerned with recovering the shape of the inclusion, namely D\partial D, independent of its content σ0\sigma_{0}, by measuring the perturbed electric field (uH)(u-H) away from the inclusion. This is one of the central problems in EIT, which forms the fundamental basis for the electric prospecting. The case with a single measurement, namely the use of a single probing field HH, is a longstanding problem in the literature. The existing results for the single-measurement case are mainly concerned with specific shapes including discs/balls and polygons/polyhedrons [12, 13, 18, 25, 33, 38, 42] as well as the other general shapes but with a-priori conditions; see [1, 2, 3, 16, 27, 30]. As discussed earlier, in [39], the local recovery of the highly-curved part of D\partial D was also considered. Next, using the asymptotic result quantitative result in Theorem 3.2, we shall show that one can uniquely determine a conductive inclusion up to an error level δ1\delta\ll 1.

Theorem 4.1.

Let Dj=Dj(Lj,δj,𝐳0(j),σ0(j))D_{j}=D_{j}(L_{j},\delta_{j},\mathbf{z}_{0}^{(j)},\sigma_{0}^{(j)}), j=1,2j=1,2, be two conductive rods such that Lj1,δjδ1L_{j}\sim 1,\delta_{j}\sim\delta\ll 1 and σ0(j)1\sigma_{0}^{(j)}\sim 1 for j=1,2j=1,2. Let uju_{j} be the corresponding solution to (1.1) associated with DjD_{j} and a given nontrivial H(𝐱)=𝐚𝐱H(\mathbf{x})=\mathbf{a}\cdot\mathbf{x}. Suppose that

u1=u2onΣ,u_{1}=u_{2}\quad\mbox{on}\ \ \partial\Sigma, (4.1)

where Σ\Sigma is a bounded simply-connected Lipschitz domain enclosing DjD_{j}. Then it cannot hold that

dist(D1,D2)δ.\mathrm{dist}(D_{1},D_{2})\gg\delta. (4.2)
Proof.

First, by (4.1), we know that u1=u2u_{1}=u_{2} in 2\Σ\mathbb{R}^{2}\backslash\Sigma and hence by unique continuation, we also know that u1=u2u_{1}=u_{2} in 2\(D1D2)\mathbb{R}^{2}\backslash(D_{1}\cup D_{2}). Next, since the Laplacian is invariant under rigid motions, we note that the quantitative result in Theorem 3.2 still holds for DjD_{j}. By contradiction, we assume that (4.2) holds. It is easily seen that there must be one cap point, say Θ0D1\Theta_{0}\in\partial D_{1}, which lies away from D2D_{2} and dist(Θ0,D2)δ\mathrm{dist}(\Theta_{0},D_{2})\gg\delta. Hence, one has u1(Θ0)=u2(Θ0)u_{1}(\Theta_{0})=u_{2}(\Theta_{0}). Now, we arrive at a contradiction by noting that using Theorem 3.2, one has u1(Θ0)1u_{1}(\Theta_{0})\sim 1 whereas u2(Θ0)δ1u_{2}(\Theta_{0})\sim\delta\ll 1.

The proof is complete.

Acknowledgments

The work of X. Fang was supported by Humanities and Social Sciences Foundation of the Ministry of Education no. 20YJC910005, Major Project for National Natural Science Foundation of China no. 71991465, PSCF of Hunan No. 18YBQ077 and RFEB of Hunan No. 18B337. The work of Y. Deng was supported by NSF grant of China No. 11971487 and NSF grant of Hunan No. 2020JJ2038. The work of H. Liu was supported by a startup fund from City University of Hong Kong and the Hong Kong RGC General Research Funds, 12301218, 12302919 and 12301420.

References

  • [1] G. Alessandrini, Generic uniqueness and size estimates in the inverse conductivity problem with one measurement, Matematiche (Catania), 54 (1999), suppl., 5–14.
  • [2] G. Alessandrini and V. Isakov, Analyticity and uniqueness for the inverse conductivity problem, Rend. Istit. Mat. Univ. Trieste, 28 (1996), no. 1-2, 351–369.
  • [3] G. Alessandrini, V. Isakov and J. Powell, Local uniqueness in the inverse conductivity problem with one measurement, Trans. Amer. Math. Soc., 347 (1995), no. 8, 3031–3041.
  • [4] H. Ammari, J. Garnier, H. Kang, W.K. Park, and K. Solna, Imaging schemes for perfectly conducting cracks, SIAM J. Appl. Math., 71 (2011), 68–91.
  • [5] H. Ammari and H. Kang, Reconstruction of Small Inhomogeneities from Boundary Measurements, Lecture Notes in Mathematics, 1846, Springer-Verlag, Berlin, 2004.
  • [6] H. Ammari, H. Kang, Polarization and Moment Tensors With Applications to Inverse Problems and Effective Medium Theory, Applied Mathematical Sciences, Springer-Verlag, Berlin Heidelberg, 2007.
  • [7] H. Ammari, G. Ciraolo, H. Kang, H. Lee, and G. Milton, Spectral analysis of a Neumann-Poincar -type operator and analysis of cloaking due to anomalous localized resonance, Arch. Ration. Mech. Anal., 208 (2013), 667–692.
  • [8] H. Ammari, G. Ciraolo, H. Kang, H. Lee, and G. Milton, Spectral analysis of a Neumann-Poincaré-type operator and analysis of cloaking due to anomalous localized resonance II, Contemporary Mathematics, 615 (2014), 1–14.
  • [9] H. Ammari, J.K. Seo, T. Zhang, and L. Zhou, Electrical impedance spectroscopy-based nondestructive testing for imaging defects in concrete structures, Sensors, 15 (2015), 10909–10922.
  • [10] K. Ando and H. Kang, Analysis of plasmon resonance on smooth domains using spectral properties of the Neumann-Poincare operator, J,. Math. Anal. Appl., 435 (2016), 162–178.
  • [11] K. Ando, H. Kang and H. Liu, Plasmon resonance with finite frequencies: a validation of the quasi-static approximation for diametrically small inclusions, SIAM J. Appl. Math., 76 (2016), 731–749.
  • [12] B. Barceló, E. Fabes and J.-K. Seo, The inverse conductivity problem with one measurement: uniqueness for convex polyhedra, Proc. AMS, 122 (1994) 183–189.
  • [13] E. Beretta, E. Francini and S. Vessella, Global Lipschitz stability estimates for polygonal conductivity inclusions from boundary measurements, Appl. Anal., https://doi.org/10.1080/00036811.2020.1775819
  • [14] E. Beretta, E. Francini and M. S. Vogelius, Asymptotic formulas for steady state voltage potentials in the presence of thin inhomogeneities. A rigorous error analysis, J. Math. Pures Appl., 82 (2003), 1277–1301.
  • [15] L. Borcea, Electrical impedance tomography, Inverse Problems, 18 (2002), no. 6, R99–R136.
  • [16] K. Bryan, Single measurement detection of a discontinuous conductivity, Comm. Partial Differential Equations, 15 (1990), 503–514.
  • [17] X. Cao, H. Diao and H. Liu, Determining a piecewise conductive medium body by a single far-field measurement, arXiv:2005.04420
  • [18] D. S. Choi, J. Helsing and M. Lim, Corner effects on the perturbation of an electric potential, SIAM J. Appl. Math., 78 (2018), no. 3, 1577–1601.
  • [19] Y. Deng, H. Li, H. Liu, On spectral properties of Neuman-Poincaré operator on spheres and plasmonic resonances in 3D elastostatics, J. Spectr. Theory, 140 (2020), 213–242.
  • [20] Y. Deng, H. Li, H. Liu, Analysis of surface polariton resonance for nanoparticles in elastic system, SIAM J. Math. Anal., 52 (2020), 1786–1805.
  • [21] Y. Deng, H. Liu and G. Uhlmann, On regularized full- and partial-cloaks in acoustic scattering, Comm. Partial Differential Equations, 42 (6) (2017), 821-851.
  • [22] Y. Deng, H. Liu and G. Uhlmann, Full and partial cloaking in electromagnetic scattering, Arch. Ration. Mech. Anal., 223 (1)(2017), 265-299.
  • [23] Y. Deng, H. Liu, G. Zheng, Mathematical analysis of plasmon resonances for curved nanorods, arXiv:2007.11181
  • [24] E. Dibenedetto, C. M. Elliott and A. Friedman, The free boundary of a flow in a porous body heated from its boundary, Nonlinear Anal: TMA, 10, (1986), no. 9, 879–900.
  • [25] E. Fabes, H. Kang and J.-K. Seo, Inverse conductivity problem with one measurement: Error estimates and approximate identification for perturbed disks, SIAM J. Math. Anal., 30 (1999), 699–720.
  • [26] X. Fang, Y. Deng and J. Li, Plasmon resonance and heat generation in nanostructures, Math. Meth. Appl. Sci, 38 (2015), 4663–4672.
  • [27] A. Friedman and V. Isakov, On the uniqueness in the inverse conductivity problem with one measurement, Indiana Univ. Math. J., 38 (1989), 563–579.
  • [28] P. Grisvard, Boundary Value Problems in Non-Smooth Domains, Pitman, London 1985.
  • [29] J. Helsing, H. Kang and M. Lim, Classification of spectral of the Neumann–Poincaré operator on planar domains with corners by resonance, Ann. I. H. Poincare-AN, 34 (2017), 991–1011.
  • [30] V. Isakov and J. Powell, On the inverse conductivity problem with one measurement, Inverse Problems, 6 (1990), 311–318.
  • [31] H. Kang, M. Lim and S. Yu, Spectral resolution of the Neumann-Poincaré operator on intersecting disks and analysis of plasmon resonance, Arch. Ratio. Mech. Anal., 226 (2017), 83–115.
  • [32] H. Kang, K. Kim, H. Lee, J. Shin and S. Yu, Spectral properties of the Neumann-Poincaré operator and uniformity of estimates for the conductivity equation with complex coefficients, J London Math Soc (2), 93 (2016), 519–546.
  • [33] H. Kang and J.-K. Seo, Inverse conductivity problem with one measurement: uniqueness of balls in 3\mathbb{R}^{3}, SIAM J. Appl. Math., 59 (1999), 1533–1539.
  • [34] A. Khelifi and H. Zribi, Asymptotic expansions for the voltage potentials with two-dimensional and three-dimensional thin interfaces, Math. Meth. Appl. Sci., 34 (2011), 2274–2290.
  • [35] H. Li and H. Liu, On anomalous localized resonance and plasmonic cloaking beyond the quasistatic limit, Proceedings of the Royal Society A, 474: 20180165
  • [36] J. Li, H. Liu, L. Rondi and G. Uhlmann, Regularized transformation-optics cloaking for the Helmholtz equation:from partial cloak to full cloak, Comm. Math. Phys., 335 (2015), 671–712.
  • [37] E. Blåsten and H. Liu, Recovering piecewise-constant refractive indices by a single far-field pattern, Inverse Problems, 36 (2020), 085005.
  • [38] H. Liu and C.-H. Tsou, Stable determination of polygonal inclusions in Calderón’s problem by a single partial boundary measurement, Inverse Problems, 36 (2020), 085010.
  • [39] H. Liu, C.-H. Tsou and W. Yang, On Calderón’s inverse inclusion problem with smooth shapes by a single partial boundary measurement, arXiv:2006.10586.
  • [40] J. C. Nédélec, Acoustic and Electromagnetic Equations: Integral Representations for Harmonic Problems, Springer-Verlag, New York, 2001.
  • [41] L. Rondi, Optimal stability of reconstruction of plane Lipschitz cracks, SIAM J. Math. Anal., 36 (2005), 1282–1292.
  • [42] F. Triki and C.-H. Tsou, Inverse inclusion problem: a stable method to determine disks, J. Differential Equations, 269 (2020), 3259–3281.
  • [43] G. Uhlmann, Electrical impedance tomography and Calderón’s problem, Inverse Problems, 25 (2009), no. 12, 123011, 39 pp.