\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2006(2006), No. 49, pp. 1--10.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2006 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2006/49\hfil Nonlinear elastic membranes] {Nonlinear elastic membranes involving the p-Laplacian operator} \author[F. Cuccu, B. Emamizadeh, G. Porru\hfil EJDE-2006/49\hfilneg] {Fabrizio Cuccu, Behrouz Emamizadeh, Giovanni Porru} % in alphabetical order \address[F. Cuccu]{Mathematics Department\\ Universit\'a di Cagliari\\ Via Ospedale 72, 09124 Cagliari, Italy} \email{fcuccu@unica.it} \address[B. Emamizadeh]{Department of Mathematics\\ The Petroleum Institute\\ P. O. Box 2533, Abu Dhabi, UAE} \email{bemamizadeh@pi.ac.ae} \address[G. Porru]{Mathematics Department\\ Universit\'a di Cagliari\\ Via Ospedale 72, 09124 Cagliari, Italy} \email{porru@unica.it} \date{} \thanks{Submitted February 2, 2006. Published April 14, 2006.} \subjclass[2000]{35J20, 35B38, 49Q10} \keywords{$p$-Laplace; rearrangements; maximum principle; existence; \hfill\break\indent uniqueness; domain derivative} \begin{abstract} This paper concerns an optimization problem related to the Poisson equation for the $p$-Laplace operator, subject to homogeneous Dirichlet boundary conditions. Physically the Poisson equation models, for example, the deformation of a nonlinear elastic membrane which is fixed along the boundary, under load. A particular situation where the load is represented by a characteristic function is investigated. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \section{Introduction} Let $\Omega$ be a bounded smooth domain in $\mathbb{R}^N$. This paper is concerned with an optimization problem related to the Poisson boundary-value problem \begin{equation} \begin{gathered} -\Delta_pu=f,\quad \mbox{in }\Omega,\\ u=0,\quad \mbox{on }\partial\Omega. \end{gathered} \label{poisson} \end{equation} Here $p\in (1,\infty)$, and $\Delta_p$ stands for the usual $p$-Laplacian; that is, $\Delta_p u=\nabla\cdot (|\nabla u|^{p-2}\nabla u)$. Let $f_0\in L^q(\Omega)$, with $q=p/(p-1)$, and let $\mathcal{R}$ be the class of rearrangements of $f_0$. We are interested in finding \begin{equation} \sup_{f\in\mathcal{R}}\int_\Omega f\, u_f\, dx \label{maximization} \end{equation} where $u_f$ is the (unique) solution of (\ref{poisson}). The $p$-Laplace operator arises in various physical contexts: non Newtonian fluids, reaction diffusion problems, non linear elasticity, electrochemical machining, elastic-plastic torsional creep, etc., see \cite{AM}, \cite{HS}, and references therein. For a theoretical develop of the theory of the p-Laplacian we refer to the monograph \cite{HKM}. The case of $p=2$ is the most important and easier to discuss: it corresponds to a first approximation, the linear case. For non ideal materials, it is often appropriate to involve a power of the gradient $|\nabla u|$ to describe the law governing the model. For example, problem (\ref{poisson}) models a nonlinear elastic membrane under load $f$. The solution $u_f$ stands for the deformation of the membrane from the rest position. Therefore, the functional $\int_{\Omega}f\, u_f\, dx$ measures the average deformation, with respect to the measure $f\, dx,$ of the membrane. Thus, any solution to (\ref{maximization}) determines an {\em optimal} load chosen from the class $\mathcal{R}$. Our interest in (\ref{maximization}) spans questions such as existence, uniqueness (in case $\Omega$ is a ball), and qualitative properties of maximizers. In case of $p=2$, the problem is well understood, see \cite{B1}, \cite{B2}, \cite{BM}, \cite{CJP}, \cite{NE}. In this case the functional $\int_\Omega f\, u_f\, dx $ is weakly sequentially continuous and strictly convex, say on $L^2(\Omega)$, so the classical results of R. Burton are available to be applied to prove existence and some qualitative properties of the maximizers. However, in the case $p\neq 2$, we will use a method which does not need the convexity of the functional. The existence in a similar situation has been discussed in \cite{CP}. \section{Preliminaries} In this section we collect some well known results. Let us begin with the definition of a weak solution of (\ref{poisson}). \noindent {\bf Definition.} A function $u\equiv u_f\in W_0^{1,p}(\Omega)$ is a weak solution of (\ref{poisson}) provided \[ \int_\Omega |\nabla u|^{p-2}\nabla u\cdot\nabla v\,dx=\int_\Omega fv\,dx,\;\quad\forall \;v\in W_0^{1,p}(\Omega). \] It is a standard result that (\ref{poisson}) has a unique weak solution $u_f$, for which the following equations hold \begin{equation} \int_\Omega fu_f\,dx=\int_\Omega |\nabla u_f|^p\,dx =\frac{1}{p-1}\sup_{u\in W_0^{1,p}(\Omega)}\int_\Omega \bigl(pf\, u- |\nabla u|^p\bigr)dx. \label{identities} \end{equation} \noindent{\bf Definition.} Suppose $f:(X,\Sigma,\mu)\to\mathbb{R}^+$ and $g:(X ',\Sigma ',\mu ')\to\mathbb{R}^+$ are measurable functions. We say $f$ and $g$ are rearrangements of each other if and only if \[ \mu (\{x\in X|\;f(x)\geq\alpha\})=\mu ' (\{x\in X '|\;g(x)\geq\alpha\}),\quad\forall \alpha\geq 0. \] Henceforth we fix $f_0\in L_+^q(\Omega)$, with $q=p/(p-1)$. The set of all rearrangements of $f_0$ is denoted by $\mathcal{R}$. Thus, for any $f\in\mathcal{R}$, we have \[ \mathcal{L}_N(\{x\in\Omega|\;f(x)\geq\alpha\})=\mathcal{ L}_N(\{x\in\Omega|\;f_0(x)\geq\alpha\}),\quad\forall \alpha\geq 0, \] where $\mathcal{L}_N$ denotes the $N$-dimensional Lebesgue measure. For $f:\Omega\to\mathbb{R}^+$, $f^\Delta$ and $f^*$ denote the decreasing and Schwarz rearrangements of $f$, respectively. Recall that $f^\Delta$ is defined on $(0,\mathcal{ L}_N(\Omega))$, and $f^*$ is defined on $B$, the ball centered at the origin with volume equal to $\mathcal{L}_N(\Omega)$. \begin{lemma} \label{lem1} Let $q=p/(p-1)$, $f\in L^q_+(\Omega)$, $g\in L^p_+(\Omega)$. Suppose that every level set of $g$ $($that is, sets of the form $g^{-1}(\{\alpha\}))$, has measure zero. Then there exists an increasing function $\phi$ such that $\phi\circ g$ is a rearrangement of $f$. \end{lemma} \begin{lemma} \label{lem2} Suppose $\zeta\in L^p_+(\Omega)$, and $f\in L^q_+(\Omega)$. Suppose there exists an increasing function $\phi$ such that $\phi\circ \zeta\in \mathcal{R}(f)$, the set of rearrangements of $f$. Then $\phi\circ \zeta$ is the unique maximizer of the linear functional $\int_\Omega h\, \zeta\, dx$, relative to $h\in\overline {\mathcal{R}}(f)$, where $\overline{\mathcal{R}}(f)$ denotes the weak closure of $\mathcal{R}(f)$ in $L^q(\Omega)$. \end{lemma} \begin{lemma} \label{lem3} Suppose $g\in L^p_+(\Omega)$. Then there exists $\hat f\in \mathcal{ R}(f_0)$ which maximizes the linear functional $\int_\Omega hg\,dx$, relative to $h\in \overline{\mathcal{R}}(f_0)$; that is, \[ \int_\Omega \hat fg\,dx=\sup_{h\in \overline{\mathcal{R}}(f_0)}\int_\Omega hg\,dx. \] \end{lemma} For the proof of Lemma \ref{lem1}, see \cite[Lemma 2.4]{BM}, and for Lemmas \ref{lem2} and \ref{lem3}, see \cite[Lemma 2.4]{B2}. Next we recall a well known rearrangement inequality. If $u\in W^{1,p}_0(\Omega)$ is non-negative and if $u^*$ denotes the Schwarz rearrangement of $u$, then $u^*\in W^{1,p}_0(\Omega)$ and the inequality \begin{equation} \int_B|\nabla u^*|^p\,dx\leq \int_\Omega |\nabla u|^p\,dx \label{polya} \end{equation} holds. The case of equality in (\ref{polya}) has been considered in \cite{BZ}. The following result can be deduced from Lemma 3.2, Theorem 1.1, and Lemma 2.3(v), in \cite{BZ}. \begin{theorem} \label{thm1} Let $u\in W^{1,p}_0(\Omega)$ be non-negative, and suppose equality holds in $(\ref{polya})$. Then $u^{-1}(\alpha,\infty)$ is a translate of ${u^*}^{-1}(\alpha,\infty)$, for every $\alpha\in [0,M]$, where $M$ is the essential supremum of $u$ over $\Omega$, modulo sets of measure zero. Moreover, if \begin{equation} \mathcal{L}_N(\{x\in\Omega|\;\nabla u(x)=0,\;\;00\}$. Set \[ \gamma=\inf_{x\in S(\hat f)}\hat u(x),\quad \delta=\sup_{x\in \Omega\setminus S(\hat f)}\hat u(x). \] Then $\gamma\geq\delta$. \end{lemma} \begin{proof} To derive a contradiction assume $\gamma <\delta$. Let us fix $\gamma <\xi_1<\xi_2<\delta$. Since $\xi_1>\gamma$, there exists a set $A\subset S(\hat f)$, with positive measure, such that $\hat u\leq\xi_1$ on $A$. Similarly, $\xi_2<\delta$ implies that there exists a set $B\subset \Omega\setminus S(\hat f)$, with positive measure, such that $\hat u\geq\xi_2$ on $B$. Without loss of generality we may assume that $\mathcal{L}_N(A)=\mathcal{ L}_N(B)>0$ (otherwise we consider suitable subsets of $A$ and $B$ having the same measures). Next, consider a measure preserving map $T:A\to B$, see \cite{R}. Using $T$ we define a particular rearrangement of $\hat f$, denoted $\overline f$. \[ \overline f(x)=\begin{cases} \hat f(Tx), & x\in A \\ \hat f(T^{-1}x) & x\in B \\ \hat f(x) & \Omega\setminus (A\cup B). \end{cases} \] Thus \begin{align*} \int_\Omega \overline f\hat u\,dx-\int_\Omega\hat f\hat u\,dx &= \int_{A\cup B}\overline f\hat u\,dx-\int_{A\cup B}\hat f\hat u\,dx \\ &=\int_B\overline f\hat u\,dx-\int_A\hat f\hat u\,dx \\ &\geq \xi_2\int_B\overline fdx-\xi_1\int_A\hat fdx\\ &=(\xi_2-\xi_1)\int_A\hat f\,dx>0. \end{align*} Therefore, $\int_\Omega\overline f\hat u\,dx>\int_\Omega\hat f\hat u\,dx$, which contradicts the maximality of $\hat f$ (see Lemma \ref{lem4}). \end{proof} \begin{proof}[Proof of theorem \ref{thm3}] Notice that from \eqref{poisson} and \cite[Lemma 7.7]{GT}, it is clear that the level sets of $\hat u$, restricted to $S(\hat f)$, have measure zero. Therefore applying Lemma \ref{lem1}, we infer existence of an increasing function $\tilde\phi$ such that $\tilde\phi\circ\hat u$ is a rearrangement of $\hat f$ relative to the set $S(\hat f)$. Equivalently, $\tilde\phi\circ\hat u$, restricted to $S(\hat f)$, is a rearrangement of $\hat f^\Delta$, restricted to the interval $(0,s)$, where $s=\mathcal{L}_N(S(\hat f))$. Now, define \[ \phi (t)=\begin{cases} \tilde\phi (t) & t\geq\gamma \\ 0 & t<\gamma, \end{cases} \] where $\gamma=\inf_{S(\hat f)}\hat u(x)$. Note that, since $\tilde\phi$ is non-negative, $\phi$ is an increasing function. Moreover, $\phi\circ\hat u$ is a rearrangement of $\hat f^\Delta$, on $(0,\omega)$, where $\omega=\mathcal{L}_N(\Omega)$. Thus $\phi\circ\hat u\in \mathcal{R}(f_0)$, hence we can apply Lemma \ref{lem2} to deduce that $\phi\circ\hat u$ is the unique maximizer of the linear functional $\int_\Omega h\hat u\,dx$, relative to $h\in \mathcal{R}(f_0)$. This obviously implies $\hat f=\phi\circ\hat u$, almost everywhere in $\Omega$. \end{proof} \noindent{\bf Remark.} The function $\phi$ in above can be extended to all of $\mathbb{R}$. Thus from (\ref{euler}) we infer $S(\hat f)=\hat u^{-1}\bigl(\phi ^{-1}(0,\infty)\bigr)$. Since $\phi$ is increasing the set $\phi ^{-1}(0,\infty)$ is either the interval $(\gamma,\infty)$ or $[\gamma,\infty)$. In both cases, since the level sets of $\hat u$ on $S(\hat f)$ have measure zero, we can write $S(\hat f)=\{\hat u>\gamma\}$. If we assume $f_0\in L^\infty(\Omega)$ then we have the continuity of the solution $\hat u$ (see \cite{T}). In this situation the boundary of $S(\hat f)$, denoted $\partial S(\hat f)$, satisfies \begin{equation} \partial S(\hat f)\subset\{\hat u=\gamma\}, \label{boundary} \end{equation} thanks to the continuity of $\hat u$. Note that $\mathcal{L}_N(S(f_0))=\mathcal{L}_N(S(\hat f))$. Therefore, if $\mathcal{L}_N(S(f_0))<\mathcal{L}_N(\Omega)$, it follows that $\gamma$ is strictly positive. An example of interest is the following. \smallskip \noindent{\bf Example.} Suppose $f_0=\chi_{E_0}$, where $\chi_{E_0}$ is the characteristic function of the measurable set $E_0\subset\Omega$, and let $\mathcal{L}_N(E_0)<\mathcal{L}_N(\Omega)$. Denoting a solution of (\ref{maximization}) by $\hat f$, it is clear that $\hat f=\chi_{\hat E}$, for some measurable set $\hat E\subset\Omega$, having the same measure as $E_0$. From the last Remark we infer that $\hat u=u_{\hat f}$ is constant on $\partial\hat E$. Also, $\partial\hat E$ does not intersect $\partial\Omega$. So physically speaking, in order to maximize the average deformation of the nonlinear elastic membrane under uniform loads (given by the appropriate rearrangement class) it is best to place the load away from the boundary (independently of the geometry of the membrane). We will return to this example in the last section. We now address the question of uniqueness in a ball. \begin{theorem} \label{thm4} Suppose $\Omega$ is a ball centered at the origin. Then the maximization problem $(\ref{maximization})$ with $f_0$ non negative and essentially bounded has a unique solution, namely, $f_0^*$. \end{theorem} For the rest of this section $\Omega$ is always a ball. We need the following result. \begin{lemma} \label{lem6} If $f\geq 0$, then \begin{equation} \frac{1}{p}\int_\Omega|\nabla u_f^*|^p\,dx+\frac{1}{q}\int_\Omega f^*u_{f^*}\, dx\geq\int_\Omega f^*u_f^*\,dx. \label{ali7} \end{equation} \end{lemma} \begin{proof} From the variational characterization of $u_{f^*}$ the following inequality is clear \begin{align*} \frac{1}{p}\int_\Omega|\nabla u_f^*|^p\,dx + \frac{1}{q}\int_\Omega f^*u_{f^*}\, dx & \geq \frac{1}{p}\int_\Omega |\nabla u_{f^*}|^p\,dx-\int_\Omega f^*u_{f^*}\,dx \\ &\quad + \frac{1}{q}\int_\Omega f^* u_{f^*}\,dx+\int_\Omega f^*u_f^*\,dx. \label{ali8} \end{align*} Since the first three terms on the right hand side of the above inequality drop out thanks to (\ref{identities}), we obtain (\ref{ali7}). \end{proof} \begin{proof}[Proof of Theorem \ref{thm4}] Suppose $f$ is a solution of (\ref{maximization}). Then, from Lemma \ref{lem6} we have \[ \frac{1}{p}\int_\Omega|\nabla u_f^*|^p\,dx+\frac{1}{q}\int_\Omega f^*u_{f^*}\, dx\geq\int_\Omega f^*u_f^*\,dx. \] Applying the Hardy-Littlewood inequality $\int_\Omega f^*u_f^*\,dx\geq\int_\Omega fu_f\,dx$ to the right hand side of the above inequality we find \begin{equation} \frac{1}{p}\int_\Omega|\nabla u_f^*|^p\,dx+\frac{1}{q}\int_\Omega f^*u_{f^*}\,dx\geq\int_\Omega fu_f\,dx=\frac{1}{p}\int_\Omega fu_f\,dx+\frac{1}{q}\int_\Omega fu_f\,dx. \label{beh1} \end{equation} Since $f$ is a solution of (\ref{maximization}) we have $\int_\Omega fu_{f}\,dx\geq\int_\Omega f^*u_{f^*}\,dx$. Moreover, an application of (\ref{identities}) yields $\int_\Omega fu_{f}\,dx=\int_\Omega |\nabla u_f|^p\,dx.$ Therefore, (\ref{beh1}) implies \begin{equation} \int_\Omega |\nabla u_f^*|^p\,dx\geq \int_\Omega |\nabla u_f|^p\,dx. \label{beh2} \end{equation} Hence, from (\ref{polya}), we obtain equality in (\ref{beh2}). Next we show that $u_f=u_f^*$. According to Theorem \ref{thm1}, we only need to show that (\ref{measure}) holds. Let us consider $x\in\Omega$ such that $00$ in $S$. Therefore, $u(z)>u(x)$ for all $z\in S$. Hence $x$ must be a boundary point of $S$. So, by the Hopf boundary lemma \cite[Theorem 5]{V} we derive $\frac{\partial u}{\partial\nu}(x)=\frac{\partial v}{\partial\nu}(x)\neq 0$, where $\nu$ stands for the outward unit normal to $\partial S$ at $x$. Thus (\ref{measure}) holds, as desired. Finally, from (\ref{euler}), we deduce that $f$ coincides with its Schwarz rearrangement, so $f=f^*=f_0^*$. This completes the proof of the theorem. \end{proof} \section{Domain derivative} This section is devoted to the example mentioned earlier. We have seen that if $\hat f=\chi _{\hat D}$ is a solution of (\ref{maximization}), then $\hat u=u_{\hat f}$ is constant on the free boundary $\partial\hat D$. A natural question arises: Does the same result hold if $\chi _{\hat D}$ is {\em any} critical point of the functional $\int_\Omega fu_f$, relative to the class of rearrangements of $\chi_{\hat D}$? We give an affirmative answer to this question under some restrictions on $\hat D$. In order to put things in the right context we need to introduce the notion of domain derivative \cite{S1}, \cite{S2}, that is specialized to our situation. Let $D$ be an open smooth subset of $\Omega$ with $\mathop{\rm dist}(D,\partial\Omega)>0$. Let $V$ be a regular (smooth) vector field with support in $\Omega$. Define $D^t=(Id +tV)(D)$, with small $t\in R^+$ such that $D^t\subset\Omega$. Here $Id$ denotes the identity map. Note that for small $t$, the operator $Id +tV$ is a diffeomorphism. In particular, $D^t$ is an open and smooth set. If $D\Delta D^t$ denotes the familiar symmetric difference of $D$ and $D^t$, then \begin{equation} \mathcal{L}_N(D\Delta D^t)\leq ct, \label{symmetric} \end{equation} where $c$ is a positive constant independent of $t$. As a consequence of (\ref{symmetric}), the function $\chi_{D^t}-\chi_D$ tends to zero in $L^q(\Omega)$ (for any $q>1$) as $t$ tends to zero. Let us define \[ I(D)=\int_Du\,dx, \] where $u$ satisfies \begin{equation} -\Delta_pu=\chi_D\quad\mbox{in }\Omega, \quad u=0\quad \mbox{on }\partial\Omega. \label{poisson2} \end{equation} Also \[ I(D^t)=\int_{D^t}u^t\,dx, \] where $u^t$ satisfies \begin{equation} -\Delta_pu^t=\chi_{D^t}\quad \mbox{in }\Omega,\quad u^t=0\quad \mbox{on }\partial\Omega. \label{poisson3} \end{equation} For the sake of the following definition we introduce $\mathcal{V}$ to be the set of all regular vector fields with support in $\Omega$. \smallskip \noindent{\bf Definition.} We say that $D$ (as above) is a critical point of the functional $I$ provided \[ dI(D;V)=c\;dVol (D;V), \] for some constant $c$ and every $V\in \mathcal{V}$. Here \begin{gather*} \mathop{\rm dI}(D;V):=\lim_{t\to 0^+}\frac{I(D^t)-I(D)}{t},\\ \mathop{\rm dVol} (D;V):=\lim_{t\to 0^+}\frac{\mathcal{L}_N(D^t) -\mathcal{L}_N(D)}{t}. \end{gather*} Of course, if we consider measure preserving vector fields $V$ then $dVol (D;V)=0$. We are now ready to state the main result of this section. \begin{theorem} \label{thm5} Suppose $D$ is an open smooth subset of $\Omega$. Suppose $\mathop{\rm dist}(D,\partial\Omega)>0$, and $D$ is a critical point of $I$, relative to $\Gamma$. Then $u$, the solution of \eqref{poisson2}, is constant on $\partial D$. \end{theorem} \begin{lemma} \label{lem7} Let $u$ and $u^t$ be the solutions of \eqref{poisson2} and \eqref{poisson3}, respectively. Then $u^t\to u$, in $W^{1,p}_0(\Omega)$, as $t\to 0^+$. \end{lemma} \begin{proof} Let us recall the well known inequality (see, for example, \cite{T}) \begin{equation} \label{simon} C(|X|^{p-2}X-|Y|^{p-2}Y,X-Y)\geq \begin{cases} |X-Y|^p, & p\geq 2 \\ \frac{|X-Y|^2}{(|X|+|Y|)^{2-p}}, & p\leq 2, \end{cases} \end{equation} where $X$ and $Y$ are vectors in $\mathbb{R}^n$, $|X|$ (similarly $|Y|$) denotes the Euclidean length of $X$, $C$ is a positive constant, and $(\cdot,\cdot)$ stands for the usual dot product in $\mathbb{R}^n$. Let us consider two cases \noindent {\em Case 1: $p\geq 2$}: Using (\ref{simon}) we have \[ \|\nabla u^t-\nabla u\|_p^p\leq C\int_\Omega (|\nabla u^t|^{p-2}\nabla u^t-|\nabla u|^{p-2}\nabla u)\cdot (\nabla u^t-\nabla u)\,dx. \] Using (\ref{poisson2}) and (\ref{poisson3}) we can rewrite the above inequality as \[ \|\nabla u^t-\nabla u\|_p^p\leq C \int_\Omega (\chi_{D^t}-\chi_D)(u^t-u)\,dx. \] So by applying the H\"older's inequality followed by the Poincar\'e inequality we obtain \[ \|\nabla u^t-\nabla u\|_p^{p-1}\leq \tilde C\Big(\int_\Omega |\chi_{D^t}-\chi_D|^q\,dx\Big)^{1/q}. \] >From the above inequality, the assertion of the lemma follows. \noindent{\em Case 2: $p\leq 2$}: Let us begin with the following observation \begin{align*} \|\nabla u^t-\nabla u\|^p_p &= \int_\Omega\frac{|\nabla u^t-\nabla u|^p}{(|\nabla u^t|+|\nabla u|)^{\frac{p(2-p)}{2}}}\;(|\nabla u^t|+|\nabla u|)^{\frac{p(2-p)}{2}}\,dx \\ &\leq \Big(\int_\Omega\frac{|\nabla u^t-\nabla u|^2}{(|\nabla u^t|+|\nabla u|)^{(2-p)}}\,dx\Big)^{p/2} \Big(\int_\Omega (|\nabla u^t|+|\nabla u|)^p\,dx\Big)^{(2-p)/2}, \end{align*} which follows from the H\"older inequality, since $2/p>1$. Note that $(u^t)$ is bounded in $W^{1,p}_0(\Omega)$. Thus from the above inequality we find \begin{equation} \|\nabla u^t-\nabla u\|^p_p \leq C\left(\int_\Omega\frac{|\nabla u^t-\nabla u|^2}{(|\nabla u^t|+|\nabla u|)^{(2-p)}} \,dx\right)^{p/2}. \label{beh4} \end{equation} Now applying (\ref{simon}) to the right hand side of (\ref{beh4}), the assertion of the lemma can be confirmed using similar arguments as in the ending part of Case 1. \end{proof} \begin{proof}[Proof of Theorem \ref{thm5}] Let us begin with the identity \begin{equation} I(D^t)-I(D)=\int_{D^t}(u^t-u)\,dx+\int_{D^t}u\,dx-\int_Du\,dx. \label{porru1} \end{equation} Following \cite{S1}, we define \begin{equation} u' (x)=\lim_{t\to 0^+}\frac{u^t(x)-u(x)}{t}. \label{porru2} \end{equation} Moreover, we have \[ \int_{D^t}u\,dx-\int_D u\,dx=\int_D\Big[ u(x+tV)\big|\det \big(\delta_{ij} +t\frac{\partial V_i}{\partial x_j}\big)\big|-u(x)\Big]\,dx. \] Since $t$ is small, we find \begin{equation} \begin{aligned} \int_{D^t}u\,dx-\int_D u\,dx &= \int_D[(u(x)+t\nabla u\cdot V+o(t))(1+t\nabla\cdot V+o(t))-u(x)]\,dx \\ &= t\int_D(\nabla u\cdot V+u\nabla\cdot V)\,dx+o(t) \\ &= t\int_D\nabla\cdot (uV)\,dx+o(t) \\ &= t\int_{\partial D}u(V\cdot\nu)\,d\sigma +o(t), \end{aligned}\label{porru3} \end{equation} where $\nu$ is the outward unit normal to $\partial D$ and $d\sigma$ is the surface measure. Inserting (\ref{porru3}) and (\ref{porru2}) into (\ref{porru1}) yields \begin{equation} \lim_{t\to 0^+}\frac{I(D^t)-I(D)}{t} =\int_{D}u'\,dx+\int_{\partial D}u(V\cdot\nu)\,d\sigma. \label{derivative} \end{equation} Now multiply (\ref{poisson2}) by $u^t$, (\ref{poisson3}) by $u$, subtract the new equations, and finally integrate over $\Omega$. We find \begin{equation} \int_\Omega\frac{|\nabla u^t|^{(p-2)}-|\nabla u|^{(p-2)}}{t}\;\nabla u^t\cdot\nabla u\,dx =\frac{1}{t}\Big[\int_{D^t}u\,dx-\int_Du\,dx\Big] +\int_D\frac{u-u^t}{t}\,dx. \label{porru4} \end{equation} Since $u^t$ tends to $u$ in $W^{1,p}_0(\Omega)$ as $t$ tends to zero, and also \[ \frac{d}{dt}|\nabla u^t|^{(p-2)}=(p-2)|\nabla u|^{(p-4)} \nabla u\cdot\nabla u',\quad \mbox{at } t=0, \] taking the limit of (\ref{porru4}), when $t$ tends to zero, we find \begin{equation} (p-2)\int_\Omega |\nabla u|^{(p-2)}\nabla u\cdot\nabla u'\,dx=\int_{\partial D}u(V\cdot\nu)\,d\sigma-\int_Du'\,dx. \label{porru5} \end{equation} If we multiply (\ref{poisson2}) by $u'$, and integrate we find (recall that the support of $V$ is in $\Omega$) \[ \int_\Omega |\nabla u|^{p-2}\nabla u\cdot\nabla u'\,dx=\int_Du'\,dx. \] Inserting this equation in (\ref{porru5}) we find \[ \int_Du'\,dx=\frac{1}{p-1}\int_{\partial D}u(V\cdot\nu)\,d\sigma. \] Finally inserting the latter estimate into (\ref{derivative}) yields \[ dI(D,V)=q\int_{\partial D}u(V\cdot\nu)\,d\sigma. \] Recalling the formula for the derivative of the volume, that is, \[ \mathop{\rm dVol}(D,V)=\int_{\partial D}(V\cdot\nu)\,d\sigma, \] and the fact that $D$ is a critical point of $I$, we derive \[ \mathop{\rm dI}(D,V)=c\mathop{\rm dVol} (D,V) \Leftrightarrow u(x)=\mbox{constant},\quad \mbox{on }\partial D. \] This obviously completes the proof of the theorem. \end{proof} \subsection*{Acknowledgement} This work was initiated when B. Emamizadeh visited University of Cagliari. He wants to thank Professors G. Porru and F. Cuccu for their great hospitality. He also thanks the Petroleum Institute for its support during this visit. \begin{thebibliography}{99} \bibitem{AM} A. Acker and R. Meyer, A free boundary problem for the p-Laplacian: uniqueness, convexity and successive approximation of solutions. {\em Electr. J. Diff. Equ.} Vol. 1995, no. 8: 1-20 (1995). \bibitem{BZ} J. E. Brothers and W. P. Ziemer, Minimal rearrangements of Sobolev functions. {\em J. Reine Angew. Math}. 384: 153-179 (1988). \bibitem{B1} G. R. Burton, Rearrangements of functions, maximization of convex functionals and vortex rings. {\em Math Ann}. 276: 225-253 (1987). \bibitem{B2} G. R. Burton, Variational problems on classes of rearrangements and multiple configurations for steady vortices. {\em Ann Inst Henri Poincare}. 6(4): 295-319 (1989). \bibitem{BM} G. R. Burton and J. B. McLeod, Maximisation and minimisation on classes of rearrangements. {\em Proc. Roy. Soc. Edinburgh Sect. A}. 119 (3-4): 287--300 (1991). \bibitem{CJP} F. Cuccu, K. Jha and G. Porru, Geometric properties of solutions to maximisation problems. {\em Electr. J. Diff. Equ.} 71: 1-8 (2003). \bibitem{CP} F. Cuccu and G. Porcu, Existence of solutions in two optimization problems. {\em Comp. Rend. de l'Acad. Bulg. des Sciences.} 54(9): 33-38 (2001). \bibitem{GT} D. Gilbarg and N. Trudinger, Elliptic partial differential equations of second order, Springer-Verlag, Berlin, 1977. \bibitem{HKM} J. Heinonen, T. Kilpel\"ainen and O. Martio, Nonlinear potential theory of degenerate elliptic equations, Oxford Mathematical Monographs, Oxford, 1993. \bibitem{HS} A. Henrot and H. Shahgholian, Existence of classical solutions to a free boundary problem for the p-Laplace operator: (I) the exterior case. {\em J. Reine Angew. Math}. 521: 85-97 (2000). \bibitem{NE} J. Nycander and B. Emamizadeh, Variational problems for vortices attached to seamounts. {\em Nonlinear Analysis}. 55: 15-24 (2003). \bibitem{R} H. L. Royden, Real analysis. Third edition. Macmillan Publishing Company, New York, 1988. \bibitem{S1} J. Simon, Differentiation with respect to the domain in boundary-value problems. {\em Numer. Funct. Anal. Optim}. 2(7-8): 649-687 (1980). \bibitem{S2} J. Simon, Regularit\'e de la solution d'un probleme aux limites non lineaires, {\em Annales Fac. Sci. Tolouse Math}. 6: 247-274 (1981). \bibitem{T} P. Tolksdorf, Regularity for more general class of quasilinear elliptic equations. {\em Journal of Differential Equations}. 51: 126-150 (1984) \bibitem{V} J. L. Vazquez, A strong maximum principle for some quasilinear elliptic equations. {\em Appl. Math. Optim.} 191-202 (1984). \end{thebibliography} \end{document}