\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2013 (2013), No. 108, 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} \thanks{\copyright 2013 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2013/108\hfil Symmetry and regularity] {Symmetry and regularity of an optimization problem related to a nonlinear BVP} \author[C. Anedda, F. Cuccu\hfil EJDE-2013/108\hfilneg] {Claudia Anedda, Fabrizio Cuccu} % in alphabetical order \address{Claudia Anedda \newline Dipartimento di Matematica e Informatica, Universit\'a di Cagliari, Via Ospedale 72, 09124 Cagliari, Italy} \email{canedda@unica.it} \address{Fabrizio Cuccu \newline Dipartimento di Matematica e Informatica, Universit\'a di Cagliari, Via Ospedale 72, 09124 Cagliari, Italy} \email{fcuccu@unica.it} \thanks{Submitted January 7, 2013. Published April 29, 2013.} \subjclass[2000]{35J20, 35J60, 40K20} \keywords{Laplacian; optimization problem; rearrangements; Steiner symmetry; \hfill\break\indent regularity} \begin{abstract} We consider the functional $$ f\mapsto\int_\Omega \big(\frac{q+1}{2} |Du_f|^2-u_f|u_f|^q f\big) dx, $$ where $u_f$ is the unique nontrivial weak solution of the boundary-value problem $$ -\Delta u=f|u|^q\quad \text{in }\Omega,\quad u\big|_{\partial\Omega}=0, $$ where $\Omega\subset\mathbb{R}^n$ is a bounded smooth domain. We prove a result of Steiner symmetry preservation and, if $n=2$, we show the regularity of the level sets of minimizers. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{definition}[theorem]{Definition} \newtheorem{remark}[theorem]{Remark} \allowdisplaybreaks \section{Introduction} Let $\Omega$ be a bounded domain $\Omega\subset \mathbb{R}^n$ with smooth boundary. We consider the Dirichlet problem \begin{equation}\label{a1} \begin{gathered} -\Delta u=f|u|^q \quad\text{in }\Omega,\\ u\big|_{\partial\Omega}=0, \end{gathered} \end{equation} where $0\leq q<\min\{1,4/ n\}$ and $f$ is a nonnegative bounded function non identically zero. We consider nontrivial solutions of \eqref{a1} in $H^1_0(\Omega)$. The equation \eqref{a1} is the Euler-Lagrange equation of the integral functional $$ v\mapsto\int_\Omega \Big(\frac{q+1}{2} |Dv|^2-v|v|^q f\Big) dx, \quad v\in H_0^{1} (\Omega). $$ By using a standard compactness argument, it can be proved that there exists a nontrivial minimizer of the above functional. This minimizer is a nontrivial solution of \eqref{a1}. From the maximum principle, every nontrivial solution of \eqref{a1} is positive. Then, by \cite[Theorem 3.2]{CPS} the uniqueness of problem \eqref{a1} follows. To underscore the dependence on $f$ of the solution of \eqref{a1}, we denote it by $u_f$. Moreover, $u_f\in W^{2,2}(\Omega)\cap C^{1,\alpha}(\overline{\Omega})$ for all $\alpha$, $ 0<\alpha<1$ (see \cite{GT,T}). Let $f_0$ be a fixed bounded nonnegative function. We study the problem \begin{equation}\label{a2} \inf_{v\in H_0^{1}(\Omega),\, f\in\mathfrak{F}(f_0)} \int_\Omega \Big(\frac{q+1}{2} |Dv|^2-v|v|^q f\Big) dx, \end{equation} where, denoting by $|A|$ the Lebesgue measure of a set $A$, \begin{equation}\label{a5} \mathfrak{F}(f_0)=\{f\in L^\infty(\Omega): |\{f\geq c\}| =|\{f_0\geq c\}|\;\forall c\in\mathbb{R}\}; \end{equation} here $\mathfrak{F}(f_0)$ is called class of rearrangements of $f_0$ (see \cite{K}). Problems of this kind are not new; see for example \cite{ATL,CGIKO,CGK,CPS}. From the results in \cite{CPS} it follows that \eqref{a2} has a minimum and a representation formula. Let \begin{equation} E(f)=\inf_{v\in H_0^1(\Omega)} \int_\Omega\Big(\frac{q+1}{2} |Dv|^2-v|v|^q f\Big) dx. \end{equation} Renaming $q'$ the constant $q$ and putting $p=2$ and $q'=q+1$ in \cite{CPS} we have $$ E(f)=\frac{q'-2}{2}\,I(f), $$ where $$ I(f)=\sup_{H^1_0(\Omega)}\frac{q'}{2-q'}\int_\Omega \Big(\frac{2}{q'} f|v|^{q'}-|Dv|^2\Big) dx $$ is defined in the same paper. By \cite[Theorem 2.2]{CPS} it follows that there exist minimizers of $E(f)$ and that, if $\overline{f}$ is a minimizer, there exists an increasing function $\phi$ such that \begin{equation}\label{a6} \overline{f}=\phi(u_{\overline{f}}). \end{equation} We denote by $\operatorname{supp}f$ the support of $f$, and we call a level set of $f$ the set $\{x\in\Omega: f(x)>c\}$, for some constant $c$. In Section 2, we consider a Steiner symmetric domain $\Omega$ and $f_0$ bounded and nonnegative, such that $|\operatorname{supp}f_0|<|\Omega|$. Under these assumptions, we prove that the level sets of the minimizer $\overline{f}$ are Steiner symmetric with respect to the same hyperplane of $\Omega$. As a consequence, we have exactly one optimizer when $\Omega$ is a ball. Chanillo, Kenig and To \cite{CKT} studied the regularity of the minimizers to the problem $$ \lambda(\alpha, A)=\inf_{u\in H_0^1(\Omega),\, \|u\|_2, |D|=A} \int_\Omega|Du|^2\,dx+\alpha\int_D u^2\,dx, $$ where $\Omega\subset\mathbb{R}^2$ is a bounded domain, $00$. In particular they prove that, if $D$ is a minimizer, then $\partial D$ is analytic. In Section 3, following the ideas in \cite{CKT}, we give our main result. We restrict our attention to $\Omega\subset\mathbb{R}^2$. Let $b_1,\ldots, b_m>0$ and $0c_1\},\quad D_2=\{u_{\overline{f}}>c_2\},\quad \dots, \quad D_m=\{u_{\overline{f}}>c_m\}, $$ for suitable constants $c_1>c_2>\dots>c_m>0$. We show regularity of $\partial D_i$ for each $i$ proving that $|Du_{\overline{f}}|>0$ in $\partial D_i$. Following the method used in \cite{CKT}, we consider \begin{equation}\label{1.3} E(s,t)=\int_{\Omega}\Big(\frac{q+1}{2} |Du_{\overline{f}}+sDv|^2-(u_{\overline{f}}+sv) |u_{\overline{f}}+sv|^q{\overline{f}_t} \Big)dx-\eta, \end{equation} where $v\in H_0^{1}(\Omega)$, $\overline{f}_t$ is a family of functions such that $\overline{f}_t\in\mathfrak{F}(f_0)$ with $\overline{f}_0=\overline{f}$, and $s\in \mathbb{R}$. We have $$ E(s,t)\geq E(0,0)=0\quad \forall s, t. $$ Therefore, $(s,t)=(0,0)$ is a minimum point; it follows that \begin{equation}\label{1.4} \begin{vmatrix} \frac{\partial^2 E}{\partial s^2} (0,0) & \frac{\partial^2 E}{\partial s \partial t} (0,0) \\ \noalign{\vskip10pt}\frac{\partial^2 E}{\partial t \partial s} (0,0) & \frac{\partial^2 E}{\partial t^2} (0,0)\end{vmatrix} \geq 0. \end{equation} Expanding \eqref{1.4} in detail and using some lemmas from \cite{CKT} we prove that the boundaries of level sets of $\overline{f}$ are regular. \begin{theorem} \label{thm1} Let $\Omega\subset\mathbb{R}^2$, $f_0=\sum_{i=1}^m b_i\chi_{G_i}$ with $m\geq 2$, and $\overline{f}=\sum_{i=1}^m b_i\chi_{D_i}$ a minimizer of \eqref{a2}. Then $|Du_{\overline{f}}|>0$ on $\partial D_i$, $i=1,\ldots,m$. \end{theorem} \section{Symmetry} In this section we consider Steiner symmetric domains. We prove that, under suitable conditions on $f_0$ in \eqref{a5}, minimizers inherit Steiner symmetry. \begin{definition} \rm Let $P\subset\mathbb{R}^n$ be a hyperplane. We say that a set $A\subset\mathbb{R}^n$ is \emph{Steiner symmetric} relative to the hyperplane $P$ if for every straight line $L$ perpendicular to $P$, the set $A\cap P$ is either empty or a symmetric segment with respect to $P$. \end{definition} To prove the symmetry, we need \cite[Theorem 3.6 and Corollary 3.9]{F}, that, for more convenience for the reader, we state here. These results are related to the classical paper \cite{GNN}. \begin{theorem}\label{t1} Let $\Omega\subset\mathbb{R}^n$ be bounded, connected and Steiner symmetric relative to the hyperplane $P$. Assume that $u:\overline{\Omega} \to\mathbb{R}$ has the following properties: \begin{itemize} \item $u\in C(\overline{\Omega})\cap C^1(\Omega)$, $u>0$ in $\Omega$, $u|_{\partial\Omega}=0$; \item for all $\phi\in C^\infty_0(\Omega)$, \begin{equation*} \int_\Omega Du\cdot D\phi\,dx=\int_\Omega\phi F(u)\,dx, \end{equation*} where $F$ has a decomposition $F=F_1+F_2$ such that $F_1: [0,\infty)\to\mathbb{R}$ is locally Lipschitz continuous, while $F_2:[0,\infty)\to\mathbb{R}$ is non-decreasing and identically $0$ on $[0,\epsilon]$ for some $\epsilon>0$. \end{itemize} Then $u$ is symmetric with respect to $P$ and $\frac{\partial u} {\partial\mathbf{v}}(x)<0$, where $\mathbf{v}$ is a unit vector orthogonal to $P$ and $x$ belongs to the part of $\Omega$ that lies in the halfspace (with origin in $P$) in which $\mathbf{v}$ points. \end{theorem} \begin{theorem} Let $\Omega$ be Steiner symmetric and $f_0$ a bounded nonnegative function. If $|\operatorname{supp}f_0|<|\Omega|$ and $\overline{f}\in\mathfrak{F}(f_0)$ is a minimizer of \eqref{a2}, then the level sets of $\overline{f}$ are Steiner symmetric with respect to the same hyperplane of $\Omega$. \end{theorem} \begin{proof} Let $u=u_{\overline{f}}$ be the solution of \eqref{a2}. Then $u\in C^0(\overline{\Omega})\cap C^1(\Omega)$ and satisfies \begin{equation*} \int_{\Omega}D u\cdot D\psi\,dx=\int_{\Omega}\psi fu^q\,dx\quad \forall \psi\in C_0^\infty(\Omega). \end{equation*} Since $u>0$ and since (from \eqref{a6}) $\overline{f}=\phi(u)$ with $\phi$ increasing function, it follows that $\phi(u)\equiv 0$ on $\{x\in\Omega: u(x)0$ on $\partial D_i$, $i=1,\ldots,m$. \end{theorem} We use the notation introduced in Section 1. Without loss of generality we can assume $m=3$; the general case easily follows. Let $\overline{f}=b_1\chi_{D_1}+b_2\chi_{D_2}+b_3\chi_{D_3}$. We will prove $|Du_{\overline{f}}|>0$ in $\partial D_2$; we omit the proof for $D_1$ and $D_3$ because it is similar. We define the family $\overline{f}_t$ by replacing only the set $D_2$ by a family of domains $D_2(t)$. First of all, we explain how to define the family $D_2(t)$.\\ In the sequel we use the notation introduced in \cite{CKT}, reorganized according to our needs. We call a curve $\gamma:[a,b]\to\mathbb{R}$, $-\infty0\}. $$ Let $J=\cup_{k=1}^pJ_k$ be a finite union of open bounded intervals $J_k\subset \mathbb{R}$, $\gamma= (\gamma_1, \gamma_2):J\to \mathcal{F}^*$ a simple curve which is regular on each interval $J_k$ and $\overline{\gamma(J)}\subset\mathcal{F}^*$. We suppose that $\operatorname{dist}\big(\gamma(J_k), \gamma(J_h)\big)>0$ for $1\leq h\neq k\leq p$. Assume also that $|\gamma'|\geq \theta$ on $J$. For each $\xi \in J$, we denote by $\mathbf{N}(\xi)= \big(N_1(\xi), N_2(\xi)\big)$ the outward unit normal with respect to $D_2$ at $\gamma(\xi)$. We also define the tangent vector to $\gamma$ $\mathbf{N}^\bot(\xi)=\big(-N_2(\xi),N_1(\xi)\big)$, and $\mathbf{N}'$ the first derivative of $\mathbf{N}$. Reversing the direction of $\gamma$ if necessary, we will assume, without loss of generality, that $\gamma'$ and $\mathbf{N}^\bot$ have the same direction; i.e., $\angle \gamma', \mathbf{N}^\bot\rangle =|\gamma'|$. We observe that, because $\gamma$ is $C^2$ and simple on $\overline{J_k}$, for each $k$ there exists $\beta_k>0$ such that the function $$ \phi_k:J_k \times[-\beta_k, \beta_k]\to\mathbb{R}^2, \; (\xi,\beta)\mapsto (x_1,x_2)=\phi_k(\xi, \beta)=\gamma(\xi)+\beta \mathbf{N}(\xi) $$ is injective. Because $\operatorname{dist}\big(\gamma(J_k), \gamma(J_h)\big)>0$ for all $h\neq k$, we can find a number $\beta_0>0$ and we can paste together the functions $\phi_k$ to obtain a function $\phi$ injective on $J\times [-\beta_0, \beta_0]$. Choose $\beta_0$ such that $\operatorname{dist}\big(\phi(J\times [-\beta_0, \beta_0]),\partial D_1\big)>0$ and $\operatorname{dist}\big(\phi(J\times [-\beta_0, \beta_0]),\partial D_3\big)>0$. Now, we define $$ K= D_2\setminus \phi\big(J\times (-\beta_0, 0]\big); $$ for $t\in(-t_0,t_0)$ we define \begin{equation}\label{2.1} D_2(t)=K\cup\{\phi(\xi,\beta): \xi \in J, \, \beta < g(\xi, t)\}, \end{equation} where $g: J\times (-t_0,t_0)\to\mathbb{R}$, $t_0>0$, is a function such that \begin{equation}\label{c1} g(\xi, t),\; g_t(\xi, t),\; g_{tt}(\xi, t)\in C(\overline{J})\quad \forall t\in (-t_0,t_0) \end{equation} and \begin{equation}\label{2.2} g(\xi, 0)\equiv 0\quad \forall \xi\in J. \end{equation} We observe that $D_2(0)=D_2$. Next we compute the measure of $D_2(t)$. Put $A(t)=|D_2(t)|$ and $A=|D_2(0)|=|D_2|$; we have $$ A(t)= |D_2| +\int_J \int_0^{g(\xi,t)} J(\xi, \beta) d\beta d\xi, $$ where \begin{align*} J(\xi, \beta) &= \frac{\partial(x_1, x_2)}{\partial(\xi,\beta)}=\begin{vmatrix} \gamma_1'+\beta N_1' & N_1 \\ \noalign{\vskip10pt}\gamma_2'+\beta N_2' & N_2 \end{vmatrix}\\ &= \big| -\langle\gamma', \mathbf{N}^\bot\rangle -\beta\langle \mathbf{N}',\mathbf{N}^\bot\rangle\big| =\big| |\gamma'|+\beta\langle \mathbf{N}',\mathbf{N}^\bot\rangle\big|. \end{align*} We show that $ |\gamma'|+\beta\langle \mathbf{N}',\mathbf{N}^\bot\rangle\geq 0$. Indeed, from the fact that $\|\gamma\|_{C^2(J)}<\infty$, we have $\|\langle\mathbf{N}',\mathbf{N}^\bot\rangle\|_{L^\infty(J)}<\infty$. Substituting $t_0$ by a smaller positive number if necessary, we can assume that $$\| g\|_{L^\infty(J\times(-t_0,t_0))}<\beta_0$$ and $$\|\langle\mathbf{N}',\mathbf{N}^\bot\rangle\|_{L^\infty(J)}\ \|g\|_{L^\infty(J\times(-t_0,t_0))}<\theta.$$ Note that the first of these assumptions guarantees that $\partial D_2(t)$ has positive distance from $\partial D_1$ and $\partial D_3$. We have $$ |\beta| \left| \langle \mathbf{N}',\mathbf{N}^\bot\rangle\right| \leq \|g\|_{L^\infty(J\times(-t_0,t_0))} \|\langle \mathbf{N}',\mathbf{N}^\bot\rangle\|_{L^\infty(J)} \leq \theta\leq |\gamma'| $$ for all $\xi\in J$ and $|\beta|\leq \|g\|_{L^\infty(J\times(-t_0,t_0))}$. Thus, $J(\xi,\beta) = |\gamma'|+\beta\langle \mathbf{N}',\mathbf{N}^\bot\rangle$. Substituting into the formula for $A(t)$ we have \begin{align*} A(t)&=A+\int_J \int_0^{g(\xi, t)} ( |\gamma'|+\beta \langle \mathbf{N}',\mathbf{N}^\bot\rangle )d\beta d\xi\\ &= A+\int_J\Big(g(\xi, t)|\gamma'|+\frac{1}{2} (g(\xi, t))^2 \langle \mathbf{N}',\mathbf{N}^\bot\rangle \Big)d\xi. \end{align*} To obtain $|D_2(t)|=|D_2|$ for all $t\in (-t_0,t_0)$, we find the further constraint on $g$: \begin{equation}\label{2.3} \int_J\Big(g(\xi, t) |\gamma'| +\frac{1}{2}\ \big(g(\xi, t)\big)^2 \langle \mathbf{N},\mathbf{N}^\bot\rangle \Big) d\xi=0 \quad \forall t\in (-t_0,t_0). \end{equation} Moreover, we calculate the derivatives of $A(t)$, that we will use later. \begin{equation}\label{2.9} \begin{gathered} A'(t)=\int_J \big(g_t(\xi, t) |\gamma'(\xi)|+g(\xi, t)g_t(\xi, t) \langle \mathbf{N}',\mathbf{N}^\bot\rangle \big) d\xi=0; \\ A''(t)=\int_J \big(g_{tt}(\xi, t) |\gamma'(\xi)|+\big(g(\xi, t)g_{tt}(\xi, t)+g^2_t(\xi, t)\big) \langle \mathbf{N}',\mathbf{N}^\bot\rangle \big) d\xi=0. \end{gathered} \end{equation} Once we have defined the family $D_2(t)$, we can go back to the functional \eqref{1.3}. The following lemma describes \eqref{1.4} with $\overline{f}_t=b_1\chi_{D_1}+b_2\chi_{D_2(t)}+b_3\chi_{D_3}$. We find an inequality corresponding to \cite[(2.3) of Lemma 2.1]{CKT}. \begin{lemma}\label{lem1} Let $\overline{f}_t=b_1\chi_{D_1}+b_2\chi_{D_2(t)}+b_3\chi_{D_3}$, where the variation of domain $D_2(t)$ is described by \eqref{2.1} and $g: J\times (-t_0,t_0)\to\mathbb{R}$, $t_0>0$, satisfies \eqref{c1}, \eqref{2.2} and \eqref{2.3}. Then, for all $v\in H_0^{1}(\Omega)$, the conditions \eqref{1.4} becomes \begin{equation}\label{a8} \begin{aligned} &\int_\Omega\big( |Dv|^2-qu_{\overline{f}}\ v^2|u_{\overline{f}}|^{q-2}\overline{f}\big) dx\cdot\int_\gamma g_{t}^2(\gamma^{-1},0) |Du_{\overline{f}}| d\sigma \\ &\geq b_2c_2^q\Big(\int_\gamma g_{t}(\gamma^{-1},0)\ v d\sigma\Big)^2. \end{aligned} \end{equation} \end{lemma} \begin{proof} We calculate the second derivative of the functional \eqref{1.3}, with respect to $s$. We have $$ \frac{\partial E}{\partial s} =(q+1) \int_\Omega\big(\langle Du_{\overline{f}}+sDv, Dv\rangle -v|u_{\overline{f}}+sv|^q \overline{f}_t\big)dx $$ and \begin{equation} \label{2.5} \frac{\partial^2 E}{\partial s^2} (0,0) = (q+1)\int_\Omega\big( |Dv|^2-qu_{\overline{f}}\ v^2|u_{\overline{f}}|^{q-2}\overline{f}\big) dx. \end{equation} Before calculating the second derivative of $E$ with respect to $t$, we rewrite \eqref{1.3} in the form \begin{align*} E(s,t)&=\int_{\Omega}\frac{q+1}{2} \ |Du_{\overline{f}}+sDv|^2dx-b_1\int_{D_1}(u_{\overline{f}}+sv)|u_{\overline{f}}+sv|^q dx\\ &\quad -b_2\int_{D_2(t)}(u_{\overline{f}}+sv)|u_{\overline{f}}+sv|^q dx-b_3\int_{D_3}(u_{\overline{f}}+sv)|u_{\overline{f}}+sv|^q dx-\eta. \end{align*} We observe that, if $F:\mathbb{R}^2\to\mathbb{R}$ is a continuous function, then $$ \int_{D_2(t)} F-\int_{D_2} F = \int_J \int_0^{g(\xi,t)} F\big(\phi(\xi, g(\xi,\beta))\big)J(\xi, \beta) d\beta d\xi; $$ whence, from the Fundamental Theorem of Calculus, $$ \frac{\partial}{\partial t} \int_{D_2(t)} F =\int_J g_t(\xi,t) F\big(\phi(\xi, g(\xi,t))\big)J(\xi,g(\xi,t)) d\xi. $$ Using the above relation with $F= (u+sv)\big|u+sv\big|^q$, we have $$ \frac{\partial E}{\partial t}= -b_2\int_J g_t(\xi,t) \big(u_{\overline{f}}+sv\big) \big|u_{\overline{f}}+sv\big|^q J(\xi, g(\xi,t))\,d\xi, $$ where, for simplicity of notation, we set $u_{\overline{f}}\big(\phi(\xi, g(\xi,t))\big)=u_{\overline{f}}$ and $v\big(\phi(\xi, g(\xi,t))\big)=v$. Moreover \begin{align*} \frac{\partial^2 E }{\partial t^2} & = -b_2\int_J |u_{\overline{f}}+ sv|^q\Big\{\big[g_{tt}(\xi,t)(u_{\overline{f}}+sv) +(q+1)g^2_{t}(\xi,t)\langle Du_{\overline{f}}+sDv,\mathbf{N}\rangle \big]\\ &\quad\times J(\xi, g(\xi,t)) + g^2_{t}(\xi,t)(u_{\overline{f}}+sv)\langle\mathbf{N}', \mathbf{N}^{\bot}\rangle \Big\}d\xi, \end{align*} where we have used that \begin{gather*} \frac{\partial }{\partial t}\ u_{\overline{f}}\big(\phi(\xi, g(\xi,t))\big) = \langle Du_{\overline{f}}\big(\phi(\xi, g(\xi,t))\big), \mathbf{N}\rangle g_{t}(\xi,t),\\ \frac{\partial }{\partial t} J(\xi,g(\xi,t)) =g_t\langle\mathbf{N}',\mathbf{N}^\bot\rangle . \end{gather*} We note that, when $t=0$, $$ u_{\overline{f}}\big(\phi(\xi,g(\xi,t))\big) =u_{\overline{f}}(\gamma(\xi))=c_2, $$ $Du_{\overline{f}}\big(\phi(\xi,g(\xi,0))\big) =-|Du_{\overline{f}}\big(\gamma (\xi)\big)| \mathbf{N}(\xi)$ and $J(\xi, g(\xi,0))=J(\xi, 0)=|\gamma'(\xi)|$. Evaluating the above expression in $(0,0)$, we find \begin{align*} \frac{\partial^2 E }{\partial t^2} (0,0) &=-b_2 c_2^{q+1}\int_J \Big[g_{tt}(\xi,0)|\gamma'(\xi)| +g^2_{t}(\xi,0)\langle\mathbf{N}',\mathbf{N}^{\bot}\rangle\Big]d\xi\\ &\quad + b_2 c_2^q(q+1)\int_J g^2_{t}(\xi,0)|Du_{\overline{f}}(\gamma(\xi))| |\gamma'(\xi)|d\xi. \end{align*} By using \eqref{2.9} with $t=0$ we find \begin{equation}\label{a7} \begin{aligned} \frac{\partial^2 E }{\partial t^2} (0,0) &= b_2 c_2^q(q+1) \int_J g^2_{t}(\xi,0)|Du_{\overline{f}}(\gamma(\xi))|\, |\gamma'(\xi)|d\xi \\ &= b_2 c_2^q(q+1)\int_\gamma g_{t}^2(\gamma^{-1},0) |Du_{\overline{f}}|\,d\sigma. \end{aligned} \end{equation} We also have $$ \frac{\partial^2 E }{\partial s\partial t}= -b_2(q+1)\int_J g_{t}(\xi,t) v|u_{\overline{f}}+sv|^q J(\xi, g(\xi,t))\,d\xi; $$ that is, \begin{equation}\label{2.7} \begin{aligned} \frac{\partial^2 E }{\partial s\partial t} (0,0) &= -b_2c_2^q(q+1)\int_J g_{t}(\xi,0)\ v\big(\gamma(\xi)\big) |\gamma' (\xi)|\,d\xi\\ &= -b_2c_2^q(q+1)\int_\gamma g_{t}(\gamma^{-1},0)\ v\,d\sigma . \end{aligned} \end{equation} Using \eqref{1.4} in the form $$ \frac{\partial^2 E }{\partial s^2}(0,0) \frac{\partial^2 E }{\partial t^2} (0,0)\geq \Big( \frac{\partial^2 E }{\partial s \partial t} (0,0)\Big)^2, $$ and using \eqref{2.5}, \eqref{a7} and \eqref{2.7} in this inequality, we obtain \eqref{a8}. \end{proof} Note that in inequality \eqref{a8} only $g(\gamma^{-1},0)$ appears. Moreover, $g(\gamma^{-1},0)$ has null integral on $\gamma$. Indeed, differentiating \eqref{2.3} with respect to $t$ and putting $t=0$, we obtain $$ \int_J g(\xi,0)|\gamma'| d\xi=0. $$ Now a natural question arises: does inequality \eqref{a8} hold for any function $h$ with null integral on $\gamma$? The answer is contained in the following result. \begin{lemma}\label{lem2} Let $J$ and $\gamma$ be the same as described. Let $h:\gamma\to\mathbb{R}$ bounded, continuous and such that $\int_\gamma h\,d\sigma=0$. Then, for all $ v\in H_0^1(\Omega)$ and for all $a\in \mathbb{R}$ we have \begin{equation} \int_\Omega\Big( |Dv|^2-qu_{\overline{f}}\ v^2|u_{\overline{f}}|^{q-2}\overline{f}\Big) dx\cdot\int_\gamma h^2|Du_{\overline{f}}|\,d\sigma \geq b_2c_2^q\Big(\int_\gamma h(v-a)\,d\sigma\Big)^2. \end{equation} \end{lemma} The proof of the above lemma is similar to that of \cite[Lemma 2.2]{CKT}; we omit it. The following lemma is an analogue to \cite[Lemma 3.1]{CKT}. \begin{lemma}\label{lem3} Let $P$ be a point on $\mathcal{F}=\partial \{u_{\overline{f}}>c_2\}$. Suppose that for all $k\in \mathbb{Z}^+$ there exist a positive number $r_k$, a bounded open interval $J_k$ and a regular curve $\gamma_k:J_k\to \mathcal{F}^*$ such that $r_1>r_2>\dots\to 0$, $\overline{\gamma_k(J_k)}\subset \mathcal{F}^* \cap B_{r_k} (P) \setminus \overline{B_{r_{k+1}}(P)}$. Then we must have $$ \sum_{k=1}^\infty \int_{\gamma(J_k)} \frac{1}{|Du_{\overline{f}}|}\,d\sigma< \infty. $$ \end{lemma} \begin{proof} Without loss of generality, we assume that $P$ is the origin. We suppose also that $J_k\cap J_h=\emptyset$ for all $k\neq h$, and denote all $\gamma_k$ with $\gamma$. We define $$ J_{k,m} =\begin{cases} J_k \cup J_{k+1}\cup\cdots\cup J_m &\text{if } m\geq k, \\ \emptyset &\text{otherwise.} \end{cases} $$ We suppose by contradiction that \begin{equation} \label{3.1} \sum_{k=1}^\infty \int_{\gamma(J_k)} \frac{1}{|Du_{\overline{f}}| }\,d\sigma =\infty. \end{equation} Let $V$ be a smooth radial function in $\mathbb{R}^2$, decreasing in $|x|$, defined by $$ \begin{cases} V(x)=2, &|x| =0\\ 10$ in $B_{r_m}$, $\gamma(J_l)\subset B_{r_m}$ for all $l\geq m$ and $v_k(x)-1 \to 1$ as $x\to0$ and \eqref{3.1}, we have $$ \int_{\gamma(J_{m,l})} \frac{v_k -1}{|Du_{\overline{f}}|}\,d\sigma \to\infty \ \ \ \text{for }l\to\infty. $$ Consequently, there must be a number $l\geq m$ such that $$ \int_{\gamma(J_{m,l-1})} \frac{v_k-1}{|Du_{\overline{f}}| }\,d\sigma \leq -\int_{\gamma(J_{1,k-1})} \frac{v_k-1}{|Du_{\overline{f}}| } \,d\sigma < \int_{\gamma(J_{m,l})} \frac{v_k-1}{|Du_{\overline{f}}| } \,d\sigma. $$ Choose a subinterval $J_l'\subset J_l$ such that $$ \int_{\gamma(J_{m,l-1})} \frac{v_k-1}{|Du_{\overline{f}}| } \, d\sigma\ + \int_{\gamma(J_{l}')} \frac{v_k-1}{|Du_{\overline{f}}| } \,d\sigma= -\int_{\gamma(J_{1,k-1})} \frac{v_k-1}{|Du_{\overline{f}}| }\,d\sigma. $$ Then we have $$ \int_{\gamma(J^k)} \frac{v_k-1}{|Du_{\overline{f}}| }\,d\sigma=0, $$ where $J^k=J_{1,k-1}\cup J_{m,l-1}\cup J_l'$. Now we can apply Lemma \ref{lem2} to $J^k$, $\gamma$, $v_k$, $a=1$ and $h=\frac{v_k -1}{|Du_{\overline{f}}|}\ $ and, after rearranging, obtain $$ \int_\Omega\left( |Dv_k|^2-qu_{\overline{f}}\ v_k^2|u_{\overline{f}}|^{q-2}\overline{f}\right) dx\geq b_2c_2^q\int_{\gamma(J^k)}\frac{(v_k -1)^2}{|Du_{\overline{f}}|} \,d\sigma. $$ We find that $$ \int_\Omega\Big(|Dv_k|^2-qu_{\overline{f}}\ v_k^2|u_{\overline{f}}|^{q-2} \overline{f}\Big) dx\leq \int_{B_1(0)}|DV|^2dx. $$ By the above estimate, for a suitable constant $C$, we have \begin{align*} C\int_{B_1(0)} |D V|^2 dx &\geq\int_{\gamma(J^k)} \frac{(v_k -1)^2}{|Du_{\overline{f}}|}\,d\sigma \\ & \geq \int_{\gamma(J_{1,k-1})} \frac{(v_k -1)^2}{|Du_{\overline{f}}|} \,d\sigma \\ &= \sum_{h=1}^{k-1} \int_{\gamma(J_h)} \frac{(v_k -1)^2}{|Du_{\overline{f}}|}\, d\sigma. \end{align*} Then, when $k\to \infty$, we have $$ C\int_{B_1(0)} |D V|^2 dx\geq\sum_{h=1}^{\infty} \int_{\gamma(J_h)} \frac{(v_k -1)^2}{|Du_{\overline{f}}|}\,d\sigma =+\infty ,$$ which is a contradiction. So we must have $$ \sum_{k=1}^\infty \int_{\gamma(J_k)} \frac{d\sigma}{|Du_{\overline{f}}|}\ < \infty, $$ as desired. \end{proof} \begin{lemma}\label{lem4} Let $P$ be a point on $\mathcal{F}=\partial \{u_{\overline{f}}>c_2\}$. Suppose that there are numbers $K\in\mathbb{Z}$ and $\overline{\sigma}>0$ such that, for each $k\geq K$, there exists a regular curve $\gamma_k: J_k\to \mathcal{F}^*$ with the following two properties: \begin{gather*} \overline{\gamma(J_k)}\subset\mathcal{F}^*\cap B_{2^{-k}} (P) \setminus \overline{B_{2^{-(k+1)}}(P)},\\ \mathcal{H}^1 (\gamma_k(J_k))=\int_{J_k}|\gamma'(\xi)| d\xi >\overline{\sigma}\,2^{-k}. \end{gather*} Then $|Du_{\overline{f}}(P)|>0$. \end{lemma} For a proof of the above lemma, see \cite[Lemma 3.2]{CKT}. From an intuitive point of view, this lemma says that, if the set $\partial \{u_{\overline{f}}>c_2\}\cap\{|Du_{\overline{f}}|>0\}$ is big enough around a point of $\partial\{u_{\overline{f}}>c_2\} $, then $|Du_{\overline{f}}|>0$ at this point. Now, we are able to prove our main theorem. \begin{proof}[Proof of Theorem \ref{thm6}] By using the previous Lemmas and superharmonicity of $u_{\overline{f}}$ the Theorem follows from the results of sections 5 and 6 in \cite{CKT}. \end{proof} \subsection*{Open problems} The method used in this paper to prove regularity does not work when the number of level sets of $\overline f$ is infinite. Therefore it remains to study the boundaries of level sets of $\overline f$ in the case of the rearrangement class $\mathfrak{F}(f_0)$ of a general function $f_0$. We can obtain an analogous result to Lemma \ref{lem3} for the $p$-Laplacian operator, but we cannot go further because we lack a suitable regularity theory for the $p$-Laplacian operator and its solutions. We think that it is reasonable to guess that a regularity result of the type that we have proven in this work will hold for the situation with the $p$-Laplacian when $p < 2$. \subsection*{Acknowledgments} The authors want to thank the anonymous referees for their valuable comments and suggestions. \begin{thebibliography}{99} \bibitem{ATL} A. Alvino, G. Trombetti and P. L. Lions; \emph{On optimization problems with prescribed rearrangements}, Nonlinear Analysis, TMA \textbf{13} (1989), 185--220. \bibitem{CGIKO} S. Chanillo, D. Grieser, M. Imai, K. Kurata and I. Ohnishi; \emph{Symmetry breaking and other phenomena in the optimization of eigenvalues for composite membranes}, Commun. Math. Phys.\textbf{214} (2000), 315--337. \bibitem{CGK} S. Chanillo, D. Grieser and K. Kurata; \emph{The free boundary problem in the optimization of composite membranes}, Contemporary Math. \textbf{268} (1999), 61--81. \bibitem{CKT} S. Chanillo, C.E. Kenig and G. T. To; \emph{Regularity of the minimizers in the composite membrane problem in $\mathbb{R}^2$}, Journal of Functional Analysis \textbf{255} (2008), 2299--2320. \bibitem{CPS} F. Cuccu, G. Porru and S. Sakaguchi; \emph{Optimization problems on general classes of rearrangements}, Nonlinear Analysis \textbf{74} (2011), 5554--5565. \bibitem{F} L.E. Fraenkel; \emph{An Introduction to maximum principles and symmetry in elliptic problems}, Cambridge Tracts in Mathematics, Cambridge Univ. Press. (2000). \bibitem{GNN} B. Gidas, W. M. Ni and L. Nirenberg; \emph{Symmetry and related properties via the maximum principle}, Comm. Math. Phys. \textbf{68}, N. 3 (1979), 209--243. \bibitem{GT} D. Gilbarg, N. S. Trudinger; \emph{Elliptic partial differential equations of second order}, Springer Verlag, Berlin, (1977). \bibitem{K} B. Kawohl; \emph{Rearrangements and convexity of level sets in PDE's}, Lectures Notes in Mathematics, Springer \textbf{1150} (1985). \bibitem{T} P. Tolksdorff; \emph{Regularity for a more general class of quasilinear elliptic equations}, Journal of Differential Equations \textbf{51} (1984), 126--150. \end{thebibliography} \end{document}