\documentstyle[twoside,amssymb]{article} % amssymb is used for R in Real numbers \pagestyle{myheadings} \markboth{\hfil Stable multiple-layer stationary solutions \hfil EJDE--1997/22}% {EJDE--1997/22\hfil Arnaldo Simal do Nascimento \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol.\ {\bf 1997}(1997), No.\ 22, pp. 1--17. \newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp (login: ftp) 147.26.103.110 or 129.120.3.113} \vspace{\bigskipamount} \\ Stable multiple-layer stationary solutions of a semilinear parabolic equation in two-dimensional domains \thanks{ {\em 1991 Mathematics Subject Classifications:} 35K20, 35K57, 35B25. \hfil\break\indent {\em Key words and phrases:} Diffusion equation, Gamma-convergence, transition layers, \hfil\break\indent stable equilibria. \hfil\break\indent \copyright 1997 Southwest Texas State University and University of North Texas. \hfil\break\indent Submitted May 13, 1997. Published December 1, 1997.} } \date{} \author{Arnaldo Simal do Nascimento\\ \\ Dedicated to the memory of Ennio De Giorgi (1928-1996)} \maketitle \newcommand{\diver}{\mathop{\rm div}} \begin{abstract} We use $\Gamma$--convergence to prove existence of stable multiple--layer stationary solutions (stable patterns) to a reaction--diffusion equation. Given nested simple closed curves in ${\Bbb R}^2$, we give sufficient conditions on their curvature so that the reaction--diffusion problem possesses a family of stable patterns. In particular, we extend to two-dimensional domains and to a spatially inhomogeneous source term, a previous result by Yanagida and Miyata. \end{abstract} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}{Lemma}[section] \newtheorem{coro}{Corollary}[section] \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \newcommand{\quina}{\mathbin{\hbox{\vrule height8pt}\vbox{\hrule width 5pt}}} \section{Introduction} This paper is a contribution to the investigation of some diffusion processes, wherein the interplay between diffusivity and a source term gives rise to stable multiple-layer stationary solutions. This kind of solution will be referred to as a multiple-layer pattern. At the same time, along with \cite{N2} and \cite{N3}, it consolidates the use of the $\Gamma$--convergence technique to show the existence of multiple-layer patterns of some semilinear parabolic equations which involve a small parameter. Specifically we will be focusing on processes governed by the evolution problem \begin{eqnarray} \frac{\partial v_\varepsilon}{\partial t} &=& \varepsilon^2\, \diver\left( k_1(x)\nabla v_\varepsilon\right) + k_2(x)(v_\varepsilon -\alpha) (\beta-v_\varepsilon) (v_\varepsilon -\gamma_\varepsilon(x))\,,\nonumber \\ && \mbox{ for } (x,t)\in\Omega\times{\Bbb R}^+ \\ &&v_\varepsilon(x,0) = v_0 \quad \frac{\partial v_\varepsilon}{\partial \widehat{n}} = 0\,, \quad\mbox{for } x\in \partial\Omega\,, \ t >0\nonumber \end{eqnarray} where $\widehat{n}$ is the inward normal to $\partial\Omega$; $\alpha,\beta\in{\Bbb R}$, $\alpha<\beta$, with $\varepsilon$ a small positive parameter; $k_i(x)$, $i=1,2$, are positive functions in $C^1(\Omega)$; $\Omega\subset{\Bbb R}^2$ has smooth boundary $\partial\Omega$, say of class $C^3$; and \begin{eqnarray} &\gamma_\varepsilon(x) =\left(\frac{\alpha+\beta}{2}\right)+g_\varepsilon(x), \ \alpha< \beta, \ \gamma_\varepsilon\in C(\overline{\Omega})& \nonumber \\ &g_\varepsilon(x) = o(\varepsilon), \ \mbox{ uniformly in $\Omega$ as $\varepsilon\to 0$}\,. & \end{eqnarray} In \cite{MY}, the existence of multiple-layer patterns has been proved for the one-dimensional case, i.e, $\Omega=[0,1]$, with the hypotheses that $k_2(x)\equiv 1$, $\gamma_\varepsilon=1/2 + a_1\varepsilon + o(\varepsilon^2)$,$a_1$ constant, $\alpha=0$, and $\beta=1$. Here, besides considering any two-dimensional domain $\Omega$, we allow for spatially inhomogeneous perturbations of the state $v = \left[ (\alpha+\beta)/2\right]$. The procedure used in \cite{MY} was to construct super and sub-solutions by modifying a traveling wave solution of (1.1), with the aforementioned restrictions, to obtain a multiple-layer pattern. However, this procedure is not suitable for a generalization to two-dimensional domains. Herein a technique known as $\Gamma$--convergence devised by De Giorgi \cite{GiF} and developed by many others is used. When seeking stable stationary solutions to some classes of semilinear parabolic equations, the $\Gamma$--convergence approach turns out to be very useful. This approach replaces the original problem of minimizing a family of functionals by a more tractable problem in the space of functions of bounded variation, $BV(\Omega$), which usually yields more precise information on the geometric structure of the minimizers. The prototype for the source term in (1.1) is the case $\alpha=0$, $\beta=1$ and $0<\gamma_\varepsilon<1$, which stems from the theory of population genetics where $\alpha,\beta$ and $\gamma_\varepsilon$ denote some probability measures. The case in which $\Omega=[0,1]$, $\alpha=0$, $\beta=1$ and $\gamma_\varepsilon(x)=a(x)$, $0 0$ and $\Sigma_i(s,0)=\gamma_i(s)$. Let us also set $\widetilde{k}_1(s,d)=k_1(\Sigma(s,d))$ , $\widetilde{k}_2(s,d)=k_2(\Sigma(s,d))$ and regarding (1.1) put $$k(x)\stackrel{\rm def}{=} \left[ k_1(x)k_2(x)\right]^{1/2}\,.$$ We now state our main result. \begin{theorem} Suppose that $\gamma_i$, $i=1,\ldots,p$, is a $\nu_i$--level curve of $k$, \ i.e., \[ k(x) = \nu_i\,, \ \ \mbox{for} \ \ x\in \gamma_i\,, \ i=1,\ldots,p\,. \] Let $\widetilde{k}(s,d)=k(\Sigma(s,d))$ and $\Lambda_i(s,d)=\widetilde{k}(s,d) J_{\Sigma_i}(s,d)$, $(s,d)\in [0,L_i]\times (-\delta,\delta)$. Suppose that $$ \Lambda_i(s,d) > \Lambda_i(s,0) = \nu_i\,, \ \ \mbox{for} \ \ d\in(-\delta,\delta), \ d\neq 0, \ i=1,\ldots,p \eqno(H1) $$ Then there is a family ${\{ v_\varepsilon\}}_{0<\varepsilon\leq\varepsilon_0}$, with $\varepsilon_0$ small, of stationary solutions of (1.1) such that for each $\varepsilon\in (0,\varepsilon_0)$: \begin{description} \item[{\rm (2.1.i)}] $v_\varepsilon\in C^{2,\sigma}(\overline{\Omega})$, $0<\sigma <1$ and $\alpha < v_\varepsilon(x) < \beta$, \ $\forall \, x\in\overline{\Omega}$. \item[{\rm (2.1.ii)}] $\| v_\varepsilon - v_0\|_{L^1(\Omega)}\longrightarrow 0$, as $\varepsilon\to 0$. \item[{\rm (2.1.iii)}] For each $\lambda\in (0,(\beta-\alpha)/2)$ and $\Omega^\lambda_\varepsilon = \left\{ x\in \Omega : \alpha +\lambda < v_\varepsilon(x) < \beta-\lambda\right\}$ it holds that $\left| \Omega^\lambda_\varepsilon\right| \longrightarrow 0$, as $\varepsilon\to 0$, where $|\cdot |$ is the 2-dimensional Lebesgue measure. \item[{\rm (2.1.iv)}] $v_\varepsilon$ is a stable stationary solution of (1.1). \end{description} \end{theorem} Next we give necessary conditions for (H1) to be satisfied. \begin{lemma} Suppose that each $\gamma_i$, $i=1,\ldots,p$, satisfies $$ \begin{array}{rl} \displaystyle \frac{\partial \widetilde{k}(s,0)}{\partial d} = \nu_i\,\kappa_i(s), & 0\leq s \leq L_i \\ \displaystyle \frac{\partial^2\,\widetilde{k}(s,0)}{\partial d^2} > 2\,\nu_i\,\kappa^2_i(s), & 0\leq s \leq L_i \end{array} \eqno(H2) $$ Then hypothesis (H1) is satisfied. \end{lemma} \paragraph{Remark.} An example of a function which satisfies (H2) can be easily constructed. For instance, let $\gamma_i\subset\Omega\subset{\Bbb R}^2$ with $\gamma_i=\{ x\in\Omega :\; \| x\| =1\}$. Then the function $\tilde{k}(s,t)=1 + t + 2\,t^2$ with $t\in(-\delta,\delta)$, $\delta$ small, and $s\in[0,2\pi)$, satisfies (H2) with $\nu_i =1$. \paragraph{Remark.} The hypothesis on $\gamma_i$ could have been slightly more general. It would suffice to have $\gamma_i$, $i=1,\ldots,p$, a subset of a $\nu^1_i$--level set of $k_1$ and also a subset of a $\nu^2_i$--level set of $k_2$. Then $\gamma_i$ would be a $\nu_i$--level curve of $k(x)$ where $\nu_i=\left( \nu^1_i\nu^2_i\right)^{1/2}$. \paragraph{Remark.} The two conditions in (H2) dictate the behavior of $k$ in a neighbourhood of the limiting transition-phase curve $\gamma_i$. The first condition requires that $k(x)$ increase or decrease as $x$ crosses $\gamma_i$ along the direction of $\widehat{n}_i$ according to the sign of the curvature $\kappa_i$, its slope being proportional to $|\kappa_i|$ at the crossing-point, provided $\nu_i \neq 0$. The second one, for a fixed $s$ relates the concavity of $k(\Sigma(s,d))$ with the curvature $\kappa_i(s)$ of $\gamma_i$ at $s$. Note that $\frac{\partial\widetilde{k}}{\partial d}$ changes sign whenever the curvature of $\gamma_i$ does so. Each limiting phase-transition curve $\gamma_i$ acts through (H2) as a barrier which prevents a, say, diffusing substance whose initial concentration evolves in time according to (1.1), to spread homogeneously in space and eventually settling down in a uniform concentration. Actually this would be the case if for instance we had $\Omega$ convex, $k_1(x)\equiv\mbox{const.}$, $k_2(x)\equiv\mbox{const.}$ and $g_\varepsilon(x)\equiv 0$. See \cite{M,CH} for this matter. An example where this occurs, in despite of the fact that the diffusion function $k_2$ is not constant, can be found in \cite{N1}. \paragraph{Remark.} The stability referred to in (2.1.iv) should be understood in the following sense: a stationary solution $v(x)$ of (1.1) is said to be stable in the $H^1(\Omega)$-norm, say, if for any $\mu>0$ there exists $\delta>0$ such that $T(t)\psi$ exists for all $t>0$ (here $T(t)$ denotes the nonlinear semigroup generated by (1.1)\,) and \[ \| T(t)\psi - v\|_{H^1(\Omega)} < \mu\,, \ \ 0 0 \ \ \mbox{for any} \ \ v\in{\Bbb R}\,, \ v\neq\alpha\,, \ v \neq\beta &\\ &\widetilde{F}'(\alpha) = \widetilde{F}'(\beta) = 0\,, \ \widetilde{F}''(\alpha) > 0\,, \ \widetilde{F}''(\beta) > 0\,.\nonumber \end{eqnarray} It is easy to see that any local minimizer $v_\varepsilon$ of $E_\varepsilon$ will be a solution to (H1) and, by regularity, $v_\varepsilon\in C^{2,\nu}(\Omega), \ 0<\nu <1$. Therefore, it is our aim now to find a family of local minimizers of $E_\varepsilon$. To that end we will need the concept of $\Gamma$--convergence and of a local isolated minimizer of $E_\varepsilon$. See for example \cite{Ma}. \paragraph{Definition} A family $\{ E_\varepsilon\}_{0<\varepsilon\leq\varepsilon_0}$ of real-extended functionals defined in $L^1(\Omega)$ is said to $\Gamma$--converge, as $\varepsilon\to 0$, to a functional $E_0$, at $v$ and we write \[ \Gamma(L^1(\Omega)^-) - \lim_{\varepsilon\to 0} E_\varepsilon(v)=E_0(v) \] if \begin{itemize} \item For each $v\in L^1(\Omega)$ and for any sequence $\{ v_\varepsilon\}$ in $L^1(\Omega)$ such that $v_\varepsilon\to v$ in $L^1(\Omega)$, as $\varepsilon\to 0$, it holds that $E_0(v)\leq\displaystyle\liminf_{\vspace*{-6mm}\varepsilon\to 0} E_\varepsilon(v_\varepsilon)$. \item For each $v\in L^1(\Omega)$ there is a sequence $\{ w_\varepsilon\}$ in $L^1(\Omega)$ such that $w_\varepsilon\to v$ in $L^1(\Omega)$, as $\varepsilon\to 0$ and also $E_0(v)\geq \displaystyle \limsup_{\vspace*{-6mm}\varepsilon\to 0} E_\varepsilon(w_\varepsilon)$. \end{itemize} \paragraph{Definition.} We say that $v_0\in L^1(\Omega)$ is an $L^1$-local minimizer of $E_0$ if there is $\rho>0$ such that \[ E_0(v_0)\leq E_0(v) \ \ \mbox{whenever} \ \ 0 < \| v-v_0\|_{L^1(\Omega)} < \rho\,.\] Moreover, if $E_0(v_0)0}$ such that $E_\varepsilon(u_\varepsilon)\leq\mbox{const.}<\infty$, $\varepsilon>0$, is compact in $L^1(\Omega)$. \item[{\rm (4.1.ii)}] There exists an isolated $L^1$-local minimizer $v_0$ of $E_0$. \end{description} Then there exists $\varepsilon_0>0$ and a family $\{ v_\varepsilon\}_{0<\varepsilon\leq\varepsilon_0}$ such that: \begin{itemize} \item $v_\varepsilon$ is an $L_1$-local minimizer of $E_\varepsilon$ \item $\|v_\varepsilon-v_0\|_{L^1(\Omega)} \to 0$, \, as \, $\varepsilon\to 0$. \end{itemize} \end{theorem} Let us consider the family of functionals defined by (4.2). For the case in which $g_\varepsilon(x)\equiv 0$ and $\widetilde{F}$ satisfies (4.3), the $\Gamma$--convergence of this type of functionals has been treated in \cite{S,MM1,Mo}. In (4.2) the presence of the term $\varepsilon^{-1}g_\varepsilon(x)k_2(x)F^1(v)$ adds no additional difficulty because its smoothness makes of it a continuous perturbation with respect to $L^1$--convergence. Moreover by virtue of (1.2) the perturbation term $\int_\Omega\varepsilon^{-1}g_\varepsilon(x)k_2(x)F^1(v_\varepsilon)dx$ vanishes when one takes the $\Gamma$--limit. Hence the above results can be evoked thus yielding the following theorem. \begin{theorem} Consider the family of functionals given by (4.2). Then \[ \Gamma(L^1(\Omega)^-) - \lim_{\varepsilon\to 0} E_\varepsilon(v) = E_0(v) \] where \begin{equation} E_0(v) = \left\{\begin{array}{ll} \displaystyle C_0 \int_\Omega\,k\left | D{\chi}_{\{v=\beta\}}\right |, &\mbox{if } v\in BV(\Omega,\{\alpha,\beta\})\,, \\ & 0 < |\{v=\beta\}| < |\Omega| ;\\ \infty\, & \mbox{otherwise.} \end{array}\right. \end{equation} and \ $C_0 = (\beta -\alpha)\int^\beta_\alpha\,(\widetilde{F}(t))^{1/2}dt$. \end{theorem} Our next goal is to apply Theorem 4.1 to $\{ E_\varepsilon\}$, $0<\varepsilon\leq\varepsilon_0$ and to $E_0$, as defined in (4.2) and (4.4). We only have to worry about (4.1.ii) since (4.1.i) has essentially been proved in \cite{KS,Mo}. Their proof can be easily adapted to fit our case. As for (4.1.ii) it has been proved in \cite{N2} and \cite{N3} for the case of just one limiting transition--phase curve. We give here a neater and more geometric proof for this case and then consider the case of $p$ limiting transition--phase curves. For the single curve case the proof presented in the references above uses the approximation result mentioned at the beginning of the proof of Lemma 3.2. This approach is somehow cumbersome since it has to deal with sequences and subsequences throughout the proof and unnatural since it avoids to work with the geometry of $BV$--spaces which is the setting our problem is naturally casted into. This approximation approach can be avoided by resorting to Lemma 3.1. In the next theorem we will be dealing with the single curve case, i.e., $p=1$, and therefore we set for simplicity $\gamma=\gamma_1$. Thus $N_\delta=N_\delta(\gamma)$ will denote its tubular neighbourhood,\, $\Sigma$\, the corresponding coordinate map and we drop the subindex 1 in all other definitions. The proof is a modification of a similar proof given in \cite{KS}. \begin{theorem} With the notation of Section 2, suppose that $\gamma\subset \Omega$ is a simple closed $\nu$--level curve of $k$ and that {\rm (H1)} is satisfied with $\Omega=\Omega_1\cup\gamma\cup\Omega_2$. Then \[ v_0(x) = \alpha\,{\chi}_{\Omega_1}(x) + \beta\,{\chi}_{\Omega_2}(x)\,, \ \ x\in \Omega \] is an $L^1$-local isolated minimizer of $E_0$. \end{theorem} \paragraph{ Proof:} It suffices to prove that, if $v \in BV(\Omega ;\{\alpha,\beta\})$ and $0 < \|v - v_0\|_{L^1(N_\delta)} <\rho$ for a suitable $\rho > 0$, then \[ \int_{N_\delta(\gamma)}k |Dv| > \int_{N_\delta(\gamma)}k|Dv_0|.\] Let us start by computing $E_0(v_0)$. By the coarea formula (see \cite{F}) we obtain \begin{eqnarray*} E_0(v_0)&=&\int_{N_\delta(\gamma)} k|Dv_0| = \int_{\alpha}^{\beta}\int_{N_\delta\cap\partial_*\{{v_0>\xi}\}}kd {\cal H}^1 d\xi \\ &=& (\beta -\alpha)\int_{\gamma}kd{\cal H}^1 = (\beta -\alpha)\nu\,L\,, \end{eqnarray*} where $L$ is the total arc--length of $\gamma$. For a fixed $d\in (-\delta,\delta)$ define \[ \ell_d = \left\{ (s,d)\,,\,0 < s < L\right\}\,. \] Then the trace of $\widetilde{v}(\cdot ,d)$ is well-defined on $\ell_d$, a.e. in $(-\delta,\delta)$; because each $\ell_d$ is $C^1$ (actually Lipschitz would suffice). Suppose that \begin{description} \item{i)} \ $\widetilde{v}=\widetilde{v}_0$, along $\ell_{\widetilde{d}}\cup\ell_{-\widetilde{d}}$, in the sense of traces, for some $\widetilde{d}\in (\delta/2,\delta)$. \end{description} Recall that $\Lambda(s,d) = \widetilde{k}(s,d)J_\Sigma(s,d)$. We say that $v$ is an admissible function if $\|v-v_0\|_{L^1(N_\delta)}>0$ and $v\in BV(N_\delta(\gamma);\{\alpha,\beta\})$. For any such function $v$ and with the use of Lemma 3.2 we have: \begin{eqnarray*} E_0(v)&=&\int_{N_\delta}k|Dv| \\ &\geq & \int_{\widetilde{N}_\delta}\widetilde{k}\left( |\widetilde{\mu}_1|^2+ |\widetilde{\mu}_2|^2\right)^{1/2} \geq \int_{\widetilde{N}_\delta}\widetilde{k}|\widetilde{\mu}_2| \\ & \geq & \nu\,\int^L_0\int^\delta_{-\delta}\left|\frac{\partial\widetilde{v}}{\partial d}\right| \; \mbox{(by $(H_1)$ and definition of $\widetilde{\mu}_2$)} \\ & \geq & \nu(\beta-\alpha)L\,, \end{eqnarray*} since \ i) \ implies that $\int_{-\delta}^{\delta}\left |\frac{\partial\widetilde{v}}{\partial d}\right | \geq (\beta - \alpha)$ for any $s\in [a,b]$. We claim that $E_0(v_0)\xi\}} k|Dv|\right) d\xi \nonumber \\ & = & \int^\infty_{-\infty}\left( \int_{N_\delta\cap\partial_*\{ v>\xi\}} kd{\cal H}^1\right) d\xi \nonumber \\ & = & \int^\infty_{-\infty}\left( \int_{\widetilde{N}_\delta\cap\partial_*\{\tilde{v}>\xi\}} \Lambda(s,t)d{\cal H}^1\right) d\xi \nonumber \\ & = & (\beta -\alpha)\int_{\widetilde{N}_\delta\cap\partial_* \{\widetilde{v}=\beta\}\cap\partial_*\{\widetilde{v}=\alpha\}} \Lambda(s,t)d{\cal H}^1 \end{eqnarray} To deduce (4.5) we use that \[ |D\widetilde{v}| = {\cal H}^1 \quina \left( \partial_*\{\widetilde{v}=\alpha\} \cap \partial_*\{\widetilde{v}=\beta\}\right) \] taking into account that $\partial\{\widetilde{v}=\alpha\} \cap \partial\{\widetilde{v}=\beta\}$ is a set of finite perimeter in $\widetilde{N}_\delta$. Also our hypothesis implies \[ \int_{\widetilde{N}_\delta}\widetilde{k}\left( |\widetilde{\mu}_1|^2 + |\widetilde{\mu}_2|^2 \right)^{1/2} = \int_{\widetilde{N}_\delta}\widetilde{k}|\widetilde{\mu}_2|. \] But the above equality holds if and only if $|\widetilde{\mu}_1| \equiv 0$ on $\widetilde{N}_\delta$, what is to say that $\widetilde{v}(\cdot , d)$ is ${\cal H}^1$-a.e. constant along each $\ell_d$, for a.e. $d\in (-\delta,\delta)$. Hence, $$\widetilde{N}_\delta\cap\partial_*\{\widetilde{v}= \alpha\}\cap\partial_*\{\widetilde{v}=\beta\} = \cup^m_{j=1}\ell_{d_{\scriptstyle j}}\,,$$ for some $m\in{\Bbb N}$ and $-\delta 0$. This takes care of our theorem in the case in which \, i-) \, holds. Now if \, i) \, does not hold then one of the following cases would occur, in the sense of traces of $BV$-functions: \begin{description} \item[ii)] $\widetilde{v}$ is not const. ${\cal H}^1$--a.e. along $\ell_{\widetilde{d}}$, for a.e. $\widetilde{d}\in\left( \delta/2\,,\,\delta\right)$ \item[iii)] $\widetilde{v}$ is not const. ${\cal H}^1$--a.e. along $\ell_{-\widetilde{d}}$, for a.e. $\widetilde{d}\in\left( \delta/2\,,\,\delta\right)$ \item[iv)] $\widetilde{v}\equiv\alpha\;{\cal H}^1$-a.e. along $\ell_{\widetilde{d}}$\, and $\widetilde{v}\equiv\beta\, ,{\cal H}^1$-a.e. along $\ell_{-\widetilde{d}}$\, , for a. e. $\widetilde{d}\in (\delta/2,\delta)$. \end{description} Following ideas set forth in \cite{KS}, we define a set $\Delta \subset (0,\delta)$, $$ \Delta\stackrel{\rm def}=\left\{ \begin{array}{l} \displaystyle \widetilde{d}\in(0,\delta) : \int^L_0\big\{ |\widetilde{v}-\widetilde{v}_0|(s,\widetilde{d}) J_\Sigma(s,\widetilde{d}) + |\widetilde{v}-\widetilde{v}_0|(s,-\widetilde{d}) J_\Sigma(s,-\widetilde{d})\big\}\, ds \\ \displaystyle \quad > 4\rho/\delta, \mbox{ where }\int_{\widetilde{N}_\delta}|\widetilde{v}-\widetilde{v}_0| J_\Sigma < \rho \end{array}\right\} $$ Hence $|\Delta| < \delta/4$. Now if \ iv) does not hold, then by choosing $ \rho < \delta L(\beta-\alpha)/ 2$ we obtain \[ \int^L_0 \left\{ |\widetilde{v}-\widetilde{v}_0|(s,\widetilde{d}) J_\Sigma(s,\widetilde{d}) + |\widetilde{v}-\widetilde{v}_0|(s,-\widetilde{d}) J_\Sigma(s,-\widetilde{d})\right\} = 2\,L\,(\beta-\alpha) > (4\rho/\delta)\,, \] which by definition implies that $\widetilde{d}\in \Delta$. But then for a.e. $\widetilde{d}\in \left( (\delta/2,\delta)\backslash\Delta\right)$, either \, ii) or \, iii) holds. For any admissible $v$, Lemma 3.1 allows us to conclude that \begin{eqnarray*} E_0(v) & = & \int_{N_\delta}k\,|Dv| \geq \int_{N_\delta\backslash N_{\delta/2}}k\,|Dv| + \int_{N_{\delta/2}}k\,|Dv| \\ & = & \int_{\widetilde{N}_\delta\backslash \widetilde{N}_{\delta/2}} \widetilde{k}\,\left( |\widetilde{\mu}_1|^2 +|\widetilde{\mu}_2|^2\right)^{1/2} + \int_{\widetilde{N}_{\delta/2}}\widetilde{k}\,\left( |\widetilde{\mu}_1|^2 +|\widetilde{\mu}_2|^2\right)^{1/2} \\ &\geq & \int_{\widetilde{N}_\delta\backslash\widetilde{N}_{\delta/2}}\widetilde{k} \,\left |\frac{\partial\widetilde{v}}{\partial s}\right | + \int_{\widetilde{N}_{\delta/2}}\Lambda\, \left |\frac{\partial\widetilde{v}}{\partial d}\right | \\ &=& I_1+I_2 \end{eqnarray*} where $I_1$ and $I_2$ denote respectively the first and the second integrals in the last term of the above inequalities. If $ k_m = \min_{x\in\overline{\Omega}}\,k(x) $ then by virtue of \, ii) and \, iii), for a.e. $\widetilde{d}\in (\delta/2,\delta)\backslash\Delta$ it holds that \begin{eqnarray*} \lefteqn{\int^L_0\left\{\widetilde{k}(s,d)\,|\widetilde{\mu}_1| (s,\widetilde{d}) + \widetilde{k}(s,-\widetilde{d})\,|\widetilde{\mu}_1| (s,-\widetilde{d})\right\} }\\ & \geq & k_m\left\{ \mbox{ess }V^L_0\left[\widetilde{v}(\cdot,\widetilde{d})\right] + \mbox{ess }V^L_0\left[\widetilde{v}(\cdot,-\widetilde{d})\right]\right\} \\ & \geq & k_m(\beta -\alpha), \end{eqnarray*} where ess $V^L_0\left[\widetilde{v}(\cdot,\widetilde{d})\right]$ stands for the essential variation of $\widetilde{v}(\cdot,\widetilde{d})$ on $[0,L]$. Note that since $\tilde{v}\in BV(\tilde{N}_{\delta},\{\alpha,\beta\})$, the function $$ \tilde{d}\longrightarrow\;\mbox{ess}\;V^L_0 [\tilde{v}(\cdot,\tilde{d})] = \left |\frac{\partial \tilde{v}(\cdot,\tilde{d})}{\partial s} \right | (0,L) $$ is integrable in $(\delta/2,\delta)$. See \cite{Z,EG}. Then by integrating over $(\delta/2,\delta)\backslash\Delta$, \begin{eqnarray*} I_1 & \geq & k_m\int^\delta_{\delta/2}\left\{ \mbox{ess }V^L_0\left[\widetilde{v} (\cdot,\widetilde{d})\right]+\mbox{ess }V^L_0\left[\widetilde{v}(\cdot,-\widetilde{d}) \right]\right\} \\ & \geq & (\delta/4)k_m(\beta -\alpha). \end{eqnarray*} In order to obtain a lower estimate for $I_2$ we begin by remarking that since \[ |(0,\delta/2)\backslash\Delta| \geq \delta/4 \] and\, $\widetilde{v}$\, is an admissible function then there is $\overline{d}\in (0,\delta/2)\backslash\Delta$ such that $(s,\overline{d})$ and $(s,-\overline{d})$ are points of approximate continuity of $\widetilde{v}(s,\widetilde{d})$, for a.e. $s\in[0,L]$. Also by the definition of $\widetilde{v}_0$, \[ |\widetilde{v}(s,\widetilde{d})-\widetilde{v}(s,-\widetilde{d})| \geq (\beta-\alpha) - \left\{|\widetilde{v}_0-\widetilde{v}|(s,\widetilde{d}) + |\widetilde{v}_0- \widetilde{v}|(s,-\widetilde{d})\right\} \] for any $(s,\widetilde{d})\in\widetilde{N}_\delta$ such that $\widetilde{v}$ is approximately continuous at $(s,\widetilde{d})$. Hence \begin{eqnarray*} I_2 & = & \int_{\widetilde{N}_{\delta/2}}\Lambda\, \left |\frac{\partial\widetilde{v}}{\partial d}\right | \geq \nu \int^L_0\int^{\delta/2}_{-\delta/2}\left |\frac{\partial\widetilde{v}}{\partial d} \right | \\ & = & \nu\int^L_0 \mbox{ess }V^{\delta/2}_{-\delta/2}\left[\widetilde{v}(s,\cdot) \right] ds \geq \nu\int^L_0 |\widetilde{v}(s,\overline{d}) - \widetilde{v}(s,-\overline{d})|ds \\ &\geq & \nu\int^L_0 \left\{ (\beta-\alpha)-\left[ |\widetilde{v}- \widetilde{v}_0|(s,\overline{d}) + |\widetilde{v}-\widetilde{v}_0|(s,-\overline{d})\right]\right\} ds \\ &\geq & \left\{ \nu\,L(\beta-\alpha)-\frac{4\,\rho\,\nu}{\delta\,J_m(-\delta,\delta)} \right\} \,, \end{eqnarray*} where \[ J_m(-\delta,\delta) = \min_{-\delta\leq d\leq\delta}\left\{ \min_{0\leq s\leq L}\, J_\Sigma(s,d)\,,\,\min_{0\leq s\leq L}\, J_\Sigma(s,-d)\right\}\,. \] These estimates finally yield \begin{eqnarray*} E_0(v) \geq I_1+I_2 & \geq & \left[ \frac{\delta}{4}\,k_m(\beta-\alpha) + \nu\,L(\beta-\alpha) - \frac{4\,\rho\,\nu}{\delta\,J_m(-\delta,\delta)}\right] \\ & > & \nu(\beta-\alpha)L = E_0(v_0)\,, \end{eqnarray*} as long as we take \begin{equation} \rho < \min\left\{ \frac{k_m(\beta-\alpha)\delta^2J_m(-\delta,\delta)}{16\nu}\,, \,{\delta\,L(\beta-\alpha)\over 2}\right\} \end{equation} Now our claim follows by extending $v_0$ to be constant on each connected component of $\Omega\backslash\gamma$ and observing that $ |Dv_0|\left( \Omega\backslash\gamma\right)=0$. \hfil$\Box$ The next task is to generalize Theorem 3.1 to the case of $p$ limiting phase-transition curves. Recall the notation set forth in Section 2 where $N_{\delta,i}$ is a $\delta$-tubular neighbourhood around each limiting phase-transition curve $\gamma_i$, $i=1,2,\ldots,p$. \begin{coro} Let $\gamma_i$, $i=1,\ldots,p$, $\Omega_i$, $k$ be as in Section 2. If (H1) is satisfied, then the function $v_0$ is a $L^1$--local minimizer of $E_0$. \end{coro} \paragraph{ Proof:} For any $v\in BV(\Omega ,\{\alpha,\beta\})$, $||v - v_0||_{L^1(\Omega)}< \rho$, where $\rho$ satisfies (4.6) with $\nu = \max\{\nu_i, i=1,...,p\}$, let \[ \Omega_\alpha \stackrel{\rm def}{=}\{ x\in \Omega:v(x)=\alpha\}\,, \quad \Omega_\beta \stackrel{\rm def}{=}\{ x\in\Omega:v(x)=\beta\} \] and consider the sets $ \Omega^0_\alpha \;\mbox{and}\; \Omega^0_\beta$ defined by (2.2). Then the set $(\Omega_\alpha \cup \Omega_\beta )\cap \Omega$ has finite perimeter. By applying Theorem 4.3 to each $\gamma_i$, $i=1,\ldots,p$ we obtain, using the coarea formula \begin{eqnarray*} E_0(v) &=& \int_\Omega k |Dv| = (\beta-\alpha)\,\int_{\partial_*\Omega_\alpha\cap\partial_*\Omega_\beta\cap\Omega} k\,d{\cal H}^1 \\ & \geq & (\beta-\alpha)\,\sum^p_{i=1}\int_{\partial_*\Omega_\alpha\cap\partial_*\Omega_\beta\ cap N_{\delta,i}}k\,d\,{\cal H}^1 \\ & > & (\beta-\alpha)\sum^p_{i=1}\int_{\partial_*\Omega^0_\alpha\cap\partial_*\Omega^0_\alpha \cap N_{\delta,i}}k\,d\,{\cal H}^1 \\ & = & (\beta-\alpha)\int_{\bigcup^{p}_{i=1}\!\gamma_i}k\,d\,{\cal H}^1 = \int_\Omega k\,|Dv_0| = E_0(v_0)\,. \end{eqnarray*} \section{Proof of Theorem 2.1} \setcounter{equation}{0} In this section we show how the previous sequence of lemmas can be used to accomplish the proof of Theorem 2.1. In view of Corollary 4.1 and previous remarks, Theorem 4.1 yields a family $\{v_\varepsilon\}_{0<\varepsilon \leq\varepsilon_0}$ of $L^1$--local minimizers of $E_\varepsilon$, where $E_\varepsilon$ is given by 4.2. Clearly any such minimizer $v_\varepsilon$ is a weak solution ($H^1$-sense) of (4.1) and regularity theory implies $v_\varepsilon \in C^{2,\sigma}(\overline{\Omega})$, $0<\sigma <1$. An application of the maximum principle then gives $\alpha0$ such that \[ \left|\Omega^\lambda_{\varepsilon_{\scriptstyle j}}\right| \geq \tau > 0 \,, \] for a fixed $\lambda$ and $\forall \, j$. Since $\| v_{\varepsilon_{\scriptstyle j}} - v_0\|_{L^1(\Omega)} \stackrel{j\to\infty}{\longrightarrow} 0$ it follows that $\{ \varepsilon_j\}$ has a subsequence, still denoted by $\{ \varepsilon_j\}$ such that $ v_{\varepsilon_{\scriptstyle j}}\to v_0 , \ \ \ \mbox{a.e. in}\ \Omega$. Thus, (4.3) allows us to invoke the Lebesgue convergence theorem to conclude that \[\lim_{j\to 0} \, \int_\Omega \widetilde{F}(v_{\varepsilon_{\scriptstyle j}})dx =0\,. \] But this is a contradiction to \[ \int_\Omega \widetilde{F}(v_{\varepsilon_{\scriptstyle j}})dx \geq \min \left\{ \widetilde{F}(\alpha+\lambda),\,\widetilde{F}(\beta+\lambda)\right\}\, \left|\Omega^\lambda_ {\varepsilon_{\scriptstyle j}}\right| \geq \widetilde{F}(\alpha+\lambda)\,\tau > 0 \,, \] for a fixed $\lambda$ and $\forall \, j$. Note that $\widetilde{F}(\alpha+\lambda) = \widetilde{F}(\beta-\lambda)$. It remains to prove \, 2.1.iv)\,, \, which concerns stability of the family $\{ v_\varepsilon\}$, ${0<\varepsilon\leq\varepsilon_0}$. From the fact that $v_\varepsilon$ is also a $H^1$-local minimizer of $E_\varepsilon$, ${0<\varepsilon\leq\varepsilon_0}$, it follows that \begin{equation}\langle E''_\varepsilon(v_\varepsilon)\psi\,,\,\psi\rangle_{H^1,H^*}\geq 0 \,, \ \ \ \forall \, \psi \in H^1(\Omega) \end{equation} where $H^*$ stands for the dual of $H^1(\Omega)$. Consider now the linearized eigenvalue problem \begin{eqnarray} &\varepsilon^2\diver\left[ k_1(x)\nabla \psi\right] + k_2(x)f'_\varepsilon(x,v_\varepsilon)\psi = \lambda\psi\,, \ \ x\in\Omega &\nonumber \\ &\frac{\partial\psi}{\partial n} = 0 \ \ \mbox{ on } \ \ \partial\Omega & \end{eqnarray} where $\displaystyle f'_\varepsilon(x,v)=\frac{\partial f_\varepsilon(x,v)}{\partial v}$. Denoting by $\{\lambda_n\}_n$, $n=1,2,\dots$, the sequence of eigenvalues of (5.2) then, taking into account (5.1), and using the variational characterization of the eigenvalues, we infer that $\lambda_n \leq 0$, $n=1,2,\ldots$. By well-known results from linearized stability and semigroup theory we conclude that in the case that the first eigenvalue $\lambda_1$ is negative then $v_\varepsilon$ is a strongly stable stationary solution of (1.1). Now if $\lambda_1=0$ a classical application of the Krein-Rutman theorem gives that zero is a simple eigenvalue of (5.2). In this case there is a local one dimensional critical manifold $M_c(v_\varepsilon)$, tangent to $[\psi_1]$ (the eigenspace spanned by the principal eigenfunction $\psi_1$\,), at $v_\varepsilon$, such that if $v_\varepsilon$ is stable in $M_c(v_\varepsilon)$ then it is also stable in $H^1(\Omega)$. For this matter we refer to Theorem 6.2.1 in \cite{H}, which proof can be adapted to fit our case. But now the stability of $v_\varepsilon$ in $M_c(v_\varepsilon)$ follows from the fact that the semigroup $\{T(t)\}_{t\geq 0}$ generated by (1.1) defines a gradient flow in $H^1(\Omega)$. To be more specific the functionals $E_\varepsilon[v_\varepsilon(x,t)]$ defines a Lyapunov function and along each solution $v_\varepsilon(x,t)$ it holds that \[ \frac{d}{dt}\,E_\varepsilon[v_\varepsilon(x,t)] \leq 0, \ \ \ t\geq 0 \;.\] This concludes the proof of Theorem 2.1. \hfil$\Box$ \paragraph{Remark.} Condition (2.1.iii) actually shows the multiple--layer profile of $v_\varepsilon$, for small $\varepsilon$, and it should hold that $v_\varepsilon\to v_0$, as $\varepsilon\to 0$, uniformly on compact sets of $\Omega\backslash\cup^p_{i=1}\gamma_i$. This should be done following ideas set forth in \cite{CC}. \paragraph{Remark.} The difficulties when trying to generalize our results to higher space dimensions are of technical nature and inherent to the proof of Theorem 4.3. For instance, in this case one would no longer have a global parametrization of the limiting phase-transition hypersurface. \paragraph{Remark.} Other types of patterns can be considered rather than just the case of nested limiting phase--transition curves. For instance, let $\gamma_i$, $i=1,\dots,p$, be smooth simple closed curves in $\Omega$ with $O_i$ denoting the open region enclosed by $\gamma_i$, and $\overline{O}_i$ its closure, then $\cap_{i=1}^p \overline{O}_i$ is empty. Then Theorem 4.3 and Corollary 4.1 still hold with $$v_0 = \alpha\chi_{\cup_{i=1}^p O_i}+\beta\chi_{\Omega \backslash\cup_{i=1}^p \overline{O}_i}\,. $$ A suitable combination of the two patterns referred to above could also be considered. \begin{thebibliography}{99} \bibitem{AMP} Angenent, S.B, Mallet-Paret, J. and Peletier, L.A., ``Stable transition layers in a semilinear boundary value problem''. {\em J. Diff. Eqns.} 67 (1987), 212-242. \bibitem{CC} Caffarelli, L.A. and Cordoba, A., ``Uniform convergence of a singular perturbation problem''. {\em Communications on Pure and Applied Math.,} vol. XLVIII (1995), 1-12. \bibitem{CH} Casten, R.G. and Holland, C.J., ``Instability results for reaction-diffusion equation with Neumann boundary conditions''. {\em J. Diff. Eqns.}, vol. 108, No. 2 (1994), 296-325. \bibitem{EG} Evans, L. and Gariepy, R., {\em Measure Theory and Fine Properties of Functions}. Studies in Advanced Mathematics, CRC Press (1992). \bibitem{ES} Evans, L. C. and Souganidis, P. E., ``A PDE approach to geometric optics for certain reaction-diffusion equations''. {\em Ind. Univ. Math. J.} 38 (1989), 141-172. \bibitem{F} Federer, H., {\em Geometric Measure Theory}. Springer Verlag, N. York, 1969. \bibitem{FH} Fusco, G. and Hale, J. K., ``Stable equilibria in a scalar parabolic equation with variable diffusion''. {\em SIAM J. Math. Anal.} 16, No. 6 (1985), 1152-1164. \bibitem{FiH} Fife, P. C. and Hsiao, L., ``The generation and propagation of internal layers''. {\em Nonlinear Anal., T., M. \& A.} 12, No. 1 (1988), 19-41. \bibitem{G} Giusti, E., {\em Minimal Surfaces and Functions of Bounded Variation}. Birkha\"user (1984). \bibitem{GT} Gilbarg, D. and Trudinger, N.S., ``Elliptic Partial Differential Equation of Second Order''. {\em A Series of Comp. Stud. in Math}, 224, Springer-Verlag (1983). \bibitem{GiF} Giorgi, E. De and Franzoni, T., ``Su un tipo di convergenza variazionale''. {\em Atti Accad. Naz. Lincei Rend. Cl. Sci. Fis. Mat. Natur.}, 58 (1975), 842-850. \bibitem{H} Henry, D., {\em Geometric Theory of Semilinear Parabolic Equations}. Springer-Verlag, 1984. Lectures Notes in Mathematics, 840. \bibitem{HR} Hale, J. K. and Rocha C., ``Bifurcations in a parabolic equation with variable diffusion''. {\em Nonlinear Anal., T., M. \& A.} 9, No. 5 (1985), 479-494. \bibitem{HS} Hale, J. K. and Sakamoto, K., ``Existence and stability of transition layers''. {\em Japan J. Appl. Math.} 5 (1988), 367-405. \bibitem{KS} Kohn, R.V. and Sternberg, P., ``Local minimizers and singular perturbations''. {\em Proceedings of the R. Soc. of Edinburgh}, {\bf 111} A (1989), 69-84. \bibitem{Ma} Maso, G. Dal, {\em An Introduction to $\Gamma$--convergence}. Birkha\"user, Boston (1993). \bibitem{M} Matano, H., ``Convergence of solutions of one-dimensional semilinear pa-rabolic equation''. {\em J. Math. Kyoto University} {\bf 18} (1978), 224-243. \bibitem{MM} Modica, L. and Mortola, S., ``Un esempio di $\Gamma$-convergenza''. {\em Boll. Un. Mat. Ital.}, 14-B (1977), 285-299. \bibitem{MM1} Modica, L. and Mortola, S., ``Il limite nella $\Gamma$-convergenza di una famiglia di funzionali ellittichi ''. {\em Boll. Un. Mat. Ital.},A(3) 14 (1997), 526-529. \bibitem{Mo} Modica, L., ``The gradient theory of phase transitions and minimal interface criterion''. {\em Arch. Rat. Mech. Anal.} {\bf 98}, No. 2 (1987), 123-142. \bibitem{MY} Miyata, S. and Yanagida, E., ``Stable Stationary Solution with Multiple Layers in a Scalar Parabolic Equation with Variable Diffusion''. {\em Funkcialaj Ekvacioj}, 38 (1995), 367-380. \bibitem{N1} Nascimento, A.S. Do, ``Bifurcation and stability of radially symmetric equilibria of a parabolic equation with variable diffusion''. {\em J. Diff. Eqns.}, vol. 77, No. 1 (1989), 84-103. \bibitem{N2} Nascimento, A.S. Do, ``Local Minimizers Induced by Spatial Inhomogeneity with Inner Transition Layer''. {\em J. Diff. Eqns.}, vol. 133, No. 2 (1997), 203-223. \bibitem{N3} Nascimento, A.S. Do, ``Stable Stationary Solutions Induced by Spatial Inhomogeneity Via $\Gamma$-Convergence''. To appear in Bulletin of the Brazilian Mathematical Society. \bibitem{S} Sternberg, P., ``The effect of a singular perturbation on nonconvex variational problems''. {\em Arch. Rat. Mech. Anal.}, {\bf 101} (1988), 209-260. \bibitem{Z} Ziemer, W.P., {\em Weakly Differentiable Functions}. Springer-Verlag (1989). \end{thebibliography} \bigskip {\sc Arnaldo Simal do Nascimento}\\ Universidade Federal de S\~ao Carlos, D.M. \\ 13565-905 - S\~ao Carlos, S.P. Brazil \\ E-mail address: dasn@power.ufscar.br \\ \end{document}