0810.4631/LY10.tex
1: \documentclass[10pt,a4paper]{article}
2: \usepackage{amsmath, amssymb}
3: \usepackage{latexsym, color}
4: \usepackage{epsfig}
5: 
6: \setlength{\textwidth}{140mm} \setlength{\textheight}{200mm}
7: \setlength{\oddsidemargin}{11mm} \setlength{\evensidemargin}{11mm}
8: 
9: 
10: 
11: 
12: \begin{document}
13: 
14: \newtheorem{thm}{Theorem}[section]
15: \newtheorem{cor}[thm]{Corollary}
16: \newtheorem{lem}[thm]{Lemma}
17: \newtheorem{prop}[thm]{Proposition}
18: \newtheorem{defn}[thm]{Definition}
19: \newtheorem{rem}[thm]{Remark}
20: \newtheorem{Ex}[thm]{EXAMPLE}
21: \def\nm{\noalign{\medskip}}
22: 
23: \bibliographystyle{plain}
24: 
25: %\numberwithin{equation}{section}
26: 
27: 
28: \newcommand{\qed}{\hfill \ensuremath{\square}}
29: \newcommand{\ds}{\displaystyle}
30: \newcommand{\pf}{\medskip \noindent {\sl Proof}. ~ }
31: \newcommand{\p}{\partial}
32: \renewcommand{\a}{\alpha}
33: \newcommand{\z}{\zeta}
34: \newcommand{\pd}[2]{\frac {\p #1}{\p #2}}
35: \newcommand{\norm}[1]{\| #1 \|}
36: \newcommand{\dbar}{\overline \p}
37: \newcommand{\eqnref}[1]{(\ref {#1})}
38: \newcommand{\na}{\nabla}
39: \newcommand{\Om}{\Omega}
40: \newcommand{\ep}{\epsilon}
41: \newcommand{\tmu}{\widetilde \mu}
42: \newcommand{\vep}{\varepsilon}
43: \newcommand{\tlambda}{\widetilde \lambda}
44: \newcommand{\tnu}{\widetilde \nu}
45: \newcommand{\vp}{\varphi}
46: \newcommand{\RR}{\mathbb{R}}
47: \newcommand{\CC}{\mathbb{C}}
48: \newcommand{\NN}{\mathbb{N}}
49: \renewcommand{\div}{\mbox{div}~}
50: \newcommand{\bu}{{\bf u}}
51: \newcommand{\la}{\langle}
52: \newcommand{\ra}{\rangle}
53: \newcommand{\Scal}{\mathcal{S}}
54: \newcommand{\Lcal}{\mathcal{L}}
55: \newcommand{\Kcal}{\mathcal{K}}
56: \newcommand{\Dcal}{\mathcal{D}}
57: \newcommand{\tScal}{\widetilde{\mathcal{S}}}
58: \newcommand{\tKcal}{\widetilde{\mathcal{K}}}
59: \newcommand{\Pcal}{\mathcal{P}}
60: \newcommand{\Qcal}{\mathcal{Q}}
61: \newcommand{\id}{\mbox{Id}}
62: %%%%%%%%%%
63: \newcommand{\be}{\begin{equation}}
64: \newcommand{\ee}{\end{equation}}
65: 
66: 
67: 
68: 
69: 
70: 
71: \title{Strong Influence of a Small Fiber on Shear Stress in Fiber-Reinforced Composites}
72: 
73: \author{Mikyoung Lim\thanks{Department of Mathematics, Colorado State
74: University, Fort Collins, CO 80523, USA (lim@math.colostate.edu)}
75: \and KiHyun Yun\thanks{\footnotesize Department of Mathematics,
76: Michigan State University, East Lansing, MI 48824,
77: USA (kyun@math.msu.edu)}}
78:  \maketitle
79: 
80: \begin{abstract}
81: In stiff fiber-reinforced material, the high shear stress
82: concentration occurs in the narrow region between fibers. With the
83: addition of a small geometric change in cross-section, such as a
84: thin fiber or a overhanging part of fiber, the concentration is
85: significantly increased. This paper presents mathematical analysis
86: to explain the rapidly increased growth of the stress by a small
87: particle in cross-section. To do so, we consider two crucial cases
88: where a thin fiber exists between a pair of fibers, and where one of
89: two fibers has a protruding small lump in cross-section. For each
90: case, the optimal lower and upper bounds on the stress associated
91: with the geometrical factors of fibers is established to explain the
92: strongly increased growth of the stress by a small particle.
93: \end{abstract}
94: 
95: 
96: MSC-class:  35J25, 73C40
97: 
98: 
99: 
100: 
101: 
102: 
103: \section{Introduction}
104: In this paper, we concern ourselves with the high stress
105: concentration occurring in the stiff fiber-reinforced composites
106: when fibers are located closely. The primary investigation focuses
107: on the case when a smaller fiber is located in-between area of two
108: fibers, see Figure \ref{caseAB} and Figure \ref{caseCD}. This paper
109: reveals that, with the addition of a smaller fiber, the growth of
110: stress is significantly increased: if the diameter $d$ of the fiber
111: in the middle is sufficiently small and the distance between
112: adjoining fibers is $\epsilon$, then the stress blows up at the rate
113: of $\frac 1 {\sqrt{d \epsilon}}$ in the narrow region, even though
114: the blow-up rate has been known as $\frac 1 {\sqrt {\epsilon}}$ as
115: in the case of a pair of fibers. This means that the defect of fiber
116: as a protrusion causes much lower strengths in composites than had
117: been thought. To derive it, we estimate the optimal lower and upper
118: bounds of the stress concentration in terms of the diameters of
119: fibers and the distances between them. These bounds explain the
120: dramatic change of the growth of stress when the diameter of the
121: fiber placed in the middle is relatively smaller than other two
122: fibers.
123: 
124: 
125: 
126: \par In the anti-plane shear model, the stress tensor represents the
127: electric field in the two dimensional space, where the out-of-plane
128: elastic displacement satisfies a conductivity equation, and the
129: cross-section of stiff fibers corresponds to the embedded
130: conductors. In this respect, we consider the gradient of the
131: solution to a conductivity problem to estimate the stress. Adjacent
132: stiff fiber-reinforcement induces the high stress concentration in
133: the narrow region between the fibers. This implies the blow-up of
134: the gradient of the solution between adjoining conductors, see
135: \cite{ADKL, AKL, AKLLL, BC, BLY, K,Y,Y2}.
136: 
137: \par  Meanwhile, the extreme conductivities are indispensable to the
138: blow-up phenomena: when the inclusion's conductivity is away from
139: zero and infinity, the boundedness of the stress function has been
140: derived by Li and Vogelius \cite{LV}, see also \cite{BV}, and it was
141: generalized to elliptic systems by Li and Nirenberg \cite{LN}. In
142: \cite{AKL,AKLLL}, for conductivities including both bounded and
143: extreme cases, Ammari et al.
144:  have established the optimal bounds of the gradient of solutions to the conductivity equation,
145:  when conductors are of circular shape in two dimensions, and the optimal bounds provides $\epsilon^{-1/2}$
146:   blow-up rate, where $\epsilon$ is the distance between two conductors. Yun \cite{Y, Y2} has extended
147:   this blow-up result for the case of two adjacent perfect conductors of a sufficiently general shape
148:    in two dimensions. In Bao, Li and Yin's paper \cite{ BLY},
149:    it has been also investigated as the blow-up phenomena in higher dimensional spaces, also see \cite{ADKL, LY}.
150: They has also done a natural follow-up in \cite{Bao_thesis, BLY2}
151: that the blow-up rate known only for a pair of fibers is still valid
152: for the multiple inclusions in any dimensions.
153: \par In contrast, our paper witnesses an unexpected fact on multiple inclusions that the
154: growth of stress can be significantly increased by a little
155: geometric change of an inclusion, even though the blow-up rate is
156: still $\epsilon^{-1/2}$.
157: 
158: 
159: 
160: 
161: \begin{figure}[h!]
162: \begin{center}
163: \epsfig{figure=aaa.eps,width=7.0cm}
164: \end{center}
165: \caption{Case (A) and case (B)}\label{caseAB}\end{figure}
166: 
167: 
168: \begin{figure}[h!]
169: \begin{center}\includegraphics[width=4 cm]{caseC.eps}
170: %\hbox{\setlength{\epsfxsize}{3.2 in} \epsfbox{96-emmer.eps}}
171: %\caption{a soup bubble in the tetrahedron}\label{fig.1}
172: ~~~\includegraphics[width=4 cm]{caseD.eps}
173: %\hbox{\setlength{\epsfxsize}{3.2 in} \epsfbox{96-emmer.eps}}
174: %\caption{caption Caption Caption Caption.} \label{fig.1}
175: \caption{Case (C) and case (D)}\label{caseCD}
176: \end{center}
177: \end{figure}
178: 
179: 
180: 
181: For $l=1,\dots,L$, let $D_l$ be conducting inclusions in $\RR^2$,
182:  that is cross-sections of stiff fibers. Then, under the action of the applied field $H$, the electric potential $u$ satisfies the following conductivity equation:
183: \begin{equation} \label{eq:conductivity}
184: \quad \left\{
185: \begin{array}{ll}
186: \ds\Delta u  = 0,\quad&\mbox{in }{\mathbb{R}^2 \backslash \overline{\cup_{l=1}^L D_l}},\\
187: \ds u(\textbf{x})- H(\textbf{x}) = O(|\textbf{x}|^{-1}),\quad&\mbox{as } |\textbf{x}| \rightarrow \infty,\\
188: \ds u |_{\partial D_l} = C_l~ \mbox{(constant)},\quad&\mbox{for }l=1,\dots,L,\\
189:  \int_{\partial D_l} {\partial_{  \nu} u }~dS = 0,\quad
190: &\mbox{for } l =1,\dots,L,
191: \end{array}
192: \right.
193: \end{equation}
194: where $H$ is an entire harmonic function $H$ in $\mathbb{R}^2$ and
195: $\textbf{x} =(x_1,x_2)$. In this paper, we only consider the case of
196: $L=2,3$,
197: 
198: 
199: 
200: 
201: 
202: As we mentioned previously, for the two fibers with the circular
203: cross-sectional shape, Ammari et al. \cite{AKL, AKLLL} obtained the
204: optimal blow-up rate $\epsilon^{-1/2}$ for $\nabla u$, and this
205: result is extended by Yun \cite{Y, Y2} to general shaped fibers.
206: Building on the prior results, we extend these into the the
207: following interesting direction: we first consider the circular
208: inclusions in Case (A) and Case (B), and second extend the result
209: into general shaped ones in Case (C) and Case (D). In Case (A) and
210: Case (B), we add a small circular inclusion between two others so
211: that three disk centers are lined up in one straight line. The
212: additional disk can be embedded disjointly from other disks, or it
213: partially overlap one of two disks, and we formulate these two cases
214: as follows.
215: \begin{itemize}
216:  \item[(A)] One disk and a pair of partially overlapping disks, Figure \ref{caseAB}: there is a portion of disk protruding from one of circular inclusions, i.e.,
217: $L=2$, and $D_1$ and $D_2$ are $\epsilon$-distanced
218: domains defined as
219: \begin{equation}\label{def:proballs}D_1=B_{r_1}(\mathbf{c}_1)\mbox{ and }D_2=B_{r_2}(\mathbf{c}_2)\cup B_{r_3}(\mathbf{c}_3),
220: \end{equation}
221: where $B_{r_l}(\mathbf{c}_l)$ is the disk with the radius $r_i$ and centered at $\mathbf{c}_l$,
222: and \begin{equation}\label{eqn:lumpcenters}\mathbf{c}_1=(-r_1-\frac{\epsilon}{2},0),\ \mathbf{c}_2=(r_2+\frac{\epsilon}{2},0),\mbox{ and } \mathbf{c}_3=(r_3+a+\frac{\epsilon}{2},0).\end{equation}
223: Here, $B_{r_2}(\mathbf{c}_2)$ is a small disk protruding from $B_{r_3}(\mathbf{c}_3)$, and we assume
224: $$
225: B_{r_2}(\mathbf{c}_2)\cap B_{r_3}(\mathbf{c}_3)\neq\emptyset,\mbox{
226: i.e., }0<a<2r_2,$$  $$\mbox{dist} (B_{r_1}(\mathbf{c}_1),
227: B_{r_3}(\mathbf{c}_3)) \simeq r_2$$ and $$0<\epsilon\ll r_2\ll
228: r_1\simeq\ r_3.$$
229: 
230: 
231: \item[(B)] Three disjoint disks, Figure \ref{caseAB}: a small disk is disjointly embedded into the in-between area of two disks, i.e.,
232: $L=3$, and
233: \begin{equation}\label{def:Bthree}D_l = B_{r_l}(\mathbf{c}_l),\
234: l=1,2,3,\end{equation}  where $\mathbf{c}_1=(-r_1 - \frac
235: {\epsilon_1} 2, 0)$, $\mathbf{c}_2=(r_2+ \frac {\epsilon_1} 2,0)$
236: and $\mathbf{c}_3=(r_3+r_2+\frac {\epsilon_1} 2 + \epsilon_2,0)$.
237: Hence, the distance between $D_1$ and $D_2$ is $\epsilon_1$, and the
238: distance between $D_2$ and $D_3$ is $\epsilon_2$. Here, $D_2$ is
239: regarded as the cross-section of the thin fiber between a pair of fibers with the cross-section $D_1$ and $D_3$. Thus, we assume
240: that
241: $$0<\epsilon_{i}\ll r_2\ll r_1\simeq r_3~~\mbox{for}~i =1,2.$$
242: \end{itemize}
243: 
244: 
245: 
246: 
247: 
248: In both cases, the blow-up rate is remarkably increased due to the
249: existence of $B_{r_2}(\mathbf{c_2})$ as follows:
250: \begin{thm}[Case A: Protruding small disk]\label{thm:A}
251: Let $D_1$ and $D_2$ be defined as \eqref{def:proballs}. Then there
252: is a positive constant $C$ independent of $\epsilon$, $r_1$, $r_2$
253: and $r_3$ such that
254: $$u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}\geq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\sqrt{\epsilon},$$
255: where $u$ is the solution to \eqref{eq:conductivity} with $H(x_1,x_2)=x_1$.
256: As a result, by the Mean Value Theorem, there is a point $\mathbf{x}_0$ in the narrow region between $D_1$ and $D_2$ such that
257: $$|\nabla u (\mathbf{x}_0)|\geq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon}}.$$
258: \par For any entire harmonic function $H$, let $u$ be the solution
259: to \eqref{eq:conductivity} with $H$. Then, there is a positive
260: constant $C$ independent of $\epsilon$, $r_1$, $r_2$ and $r_3$ such
261: that
262: $$|\nabla u (\mathbf{x})|\leq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon}}
263: $$ in the narrow region between $D_1~\mbox{and}~D_2$.
264: 
265: 
266: 
267: \end{thm}
268: 
269: \begin{thm}[Case B:  Disjointly embedded small disk] \label{thm:B}
270: Let $D_i$, $i=1,2,3$, be balls defined as \eqref{def:Bthree}. Then
271: there is a positive constant $C$ independent of $\epsilon_1$,
272: $\epsilon_2$, $r_1$, $r_2$ and $r_3$ such that
273: $$u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}\geq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\sqrt{\epsilon_1},$$
274: and
275: $$u\Bigr|_{\p D_3}-u\Bigr|_{\p D_2}\geq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\sqrt{\epsilon_2},$$
276: where $u$ is the solution to \eqref{eq:conductivity} with $H(x_1,x_2)=x_1$. As a result, by the Mean Value Theorem, there exists points; $\mathbf{x}_1$ in the narrow region between $D_1$ and $D_2$; $\mathbf{x}_2$ in the narrow region between $D_2$ and $D_3$, which satisfy that
277: $$|\nabla u (\mathbf{x}_i)|\geq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon_i}}~~\mbox{for}~i=1,~2.$$
278: 
279: 
280: \par For any entire harmonic function $H$, let $u$ be the solution
281: to \eqref{eq:conductivity} with $H$. Then, there is a positive
282: constant $C$ independent of $\epsilon_1$, $\epsilon_2$, $r_1$, $r_2$
283: and $r_3$ such that
284: $$|\nabla u (\mathbf{x})|\leq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon_i}}~~\mbox{for}~i=1,~2.$$
285:  in the narrow regions between $D_1~\mbox{and}~D_2$, and between $D_2~\mbox{and}~D_3$, respectively.
286: 
287: 
288: 
289: \end{thm}
290: 
291: In this paper, we first estimate the lower-bounds in terms of the
292: radii of inclusions. Based on this estimates we derive the
293: remarkable blow-up rate increasing phenomena when a small conducting
294: inclusion is located in-between region of two inclusions. This paper
295: is organized as follows: In section 2, we explain the method to
296: calculate the potential difference in the case of two disks. We then
297: derive the lower bound of Case (A) in section 3; Case (B) in section
298: 4. In the case of the upper bounds, the major part of derivation
299: overlaps in Case (A) and Case (B). Thus, the derivation is presented
300: in Subsection \ref{upp}. Based on the similar derivation, we can
301: also obtain the analogues of Theorem \ref{thm:A} and \ref{thm:B} for
302: the inclusions associated by a sufficiently general class of shapes.
303: 
304: 
305: \subsubsection* {Analogues of Theorem \ref{thm:A} and \ref{thm:B} for a sufficiently general class of shapes}
306: The proofs of Theorem \ref{thm:A} and \ref{thm:B} are flexible
307: enough even though the results are restricted to circular
308: inclusions. The estimates presented in Theorem \ref{thm:A} and
309: \ref{thm:B} can be extended to the inclusions associated by a
310: sufficiently general class of shapes. To consider a large class of
311: shapes, we make the geometric assumptions more precise. To define
312: $D_1$, $D_2$ and $D_3$, we consider three domains
313: $D_{\mbox{right}}$, $D_{\mbox{center}}$ and $D_{\mbox{left}}$ in $
314: \mathbb{R}^2$. In addition, we assume that $\varphi_{\mbox{right}} :
315: \mathbb{C} \backslash B_1(0) \rightarrow \mathbb {R}^2 \backslash
316: D_{\mbox{right}}$, $\varphi_{\mbox{center}} : \mathbb{C} \backslash
317: B_1(0) \rightarrow \mathbb {R}^2 \backslash D_{\mbox{center}}$ and
318: $\varphi_{\mbox{left}} : \mathbb{C} \backslash B_1(0) \rightarrow
319: \mathbb {R}^2 \backslash D_{\mbox{left}}$ are conformal mappings in
320: $C^2 ( \mathbb{C} \backslash B_1(0))$ such that
321: $\varphi_{\mbox{right}}'(z)\neq 0~\mbox{ and}
322: ~\varphi_{\mbox{left}}'(z)\neq 0$ for $z \in
323: \partial B_1 (0)$. Here, we do not distinguish
324: $\mathbb{R}^2$ from $\mathbb{C}$.  The $C^2$ regularity condition of
325: these conformal mappings doses not allow non-smooth inclusions such
326: as polygons, but Riemann mapping theorem yields a sufficiently
327: general class of shapes: refer to Ahlfors \cite{A}. Now, we consider
328: the analogues of Theorem \ref{thm:A} and \ref{thm:B} for two cases
329: as follows:
330: \begin{itemize}
331: 
332: 
333:  \item[(C)]  One domain and a pair of partially overlapping domains, similarly to Figure \ref{caseCD}: there is a small portion of another domain protruding from a inclusion, i.e.,
334: $L=2$, and $D_1$ and $D_2$ are $\epsilon$-distanced domains defined
335: as
336: \begin{equation}\label{thC} D_1=D_{\mbox{left}}\mbox{ and } D_2= \left ( r_2 D_{\mbox{center}}\right )\cup D_{\mbox{right}},
337: \end{equation}
338: where $r_2 D_{\mbox{center}}$ is the $r_2$ times diminished domain
339: of $D_{\mbox{center}}$. We suppose that $D_2$ is a connected domain,
340: $\mbox{dist}(D_1,D_2)=\mbox{dist}(D_1,r_2 D_{\mbox{center}})$, $$
341: \mbox{dist}(D_1,D_{\mbox{right}})\backsimeq r_2,$$ $$D_1 \subset
342: \mathbb{R_{-}}\times \mathbb{R}~ \mbox{and} ~D_2 \subset
343: \mathbb{R_{+}}\times \mathbb{R}.$$ In addition, we also assume that
344: $r_2$ is small enough and $$0< \epsilon \ll r_2,$$ and that the
345: boundaries $\partial D_1$, $\partial D_2$ and $\partial
346: D_{\mbox{right}}$ are strictly convex in the narrow region between
347: $D_1$ and $D_2$.
348: 
349: 
350:  \item[(D)]  Three disjoint domains $D_1$, $D_2$ and $D_3$, Figure \ref{caseCD}:
351:  a small inclusion $D_2$ is disjointly embedded into the in-between area of two other domains, i.e.,
352: $L=3$, and
353: \begin{equation}\label{thD} D_1=D_{\mbox{left}},~D_2= r_2 D_{\mbox{center}}~\mbox{ and }
354: D_3= D_{\mbox{right}} \end{equation}where $r_2 D_{\mbox{center}}$ is
355: the $r_2$ times diminished domain of $D_{\mbox{center}}$.  We assume
356: that
357: 
358:  $D_1$ and $D_2$ are $\epsilon_1$ apart, $D_2$ and $D_3$ are $\epsilon_2$
359:  apart, and $D_1$ distances enough from $D_3$ that $r_2$ is
360:  sufficiently small and
361: $$0<\epsilon_{i}\ll r_2~~\mbox{for}~i =1,2,$$
362: since $D_2$ is regarded as the cross-section of the thin fiber, that
363: the boundaries $\partial D_1$, $\partial D_2$ and $\partial
364: D_{\mbox{right}}$ are strictly convex in the narrow region between
365: $D_1$ and $D_2$,  $$ D_1 \subset \mathbb{R}_- \times \mathbb{R}
366: ~\mbox{and}~ D_2 \cup D_3 \subset \mathbb{R}_+ \times \mathbb{R} .$$
367: 
368: \end{itemize}
369: 
370: 
371: \begin{thm}[Case C: Protruding small lump]\label{thm:C}
372: Let $D_1$ and $D_2$ be defined as \eqref{thC}. Then there is a
373: positive constant $C$ independent of $\epsilon$ and $r_2$ such that
374: $$u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}\geq C\frac{1}{\sqrt{r_2}}\sqrt{\epsilon},$$
375: where $u$ is the solution to \eqref{eq:conductivity} with
376: $H(x_1,x_2)=x_1$. As a result, by the Mean Value Theorem, there is a
377: point $\mathbf{x}_0$ in the narrow region between $D_1$ and $D_2$
378: such that
379: $$|\nabla u (\mathbf{x}_0)|\geq C\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon}}.$$
380: 
381: \par For any entire harmonic function $H$, let $u$ be the solution
382: to \eqref{eq:conductivity} with $H$. Then, there is a positive
383: constant $C$ independent of  $r_2$ and $\epsilon$ such that
384: $$|\nabla u (\mathbf{x})|\leq C\frac{1}{\sqrt{r_2}}\frac 1
385: {\sqrt{\epsilon}}$$ in the narrow region between
386: $D_1~\mbox{and}~D_2$.
387: \end{thm}
388: 
389: \begin{thm}[Case D: Disjointly embedded small inclusion] \label{thm:D}
390: Let $D_i$, $i=1,2,3$, be balls defined as \eqref{thD}. Then there is
391: a positive constant $C$ independent of $r_2$, $\epsilon_1$ and
392: $\epsilon_2$ such that
393: $$u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}\geq C\frac{1}{\sqrt{r_2}}\sqrt{\epsilon_1},$$
394: and
395: $$u\Bigr|_{\p D_3}-u\Bigr|_{\p D_2}\geq C\frac{1}{\sqrt{r_2}}\sqrt{\epsilon_2},$$
396: where $u$ is the solution to \eqref{eq:conductivity} with
397: $H(x_1,x_2)=x_1$. As a result, by the Mean Value Theorem, there
398: exists points; $\mathbf{x}_1$ in the narrow region between $D_1$ and
399: $D_2$; $\mathbf{x}_2$ in the narrow region between $D_2$ and $D_3$,
400: which satisfy that
401: $$|\nabla u (\mathbf{x}_i)|\geq C\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon_i}}~~\mbox{for}~i=1,~2.$$
402: 
403: 
404: \par For any entire harmonic function $H$, let $u$ be the solution
405: to \eqref{eq:conductivity} with $H$. Then, there is a positive
406: constant $C$ independent of $r_2$,  $\epsilon_1$ and $\epsilon_2$
407: such that
408: $$|\nabla u (\mathbf{x}_i)|\leq C\frac{1}{\sqrt{r_2}}\frac 1 {\sqrt{\epsilon_i}}~~\mbox{for}~i=1,~2.$$
409:  in the narrow regions between
410: $D_1~\mbox{and}~D_2$, and $D_2~\mbox{and}~D_3$, respectively.
411: \end{thm}
412: We derive the lower bound of Case (C) in section 3; Case (D) in
413: section 4. In the case of the upper bounds, the major part of
414: derivation overlaps in Case (A), Case (B), Case (C) and Case (D).
415: Thus, the main idea is presented in Subsection \ref{upp}.
416: 
417: \section{Preliminary}
418: \subsection{Calculation of the potential difference}\label{section:difference}
419: We explain the main idea to calculate the difference of potential
420: between two adjacent, possibly disconnected, conductors.
421: 
422: In this section, differently from \eqref{eq:conductivity}, $D_i$, $i=1,2$, could be also the union of two disjoint domains.
423: Define $u$ as the solution to \eqref{eq:conductivity}, where it is assigned one constant value throughout $D_i$ even when $D_i$ is disconnected.
424: Now, define $h$ as the solution to
425: \begin{equation} \label{def:h}
426: \quad \left\{
427: \begin{array}{ll}
428: \ds\Delta h  = 0,\quad&\mbox{in }{\mathbb{R}^2 \backslash \overline{(D_1 \cup D_2)}}, \\
429: \ds h= O(|\textbf{x}|^{-1}),\quad&\mbox{as } |\textbf{x}| \rightarrow \infty,\\
430: \ds h |_{\partial D_i} = k_i\mbox{ (constant)},\quad&\mbox{for }  i = 1, 2,\\
431: \ds\int_{\partial D_i} {\partial_{  \nu} h }~dS =(-1)^i,\quad&\mbox{for }  i = 1, 2,
432: \end{array}
433: \right.
434: \end{equation}
435: where $\nu$ is the outward unit normal vector of ${\mathbb{R}^n
436: \backslash \overline{(D_1 \cup D_2)}}$, i.e., directed inward of $D_i$.
437: To indicate the dependence of $u$ and $h$ on $D_1$ and
438: $D_2$, we denote them as
439: \begin{equation}\label{def:Phi}
440: u=\Phi[D_1,D_2],
441: \end{equation}
442: \begin{equation}\label{def:Psi}
443: h=\Psi[D_1,D_2].
444: \end{equation}
445: 
446: The potential difference of $u$ in $D_1$ and $D_2$ is represented in terms of $h$ as follows.
447: \begin{lem}\label{lemma:h}(\cite{Y})
448: \begin{equation}\label{eq:102}
449: u\Bigr|_{\partial D_2} - u\Bigr|_{\partial D_1}=\int_{\partial D_1}
450: H{\partial_{  \nu} h }\ dS+ \int_{\partial D_2}H {\partial_{ \nu} h
451: } \ dS.\end{equation}
452: \end{lem}
453: The lemma above can be derived by the Divergence Theorem, see
454: \cite{Y}.
455: 
456: 
457: \subsection{Two disks in $\RR^2$}\label{section:twodisks}
458: Using Lemma \ref{lemma:h}, we can easily calculate the potential difference $u|_{D_2}-u|_{D_1}$ of the solution $u$ to \eqnref{eq:conductivity} when
459: \begin{equation}\label{twodisks}
460: D_1 = B_{r_1}(\mathbf{c}_1)~\mbox{and}~D_2 =
461: B_{r_2}(\mathbf{c}_2),
462: \end{equation}
463: where $\mathbf{c}_1=(-r_1 - \frac{\epsilon}{2}, 0)$ and $\mathbf{c}_2=(r_2+ \frac{\epsilon}{2},0)$.
464: 
465: Let $R_i$ be the reflection with respect to $D_i$, in other words,
466: $$R_i(\mathbf{x})=\frac{r_i^2(\mathbf{x}-\mathbf{c}_i)}{|\mathbf{x}-\mathbf{c}_i|^2}+ \mathbf{c}_i,\ i=1,2,$$
467: and $\textbf{p}_1\in D_1$ be the fixed point of $R_1 \circ R_2$, then
468: $R_2(\textbf{p}_1)(=:\textbf{p}_2)$ is the fixed point of  $R_2 \circ R_1$, and
469: $$\textbf{p}_1 =
470: \Bigr(-\sqrt{2}\sqrt{\frac{r_1r_2}{r_1+r_2}}\sqrt\epsilon +O(\epsilon)
471: ,0\Bigr)\mbox{ and }\textbf{p}_2 =
472: \Bigr(\sqrt{2}\sqrt{\frac{r_1r_2}{r_1+r_2}}\sqrt\epsilon +O(\epsilon)
473: ,0\Bigr).$$
474: Moreover, we can easily show that
475: \begin{equation}\label{h:twodisks}\Psi[D_1,D_2]=\frac 1 {2\pi} \left( \log |\textbf{x}-\textbf{p}_1|- \log |\textbf{x}-\textbf{p}_2|\right).\end{equation}
476: By an elementary calculation, it can be shown that the middle point
477: $\frac {\textbf{p}_1 + \textbf{p}_2}{2}$ exists between two
478: approaching points $\left(-\frac {\epsilon} 2, 0 \right)$ and
479: $\left(\frac {\epsilon} 2, 0 \right)$. Applying the middle point
480: property to estimate for $\Psi[D_1,D_2]\left(\pm\frac {\epsilon} 2,
481: 0 \right)$, we can get the following lemma.
482: \begin{lem}\label{lemma:twodisks} There is a constant $C >0$
483: independent of $\epsilon$, $r_1$ and $r_2$ such that
484: $$\frac 1 C \sqrt{\frac {r_1+r_2} {r_1r_2}}\sqrt\epsilon \leq \Psi[D_1,D_2]\big|_{\partial D_2} - \Psi[D_1,D_2]\big|_{\partial D_1} \leq  C \sqrt{\frac {r_1+r_2} {r_1r_2}}\sqrt\epsilon $$
485: for small $\epsilon >0$. \end{lem} From Lemma \ref{lemma:h}, we
486: calculate the potential difference of $u$.
487: \begin{lem} \label{lem++}
488: Let $H(x_1,x_2)$ be an entire harmonic function. The solution $u$ to
489: \eqnref{eq:conductivity} where $L=2$ and $D_l$, $l=1,2$, are given
490: as \eqref{twodisks}
491:  satisfies
492:  \be
493:  \begin{array}{ll}u|_{\partial D_2} - u|_{\partial D_1} &=
494: H(\textbf{p}_2)-H(-\textbf{p}_1)\\ &= 2\sqrt{2} \partial_{x_1} H (0,0)
495: \sqrt{\frac{r_1r_2}{r_1+r_2}}\sqrt\epsilon+O(\epsilon).
496:  \label{eq:107}\end{array}\ee
497: 
498:  \end{lem}
499:  \begin{rem}
500:  Referring to the mean value theorem, there exists a point
501: $\textbf{x}_2$ between $\partial D_1$ and $\partial D_2$ such that
502: \be |\nabla u (\textbf{x}_2)|\geq2\sqrt{2} |\partial_{x_1} H(0,0)|
503: \sqrt{\frac{r_1r_2}{r_1+r_2}} \frac 1 {\sqrt{\epsilon}}.
504: \label{eq:108}\ee for any sufficiently small $\epsilon >0$.
505: Moreover, as a result in \cite{AKLLL}, there is a constant $C$
506: independent of $\epsilon$, $r_1$ and $r_2$ such that
507: $$ \norm{\nabla u }_{L^{\infty}(\Omega \setminus (D_1 \cup D_2)) }\leq C \norm {\nabla H}_{L^{\infty}(\Omega)}
508: \sqrt{\frac{r_1r_2}{r_1+r_2}} \frac 1 {\sqrt{\epsilon}} $$ where
509: $\Omega=B_{4(r_1+r_2)}(0,0).$
510: \end{rem}
511: 
512: 
513: 
514: 
515: 
516: 
517: \section{One disk and a pair of partially overlapping disks}
518: In this section, we consider two $\epsilon$-distanced domains $D_1$
519: and $D_2$, see Case (A) at Figure \ref{caseAB}, where
520: \begin{equation*}D_1=B_{r_1}(\mathbf{c}_1)\mbox{ and }D_2=B_{r_2}(\mathbf{c}_2)\cup B_{r_3}(\mathbf{c}_3),
521: \end{equation*}
522: where \begin{equation*}\mathbf{c}_1=(-r_1-\frac{\epsilon}{2},0),\ \mathbf{c}_2=(r_2+\frac{\epsilon}{2},0),\mbox{ and } \mathbf{c}_3=(r_3+a+\frac{\epsilon}{2},0).\end{equation*}
523: Here, $B_{r_2}(\mathbf{c}_2)$ is a small lump of $B_{r_3}(\mathbf{c}_3)$, and we assume
524: $$
525: B_{r_2}(\mathbf{c}_2)\cap B_{r_3}(\mathbf{c}_3)\neq\emptyset,\mbox{ i.e., }0<a<2r_2,$$
526: and $$\mbox{ and }0<\epsilon\ll r_2\ll \min(r_1,\ r_3).$$
527: 
528: Define
529: \begin{equation}\label{h:D1D2}h=\Psi[D_1, D_2],\mbox{ and } u=\Phi[D_1, D_2],
530: \end{equation}
531: and
532:  \begin{equation}\label{hj:D1D2}h_j=\Psi[D_1, B_{r_j}(\mathbf{c}_j)],\mbox{ and }u_j=\Phi[D_1, B_{r_j}(\mathbf{c}_j)],\qquad j=2,3,
533: \end{equation}
534: where $\Psi$ and $\Phi$ defined in section \ref{section:difference}.
535: 
536: 
537: \subsection{Properties of $h$ and $h_j$, $j=2,3$}
538: 
539: \begin{lem}\label{property:hi}
540: Let $h=\Psi[D_1, D_2]$, then we have
541: \begin{equation}\label{nuh:small}
542: \p_\nu h(x)=O(\sqrt{\epsilon}),\quad x\in \p D_2\setminus B_{r_2}(\mathbf{c}_2).
543: \end{equation}
544: \end{lem}
545: \pf
546: Set $$B_i=B_{r_i}(\mathbf{c}_i),\quad\mbox{for }i=1,2,3.$$
547: 
548: We choose a smooth domain $\widetilde{\Omega}$ as
549: follows: \begin{align*}& \widetilde{\Omega} \subset (B_2 \cup
550: B_3),~B_3 \subset \widetilde{\Omega}\notag\\&\partial
551: \widetilde{\Omega} \setminus B_2=\partial B_3 \setminus B_2\notag\\
552: & (\partial \widetilde{\Omega} \cap \partial B_2)\setminus
553: \partial B_3 = (\frac 1 2\epsilon,0)\notag\end{align*}
554: Let $\widetilde{h} = \Psi[B_1, \widetilde{\Omega}]$. Then, we
555: consider $V$ defined in $\mathbb{R}^2 \setminus (B_1 \cup B_2 \cup
556: B_3)$ as follows:
557: $$V = h - \frac {h|_{\partial B_1} - h|_{\partial (B_2 \cup B_3)}}{\widetilde{h}|_{\partial B_1} - \widetilde{h}|_{\partial \widetilde{\Omega}}} \widetilde{h}.$$
558: Then, it follows that
559: $$V|_{\partial B_1} = V|_{\partial \widetilde{\Omega}\setminus B_2}=\mbox{a constant} .$$
560: Since $h|_{\partial (B_2 \cup B_3)}>h|_{\partial B_1}$ and $
561: \widetilde{h}|_{\partial
562: \widetilde{\Omega}}>\widetilde{h}|_{\partial B_1}$, the minimum of
563: $V$ attains on $\partial \widetilde{\Omega} \setminus B_2$ and
564: $\partial B_1$. Thus, we have \be\partial_{\nu} h - \frac
565: {h|_{\partial B_1} - h|_{\partial (B_2 \cup
566: B_3)}}{\widetilde{h}|_{\partial B_1} - \widetilde{h}|_{\partial
567: \widetilde{\Omega}}} \partial_{\nu} \widetilde{h}\leq
568: 0~\mbox{on}~\partial B_1 \cup (\partial \widetilde{\Omega}\setminus
569: B_2).\label{eq:103}\ee By the integration on $\partial B_1$, we have
570: $$0<\frac
571: {h|_{\partial B_1} - h|_{\partial (B_2 \cup
572: B_3)}}{\widetilde{h}|_{\partial B_1} - \widetilde{h}|_{\partial
573: \widetilde{\Omega}}}\leq 1.$$ Using the bound \eqref{eq:103} once
574: more, we have
575: $$\partial_{\nu} h  \leq  \partial_{\nu}\widetilde{h}|_{\partial
576: \widetilde{\Omega}} ~\mbox{on}~\partial \widetilde{\Omega}\setminus
577: B_2=\partial B_3 \setminus B_2.$$
578: 
579: The domain $\widetilde{ \Omega}$ is smooth so that we can use the
580: method presented by Yun \cite{Y,Y2}. Then, up to a conformal mapping
581: to a circle, $\partial_{\nu} h$ is bounded by constant times the
582: Poisson Kernel with respect to a interior point $\sqrt \epsilon$
583: distanced from the boundary (refer to the inequality (9) in
584: \cite{Y2}). Note that $\partial B_3 \setminus B_2$ distances enough
585: from $(\epsilon,0)$. Thus, we have
586: $$\partial_{\nu} h  \leq  \partial_{\nu}\widetilde{h}|_{\partial
587: \widetilde{\Omega}} \leq C \sqrt {\epsilon} ~\mbox{on}~\partial B_3
588: \setminus B_2.$$ Therefore, we have completed the proof of the
589: lemma. \qed
590: 
591: \begin{lem}\label{lem:Anuh}
592: \begin{equation}
593: \p_\nu h (x)\leq M\p_\nu h_3(x),\quad x\in\p D_1,
594: \end{equation}
595: where\begin{equation}\label{def:M}M=\frac{h\bigr|_{\p
596: D_2}-h\bigr|_{\p D_1}}{h_3\bigr|_{\p
597: B_{r_3}(\mathbf{c}_3)}-h_3\bigr|_{\p D_1}}.\end{equation}
598: \end{lem}
599: \pf
600:  Define $$W=h-Mh_3,\qquad \mbox{in }\RR^2\setminus(D_1\cup D_2).$$
601: Since $h$ is constant on $\p  D_2$, $M>0$, and $h_3$ takes it's
602: maximum on $\p B_{r_3}(\mathbf{c}_3)$,
603: $$W\Bigr|_{\p D_2\setminus B_{r_2}(\mathbf{c}_2)}-W\Bigr|_{\p D_2\setminus B_{r_3}(\mathbf{c}_3)}=-M(h_3\Bigr|_{\p B_{r_3}(\mathbf{c}_3)}-h_3\Bigr|_{\p D_2\setminus B_{r_3}(\mathbf{c}_3)})<0,$$
604: and $$W\Bigr|_{\p D_2\setminus B_{r_2}(\mathbf{c}_2)}-W\Bigr|_{\p
605: D_1}=h\Bigr|_{\p D_2}-h\Bigr|_{\p D_1}-M(h_3\Bigr|_{\p
606: B_{r_3}(\mathbf{c}_3)}-h_3\Bigr|_{\p D_1})=0.$$
607: 
608: Therefore, $W$ takes its minimum on $\p D_1$, and
609: $$\p_\nu W\leq0,\quad \mbox{on }\p D_1.$$
610: \qed
611: 
612: \begin{lem}\label{lem:Asamediffer}
613: \begin{equation}\label{samediffer}
614: \ds h\Bigr|_{\p D_2}-h\Bigr|_{\p  D_1}=h_2\Bigr|_{\p B_{r_2}(\mathbf{c}_2)}-h_2\Bigr|_{\p D_1}+O(\epsilon).
615: \end{equation}
616: \end{lem}
617: \pf
618:  Note that $$\int_{\p D_1}\p_\nu (h-h_2)\ dS=0,$$ and
619:  \begin{align*}\ds\int_{\p D_2}\p_\nu(h-h_2)\ dS&=\ds\int_{\p D_2}\p_\nu h\ dS-\int_{\p B_{r_2}(\mathbf{c}_2)}\p_\nu h_2\ dS-\int_{\p (D_2\setminus B_{r_2}(\mathbf{c}_2))}\p_\nu h_2 \ dS\\&=\ds 1-1-0=0.\end{align*}
620: With the fact that $h|_{\p D_1}$ and $h|_{\p D_2}$ are constants and
621: the (exterior) Divergence Theorem, we have that
622: \begin{align*}
623: \ds0&=\int_{\p D_1}\p_\nu(h-h_2)h\ dS+\int_{\p D_2}\p_\nu (h-h_2)h\ dS\\
624: \ds&=\int_{\p D_1}(h-h_2)\p_\nu h\ dS+\int_{\p D_2}(h-h_2)\p_\nu h\ dS.
625: \end{align*}
626: Hence,
627: \begin{align*}
628: h\Bigr|_{\p D_1}-h\Bigr|_{\p  D_2}
629: &=\int_{\p D_1}h\p_\nu h\ d S+\int_{\p D_2}h\p_\nu h\ d S\\
630: &=\int_{\p D_1}h_2\p_\nu h \ dS+\int_{\p D_2}h_2\p_\nu h\ dS\\
631: &=h_2\Bigr|_{\p D_1}-h_2\Bigr|_{\p B_{r_2}(\mathbf{c}_2)}+\int_{\p D_2}(h_2-h_2\Bigr|_{\p B_{r_2}(\mathbf{c}_2)})\p_\nu h\ dS
632: \end{align*}
633: From \eqref{h:twodisks}, there is a constant $C$ dependent of $a$, see \eqref{eqn:lumpcenters}, such that
634: $$\Bigr|(h_2-h_2\Bigr|_{\p B_{r_2}(\mathbf{c}_2)})(x)\Bigr|\leq C\sqrt{\epsilon},\quad\mbox{for all }x\in \p  D_2\setminus B_{r_2}(\mathbf{c}_2).$$
635: Therefore, with \eqref{nuh:small} as well, we obtain \eqref{samediffer}.
636: \qed
637: 
638: 
639: 
640: 
641: 
642: 
643: \subsection{Proof Theorem \ref{thm:A}}
644: Let $H(x_1,x_2)=x_1$ and  $\nu$ be the unit normal vector of ${\mathbb{R}^2\backslash \overline{(D_1 \cup D_2)}}$, i.e., directed inward to $D_i$, $i=1,2$. Remind that we defined
645: \begin{equation}\label{h:D1D2}h=\Psi[D_1, D_2],\mbox{ and } u=\Phi[D_1, D_2],
646: \end{equation}
647: and
648:  \begin{equation}\label{hj:D1D2}h_j=\Psi[D_1, B_{r_j}(\mathbf{c}_j)],\mbox{ and }u_j=\Phi[D_1, B_{r_j}(\mathbf{c}_j)],\qquad j=2,3,
649: \end{equation}
650: where $\Psi$ and $\Phi$ defined in section \ref{section:difference}.
651: 
652: 
653: 
654: 
655: 
656: Note that $\partial_{ \nu}h\Bigr|_{\p D_2}<0$, $H <0$ on $\p D_1$
657: and $H >0$ on $\p D_2$, and, as a result, from Lemma \ref{lemma:h},
658: we have
659: \begin{align}
660: u\Bigr|_{\partial D_2} - u\Bigr|_{\partial D_1}&=\int_{\partial D_1}\nonumber
661: ({\partial_{  \nu} h }) H\ dS+ \int_{\partial D_2}( {\partial_{ \nu}
662: h }) H\ dS\\
663: &\geq\int_{\p D_1}H\p_\nu h\ dS.\label{ud1d2_first}
664: \end{align}
665: 
666: 
667: 
668: 
669: Applying the lemma \ref{lem:Anuh}, Lemma \ref{lem:Asamediffer}, \eqref{ud1d2_first} becomes \begin{align*}
670: \ds u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}&\geq \frac{h\bigr|_{\p D_2}-h\bigr|_{\p D_1}}{h_3\bigr|_{\p B_{r_3}(\mathbf{c}_3)}-h_3\bigr|_{\p D_1}}\int_{\p D_1} H\p_\nu h_3\ dS\\
671: \ds&\geq \frac{h_2\bigr|_{\p B_{r_2}(\mathbf{c}_2)}-h_2\bigr|_{\p D_1}+O(\epsilon)}{h_3\bigr|_{\p B_{r_3}(\mathbf{c}_3)}-h_3\bigr|_{\p D_1}}\sqrt{2}\sqrt{\frac{r_1r_3}{r_1+r_3}r_2}.
672: \end{align*}
673: It follows  from Lemma \ref{lemma:twodisks} that
674: $$h_2\Bigr|_{\p B_{r_2}(\mathbf{c}_2)}-h_2\Bigr|_{\p D_1} \geq C \sqrt{\frac{r_1+r_2} {r_1r_2} } \sqrt{\epsilon}+O(\epsilon)$$
675: and
676: $$h_3\Bigr|_{\p B_{r_3}(\mathbf{c}_3)}-h_3\Bigr|_{\p D_1} \leq C\sqrt{\frac{r_1+r_3} {r_1r_3} } \sqrt{r_2}+O(r_2).$$
677: Therefore,
678: \begin{align*}
679: \ds u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}&\geq C\frac{r_1r_3}{r_1+r_3}\frac{1}{\sqrt{r_2}}\sqrt{\epsilon}.
680: \end{align*}
681: This proves  Theorem \ref{thm:A}.
682: \qed
683: 
684: 
685: 
686: \subsection{Proof Theorem \ref{thm:C}}
687: We consider the general shaped domain in Theorem \ref{thm:C}. But,
688: we take an advantage of the properties of circular inclusions. To
689: make a connection between circular domains and general shaped
690: domains, we need to establish the monotonic property of $\Psi$ as
691: follows:
692: 
693: \begin{lem} \label{lem:mono} [Monotonic property of $\Psi$] Let $D_A$, $D_B$, $\widetilde{D}_A$ and
694: $\widetilde{D}_B$ be  domains. Assume that $$D_A \subseteq
695: \widetilde{D}_A~\mbox{and}~D_B \subseteq \widetilde{D}_B. $$ Then,
696: we have
697: $$ 0 \leq \Psi [\widetilde{D}_A, \widetilde{D}_B]\big|_{\partial \widetilde{D}_B} -  \Psi [\widetilde{D}_A, \widetilde{D}_B]\big|_{\partial \widetilde{D}_A}
698:  \leq \Psi [D_A, D_B]\big|_{\partial D_B} -  \Psi [D_A, D_B]\big|_{\partial D_A}.$$
699: \end{lem}
700: \pf Without any loss of generality, we consider only the case of
701: $D_A = \widetilde{D}_A.$ Let $$G =\Psi [{D}_A, \widetilde{D}_B] - M
702: \Psi [{D}_A, {D}_B] $$ where
703: $$M = \frac {\Psi [D_A, \widetilde{D}_B]\big|_{\partial D_B} -  \Psi [D_A, \widetilde{D}_B]\big|_{\partial D_A} }{\Psi [{D}_A, {D}_B]\big|_{\partial {D}_B} -  \Psi [{D}_A,{ D}_B]\big|_{\partial {D}_A} } .$$
704: The minimum  of $G$ attains on $\partial D_A$. By the Hopf's Lemma,
705: we have
706: $$\partial_{\nu} G \leq 0~\mbox{on}~\partial D_A.$$
707: Integrating $\p_{\nu} G$ on $\p D_A$, we have $-1 + M \leq 0$.
708: Therefore, we have
709: $$ 0 \leq \Psi [{D}_A, \widetilde{D}_B]\big|_{\partial \widetilde{D}_B} -  \Psi [{D}_A, \widetilde{D}_B]\big|_{\partial {D}_A}
710:  \leq \Psi [D_A, D_B]\big|_{\partial D_B} -  \Psi [D_A, D_B]\big|_{\partial D_A}.$$
711: Repeating the same argument again, we can obtain the disable
712: inequality.
713: 
714: \qed
715: 
716: 
717: 
718: 
719: Applying $D_{\mbox{left}}$, $r_2 D_{\mbox{center}}$ and
720: $D_{\mbox{right}}$ instead of $B_{r_i}(\mathbf{c}_i)$, $i=1,2,3$, to
721: the argument presented in the proof of Theorem \ref{thm:A}, we can
722: obtain
723: $$
724: \ds u\Bigr|_{\p r_2 D_{\mbox{center}}}-u\Bigr|_{\p
725: D_{\mbox{left}}}\geq C \frac{h_2\bigr|_{\p\left( r_2
726: D_{\mbox{center}} \right) }-h_2\bigr|_{\p
727: D_{\mbox{left}}}+O(\epsilon)}{h_3\bigr|_{\p
728: D_{\mbox{right}}}-h_3\bigr|_{\p D_{\mbox{left}}}}\sqrt{r_2} $$when
729: $H(x_1,x_2) = x_1$. Here, $h_1 = \Psi [D_{\mbox{left}}, r_2
730: D_{\mbox{center}}]$ and $h_2 = \Psi [D_{\mbox{left}},
731: D_{\mbox{right}}]$.
732: 
733:  It follows that from Yun
734: \cite{Y, Y2} that
735: $$h_3\Bigr|_{\p D_{\mbox{right}}}-h_3\Bigr|_{\p D_{\mbox{left}}}\simeq  \sqrt{r_2}.$$
736: To estimate $h_2\Bigr|_{\p r_2 D_{\mbox{center}}}-h_2\Bigr|_{\p
737: D_{\mbox{left}}}$, we choose two disks $B_{\mbox{left}}$ and
738: $B_{\mbox{center}}$ containing $D_{\mbox{left}}$ and
739: $D_{\mbox{center}}$ such that the distance between $B_{\mbox{left}}$
740: and $r_2 B_{\mbox{center}}$ is $\epsilon$. Using Lemma
741: \ref{lem:mono} and \ref{lemma:twodisks}, we have
742: $$h_2\Bigr|_{\p \left(r_2 D_{\mbox{center}}\right)}-h_2\Bigr|_{\p D_{\mbox{left}}} \gtrsim  \sqrt{\frac {\epsilon}{r_2}}.$$
743: Note that $D_1= D_{\mbox{left}}$, $D_2 = r_2 D_{\mbox{center}}$ and
744: $D_3 = D_{\mbox{right}}$ in this theorem.  Therefore,
745: \begin{align*}
746: \ds u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1}&\geq C
747: \frac{1}{\sqrt{r_2}}\sqrt{\epsilon}.
748: \end{align*}
749: This proves  Theorem \ref{thm:C}. \qed
750: 
751: 
752: 
753: 
754: 
755: 
756: \smallskip
757: 
758: 
759: \section{Three disjoint smooth domains}\label{section:threedomains}
760: We consider three disjoint inclusion case, see Figure \ref{caseAB}
761: and \ref{caseCD}, a small one is disjointly embedded into the
762: in-between area of two others, and prove Theorem \ref{thm:B} and
763: \ref{thm:D}. We assume that $D_1$ and $D_2$ are closely spaced with
764: the distance $\epsilon_1$, and $D_2$ and $D_3$ are closely space
765: with $\epsilon_2$, but $D_1$ and $D_3$ are not close, and that
766: $D_1$, $D_2$ and $D_3$ have the boundary regularity given in Theorem
767: \ref{thm:D}.
768: 
769: 
770: \subsection{Solution representation of $u$}\label{subsec}
771: Let $H^c$ be a harmonic function outside of $\cup_{i=1}^3 D_i$ and
772: have the same constant value in $\cup_{i=1}^3 D_i$ satisfying that
773: \begin{equation} \label{def:Hc}
774: \quad \left\{
775: \begin{array}{ll}
776: \ds\Delta H^c  = 0,\quad&\mbox{in }{\mathbb{R}^2 \backslash \overline{\cup_{i=1}^3 D_i}},\\
777: \ds H^c(\textbf{x})- H(\textbf{x}) = O(|\textbf{x}|^{-1}),\quad&\mbox{as } |\textbf{x}| \rightarrow \infty,\\
778: \ds H^c |_{\cup_{i=1}^3\partial D_i} = C_H~ \mbox{(constant)}.
779: \end{array}
780: \right.
781: \end{equation}
782: Since $H^c-H$ is harmonic at infinity, $H^c-H$ attains maximum only at the boundary points of $D_i$, $i=1,2,3$.
783: To make $H^c-H$ attains zero at infinity, $C_H$ should satisfy \begin{equation}\label{range:c}
784: \ds-\bigr\|H\bigr\|_{L^\infty{(\cup_{i=1}^3 D_i})}\leq C_H\leq\bigr\|H\bigr\|_{L^\infty{(\cup_{i=1}^3 D_i})}.
785: \end{equation}
786: Moreover, $H^c$ satisfies $\sum_{i=1}^3\int_{\partial D_i} {\partial_{  \nu} H^c }~dS = 0.$
787: 
788: The solution $u$ to \eqref{eq:conductivity} is represented as
789: \begin{equation}\label{rep:u3balls}
790: \ds u(\textbf{x})=H^c(\textbf{x})+c_1h_1(\textbf{x})+c_2h_2(\textbf{x}),
791: \end{equation}
792: where
793: $$h_1=\Psi\bigr[D_1,(D_2\cup D_3)\bigr],\ h_2=\Psi\bigr[(D_1\cup D_2),D_3\bigr],$$
794: and
795: \begin{equation}\label{def:c}
796: \left(\begin{array}{c}
797: \ds c_1\\
798: \ds c_2
799: \end{array}\right)
800: =\ds-\left( \begin{array}{cc}
801: \ds -1 &\ds\int_{\p D_1}\p_\nu h_2\ d S\\
802: \ds \int_{\p D_2}\p_\nu h_1\ dS&\ds\int_{\p D_2}\p_\nu h_2\ d S \end{array} \right)^{-1}
803: \left(\begin{array}{c}
804: \ds\int_{\p D_1}\p_\nu H^c\ dS\\
805: \ds\int_{\p D_2}\p_\nu H^c\ dS
806: \end{array}\right),
807: \end{equation}
808: where $\Psi$ is defined as \eqref{def:h} and \eqref{def:Psi}. The
809: equality \eqref{def:c} is from the integration of $\partial_{\nu } u
810: $ on $\p D_1$ and  $\p D_2$.
811: 
812: 
813:  Applying the upper bound on the gradient of  solution without the potential difference among the boundaries to conductivity equation derived in Bao et al.
814: \cite{BLY}, we can show that $\nabla H^c$ does not blow-up (also
815: refer to \cite{Y}). Using Lemma \ref {lem:B} in the following
816: section, we have $\int_{\p D_2}\p_\nu h_1\ dS = 1+  O(\sqrt
817: {\epsilon_1})$. This implies that
818: \begin{equation*}
819: \left(\begin{array}{c}
820: \ds c_1\\
821: \ds c_2
822: \end{array}\right)
823: \thickapprox\ds-\left( \begin{array}{cc}
824: \ds -1 &\ds 0 \\
825: \ds 1&\ds-1
826: \end{array} \right)^{-1} \left(\begin{array}{c}
827: \ds\int_{\p D_1}\p_\nu H^c\ dS\\
828: \ds\int_{\p D_2}\p_\nu H^c\ dS
829: \end{array}\right).
830: \end{equation*}
831: Thus, the coefficient $c_i$, $i=1,2$, is bounded independently of
832: $\epsilon_1$ and $\epsilon_2$. Therefore, the blow-up rate of
833: $\nabla u$ essentially relies on $\nabla h_i$. In this respect, we
834: consider the properties of $h_i$ in the following section.
835: 
836: \subsection{Properties of $h_1$ and $h_2$}
837: We build the optimal bounds of $u$ based on \eqref{rep:u3balls}; it
838: is essential to drive properties of $h_1$ and $h_2$ in the narrow
839: regions between inclusions. Let $h_1$ and $h_2$ be as follows:
840: $$h_1 = \Psi\bigr[D_1,(D_2\cup
841: D_3)\bigr]~\mbox{and}~h_2=\Psi\bigr[(D_1\cup D_2),D_3\bigr]. $$
842: 
843: 
844: 
845: 
846: \begin{prop}\label{prop:nablabound} There are the
847: following  estimates for $h_1$ and $h_2$:
848: \begin{itemize}
849: \item[(i)]  In the narrow region between $D_1$ and $D_2$, we have
850: $$\nabla h_1=O\Bigr(\frac{1}{\sqrt{\epsilon_1}}\Bigr)~\mbox{and}~\nabla h_2=O(\sqrt{\epsilon_2}).$$
851: \item[(ii)] In the narrow region between $D_2$ and $D_3$, we have
852: $$\nabla h_1=O(\sqrt{\epsilon_1})~\mbox{and}~\nabla h_2=O\Bigr(\frac{1}{\sqrt{\epsilon_2}}\Bigr).$$
853: \item[(iii)] $$h_1 |_{\partial D_2 \cup \partial D_3 } - h_1 |_{\partial D_1} \simeq \sqrt {\epsilon_1}$$ and $$  h_2 |_{\partial D_3} - h_2 |_{\partial D_1 \cup \partial D_2 } \simeq \sqrt {\epsilon_2}.$$
854: \end{itemize}
855: \end{prop}
856: 
857: \pf We consider $\nabla h_1$. By Lemma \ref{lem:A} and \ref{lem:C},
858: we have
859: 
860: $$0>
861: \partial_{\nu} h_1 \geq C\p_{\nu}\Psi[D_1, D_4]~\mbox{on}~\partial D_1$$
862: and
863: $$0 <
864: \partial_{\nu} h_1 \leq \partial_{\nu}
865: \Psi[D_1,D_2]~\mbox{on}~\partial D_2,$$ and by Lemma \ref{lem:B},
866: $$0 <
867: |\partial_{\nu} h_1| \leq C \sqrt {\epsilon_1}~\mbox{on}~\partial
868: D_3,$$ where $D_4$ is defined in Lemma \ref{lem:C}.  Without any
869: loss of generality, we assume that
870: $$\left(-\frac {\epsilon_1} 2, 0\right) \in \p D_1,~\left(\frac {\epsilon_1} 2, 0\right) \in \p D_2~\mbox{and}~\mbox{dist}(D_1, D_2) =\epsilon_1.$$
871: Let
872: $$p(\mathbf{x})= \log |\mathbf{x} - (\sqrt {\epsilon_1},0)| - \log |\mathbf{x} + (\sqrt {\epsilon_1},0) |. $$
873: Referring to the inequality (9) in \cite{Y2}, there is a constant
874: $C_1$ such that
875: $$0< |\nabla h_1| \leq C_1 |\nabla p|~\mbox{on}~\partial (D_1 \cup D_2 \cup D_3).$$
876: Regarding $(x_1,x_2)$ as a complex number $z=x_1 + x_2 i$, we
877: consider
878: $$\rho (z) = \frac {\p_1 h_1(z) - \p_2 h_1(z) i }{C_1 (\p_1 p(z) - \p_2 p(z) i) }.$$
879: Then, $\rho(z)$ can be extended to $\infty$ as an analytic function.
880: From definition, $|\rho(z)| < 1 $ on $\partial D_1 \cup \partial D_2
881: \cup\partial D_3 $. By the maximum principle,  $$|\rho(z)| < 1
882: ~\mbox{in}~  \mathbb{C} \setminus (D_1 \cup  D_2 \cup D_3).$$ Thus,
883: we have
884: $$|\nabla h | \leq C_1 |\nabla p | ~\mbox{in}~\mathbb{R}^2 \setminus (D_1 \cup  D_2 \cup D_3). $$
885: Therefore, $\nabla h_1=O\Bigr(\frac{1}{\sqrt{\epsilon_1}}\Bigr)$ in
886: the narrow region between $D_1$ and $D_2$, and $\nabla
887: h_1=O(\sqrt{\epsilon_1})$ in the narrow region between $D_2$ and
888: $D_3$. Similarly, we have $\nabla
889: h_2=O\Bigr(\frac{1}{\sqrt{\epsilon_2}}\Bigr)$ in the narrow region
890: between $D_2$ and $D_3$, and $\nabla h_2=O(\sqrt{\epsilon_2})$ in
891: the narrow region between $D_1$ and $D_2$. We have proven (i) and
892: (ii).
893: \par The estimate (iii) is presented by Lemma \ref{lem:C}.
894: 
895: \qed
896: 
897: 
898: \begin{lem}\label{lem:A} We have the following properties:
899: \begin{itemize}
900: \item[(i)]$$0< h_1 |_{\partial D_2} - h_1 |_{\partial D_1} \leq \Psi[D_1, D_2]\Big|_{\partial D_2} - \Psi[D_1, D_2]\Big|_{\partial D_1}.$$
901: \item[(ii)] $$0 < \partial_{\nu} h_1 \leq \partial_{\nu} \Psi[D_1,D_2]~\mbox{on}~\partial D_2.$$
902: \end{itemize}
903: \end{lem}
904: \pf Let $$M=\frac{\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p
905: D_1}-\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p D_2\cup\p
906: D_3}}{\Psi[D_1,D_2]\Bigr|_{\p D_1}-\Psi[D_1,D_2]\Bigr|_{\p D_2}},$$
907: and
908: \begin{align*}\ds G(\mathbf{x})&=\Psi[D_1,(D_2\cup D_3)](\mathbf{x})-\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p D_2\cup \p D_3}\\
909: &\qquad\qquad\ds-M\left(\Psi[D_1,D_2](\mathbf{x})-\Psi[D_1,D_2]\Bigr|_{\p
910: D_2}\right).\end{align*} Then, $G= 0$ on $\p D_1\cup \p D_2$, and
911: $G>0$ on $\p D_3$. By Hopf's lemma, $$\p_\nu G<0\quad\mbox{on }\p
912: D_1.$$ This means that \be\partial_{\nu} h_1 \leq M
913: \partial_{\nu} \Psi[D_1, D_2]~\mbox{on}~\partial D_1.\label{eq:3}\ee
914:  Note that $h_1 = \Psi[D_1,(D_2\cup
915: D_3)]$. By integrating $G$ on $\p D_1$, we have the inequality (i).
916: 
917: 
918: \par On the other hand, by Hopf's lemma, $$\p_\nu G<0\quad\mbox{on
919: }\p D_2.$$  This means that
920: $$\partial_{\nu} h_1 \leq M \partial_{\nu} \Psi[D_1, D_2]~\mbox{on}~\partial D_2.$$
921: From the inequality (i), $M<1$. Therefore, we have (ii).
922: 
923: \qed
924: 
925: 
926: 
927: \begin{lem} \label{lem:B} There is a constant $C$ such that
928: $$0\leq \partial_{\nu} h_1 \leq C \sqrt \epsilon_1 ~\mbox{on}~\partial D_3.$$
929: \end{lem}
930: \pf We use the method similar to Lemma \ref{lem:A}. Let
931: $$M=\frac{\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p D_1}-\Psi[D_1,(D_2\cup
932: D_3)]\Bigr|_{\p D_2\cup\p D_3}}{\Psi[D_1,D_3]\Bigr|_{\p
933: D_1}-\Psi[D_1,D_3]\Bigr|_{\p D_3}},$$ and
934: \begin{align*}\ds G(\mathbf{x})&=\Psi[D_1,(D_2\cup D_3)](\mathbf{x})-\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p D_2\cup \p D_3}\\
935: &\qquad\qquad\ds-M\left(\Psi[D_1,D_3](\mathbf{x})-\Psi[D_1,D_3]\Bigr|_{\p
936: D_3}\right).\end{align*} Then, $G= 0$ on $\p D_1\cup \p D_3$, and
937: $G>0$ on $\p D_2$. By Hopf's lemma, $$\p_\nu G<0\quad\mbox{on }\p
938: D_3.$$ Since $h_1 = \Psi[D_1,(D_2\cup D_3)]$, this inequality means
939: that
940: $$0\leq \partial_{\nu} h_1 \leq M \p_{\nu} \Psi[D_1,D_3]~\mbox{on}~\p D_3.$$
941: 
942: \par Now, we estimate the gradient of $M \Psi[D_1,D_3]$. To do so,
943: we consider the potential difference between $\p D_1$ and $\p D_3$
944: as follows:
945: \begin{align*}
946: M \Psi[D_1,D_3]\Big|_{\partial D_3} -M \Psi[D_1,D_3]\Big|_{\partial
947: D_1} &= h |_{\partial D_3} -
948: h|_{\partial D_1}\\&=h |_{\partial D_2} - h|_{\partial D_1}\\
949: &\leq \Psi[D_1,D_2]\Big|_{\partial D_2} -
950: \Psi[D_1,D_2]\Big|_{\partial D_1}\\
951: &\leq C \sqrt {\epsilon_1}\end{align*} The last inequality above was
952: proven by Yun in his paper \cite{Y,Y2}, since $\Psi[D_1,D_2]$ is
953: only for two domains. Note that $D_3$ is not close to $D_2$. Owing
954: to the method in Bao et al. \cite{BLY}, we have
955: $$\norm {\p_{\nu} M \Psi[D_1,D_3] }_{L^{\infty} (\partial D_3)} \leq C \sqrt {\epsilon_1}.$$
956: Therefore, we can obtain the result.
957:  \qed
958: 
959: 
960: 
961: \begin{lem} \label{lem:C} Let $D_4$ is a disk containing $D_2$ and
962: $D_3$ with $$\mbox{dist}(D_1,D_4) = \mbox{dist}(D_1,D_2).$$
963: \begin{itemize}
964: \item[(i)]There is a positive constant $C$ such that $$0> \partial_{\nu} h_1 \geq C\partial_{\nu} \Psi[D_1, D_4]~\mbox{on}~\partial D_1.$$
965: \item[(ii)] $$h_1 |_{\partial D_2} - h_1 |_{\partial D_1} \geq \Psi[D_1, D_4]\Big|_{\partial D_4} - \Psi[D_1, D_4]\Big|_{\partial D_1}.$$
966: 
967: \item[(iii)] $$h_1 |_{\partial D_2 \cup \partial D_3 } - h_1 |_{\partial D_1} \simeq \sqrt {\epsilon_1}.$$
968: \end{itemize}
969: \end{lem}
970: \pf To prove (i) and (ii), we use the same derivation to Lemma
971: \ref{lem:A}. So, we set $$M=\frac{\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p
972: D_1}-\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p D_2\cup\p
973: D_3}}{\Psi[D_1,D_4]\Bigr|_{\p D_1}-\Psi[D_1,D_4]\Bigr|_{\p D_4}},$$
974: and
975: \begin{align*}\ds G(\mathbf{x})&=\Psi[D_1,(D_2\cup D_3)](\mathbf{x})-\Psi[D_1,(D_2\cup D_3)]\Bigr|_{\p D_2\cup \p D_3}\\
976: &\qquad\qquad\ds-M\left(\Psi[D_1,D_4](\mathbf{x})-\Psi[D_1,D_4]\Bigr|_{\p
977: D_4}\right).\end{align*} Then $G \Big|_{\p D_1}= 0$ and $G \leq 0$
978: on ${\p D_4}$. By Hopf's lemma, we have
979: $$\p_{\nu}G >0~\mbox{on}~\p D_1.$$
980: By the integration on $\p D_1$, we have (ii) and $M<1$. Therefore,
981: the inequality $\p_{\nu}G >0$ can also yield (i).
982: 
983: \par From (i) of Lemma \ref{lem:A} and (ii) in this lemma, we have
984: $$ h_1 |_{\partial D_2} - h_1 |_{\partial D_1} \geq \Psi[D_1,
985: D_4]\Big|_{\partial D_4} - \Psi[D_1, D_4]\Big|_{\partial D_1}$$ and
986: $$ h_1 |_{\partial D_2} - h_1 |_{\partial D_1} \leq \Psi[D_1,
987: D_2]\Big|_{\partial D_2} - \Psi[D_1, D_2]\Big|_{\partial D_1}.$$ The
988: potential $\Psi[D_1, D_i] ~(i=1,4)$ is only for two domains and
989: thus, its difference between $D_1$ and $D_i$ ($i=1,2$) was already
990: estimated in Yun \cite{Y, Y2} as follows: for $i=1,2$,
991: $$\Psi[D_1,
992: D_i]\Big|_{\partial D_i} - \Psi[D_1, D_i]\Big|_{\partial D_1} \simeq
993: \sqrt {\epsilon_1}.$$ Therefore, we have (iii). \qed
994: 
995: 
996: 
997: 
998: \begin{lem}\label{sum:Hnuh}
999: We have
1000: $$\left|\int_{\cup_{i=1}^3\p D_i}H\p_\nu h_1\ dS\right|\leq C\sqrt{\epsilon_1}.$$
1001: \end{lem}
1002: \pf Without loss of generality, we assume that
1003: $$\left(-\frac {\epsilon_1} 2, 0\right) \in \p D_1,~\left(\frac {\epsilon_1} 2, 0\right) \in \p D_2,~\mbox{dist}(D_1, D_2) =\epsilon_1~\mbox{and}~(-1,0) \in D_2.$$
1004: We consider $\widetilde{H}$ as follows:
1005: $$\widetilde{H}= H - \p_{2} H(0,0) \frac {x_2}{|\mathbf{x}-(1,0)|^2}.  $$
1006: It follows from the Divergence Theorem that
1007: $$\int_{\p D_1 \cup \p D_2\cup \p D_3} \frac {x_2}{|\mathbf{x}-(1,0)|^2} \p_{\nu} h dS=\int_{\p D_1 \cup \p D_2 \cup \p D_3} \p_{\nu}\left( \frac {x_2}{|\mathbf{x}-(1,0)|^2}\right)  h ds =0,$$
1008: since $\frac {x_2}{|\mathbf{x}-(1,0)|^2} = O (|x|^{-1})$ as
1009: $|x|\rightarrow \infty$. Hence, we have
1010: $$\int_{\cup_{i=1}^3\p D_i}H\p_\nu h_1\ dS = \int_{\p D_1 \cup\p D_2}\widetilde{H}\p_\nu h_1\ dS + \int_{\p D_3}\widetilde{H}\p_\nu h_1\ dS.$$
1011: 
1012: 
1013: \par We first consider $\int_{\p D_1 \cup\p D_2}\widetilde{H}\p_\nu h_1\
1014: dS$. By Lemma \ref{lem:A} and \ref{lem:C}, we have $$0>
1015: \partial_{\nu} h_1 \geq C\p_{\nu}\Psi[D_1, D_4]~\mbox{on}~\partial D_1$$
1016: and
1017: $$0 <
1018: \partial_{\nu} h_1 \leq \partial_{\nu}
1019: \Psi[D_1,D_2]~\mbox{on}~\partial D_2.$$ From definition, $\p_{2}
1020: \widetilde{H} =0$. Hence, we can use Lemma 3.2 in \cite{Y2} so that
1021: $$\left|\int_{\p D_1}\widetilde{H}\p_\nu h_1\
1022: dS \right| \leq C \int_{\p D_1}\left|\widetilde{H}\Psi[D_1,
1023: D_4]\right|\ dS \leq C \sqrt {\epsilon_1}$$ and
1024: $$\left|\int_{\p D_2}\widetilde{H}\p_\nu h_1\
1025: dS \right| \leq  \int_{\p D_2}\left|\widetilde{H}\Psi[D_1,
1026: D_2]\right|\ dS \leq C \sqrt {\epsilon_1}.$$ We second consider
1027: $\int_{\p D_3}\widetilde{H}\p_\nu h_1\ dS$. By Lemma \ref{lem:B}, we
1028: can have
1029: $$\left|\int_{\p D_3}\widetilde{H}\p_\nu h_1\
1030: dS\right| \leq C \sqrt {\epsilon_1}.$$ Therefore, we have done it.
1031: 
1032: \qed
1033: 
1034: \begin{rem}
1035: We draw attention of readers to the independent work of Bao, Li and
1036: Yin in \cite{Bao_thesis} and \cite{BLY2}.  Bao et al. have shown
1037: that  the  blow-up rate know only for a pair of inclusion is still
1038: valid to the multiple inclusions cases. As a byproduct of our work,
1039: the blow-up rate of the gradient for three inclusions is established
1040: in Theorem \ref{threedomains}.
1041: \end{rem}
1042: 
1043: 
1044: \begin{thm}\label{threedomains}
1045: Let $D_1$, $D_2$ and $D_3$ be as assumed in the beginning of Section
1046: \ref{section:threedomains}. Note that $D_2$ is not assumed to be
1047: smaller than the others.
1048: \begin{itemize}
1049: \item[\rm(i)] Optimal upper bounds: For any entire harmonic function $H(x_1,x_2)$, we have the following:
1050: in the narrow region between $D_1\cup D_2$, $$|\nabla u|\leq C\frac{1}{\sqrt{\epsilon_1}},$$
1051: and, in the narrow region between $D_2\cup D_3$, $$|\nabla u|\leq C\frac{1}{\sqrt{\epsilon_2}}.$$
1052: \item[\rm(ii)] Existence of blow-up: Without loss of generality, we assume that
1053: $$\left(-\frac {\epsilon_1} 2, 0\right) \in \p D_1,~\left(\frac {\epsilon_1} 2, 0\right) \in \p D_2 ~\mbox{and}~\mbox{dist}(D_1, D_2) =\epsilon_1.$$
1054: For $H(x_1,x_2)=x_1$, there exist $\mathbf{x}_0$  in the narrow
1055: region between $D_1$ and $D_2$ such that
1056: $$|\nabla u(\mathbf{x}_0)|\geq C\frac{1}{\sqrt{\epsilon_1}},$$
1057: and, similarly, there is a linear function $H (x_1,x_2)$  with
1058: $\mathbf{y}_0$ between $D_2$ and $D_3$ such that
1059: $$|\nabla u(\mathbf{y}_0)|\geq C\frac{1}{\sqrt{\epsilon_2}}.$$
1060: \end{itemize}
1061: \end{thm}
1062: 
1063: \pf From Subsection \ref{subsec}, we have a representation
1064: \eqref{def:c} for $u$ and the coefficient $c_i$, $i=1,2$, is bounded
1065: independently of $\epsilon_1$ and $\epsilon_2$. Proposition
1066: \ref{prop:nablabound} yields the upper bound of Theorem
1067: \ref{threedomains}.
1068: 
1069: \par Now, we consider the existence of the blow-up. Using the result
1070: of Subsection \ref{subsec} again, we have a constant $C$ independent
1071: of $\epsilon$ such that
1072: $$\ds\norm{u}_{L^{\infty} (\cup_{i=1}^3 \p D_i)} \leq C \norm{H}_{L^{\infty} (\cup_{i=1}^3 D_i)}.$$
1073: Applying the Green's identity to $\int_{\cup_{i=1}^3\p D_i}u\p_\nu
1074: h_1\ dS$, we have
1075: \begin{align}
1076: \ds&\int_{\cup_{i=1}^3\p D_i}H\p_\nu h_1\ dS = \int_{\cup_{i=1}^3\p
1077: D_i}u\p_\nu h_1\ dS\nonumber\\\nonumber &=-u\bigr|_{\p
1078: D_1}+u\bigr|_{\p D_2}\bigr(\int_{\p D_2}\p_\nu h_1\
1079: dS\bigr)+u\bigr|_{\p D_3}\bigr(\int_{\p D_3}\p_\nu h_1\
1080: dS\bigr)\\\label{eqn:thmC} \ds&=-u\bigr|_{\p D_1}+u\bigr|_{\p
1081: D_2}\bigr(1-\int_{\p D_3}\p_\nu h_1\ dS\bigr)+u\bigr|_{\p
1082: D_3}\bigr(\int_{\p D_3}\p_\nu h_1\ dS\bigr).
1083: \end{align}
1084: By Lemma \ref {lem:B}, we have
1085: \begin{align}
1086: \ds\label{diff:D1D2}u\bigr|_{\p D_2}-u\bigr|_{\p D_1}\leq C
1087: \sqrt{\epsilon_1},\end{align} where the constant $C$ above depends
1088: on $\|H\|_{L^\infty(\cup_{i=1}^3\p D_i)}$. Similarly, we have
1089: \begin{equation}\label{diff:D2D3}u\bigr|_{\p D_3}-u\bigr|_{\p D_2}\leq C \sqrt{\epsilon_2}.\end{equation}
1090: Using \eqref{eqn:thmC} again, we have
1091: \begin{align} u\bigr|_{\p D_2}-u\bigr|_{\p D_1} + O (\sqrt {\epsilon_1
1092: \epsilon_2})&= \int_{\cup_{i=1}^3\p D_i}H\p_\nu h_1\ dS
1093: \notag\\&\geq\int_{\p D_1}H\p_\nu h_1\ dS. \label{eq:4}
1094: \end{align}
1095: The last inequality can be derived from the fact that $H >0$ on $\p
1096: D_2 \cup \p D_3$.
1097: 
1098: 
1099:  To get the last inequality above, we
1100: took an advantage of $H=x_1$. By \eqref{eq:3}, $$  \partial_{\nu} h
1101: \leq \left(\frac{ h_1\Bigr|_{\p D_1}-h_1\Bigr|_{\p D_2\cup\p
1102: D_3}}{\Psi[D_1,D_2]\Bigr|_{\p D_1}-\Psi[D_1,D_2]\Bigr|_{\p
1103: D_2}}\right) \partial_{\nu} \Psi[D_1,D_2]<0~\mbox{on}~\partial D_1.
1104: $$ By (iii) in Proposition \ref{prop:nablabound}, we have
1105: $$ h_1\Bigr|_{\p D_1}-h_1\Bigr|_{\p D_2\cup\p
1106: D_3} \backsimeq \sqrt {\epsilon_1}$$and$${\Psi[D_1,D_2]\Bigr|_{\p
1107: D_1}-\Psi[D_1,D_2]\Bigr|_{\p D_2}}\backsimeq \sqrt {\epsilon_1}.$$
1108: The inequality \eqref{eq:4} implies
1109: $$ u\Bigr|_{\p D_2}-u\Bigr|_{\p D_1} \gtrsim \sqrt {\epsilon}.$$
1110: By the Mean Value Theorem, we have the desirable lower bound in the
1111: narrow region between $D_1$ and $D_2$. Similarly, we can also obtain
1112: the other lower bound.
1113:  \qed
1114: 
1115: \subsection{Proof of Theorem \ref{thm:B}}
1116: We derive the optimal bounds of the gradient of the solution to \eqref{eq:conductivity}, when there are adjacent three disks:
1117:  \begin{equation}\label{def:Bi}D_l = B_{r_l}(\mathbf{c}_l),\
1118: l=1,2,3,\end{equation}  where $\mathbf{c}_1=(-r_1 - \frac
1119: {\epsilon_1} 2, 0)$, $\mathbf{c}_2=(r_2+ \frac {\epsilon_1} 2,0)$
1120: and $\mathbf{c}_3=(r_3+r_2+\frac {\epsilon_1} 2 + \epsilon_2,0)$. As
1121: defined before, $h_1 = \Psi [D_1, (D_2\cup D_3)]$. Let $w_1 =\Psi
1122: [D_1, D_2] $.
1123: 
1124: \par We begin the proof by showing  that
1125: \be w_1 \big|_{\partial D_2} -  w_1 \big|_{\partial D_1} \simeq h_1
1126: \big|_{\partial D_2} -  h_1 \big|_{\partial D_1} .\label{eq:5}\ee By
1127: the monotonic property of Lemma \ref{lem:mono}, we have
1128: $$ h_1 \big|_{\partial D_2} -  h_1 \big|_{\partial D_1} \leq
1129: w_1 \big|_{\partial D_2} -  w_1 \big|_{\partial D_1}.$$ Considering
1130: $$h_1 - \left({\frac {h_1 \big|_{\partial D_2} -  h_1 \big|_{\partial D_1} }{w_1 \big|_{\partial D_2} -  w_1 \big|_{\partial D_1}}}\right)w_1,$$
1131: we can obtain, from the Hopf's Lemma,
1132: $$\int_{\p D_2} \p_{\nu}h_1 dS \leq \left({\frac {h_1 \big|_{\partial D_2} -  h_1 \big|_{\partial D_1} }{w_1 \big|_{\partial D_2} -  w_1 \big|_{\partial D_1}}}\right)\int_{\p D_2} \p_{\nu}w_1 dS.$$
1133: By Lemma \ref{lem:B}, we have
1134: $$ \int_{\p D_3} \p_{\nu}h_1 dS = O(\sqrt {\epsilon_1}).$$
1135: Since $\int_{\p D_2 \cup \p D_3} \p_{\nu}h_1 dS =1$, we have
1136: $$ \left(w_1 \big|_{\partial D_2} -  w_1 \big|_{\partial D_1}\right)
1137: (1+ O(\sqrt {\epsilon_1})) \leq h_1 \big|_{\partial D_2} -  h_1
1138: \big|_{\partial D_1} .$$ Therefore, we can obtain \eqref{eq:5}.
1139: Owing to the estimate for $w_1 \big|_{\partial D_2} -  w_1
1140: \big|_{\partial D_1}$ in Lemma \ref{lemma:twodisks}, we have
1141: 
1142: \be h_1 \big|_{\partial D_2} -  h_1 \big|_{\partial D_1} \simeq
1143: \sqrt {\frac {r_1 + r_2}{r_1 r_2}} \sqrt
1144: {\epsilon_1}.\label{eq:6}\ee
1145: 
1146: 
1147: \par Let $w_2 =\Psi[D_1, D_3] $. Considering
1148: 
1149: $$h_1 - \left({\frac {h_1 \big|_{\partial D_3} -  h_1 \big|_{\partial D_1} }{w_2 \big|_{\partial D_3} -  w_2 \big|_{\partial D_1}}}\right)w_2,$$
1150:  from the Hopf's Lemma, we obtain
1151: $$\p_{\nu} h_1 \leq  \left({\frac {h_1 \big|_{\partial D_3} -  h_1 \big|_{\partial D_1} }{w_2 \big|_{\partial D_3} -  w_2 \big|_{\partial D_1}}}\right)\p_{\nu}w_2 \leq 0~\mbox{on}~\partial D_1.$$
1152: Here, we estimate the coefficient in the right hand side. Note that
1153: $h_1 \big|_{\partial D_2} =  h_1 \big|_{\partial D_3} $. Thus, we
1154: have $$h_1 \big|_{\partial D_3} -  h_1 \big|_{\partial D_1} \simeq
1155: \sqrt {\frac {r_1 + r_2}{r_1 r_2}} \sqrt {\epsilon_1}.$$ Since $r_2
1156: \ll r_1$ and $r_2 \ll r_3$, we also have
1157: $$w_2 \big|_{\partial D_3} -  w_2 \big|_{\partial D_1} \simeq
1158: \sqrt {\frac {r_1 + r_3}{r_1 r_3}} \sqrt {r_2}.$$ This implies that
1159: $$\p_{\nu} h_1 \lesssim  \frac {\sqrt {\frac {r_1 + r_2}{r_1 r_2}} \sqrt {\epsilon_1}}  {\sqrt {\frac {r_1 + r_3}{r_1 r_3}} \sqrt {r_2}} \p_{\nu} w_2 \leq 0~\mbox{on}~\p D_1.$$
1160: Therefore, we have
1161: \begin{align}
1162: \int_{\p D_1} H \p_{\nu} h_1 dS &\gtrsim \sqrt {\frac {r_1 +
1163: r_2}{r_1 + r_3}} \frac {\sqrt {r_3}}{r_2} \sqrt {\epsilon_1}
1164: \int_{\partial D_1} H \p_{\nu} w_2 dS\notag\\&\gtrsim
1165:  \sqrt {\frac {r_1 +
1166: r_2}{r_1 + r_3}} \frac {\sqrt {r_3}}{r_2} \sqrt {\epsilon_1}
1167: \sqrt{\frac {r_1 r_3}{r_1 + r_3}} \sqrt {r_2}\notag\\
1168: &\geq \frac {r_1 r_3}{r_1 + r_3} \frac 1 {\sqrt {r_2}}
1169: \sqrt{\epsilon_1} \geq 0 \label{eq:8}.
1170: \end{align}
1171: Owing to \eqref{eqn:thmC}, \eqref{diff:D1D2} and \eqref{diff:D2D3},
1172: we have
1173: \begin{align}
1174: \int_{\cup_{i=1}^3\p D_i}u \p_\nu h_1\ dS =&\left(1- \int_{\p
1175: D_3}\p_\nu h_1\ dS\right)\left( u\bigr|_{\p D_2} -u\bigr|_{\p
1176: D_1}\right)\notag\\&~~~~~+  \left( \int_{\p D_3}\p_\nu h_1\ dS
1177: \right) \left( u\bigr|_{\p D_3} -u\bigr|_{\p
1178: D_1}\right) \notag\\
1179: =& \left(1- O (\sqrt{\epsilon_1})\right)\left( u\bigr|_{\p D_2}
1180: -u\bigr|_{\p D_1}\right)+  O (\sqrt{\epsilon_1})\left( O
1181: (\sqrt{\epsilon_1}) + O (\sqrt{\epsilon_2})\right). \notag
1182: \end{align}
1183: Therefore, we have
1184: \begin{align}
1185: u\bigr|_{\p D_2} -u\bigr|_{\p D_1} & \geq  \frac 1 2
1186: \int_{\cup_{i=1}^3\p D_i}u \p_\nu h_1\ dS + O
1187: (\sqrt{\epsilon_1})\left( O (\sqrt{\epsilon_1}) + O
1188: (\sqrt{\epsilon_2})\right) \notag\\
1189: &=  \frac 1 2 \int_{\cup_{i=1}^3\p D_i} H \p_\nu h_1\ dS + O
1190: (\sqrt{\epsilon_1})\left( O (\sqrt{\epsilon_1}) + O
1191: (\sqrt{\epsilon_2})\right) \notag\\
1192: &\geq \frac 1 2 \int_{\p D_1} H \p_\nu h_1\ dS + O
1193: (\sqrt{\epsilon_1})\left( O (\sqrt{\epsilon_1}) + O
1194: (\sqrt{\epsilon_2})\right) \notag\\
1195: &\geq C {\frac {r_1 r_3}{r_1 + r_3}} {\frac 1 {\sqrt {r_2}}} \sqrt
1196: {\epsilon_1} + O (\sqrt{\epsilon_1})\left( O (\sqrt{\epsilon_1}) + O
1197: (\sqrt{\epsilon_2})\right).\notag
1198: \end{align}
1199: Therefore, we have completed the proof. \qed
1200: 
1201: \subsection{Proof of Theorem \ref{thm:D}}
1202: We pursuit the proof of Theorem \ref{thm:B}, taking an advantage of
1203: the monotonic property of Lemma \ref{lem:mono}. The domains $D_1$,
1204: $D_2$ and $D_3$ are as assumed in Theorem \ref{thm:D}. As assumed
1205: before, $h_1 =\Psi [D_1, (D_2\cup D_3)]  $. Let $w_1 =\Psi [D_1,
1206: D_2] $. By the same way as Theorem \ref{thm:B}, we have $$ w_1
1207: \big|_{\partial D_2} -  w_1 \big|_{\partial D_1} \simeq h_1
1208: \big|_{\partial D_2} - h_1 \big|_{\partial D_1} .$$ Here, we use the
1209: monotonic property of Lemma \ref{lem:mono} to estimate the
1210: difference between domains. Choosing two pairs of proper disks
1211: containing $D_1$ and $D_2$, and contained $D_1$ and $D_2$,
1212: respectively, we can obtain $$ h_1 \big|_{\partial D_2} - h_1
1213: \big|_{\partial D_1} \simeq \sqrt{\frac {\epsilon_1}{ { r_2}}}$$
1214: under the assumption that $r_2$ is small.
1215: \par  Let $w_2 =\Psi[D_1, D_3] $. Choosing two
1216: pairs of proper disks containing $D_1$ and $D_3$, and contained
1217: $D_1$ and $D_3$, respectively, Then, we have
1218: $$w_2 \big|_{\partial D_3} -  w_2 \big|_{\partial D_1} \simeq
1219: \sqrt {r_2}.$$ By the same argument as Theorem \ref{thm:B}, we have
1220: $$
1221: \int_{\p D_1} H \p_{\nu} h_1 dS \gtrsim \sqrt {\frac{\epsilon_1}
1222: {r_1}} \geq 0.
1223: $$
1224: Note that $D_1 \subset \mathbb{R}_{-}\times \mathbb {R}$ and
1225: $D_2\cup D_3 \subset \mathbb{R}_{+}\times \mathbb {R}.$ Continuing
1226: to follow the proof of Theorem \ref{thm:B}, we can obtain
1227: $$
1228: u\bigr|_{\p D_2} -u\bigr|_{\p D_1}  \geq C  \sqrt {\frac
1229: {\epsilon_1}{r_1}} + O (\sqrt{\epsilon_1})\left( O
1230: (\sqrt{\epsilon_1}) + O (\sqrt{\epsilon_2})\right).\notag
1231: $$
1232: Therefore, we have done the proof.
1233: 
1234: \subsection {Derivation for the optimal upper bounds}\label{upp}
1235: We consider the optimal upper bounds presented in Theorem
1236: \ref{thm:A}, \ref{thm:B}, \ref{thm:C} and \ref{thm:D}.  These proofs
1237: have essential thing in common. In this respect, we prove only the
1238: optimal upper bound presented in Theorem \ref{thm:B}. As have
1239: assumed them before, we set \begin{align*} h_1 &=\Psi [D_1, D2 \cup D_3]\\
1240:  h_2 &=\Psi [D_1, D_2]\\
1241:   h_3 &=\Psi [D_1, D_3]\\
1242:    h_4 &=\Psi [D_1, D_4].
1243: \end{align*}
1244: Here, the domain $D_4$ is given in Lemma \ref{lem:C}, which is a
1245: disk containing $D_2$ and $D_3$ with
1246: $$\mbox{dist}(D_1,D_4) = \mbox{dist}(D_1,D_2),$$
1247: and the diameter of $D_4$ is in proportion as $r_3$, because $r_2$
1248: is sufficiently small. Then, we compare $ h_1$ with $h_2$, $h_3$ and
1249: $h_4$. The proof of Lemma \ref{lem:A} contains
1250: $$0 \leq \partial_{\nu} h_1 \leq \left(\frac{ h_1\Bigr|_{\p (D_2 \cup D_3)}-h_1\Bigr|_{\p D_1}}{h_2\Bigr|_{\p D_2 }-h_2\Bigr|_{\p
1251: D_1}}\right) \partial_{\nu}h_2~\mbox{on}~\p D_2,
1252: $$ the proof of Lemma \ref{lem:B} yields
1253: $$0 \leq \partial_{\nu} h_1 \leq \left(\frac{ h_1\Bigr|_{\p (D_2 \cup D_3)}-h_1\Bigr|_{\p D_1}}{h_3\Bigr|_{\p D_3 }-h_3\Bigr|_{\p
1254: D_1}}\right) \partial_{\nu}h_3~\mbox{on}~\p D_3
1255: $$ and the proof of Lemma \ref{lem:C} implies
1256: $$0 \leq -\partial_{\nu} h_1 \leq -\left(\frac{ h_1\Bigr|_{\p (D_2 \cup D_3)}-h_1\Bigr|_{\p D_1}}{h_4\Bigr|_{\p D_4 }-h_4\Bigr|_{\p
1257: D_1}}\right) \partial_{\nu}h_3~\mbox{on}~\p D_3.
1258: $$ In the same way as Lemma \ref {sum:Hnuh}, we can consider $\widetilde{H}$ by choosing the point in $D_3$. In this respect, without any loss of generality,
1259: we can assume that
1260: $$\partial_{x_2} H (0,0) = 0.$$ The reason why we assumed above is
1261: because the integration representation for the potential difference
1262: is not good enough, refer to \cite{Y2}. The geometrical assumption
1263: of Case (B) implies that $D_1$ and $D_2 \cup D_3$ are separated by
1264: $x_1 = 0$ and they are approaching to $(0,0)$.
1265: \par Therefore, by the proof of Theorem \ref{thm:B} and Lemma \ref{lem++}, we have
1266: \begin{align*}
1267: \left| u\bigr|_{\p D_2} -u\bigr|_{\p D_1}\right| &+ O
1268: (\sqrt{\epsilon_1})\left( O
1269: (\sqrt{\epsilon_1}) + O (\sqrt{\epsilon_2})\right)\\
1270: &= \left| \int_{\partial (\cup_{i=1} ^3 D_i)} H \partial_{\nu} h_1
1271: dS \right|\\
1272: &\lesssim \left(\frac{ h_1\Bigr|_{\p (D_2 \cup D_3)}-h_1\Bigr|_{\p
1273: D_1}}{h_2\Bigr|_{\p D_2 }-h_2\Bigr|_{\p D_1}}\right)
1274: \sqrt{\frac{r_1r_2}{r_1+r_2}\epsilon_1}\\
1275: &~~+ \left(\frac{ h_1\Bigr|_{\p (D_2 \cup D_3)}-h_1\Bigr|_{\p
1276: D_1}}{h_3\Bigr|_{\p D_3 }-h_3\Bigr|_{\p D_1}}\right)
1277: \sqrt{\frac{r_1r_3}{r_1+r_3}r_2}\\&~~ +\left(\frac{ h_1\Bigr|_{\p
1278: (D_2 \cup D_3)}-h_1\Bigr|_{\p D_1}}{h_4\Bigr|_{\p D_4
1279: }-h_4\Bigr|_{\p D_1}}\right) \sqrt{\frac{r_1r_3}{r_1+r_3}\epsilon_1}
1280: \end{align*} and
1281: $$ { h_1\Bigr|_{\p (D_2 \cup D_3)}-h_1\Bigr|_{\p
1282: D_1}} \thickapprox{h_2\Bigr|_{\p D_2 }-h_2\Bigr|_{\p D_1}}.$$ Here,
1283: note that the radius of $D_4$ can be choosen between $\frac 3 2 r_3$
1284: and $2 r_3$. Lemma \ref{lemma:twodisks} implies that $$\left|
1285: u\bigr|_{\p D_2} -u\bigr|_{\p D_1}\right| \lesssim \frac {r_1 r_3}{
1286: r_1 + r_3 } \frac 1 {\sqrt {r_2}} {\sqrt {\epsilon_1}}.$$ Therefore,
1287: we establish the optimal upper bound for $\left| u\bigr|_{\p D_2}
1288: -u\bigr|_{\p D_1}\right| $.
1289: 
1290: \par Based on this, the optimal upper bound on the
1291: gradient of $u$ in the narrow region be obtained. Here, the main
1292: idea to get the gradient estimate from the potential difference has
1293: already been presented by Bao et al. (Theorem 1.3, Lemma 2.2 and 2.3
1294: in \cite{BLY}), and has been modified to fit our problem by Lim and
1295: Yun in \cite{LY}. Thurs, we give a brief description on the method.
1296: We choose a large domain $D_0$ containing $D_1$, $D_2$ and $D_3$,
1297: where $\p D_0$ is at a sufficient distance from $D_1$, $D_2$ and
1298: $D_3$. Then, $u$ can be decomposed as follows:
1299: $$u=C_0 + v_0 + C_1 v_1  + C_3 v_3$$
1300: where  for $i=0,1,3$, $v_i$ is a harmonic function in $D_0 \setminus
1301: (D_1\cup D_2 \cup D_3)$ with the boundary data
1302: $$v_i = \delta_{0j} ~\mbox{on}~\partial D_j~\mbox{for}~i=1,~3$$
1303: and
1304: $$v_0 = \delta_{ij} u  ~\mbox{on}~\partial D_j$$
1305: for any $j=0,1,2,3$. Thus, the constants $C_1$ and $C_3$ keep
1306: $$ |C_1|  \lesssim \frac {r_1 r_3}{
1307: r_1 + r_3 } \frac 1 {\sqrt {r_2}}{\sqrt {\epsilon_1}}$$ and
1308: $$ |C_3|  \lesssim \frac {r_1 r_3}{
1309: r_1 + r_3 } \frac 1 {\sqrt {r_2}} {\sqrt {\epsilon_2}}.$$
1310: 
1311: 
1312: \par To estimate $\nabla v_0$, we consider a harmonic function $\rho$
1313: in $D_0 \setminus (D_1\cup D_2 \cup D_3)$ with the boundary data
1314: $$\rho = \delta_{0j}~ \mbox{on}~\partial D_j$$
1315: for  any $j=0,1,2,3$. By comparing with the harmonic function
1316: $\rho_{i}$ in $D_0 \setminus D_i$ with $\rho_{i} = 0$ on $\p D_i$
1317: and $\rho_{i} = 1$ on $\p D_0$, the Hopf's Lemma yields
1318: 
1319: \begin{align*}&\norm { \nabla \rho }_{L^{\infty} (D_0 \setminus (D_1\cup D_2
1320: \cup D_3 ))} \\ &\leq \max \{\norm { \nabla \rho_1 }_{L^{\infty} (\p
1321: D_1 )}, \norm { \nabla \rho_2 }_{L^{\infty} (\p D_2 )}, \norm {
1322: \nabla \rho_3 }_{L^{\infty} (\p D_3 )},\norm { \nabla \rho
1323: }_{L^{\infty} (\p D_0 )}\}< C.\end{align*} Applying the Hopf's Lemma
1324: again, we can have that the gradient of $v_0$ is bounded independent
1325: of $\epsilon_1$, refer to Lemma 2.2 in \cite{BLY}.
1326: 
1327: \par We estimate $C_1 \nabla v_1$ in the narrow region between $D_1$ and
1328: $D_2$. Since $v_1$ is constat on the boundaries and the boundaries
1329: is smooth enough in the narrow region, the proof of Lemma 4.3
1330: implies that $v_1$ can be extend into the interior areas of $D_1$
1331: and $D_2$ by the distance almost $\epsilon$ from the boundaries in
1332: the narrow region, independently of $r_1$ and $r_2$. By the gradient
1333: estimate for harmonic functions allows
1334: $$ |C_1 \nabla v_1 | \lesssim \frac {r_1 r_3}{
1335: r_1 + r_3 } \frac 1 {\sqrt {r_2}}\frac 1 {\sqrt {\epsilon_1}}$$ in
1336: the narrow region between $D_1$ and $D_2$. Note that the inequality
1337: above is a local property independent of choosing $D_0$.
1338: 
1339: \par Now, we consider $ C_3 \nabla v_3$ in the narrow region between $D_1$ and
1340: $D_2$. Let $\widetilde{\rho}$ be a harmonic function in $D_0
1341: \setminus (D_2 \cup D_3)$ with the boundary data $$ \widetilde{\rho}
1342: = 0 ~\mbox{on}~\partial (D_0 \cup D_2) ~\mbox{and}~\widetilde{\rho}
1343: = 1 ~\mbox{on}~\partial D_3.
1344: $$ By the maximum principle, we have
1345: $$0\leq   v_3 \leq \widetilde{\rho}~\mbox{in}~D_0
1346: \setminus (D_1\cup D_2 \cup D_3)$$ Considering the standard estimate
1347: for $\Psi[D_2,D_3]$, we can obtain
1348: $$ |C_3 v_3| \leq C \sqrt {\epsilon_2}.$$
1349: Similarly to the estimate for $C_1 \nabla v_1$, the gradient
1350: estimate for harmonic functions yields
1351: $$ |C_3 \nabla v_3 | \lesssim \sqrt  {\frac {\epsilon_2} {{\epsilon_1}}}$$
1352: in the narrow region between $D_1$ and $D_2$. \par Therefore, we can
1353: obtain the desirable upper bound. Here, it is noteworthy that the
1354: upper bound is dominated only by the estimate for $C_2 \nabla v_1$,
1355: which is independent of choosing $D_0$. In this respect, the
1356: constant C of the upper bound in Theorem \ref{thm:B} is independent
1357: of $r_1$, $r_2$, $r_3$, $\epsilon_1$ and $\epsilon_2$.
1358:  \qed
1359: 
1360: 
1361: 
1362: \section*{Acknowledgements}
1363: The authors would like to express his gratitude to \textrm{Professor
1364: Hyeonbae Kang}, who suggested the original problem studied in this
1365: paper.  The authors is also grateful to \textrm{Professor YanYan Li}
1366: for his concern with their subject and suggestions. The second named
1367: author would also like to express his thanks to \textrm{Professor
1368: Gang Bao}, and gratefully acknowledges his hospitality during the
1369: visiting period at Michigan State University.
1370: 
1371: 
1372: \begin{thebibliography}{}\label{sec:TeXbooks}
1373: 
1374: 
1375: \bibitem{A}{\sc L. Ahlfors}, {Complex Analysis}, Third ed., McGraw-Hill, New
1376: York, 1979.
1377: 
1378: 
1379: 
1380: \bibitem{ADKL} {\sc Ammari H, Dassios G, Kang H, and Lim M}, {\em Estimates for the electric field in the presence of adjacent perfectly conducting spheres}, {Quat. Appl. Math.}, { 65} (2007), pp.~339--355
1381: 
1382: \bibitem{AKL} {\sc H. Ammari, H. Kang, and M. Lim}, {\em Gradient estimates for solutions to the conductivity problem}, Math. Ann., 332(2) (2005), pp.~277--286.
1383: 
1384: \bibitem{AKLLL} {\sc H. Ammari, H. Kang, H. Lee, J. Lee, and M. Lim}, {\em Optimal bounds on the gradient of solutions to conductivity problems}, J. Math. Pures Appl., 88 (2007), pp.~307--324.
1385: 
1386: \bibitem{Bao_thesis}{\sc S. Bao}, {\em Gradient estimates for the conductivity problems}, thesis, Rutgers University.
1387: 
1388: 
1389: \bibitem{BC}{\sc  B. Budiansky and G. F. Carrier}, {\em High shear stresses in stiff fiber composites}, J. Appl. Mech., 51 (1984), pp. 733-735.
1390: 
1391: \bibitem{BLY} {\sc E. Bao, Y.Y. Li and B. Yin} {\em Gredient estimats for the conductivity problem}, Arch. Rational Mech. Anal, to appear.
1392: 
1393: \bibitem{BLY2} {E. Bao, Y.Y. Li and B. Yin} {\em Gradient estimates for the perfect and
1394: insulated conductivity problem and elliptic systems} (in
1395: preparation)
1396: 
1397: \bibitem{BV} {\sc E. Bonnetier and M. Vogelius}, {\em An Elliptic regularity result for a composite medium with "touching" fibers of circular
1398: cross-section}, SIAM J. Math. Anal., 31, No 3 (2000), pp.~651--677.
1399: 
1400: 
1401: 
1402: 
1403: %\bibitem{J} {\sc J.D. Jackson}, {\em Classical
1404: %Electrodynamics}, Third ed., Wiley, New York, NY, 1999.
1405: %
1406: 
1407: 
1408: \bibitem{K} {\sc J.B. Keller}, {\em Stresses in narrow regions}, Trans. ASME
1409: J. Appl. Mech., 60 (1993), pp. ~1054--1056.
1410: 
1411: 
1412: 
1413: \bibitem{LN} {\sc Y.Y. Li and M. Nirenberg}, {\em Estimates for ellliptic system from composite material}, Comm. Pure Appl. Math., LVI (2003), pp.~892--925.
1414: 
1415: 
1416: \bibitem{LV} {\sc Y.Y. Li and M. Vogelius}, {\em Gradient estimates for solution to divergence form elliptic equation with discontinuous coefficients}, Arch. Rational Mech. Anal., 153 (2000),
1417: pp.~91--151.
1418: 
1419: \bibitem{LY} {\sc M. Lim and K. Yun}, {\em Blow-up of Electric Fields between Closely Spaced Spherical Perfect Conductors}, submitted.
1420: 
1421: \bibitem{Y} {\sc K. Yun},  {\em  Estimates for electric fields blown up between closely adjacent conductors with arbitrary shape}, { SIAM J. Appl. Math.}, 67, No 3 (2007), pp.~714--730.
1422: \bibitem{Y2} {\sc K. Yun}, {\em Optimal bound on high stresses occurring between stiff fibers with arbitrary shaped cross sections}, J. Math. Anal. Appl. 350, (2009), pp. 306-312
1423: \end{thebibliography}
1424: 
1425: 
1426: \end{document}
1427: