\documentclass[reqno]{amsart} \usepackage{amssymb,hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2005(2005), No. 112, pp. 1--8.\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 2005 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2005/112\hfil A resonance problem] {A resonance problem for the p-laplacian in $\mathbb{R}^N$} \author[G. Izquierdo B., G. L\'{o}pez G.\hfil EJDE-2005/112\hfilneg] {Gustavo Izquierdo Buenrostro \& Gabriel L\'{o}pez Garza} \address{Gustavo Izquierdo Buenrostro \hfil\break Dept. Mat. Universidad Aut\'{o}noma Metropolitana\\ M\'{e}xico} \email{iubg@xanum.uam.mx} \address{Gabriel L\'{o}pez Garza \hfill\break Dept. Mat. Universidad Aut\'{o}noma Metropolitana\\ M\'{e}xico} \email{grlzgz@xanum.uam.mx} \date{} \thanks{Submitted May 31, 2005. Published October 17, 2005.} \subjclass[2000]{35J20} \keywords{Resonance; $p$-Laplacian; improved Poincar\'{e} inequality} \begin{abstract} We show the existence of a weak solution for the problem $$ -\Delta_p u=\lambda_1h(x)|u|^{p-2}u+a(x)g(u)+f(x),\quad u\in\mathcal{D}^{1,p}(\mathbb{R}^N), $$ where, $2From here and henceforth the integrals and all the spaces are taken over $\mathbb{R}^N$ unless otherwise specified. The term resonance is well known in the literature, and refers to the case in which $\lambda$ is an eigenvalue of the problem \begin{equation}\label{eigen} \begin{gathered} -\Delta_p u=\lambda h(x)|u|^{p-2}u,\\ u\in \mathcal{D}^{1,p}. \end{gathered} \end{equation} In \cite{All}, Allegreto et al. show that the eigenvalue problem \eqref{eigen} possesses a sequence of eigenvalues $0<\lambda_1<\lambda_2\leqslant,\dots$ and a corresponding sequence of eigenfunctions $\{\varphi_j\}$, where $\varphi_1$ can be chosen to be positive a.e.. Moreover, we have the Rayleigh quotient characterization: \begin{equation} \label{Ray} \lambda_1=\inf\Big\{ \int |\nabla u|^p: u\in \mathcal{D}^{1,p}\mbox{ with } \int h |u|^p=1 \Big\}. \end{equation} We consider the function $\varphi_1$ to be normalized; i.e., $\int h|\varphi_1|^p=1$ and we decompose any function $u\in\mathcal{D}^{1,p}$ as a direct sum \begin{equation} \label{sum} \begin{gathered} u=\alpha \varphi_1+w\mbox{ where }\\ \alpha=\int h|\varphi_1|^{p-2}\varphi_1u\mbox{ and }\int h|\varphi_1|^{p-2}\varphi_1w=0. \end{gathered} \end{equation} Hence, we introduce the spaces \begin{equation} \label{spaces} \begin{gathered} V\stackrel{\text{def}}{=}\mathop{\rm span}\{\varphi_1\},\\ W \stackrel{\text{def}}{=}\big\{w\in\mathcal{D}^{1,p}:\int h|\varphi_1| ^{p-2}\varphi_1w=0\big\} \end{gathered} \end{equation} In order to prove our main result we use some of the results introduced by Alziary, Fleckinger and Tak\'{a}\u{c} in \cite{Takac} where the cases $10,\quad 20$ and $C>0$ such that \begin{equation} \label{H} 0=\int|\nabla u|^{p-2}\nabla u\cdot\nabla v -\lambda\int h|u|^{p-2}uv-\int ag(u)v-\int fv \end{equation} for all $u,v\in\mathcal{D}^{1,p}$. To prove theorem \ref{thm1} we use the Minimax Methods introduced by Rabinowitz \cite{R}. We recall here for the convenience of the reader some previous definitions and theorems. \noindent{\bf Palais-Smale condition.} Suppose that $E$ is a real Banach space. A functional $I\in\mathcal{C}^1(E,\mathbb{R})$ satisfies the Palais-Smale condition at level $c\in\mathbb{R}$, denoted $(PS)_c$, if any sequence $(u_n)\subset E$ for which \begin{itemize} \item[(i)] $I(u_n)\to c$ as $n\to\infty$ and \item[(ii)] $I'(u_n)\to 0$ as $n\to\infty$, \end{itemize} possesses a convergent subsequence. If $I\in\mathcal{C}^1(E,\mathbb{R})$ satisfies the $(PS)_c$ for every $c\in\mathbb{R}$, we say that $(u_n)$ satisfies the $(PS)$ condition. Any sequence for which (i) and (ii) hold is called a $(PS)_c$ sequence for $I$. Now we establish a preliminary result. \begin{proposition}\label{propPS} Let $J_{\lambda}:\mathcal{D}^{1,p}\rightarrow\mathbb{R}$ be defined as \ref{J} where $\lambda\in \mathbb{R}$. Suppose that $g$ is a continuous function with $|g(s)|\leqslant M$ for all $s\in \mathbb{R}$, $f\in L^{(p^{*})'}(\mathbb{R}^N)$, $22$ (see \cite{All} inequality (7) p.237 and subsequent inequalities) \begin{equation} \label{alegreto} \begin{aligned} \int|\nabla u_n-\nabla u|^p &\leqslant C\left\{\int (|\nabla u_n|^{p-2}\nabla u_n-|\nabla u|^{p-2} \nabla u)\cdot \nabla(u_n-u) \right\}\\ &\quad \times \Big(\int|\nabla u_n|^p+ \int|\nabla u|^p \Big). \end{aligned} \end{equation} Thus it is sufficient to show that $\lim_{n\to\infty}\int |\nabla u_n|^{p-2}\nabla u_n\cdot\nabla(u_n-u)=0$. To this aim, taking $\varphi=u_n-u$, in \eqref{PS2} we have \begin{equation} \label{PS3} \begin{aligned} &\int |\nabla u_n|^{p-2}\nabla u_n\cdot \nabla(u_n-u)\\ &=\lambda\int h|u_n|^{p-2}u_n(u_n-u)+\int ag(u_n)(u_n-u) +\int f(u_n-u)+o(1) \end{aligned} \end{equation} as $n\to\infty$. For the first integral in the right hand side, using the H\"{o}lder's inequality we have $$ \big|\int h|u_n|^{p-2}u_n(u_n-u)\big|\leqslant \Big(\int h|u_n|^p\Big)^{1/p'}\Big( \int h|u_n-u|^p\Big)^{1/p}. $$ Noting that $h\in L^{N/p}(\mathbb{R}^N)=L^{(p^{*}/p)'}(\mathbb{R}^N)$, for $1\leqslant q< p^{*}$ the functional $u\mapsto\int h|u|^q$ is weakly continuous in $\mathcal{D}^{1,p}$ (see \cite[Prop. 2.1 p. 826]{Ben}). Consequently, \begin{equation}\label{PS4} \lim_{n\to\infty} \int h|u_n|^{p-2}u_n(u_n-u)=0. \end{equation} For the integral $\int a g(u_n)|u_n-u|$ we consider the ball $B_r(0)$. Since $a,\, g$ are bounded we have $$\big|\int_{B_r(0)} ag(u_n)(u_n-u)\big|\leqslant C\int_{B_r(0)}|u_n-u|\to 0 \quad \mbox{as }n\to\infty, $$ since $u_n\to u$ strongly in $L^{1}(B_r(0))$ due to the Relich-Kondrachov theorem. Now, together with the assumption that $u_n\mbox{ and }g$ are bounded, we obtain $$ \big|\int_{\mathbb{R}^N\setminus B_r(0)}ag(u_n)|u_n-u|\big|\leqslant C\Big( \int_{\mathbb{R}^N\setminus B_r(0)}|a|^\frac{Np}{N-p}\Big) ^\frac{N+p}{Np} $$ So, by taking $r$ big enough it follows that $$ \limsup_{n\to\infty}\big|\int ag(u_n)|u_n-u|\big|\leqslant C\varepsilon $$ For arbitrary $\varepsilon$. Finally, since $f\in L^{(p^{*})'}(\mathbb{R}^N)$ we can use similar arguments as above to show that $\lim_{n\to\infty}\int f(u_n-u)=0$. \end{proof} \section{Proof of Theorem \ref{thm1}} In this section we consider the problem \begin{equation} \label{pl1} \begin{gathered} -\Delta_p u=\lambda_1h(x)|u|^{p-2}u+a(x)g(u)+f(x)\\ u\in\mathcal{D}^{1,p} \end{gathered} \end{equation} where $h$ satisfies (H). To prove the main theorem of this section we require the Saddle Point Theorem of Rabinowitz \cite{R}, which we introduce here for the convenience of the reader. \begin{theorem}[Saddle Point Theorem] \label{sadd} Let $E=V\oplus W $, where $E$ is a real Banach space and $V\not= \{0\}$ is finite dimensional. Suppose $I\in \mathcal{C}^1(E,{\mathbb{R}})$ satisfies the (PS) condition and \begin{itemize} \item[(I1)] there is a constant $\alpha$ and a bounded neighborhood $D$ of 0 in $V$ such that $I\big|_{\partial D}\leqslant\alpha$, and \item[(I2)] there is a constant $\beta>\alpha$ such that $I|_{W}\geqslant \beta$. \end{itemize} Then, $I$ possesses a critical value $c\geqslant\beta$. Moreover $c$ can be characterized as $$ c=\inf_{h\in\Gamma}\max_{u\in\overline{D}} I(h(u)), $$ where $\Gamma=\{h\in\mathcal{C}(\overline{D},E):h=\mbox{id on }\partial D\}$. \end{theorem} Now, we can show the existence of weak solutions for $J_{\lambda_1}$. \begin{proof}[Proof of Theorem \ref{thm1}] First, we show that the functional $J_{\lambda_1}$ corresponding to problem \eqref{pl1} satisfies the $(PS)_c$ condition for any $c\in\mathbb{R}$, and thereafter we verify that $J_{\lambda_1}$ satisfies the other hypotheses of the Theorem \ref{sadd}. Let $(u_n)$ be a $(PS)_c$ sequence for the functional $J_{\lambda_1}$. We claim that $(u_n)$ is bounded. For each $n\in\mathbb{N}$ write $$ u_n\stackrel{\text{def}}{=}v_n+w_n=\alpha_n\varphi_1+w_n\quad \mbox{with } \alpha_n\in\mathbb{R}\mbox{ and }w_n\in W. $$ Since $(u_n)$ is a $(PS)_c$ sequence we have $|J_{\lambda_1}(u_n)|0$. By standard calculations (see for instance \cite[p.16]{LR}), we have \begin{equation} \label{t1.3} \big|\int a(G(v_n +w_n)-G(v_n))\big|\leqslant M\int a|w_m| \leqslant C_2\|w_n\|. \end{equation} Consequently, using \eqref{t1.1}, \eqref{t1.2} and \eqref{t1.3} we have \begin{equation} \label{t1.4} \big|\int aG(v_n)+\int fv_n\big| \leqslant C_1+C_2\|w_n\|+\frac{c}{p}\|w_n\|^p. \end{equation} So, given that $\int aG(v_n)+\int fv_n\to \infty\mbox{ as }\|v_n\|=|\alpha_n|\to\infty$, we have shown that $(v_n)$ is bounded if $(w_n)$ is bounded. We show now that $(w_n)$ is bounded. In fact, note that \begin{equation} \label{t1.5} \int |\nabla u_n|^{p-2}\nabla u_n\cdot\nabla w_n \geqslant \|u_n\|^p-\int |\nabla u_n|^{p-2}u_n\cdot \nabla v_n. \end{equation} On the other hand, since $\langle J'_{\lambda_1}(u_n),v_n \rangle\stackrel{n}{\rightarrow} 0$, there exists $m_0$ such that if $n\geqslant m_0$ then, \begin{equation} \label{t1.6} \big|\int |\nabla u_n|^{p-2}\nabla u_n\cdot\nabla w_n- \int z_n \cdot w_n\big|\leqslant C\|w_n\|, \end{equation} where $z_n = \lambda_1 h|u_n|^{p-2}u_n + ag(u_n)w_n+ f$. Adding and subtracting $\lambda_1\int h|u_n|^{p-2}u_n\cdot v_n\mbox{ and }\int ag(u_n)v_n+\int f(x)v_n$, and substituting \eqref{t1.5} in \eqref{t1.6} \begin{equation}\label{t1.7} \begin{aligned} \|u_n\|^{p}-\lambda_1\int h|u_n|^p &\leqslant C\|w_n\|+\langle J'_{\lambda_1}(u_n),v_m\rangle+ \int ag(u_n)v_n+\int f(x)v_n\\ &\leqslant C\|w_n\|+\langle J'_{\lambda_1}(u_n),v_m\rangle+C'|\alpha_n| \end{aligned} \end{equation} Again, since $J'_{\lambda_1}\to 0\mbox{ as }n\to\infty$, there exist $m_1$ such that if $n\geqslant m_1$ then $\langle J'_{\lambda_1}(u_n),v_m\rangle \leqslant C\|v_n\|=C|\alpha_n|$, taking $n\geqslant\max\{m_0,m_1\}$ \begin{equation}\label{t1.8} \|u_n\|^{p}-\lambda_1\int h|u_n|^p\leqslant C\|w_n\|+C'|\alpha_n|. \end{equation} Now fix $\gamma>0$, and suppose that $(u_n)\in \mathcal{C'}_\gamma$ for all $n$. Then we have, $|\alpha_n|\leqslant (1/\gamma)\|w_n\|$ and $\int h|u_n|^p\leqslant(1/\Lambda_{\gamma})\int|\nabla u_n|^p$. Thus, by Lemma \ref{Lambda} \begin{equation} \label{t1.9} \big(1 -\frac{\lambda_1}{\Lambda_{\gamma}}\big)\|u_n\|^p\leqslant C\|w_n\| \end{equation} Since the projection $u\mapsto w$ is bounded in $\mathcal{D}^{1,p}$ we obtain \begin{equation} \label{t1.9'} \|w_n\|^p\leqslant C_\gamma\|w_n\|, \end{equation} given that $\lambda_1/ \Lambda_\gamma<1$ by Lemma \ref{Lambda}. Hence by Lemma \ref{Lambda}, $\Lambda_\gamma>\lambda_1$; therefore, $(w_n)$ is bounded if $(u_n)\in\mathcal{C'}_\gamma$. Now, set $\gamma_n =\|w_n\|/|\alpha_n|$ and define $$ \gamma \stackrel{\text{def}}{=}\liminf_n\gamma_n. $$ We have two cases: (i) $\gamma\in(0,\infty]$ and (ii) $\gamma=0$. By the above argument, if $\gamma\in(0,\infty]$ then $(w_n)$ is bounded and the proof is concluded. If $\gamma=0$, take $\varepsilon>0$ arbitrarily small, such that $\|w_n\|\leqslant\varepsilon|\alpha_n|$. Using inequality \eqref{t1.8}, Lemma \ref{Lambdatilde} with $\phi=\phi_n \stackrel{\text{def}}{=}(\|w_n\|/|\alpha_n|)\cdot w_n/\|w_n\|$, and the fact that the projection $u\mapsto \alpha$ is bounded in $\mathcal{D}^{1,p}$ we obtain \begin{gather*} |u_n|^p\big( 1-\frac{\lambda_1}{\tilde{\Lambda}} \big) \leqslant C\varepsilon|\alpha_n|+C'\|v_n\|,\\ |\alpha_n|^p\leqslant c_\gamma|\alpha_n|. \end{gather*} Therefore, $|\alpha_n|$ is bounded, and since $\|w_n\|\leqslant\varepsilon|\alpha_n|$ we have that $(u_n)$ is bounded as wanted. To verify the geometric hypotheses of the Saddle Point Theorem we note that since $\lambda_1$ is isolated (see \cite{All}) we have \begin{equation} \label{g1} \lambda_2\stackrel{\text{def}}{=}\inf\big\{\|w\|^p:w\in W,\int h|w|^p=1 \big\}, \end{equation} which satisfies $\lambda_1<\lambda_2$. As a consequence of \eqref{g1} we have \begin{equation} \label{g2} \int|\nabla w|^p\geqslant\lambda_2\int h|w|^p,\quad \forall w\in W. \end{equation} Now, if $w\in W$, \begin{equation} \label{g3} \int|\nabla w|^p-\lambda_1\int h|w|^p\geqslant \big(1-\frac{\lambda_1}{\lambda_2}\big). \end{equation} Moreover, since $|g(s)|\leqslant M$ for all $s\in\mathbb{R}$, we have that for all $w\in\mathcal{D}^{1,p}$, $$ \big|\int aG(w) \big|\leqslant M\int |a||w|\leqslant C\|w\|. $$ Therefore, $J_{\lambda_1}$ is bounded from below on $W$; i.e. (I2) in Theorem \ref{sadd} holds. Finally, if $v\in V$ we have $$ J_{\lambda_1}(v)=-\int aG(v)-\int fv. $$ Since $\int aG(v)+\int fv\to\infty$ as $\|v\|\to\infty$ by \eqref{alp} and, therefore, (I1) in the Saddle Point Theorem also holds. Hence, $J_{\lambda_1}$ has a critical point and the proof is concluded. \end{proof} \noindent\textbf{Remark.} Suppose $\lim_{s\to\infty} g(s)=g_{\infty}$ and $\lim_{s\to-\infty} g(s)=g_{-\infty}$ exist. Then, if $g_{\infty}>0$ and $g_{-\infty}<0$, $G(s)=\int^{s}_{0}g(t)dt\to\infty$ as $|s|\to\infty$. Consequently, by L' H\^{o}spital's rule, the Lebesgue dominated convergence theorem and the fact that $\varphi_1>0$ a.e. in $\mathbb{R}^N$ we have that $$ \lim_{|t|\to\infty}\frac{1}{t}\int a(x)G(t\varphi_1) =\lim_{|t|\to\infty}\int ag(t\varphi_1)\varphi_1 = \begin{cases} g_{\infty}\int a\varphi_1 &\mbox{as }t\to\infty,\\ g_{-\infty}\int a\varphi_1 &\mbox{as }t\to-\infty . \end{cases} $$ Thus, the condition \eqref{alp} in the resonance Theorem \ref{thm1} holds if \[ g_{\infty}\int a\varphi_1+\int f\varphi_1>0\quad \mbox{and}\quad g_{-\infty}\int a\varphi_1+\int f\varphi_1<0 , \] or \begin{equation} \label{LL} g_{-\infty}\int a\varphi_1<-\int f\varphi_1