\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2014 (2014), No. 151, pp. 1--9.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2014 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2014/151\hfil Existence of solutions to $p$-Laplacian equations] {Existence of solutions to $p$-Laplacian equations involving general subcritical growth} \author[Y.-Y. Lan \hfil EJDE-2014/151\hfilneg] {Yong-Yi Lan} % in alphabetical order \address{Yong-Yi Lan \newline School of Sciences, Jimei University \\ Xiamen 361021, China} \email{lanyongyi@jmu.edu.cn} \thanks{Submitted February 13, 2014. Published June 27, 2014.} \thanks{Supported by HuiZhen Huang Foundation for Subject Construction at Jimei University} \subjclass[2000]{35B33, 35J92, 35J35} \keywords{$p$-Laplacian equation; subcritical growth; variational methods; \hfill\break\indent (C) condition; mountain-pass lemma} \begin{abstract} In this article, we consider the quasilinear elliptic equation $-\Delta_p u=\mu f(x,u)$ with the Dirichlet boundary coditions, and under suitable growth condition on the nonlinear term $f$. Existence of solutions is given for all $\mu>0$ via the variational method and some analysis techniques. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{remark}[theorem]{Remark} \allowdisplaybreaks \section{Introduction and main results} In this article, we consider the Dirichlet boundary-value problem \begin{equation} \begin{gathered} -\Delta_p u=\mu f(x,u),\quad x\in\Omega,\\ u=0,\quad x\in\partial\Omega, \end{gathered} \label{e1.1} \end{equation} where $\Omega$ is an open bounded domain in $\mathbb{R}^N$ with smooth boundary $\partial\Omega $, $p>$1, $-\Delta_p u=\operatorname{div}(|\nabla u|^{p-2}\nabla u)$ is the $p$-Laplacian of $u$, $f(x,t)$ is continuous on $\overline{\Omega}\times\mathbb{R}$. We look for the weak solutions of \eqref{e1.1} which are the same as the critical points of the functional $I_{\mu}:W^{1,p}_0(\Omega)\to\mathbb{R}$ defined by \begin{equation} \label{e1.2} I_{\mu}(u)=\frac{1}{p}\int_\Omega|\nabla u|^p\,\mathrm{d}x-\mu\int_\Omega F(x,u)\,\mathrm{d}x, \end{equation} where $F(x,t)=\int_0^tf(x,s)\,\mathrm{d}s$, and $ W^{1,p}_0(\Omega)$ is the Sobolev space with the usual norm: $$ \|u\|^p=\int_\Omega|\nabla u|^p\,\mathrm{d}x. $$ In this article, the hypotheses on the nonlinearity $f(x,t)$ are the following: \begin{itemize} \item[(F1)] There exist constants $\theta\geq 1$, $\alpha>0$ such that $$ \theta G(x,t)+\alpha\geq G(x,st)\quad \text{for all } t\in\mathbb{R}, \; x\in\overline{\Omega},\; s\in[0,1], $$ where $G(x,t):=tf(x,t)-pF(x,t)$. \item[(F2)] $$ \lim_{|t|\to\infty}\frac{f(x,t)}{t|t|^{p^{\ast}-2}}=0\quad \text{uniformly a.e. } x\in\Omega, $$ where $p^*=\frac{Np}{N-p}$ if $10$, \eqref{e1.1} possesses at least a nontrivial solution if $f(x,t)$ satisfies the following conditions: \begin{itemize} \item[(1)] $p$-superlinear at $t=0$: $\lim_{t\to0}\frac{f(x,t)}{t|t|^{p-2}}=0$ uniformly a.e. $x\in\Omega$. \item[(2)] subcritical at $t=\infty$: there are positive constants $a$ and $b$ such that \[ |f(x,t)|\leq a+b|t|^{q-1}, \quad \forall t\in\mathbb{R},\; x\in\Omega. \] where $1\leq qp$, $C>0$, \begin{equation} \label{e1.3} 0<\theta F(x,t)\leq f(x,t)t,\quad \forall |t|\geq C,\;x\in\Omega. \end{equation} \end{itemize} The (AR) condition has appeared in most of the studies for quasilinear problems and plays an important role in studying the existence of nontrivial solutions of many quasilinear elliptic boundary value problems. It is quite natural and important not only to ensure that the Euler-Lagrange functional associated to problem \eqref{e1.1} has a mountain pass geometry, but also to guarantee that Palais-Smale sequence of the Euler-Lagrange functional is bounded. Since then, the (AR) condition has been used extensively in many literature sources (see \cite{Bartsch,Coti,YDing}). But this condition is very restrictive eliminating many nonlinearities. There are always many functions that do not satisfy the (AR) condition. For example, for the sake of simplicity, we consider the case $p=2$, $$ f(x,t)=2t\ln(1+|t|). $$ Many efforts have been made to extend the range of the nonlinearity. For example, Miyagaki and Souto \cite{Miyagaki} studied \eqref{e1.1} for when $p=2$ and replaced the (AR) condition by some monotonicity arguments. They assumed that there is $t_0>0$ such that \begin{equation} \label{e1.4} \frac{f(x,t)}{t} \text{ is increasing for $t\geq t_0$ and decreasing for $t\leq -t_0$, for all }x\in\Omega; \end{equation} or a weaker condition is that there exist $C>0$ such that \begin{equation} \label{e1.5} tf(x,t)-2F(x,t)\leq sf(x,s)-2F(x,s)+C,\quad \end{equation} for all $00$ such that \[ |f(x,t)|\leq C_0(1+|t|^{p-1}), \quad \forall t\in\mathbb{R},\;x\in\Omega, \] where $10$. \end{theorem} Many efforts have been made to extend these results (see \cite{Li1,Liu,Mao} and the references therein). Li and Yang extend the results from $p=2$ to $p>1$ in \cite{Li1}, they obtained the following theorem: \begin{theorem}[{\cite[Theorem 1.1]{Li1}}] \label{thmB} Under hypotheses \eqref{e1.4}, \eqref{e1.2}, {\rm (F3)} and {\rm (F4)}, problem \eqref{e1.1} has a nontrivial weak solution, for all $\lambda>0$. \end{theorem} The aim of the article is to consider the problem in a different case: based on a variant version of mountain pass theorem, we can prove the same result under more generic conditions, which generalizes Theorems \ref{thmA} and \ref{thmB}. Our main results reads as follows: \begin{theorem} \label{thm1} Suppose that {\rm (F1)--(F4)} hold. Then \eqref{e1.1} has a weak nontrivial solution, for all $\lambda>0$. \end{theorem} Note that (F3) implies that problem \eqref{e1.1} has a trivial solution $u =0$ and we are interested in the existence of nontrivial solutions. \begin{remark} \rm Theorem \ref{thm1} improves Theorem \ref{thmB} in two aspects. To show this, it suffices to compare condition (F1) with \eqref{e1.4} and \eqref{e1.5}, and to compare condition (F2) with \eqref{e1.2}. \end{remark} At first, we can easily prove that (F1) is equivalent to \eqref{e1.5} when $\theta=1$, and (F1) gives some general sense of monotony when $\theta>1$. There are functions satisfying our condition (F1) and not satisfying the condition \eqref{e1.5}. For example, for the sake of simplicity, we consider the case $p=2$, let $$ F(x,t)=t^{2}\ln(1+t^{2})+t\sin t, $$ then $$ f(x,t)=2t \ln(1+t^{2})+t^{2} \cdot \frac{2t}{1+t^{2}}+ \sin t+t \cos t, $$ it follows that $$ G(x,t)=tf(x,t)-2F(x,t)=2(t^{2}-1)+\frac{2}{1+t^{2}}+(t^{2} \cos t-t\sin t). $$ Let $\theta$ = 1000, we can prove by some simple computation that $G$ satisfies (F1) but does not satisfy the condition \eqref{e1.5} any more. Secondly, it is obvious that \eqref{e1.2} implies (F2). There are functions satisfying our growth condition (F2) and not satisfying the subcritical growth condition \eqref{e1.2}. For example, for the sake of simplicity, we consider the case $p=2$, let $$ F(t)=\frac{t^{2^{\ast}}}{\ln(e+t^{2})} $$ then $$ f(t)=\frac{2^{\ast}t^{2^{\ast}-1}(e+t^{2})\ln(e+t^{2}) -2t^{2^{\ast}+1}}{(e+t^{2})(\ln(e+t^{2}))^{2}}. $$ The proof of Theorem \ref{thm1} is much easier than that of the main results in \cite{Li1,Miyagaki}. \begin{remark} \rm In assumption (F2), we are dealing with functionals satisfying the so-called non-standard growth conditions. Due to the lack of compactness of the embeddings in $W^{1,p}_0(\Omega)\hookrightarrow L^{p^{\ast}}(\Omega)$, we cannot use the standard variational argument directly. We overcome the difficulty by Vitali convergence theorem and some analysis techniques. \end{remark} This paper is organized as follows. In section 2, we give the proof of the Theorem \ref{thm1}. In the following discussion, we denote various positive constants as $c$ or $c_i$ $(i=0, 1,2, \dots)$ for convenience. \section{Proof of Theorem \ref{thm1}} The proof consists of three steps. We prove Theorem \ref{thm1} only when $\mu=1$. The case of a general $\mu>0$ will follow immediately. In fact, if $\mu>0$ and $\mu\neq1$, we only let $g(x,t)=\mu f(x,t)$. Then \eqref{e1.1} becomes \begin{gather*} -\Delta_p u=g(x,u),\quad x\in\Omega,\\ u=0\quad x\in\partial\Omega. \end{gather*} The nonlinear term $g$ also satisfies condition (F1)--(F4), Then the same conclusion as in the case $\mu=1$ holds. \smallskip \noindent\textbf{First step:} The (C) condition. Let $\{u_{n}\}$ be any sequence in $W^{1,p}_0(\Omega)$ such that $I(u_{n})$ is bounded and $\|I'(u_{n})\|(1+\|u_{n}\|)$ converges to zero; that is, $$ I(u_{n})\to c,\quad \|I'(u_{n})\|(1+\|u_{n}\|)\to0 $$ which shows that \begin{equation} \label{e2.1} c=I(u_{n})+o(1),\quad \langle I'(u_{n}),u_{n}\rangle=o(1) \end{equation} where $o(1)\to 0$ as $n\to\infty$. We now prove that $\{u_{n}\}$ is bounded in $W^{1,p}_0(\Omega)$. By contradiction, we assume $\|u_{n}\|\to\infty$ as $n\to\infty$. Let $w_{n}=\frac{u_{n}}{\|u_{n}\|}$, then $w_{n}\in W^{1,p}_0(\Omega)$ with $\|w_{n}\|=1$. Then there exists a $w\in W^{1,p}_0(\Omega)$ such that \begin{equation} \label{e2.2} \begin{gathered} w_{n}\rightharpoonup w\quad \text{in }W^{1,p}_0(\Omega),\\ w_{n}\to w\quad \text{a.e. in } \Omega,\\ w_{n}\to w\quad \text{in }L^r(\Omega),\text{ with }1\leq r0$ such that \begin{equation} \label{e2.6} \frac{F(x,t)}{|t|^{p}}>1, \end{equation} for any $x\in\Omega$ and $|t|\geq C_0$. Since $F(x,t)$ is continuous on $\overline{\Omega}\times[-C_0,C_0]$, there is an $M>0$ such that \begin{equation} \label{e2.7} |F(x,t)|\leq M, \quad \text{for all }(x,t)\in\overline{\Omega}\times[-C_0,C_0]. \end{equation} From \eqref{e2.6}, \eqref{e2.7}, we see that there is a constant $C$ such that for any $(x,t)\in\overline{\Omega}\times\mathbb{R}$, we have \begin{equation} \label{e2.8} F(x,t)\geq C \end{equation} which shows that \[ \frac{F(x,u_{n}(x))-C}{\|u_{n}\|^{p}}\geq0. \] This implies that \begin{equation} \label{e2.9} \frac{F(x,u_{n}(x))}{|u_{n}(x)|^{p}}|w_{n}(x)|^{p}-\frac{C}{\|u_{n}\|^{p}}\geq0. \end{equation} Using \eqref{e2.1} we have $$ c=I(u_{n})+o(1)=\frac{1}{p}\|u_{n}\|^{p}-\int_{\Omega}F(x,u_{n})\,\mathrm{d}x+o(1). $$ So we see that \begin{equation} \label{e2.10} \|u_{n}\|^{p}=pc+p\int_{\Omega}F(x,u_{n})\,\mathrm{d}x+o(1). \end{equation} By \eqref{e2.1} and \eqref{e2.10}, we obtain \begin{equation} \label{e2.11} \int_{\Omega}F(x,u_{n})\,\mathrm{d}x\to+\infty. \end{equation} We claim that $|\Omega_{\neq}|=0$. In fact, if $|\Omega_{\neq}|\neq0$, then combining \eqref{e2.5} and \eqref{e2.9} with Fatou's lemma, one has \begin{equation} \label{e2.12} \begin{aligned} +\infty &= \int_{\Omega_{\neq}}\liminf_{n\to+\infty}\frac{F(x,u_{n}(x))}{|u_{n}(x)|^{p}}|w_{n}(x)|^{p}\,\mathrm{d}x- \int_{\Omega_{\neq}}\limsup_{n\to+\infty}\frac{C}{\|u_{n}\|^{p}}\,\mathrm{d}x\\ &\leq \int_{\Omega_{\neq}}\liminf_{n\to+\infty}\Big(\frac{F(x,u_{n}(x))}{|u_{n}(x)|^{p}}|w_{n}(x)|^{p}- \frac{C}{\|u_{n}\|^{p}}\Big)\,\mathrm{d}x\\ &\leq \liminf_{n\to+\infty}\int_{\Omega_{\neq}}\Big(\frac{F(x,u_{n}(x))}{|u_{n}(x)|^{p}}|w_{n}(x)|^{p}- \frac{C}{\|u_{n}\|^{p}}\Big)\,\mathrm{d}x\\ &\leq \liminf_{n\to+\infty}\int_{\Omega}\Big(\frac{F(x,u_{n}(x))}{|u_{n}(x)|^{p}}|w_{n}(x)|^{p}- \frac{C}{\|u_{n}\|^{p}}\Big)\,\mathrm{d}x\\ &= \liminf_{n\to+\infty}\int_{\Omega}\frac{F(x,u_{n}(x))}{\|u_{n}\|^{p}}\,\mathrm{d}x\\ &\leq \liminf_{n\to+\infty}\frac{\int_{\Omega}F(x,u_{n}(x)) \,\mathrm{d}x}{pc+p\int_{\Omega}F(x,u_{n})\,\mathrm{d}x+o(1)}. \end{aligned} \end{equation} So by \eqref{e2.11} and \eqref{e2.12} we deduce a contradiction. This shows that $|\Omega_{\neq}|=0$. Hence $w(x)=0$ a.e. in $\Omega$. Since $I(tu_{n})$ is continuous in $t\in[0,1]$, there exists $t_{n}\in[0,1]$ such that $$ I(t_{n}u_{n})=\max_{t\in[0,1]}I(tu_{n}). $$ Clearly, $t_{n}>0$ and $I(t_{n}u_{n})\geq 0=I(0)$. If $t_{n}<1$ we have that $\frac{\mathrm{d}}{\mathrm{d}t}I(tu_{n})|_{t=t_{n}}=0$, which gives $\langle I'(t_{n}u_{n}), t_{n}u_{n}\rangle=0$. If $t_{n}=1$, then \eqref{e2.1} gives that $\langle I'(u_{n}), u_{n}\rangle=o(1)$. So we always have $$ \langle I'(t_{n}u_{n}), t_{n}u_{n}\rangle=o(1). $$ From (F1), for $t\in[0,1]$ we have \begin{equation} \label{e2.13} \begin{aligned} p I(tu_{n}) &\leq p I(t_{n}u_{n})\\ &= p I(t_{n}u_{n})-\langle I'(t_{n}u_{n}), t_{n}u_{n}\rangle+o(1)\\ &= \int_{\Omega}[t_{n}u_{n}f(x,t_{n}u_{n})-p F(x,t_{n}u_{n})]\,\mathrm{d}x+o(1)\\ &\leq \int_{\Omega}[\theta(u_{n}f(x,u_{n})-p F(x,u_{n}))+\alpha]\,\mathrm{d}x+o(1)\\ &\leq \theta(\|u_{n}\|^{p }+p c-\|u_{n}\|^{p }+o(1))+\alpha|\Omega|+o(1)\\ &\leq p \theta c+\alpha|\Omega|+o(1). \end{aligned} \end{equation} where we used \eqref{e2.1} and \eqref{e2.10}, $\theta$ and $\alpha$ as in (F1). Furthermore, by (F2), for every $\varepsilon>0$, there exists $a(\varepsilon)>0$, such that \[ |F(x,t)|\leq \frac{1}{2C_1}\varepsilon|t|^{p^{\ast}}+a(\varepsilon), \quad \text{for $t\in\mathbb{R}$ a.e. $x\in\Omega$}. \] Let $\delta=\varepsilon/(2a(\varepsilon))>0$, $E\subseteq\Omega$, $\operatorname{meas}E<\delta$, we have \begin{align*} \big|\int_{E}F(x,w_{n})\,\mathrm{d}x\big| &\leq \int_{E}|F(x,w_{n})|\,\mathrm{d}x\\ &\leq \int_{E}a(\varepsilon)\,\mathrm{d}x +\frac{1}{2C_1}\varepsilon\int_{E}|w_{n}|^{p^{\ast}}\,\mathrm{d}x\\ &\leq \frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon, \end{align*} hence $\{\int_{\Omega}F(x,w_{n})\,\mathrm{d}x,\,n\in N\}$ is equi-absolutely-continuous. It follows easily from Vitali Convergence Theorem that \[ \int_{\Omega}F(x,w_{n})\,\mathrm{d}x\to\int_{\Omega}F(x,0)\,\mathrm{d}x=0; \] So, for any $R_0>0$, \begin{equation} \label{e2.14} p I(R_0w_{n})=\|R_0w_{n}\|^{p}-p\int_{\Omega}F(x,R_0w_{n})\,\mathrm{d}x =R_0^{p}+o(1). \end{equation} From \eqref{e2.13}, we obtain \begin{equation} \label{e2.15} p I(tu_{n}) \leq p \theta c+\alpha|\Omega|+o(1), \end{equation} for $t\in[0,1]$. So combining \eqref{e2.14} with \eqref{e2.15}, \[ R_0^{p}+o(1)=p I(R_0w_{n}) \leq p \theta c+\alpha|\Omega|+o(1). \] Letting $n\to\infty$ we obtain \[ R_0^{p} \leq p\theta c+\alpha|\Omega|+o(1). \] Letting $R_0\to\infty$, we obtain a contradiction. Hence $ \| u_{n}\|$ is bounded. By the continuity of the embedding, we have $\|u_{n}\|_{2^{\ast}}^{2^{\ast}}\leq C_{2}<\infty$ for all $n$. If necessary going to a subsequence, one obtains \[ u_{n}\rightharpoonup u\text{ in }W^{1,p}_0(\Omega),\quad \text{and}\quad u_{n}\to u\text{ in }L^r(\Omega),\quad \text{where }1\leq r0$, there exists $a(\varepsilon)>0$, such that $$ |f(x,t)t|\leq \frac{1}{2C_{2}}\varepsilon|t|^{p^{\ast}}+a(\varepsilon), \quad \text{for } t\in\mathbb{R}, \text{ a.e. }x\in\Omega. $$ Let $\delta=\varepsilon/(2a(\varepsilon))>0$, $E\subseteq\Omega$, $\operatorname{meas}E<\delta$, we have \begin{align*} \big|\int_{E}f(x,u_{n})u_{n}\,\mathrm{d}x\big| &\leq \int_{E}|f(x,u_{n})u_{n}|\,\mathrm{d}x\\ &\leq \int_{E}a(\varepsilon)\,\mathrm{d}x +\frac{1}{2C_{2}}\varepsilon\int_{E}|u_{n}|^{p^{\ast}}\,\mathrm{d}x\\ &\leq \frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon, \end{align*} hence $\{\int_{\Omega}f(x,u_{n})u_{n}\,\mathrm{d}x,\,n\in N\}$ is equi-absolutely-continuous. It follows easily from Vitali Convergence Theorem that \begin{equation} \label{e2.16} \int_{\Omega}f(x,u_{n})u_{n}\,\mathrm{d}x\to\int_{\Omega}f(x,u)u\,\mathrm{d}x. \end{equation} From (F2), for any $\varepsilon>0$ there exists $a(\varepsilon)>0$ such that $$ |f(x,t)|\leq \frac{1}{2c_1c_{2}}\varepsilon|t|^{p^{\ast}-1}+a(\varepsilon) \quad \text{for } t\in\mathbb{R},\,x\in\Omega. $$ where \[ c_1\geq\Big(\int_{\Omega}|u_{n}|^{p^{\ast}}\,\mathrm{d}x\Big) ^{\frac{p^{\ast}-1}{p^{\ast}}}\quad\forall n;\quad c_{2}:=\Big(\int_{\Omega}|u|^{p^{\ast}}\, \mathrm{d}x\Big)^{\frac{1}{p^{\ast}}}\,. \] From H\"{o}lder's inequality, for every $E\subseteq\Omega$, we have \begin{gather*} \int_{E}a(\varepsilon)|u|\,\mathrm{d}x \leq a(\varepsilon)(\operatorname{meas}E)^{\frac{p^{\ast}-1}{p^{\ast}}} \Big(\int_{E}|u|^{p^{\ast}}\,\mathrm{d}x\Big)^{\frac{1}{p^{\ast}}} \leq a(\varepsilon)(\operatorname{meas}E)^{\frac{p^{\ast}-1}{p^{\ast}}}c_1; \\ \int_{E}|u_{n}|^{p^{\ast}-1}|u|\,\mathrm{d}x \leq\Big(\int_{E}|u_{n}|^{p^{\ast}}\,\mathrm{d}x\Big)^{\frac{p^{\ast}-1}{p^{\ast}}} \Big(\int_{E}|u|^{p^{\ast}}\, \mathrm{d}x\Big)^{\frac{1}{p^{\ast}}}\leq c_1c_{2}. \end{gather*} Let $\delta=\big(\frac{\varepsilon}{2c_1a(\varepsilon)}\big) ^{\frac{p^{\ast}}{p^{\ast}-1}}>0$, $E\subseteq\Omega$, $\operatorname{meas}E<\delta$, we have \begin{align*} \big|\int_{E}f(x,u_{n})u\,\mathrm{d}x\big| &\leq \int_{E}|f(x,u_{n})u|\,\mathrm{d}x\\ &\leq \int_{E}a(\varepsilon)|u|\,\mathrm{d}x +\frac{1}{2c_1c_{2}}\varepsilon\int_{E}|u_{n}|^{p^{\ast}-1}|u|\,\mathrm{d}x\\ &\leq \frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon, \end{align*} hence $\{\int_{\Omega}f(x,u_{n})u\,\mathrm{d}x,\,n\in N\}$ is also equi-absolutely-continuous. It follows from Vitali Convergence Theorem that \begin{equation} \label{e2.17} \int_{\Omega}f(x,u_{n})u\,\mathrm{d}x\to\int_{\Omega}f(x,u)u\,\mathrm{d}x. \end{equation} Since \begin{equation} \label{e2.18} \langle I'(u_{n}),u\rangle=\int_{\Omega}\left(|\nabla u_{n}|^{p-2} \nabla u_{n}\cdot\nabla u-f(x,u_{n})u\right)\,\mathrm{d}x\to0; \end{equation} \begin{equation} \label{e2.19} \langle I'(u_{n}),u_{n}\rangle=\int_{\Omega} \left(|\nabla u_{n}|^{p-2}\nabla u_{n}\cdot\nabla u_{n}-f(x,u_{n})u_{n}\right)\, \mathrm{d}x\to0. \end{equation} It follows from \eqref{e2.16}-\eqref{e2.19} that $\|u_{n}\|\to\|u\|.$ By Kadec-Klee property, we have \[ u_{n}\to u \quad \text{in } W^{1,p}_0(\Omega). \] \smallskip \noindent\textbf{Second step:} Mountain-pass geometric structure. $I$ has a mountain pass geometry; i.e., there exist $u_1\in W^{1,p}_0(\Omega)$ and constants $r, \rho > 0$ such that $I(u_1) < 0$, $\|u_1\| > r$ and \begin{equation} \label{e2.20} I(u) \geq \rho,\quad \text{when }\| u\| = r. \end{equation} Indeed, By (F3), we have $t_0 > 0$ and $\lambda \in (0, \lambda_1)$ such that $$ \frac{f(x,t)}{t|t|^{p-2}}<\lambda,\quad \text{for }|t|0 $$ is the first eigenvalue of the operator $-\Delta_p $ with the Dirichlet boundary value in $\Omega$. This implies that $$ F(x,t)\leq\frac{\lambda}{p}|t|^{p},\quad \text{for }|t|\leq t_0. $$ This inequality with (F2) shows that $$ F(x,t)\leq\frac{\lambda}{p}|t|^{p}+C|t|^{p^{\ast}},\quad \text{for }t\in\mathbb{R} $$ with some $C > 0$. Since $\lambda_1>0$ denotes the first eigenvalue of the operator $-\Delta_p $ with the Dirichlet boundary value in $\Omega$, it follows that $\| u\|^{p}\geq\lambda_1\| u\|_p^{p}$ for $u\in W^{1,p}_0(\Omega)$. Then $I$ is estimated as $$ I(u)\geq \frac{1}{p}\| u\|^{p}-\frac{\lambda}{p}\| u\|_p^{p} -C\| u\|_{p^{\ast}}^{p^{\ast}}\geq \frac{\lambda_1-\lambda}{p\lambda_1}\| u\|^{p}-C'\| u\|^{p^{\ast}}. $$ This shows the existence of $r$ and $\rho$ satisfying: \[ I(u) \geq \rho,\quad \text{when }\| u\| = r. \] From (F4) follows that, for all $M>0$ there exists $C_{M}>0$, such that \begin{equation} \label{e2.21} F(x,t)\geq M|t|^{p}-C_{M},\quad \forall\,x\in\Omega,\; t>0. \end{equation} Let $\phi$ be a function such that $\phi\in W^{1,p}_0(\Omega)$, $\phi\geq0$, $\phi\not\equiv 0$. From \eqref{e2.21} we obtain \begin{align*} I(t\phi) &=\frac{|t|^{p}}{p}\| \phi\|^{p}-\int_{\Omega}F(x,t\phi)\,\mathrm{d}x\\ &\leq \frac{|t|^{p}}{p}\| \phi\|^{p}-t^{p}\int_{\Omega}M\phi^{p} \,\mathrm{d}x+c|\Omega|\to-\infty\quad \text{as }t\to\infty. \end{align*} We fix $t > 0$ large so that $I(t\phi) < 0$ and $t\| \phi\| > r$. Let $u_1:= t\phi\in W^{1,p}_0(\Omega)$ and then constants $r,\, \rho > 0$ such that $I(u_1) < 0,\,\|u_1\| > r$ and satisfies \eqref{e2.20}, i.e. $I$ has a mountain pass geometry. \smallskip \noindent\textbf{Third step:} Critical value of $I$. For $u_1$ in second step, we define \begin{gather*} \Gamma:=\{\gamma: C[0,1]\to W^{1,p}_0(\Omega):\gamma(0)=0,\; \gamma(1)=u_1\},\\ c_0:=\inf_{\gamma\in \Gamma}\max_{0\leq t\leq1}I(\gamma(t)) \end{gather*} As shown in \cite{Schechter1}, a deformation lemma can be proved with the (C) condition, replacing the usual Palais-Smale condition, and it turns out that the Mountain Pass Theorem still holds. Then $c_0$ is a critical value of $I$. For the proof, we refer the reader to \cite{Rabinowitz,Struwe,Willem1}. \begin{thebibliography}{99} \bibitem{Ambrosetti} A. Ambrosetti, P. Rabinowitz; \emph{Dual variational methods in critical point theory and applications}, J. Funct. Anal. 14 (1973), 349--381. \bibitem{Bartsch} T. Bartsch, Y. H. Ding; \emph{On a nonlinear Schr$\ddot{o}$dinger equation with periodic potential}, Math. Ann. 313 (1999), 15--37. \bibitem{Chen} Z. H. Chen, Y. T. Shen, Y. X. Yao; \emph{Some existence results of solutions for $p$-Laplacian}, Acta Mathematica Scientia, 23 ( 2003), B (4):487--496. \bibitem{Costa} D. G. Costa, C. A. Magalh\~aes; \emph{Variational elliptic problems which are nonquadratic at infinity}, Nonlinear Anal., Theory Methods Appl. 23 (1994), 1401--1412. \bibitem{Coti} V. Coti-Zelati, P. Rabinowitz; \emph{Homoclinic type solutions for a semilinear elliptic PDE on $\mathbb{R}^{n}$}, Commun. Pure Appl. Math 46 (1992), 1217--1269. \bibitem{LDing} L. Ding, C. L. Tang; \emph{Positive solutions for critical quasilinear elliptic equations with mixed Dirichlet-Neumann boundary conditions}, Acta Mathematica Scientia, 33 ( 2013), B (2):443--470. \bibitem{YDing} Y. H. Ding, S. X. Luan; \emph{Multiple solutions for a class of Schr$\ddot{o}$dinger equations}, J. Differential Equations 207 (2004), 423--457. \bibitem{DeFigueiredo} D. G. De Figueiredo, J. P. Gossez, P. Ubilla; \emph{Local superlinearity and sublinearity for indefinite semilinear elliptic problems}, J. Funct. Anal. 199 (2003), 452--467. \bibitem{Izydorek} M. Izydorek, J. Janczewska; \emph{Homoclinic solutions for a class of the second order Hamiltonian systems}. J. Differential Equations 219 (2005), 375--389. \bibitem{Lam} N. Lam, G. Z. Lu; \emph{Elliptic equations and systems with subcritical and critical exponential growth without the Ambrosetti--Rabinowitz condition}, J Geom Anal, 24, no. 1, (2014), 118--143. \bibitem{Li1} G. B. Li, C. Y. Yang; \emph{The existence of a nontrivial solution to a nonlinear elliptic boundary value problem of $p$-Laplacian type without the Ambrosetti-Rabinowitz condition}, Nonlinear Anal.,72 (2010), 4602--4613. \bibitem{Li} G. B. Li, H. S. Zhou; \emph{Asymptotically \lq\lq linear\rq\rq Dirichlet problem for the $p$-Laplacian}, Nonlinear Anal., Theory Methods Appl. 43 (2001), 1043--1055. \bibitem{Liu} S. B. Liu; \emph{On superlinear problems without the Ambrosetti and Rabinowitz condition}, Nonlinear Anal.,73 (2010), 788--795. \bibitem{Mao} A. M. Mao, Y. Zhu, S. X. Luan; \emph{Existence of solutions of elliptic boundary value problems with mixed type nonlinearities}, Boundary Value Problems, http://www.boundaryvalueproblems.com/content/2012/1/97, (2012). \bibitem{Miyagaki} O. H. Miyagaki, M. A. S. Souto; \emph{Superlinear problems without Ambrosetti and Rabinowitz growth condition}, J. Differential Equations 12 (2008), 3628--3638. \bibitem{Ou} Z. Q. Ou, C. Li, J. J. Yuan; \emph{Multiplicity of nontrivial solutions for quasilinear elliptic equation}, J. Math. Anal. Appl., 388 (2012), 198--204. \bibitem{Rabinowitz} P. H. Rabinowitz; \emph{Minimax methods in criticl point theory with applications to dierential equations}, CBMS (1986). \bibitem{Schechter1} M. Schechter; \emph{Superlinear elliptic boundary value problems}. Manuscr. Math. 86 (1995), 253--265. \bibitem{Schechter} M. Schechter, W. M. Zou; \emph{Superlinear problems}, Pacific J. Math. 214 (2004), 145--160. \bibitem{Struwe} M. Struwe; \emph{Variational Methods}, second edition, Springer, Berlin, (1996). \bibitem{Willem1} M. Willem; \emph{Minimax Theorems}, Birkh\"auser, Boston, (1996). \bibitem{Willem} M. Willem, W. M. Zou; \emph{On a Schr\"odinger equation with periodic potential and spectrum point zero}, Indiana Univ. Math. J. 52 (2003), 109--132. \end{thebibliography} \end{document}