\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2008(2008), No. 61, pp. 1--12.\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 2008 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2008/61\hfil Three solutions] {Three solutions for singular $p$-Laplacian type equations} \author[Z. Yang, D. Geng, H. Yan\hfil EJDE-2008/61\hfilneg] {Zhou Yang, Di Geng, Huiwen Yan} % keep this order, not by alphabetical order \address{Zhou Yang \newline School of Math. Sci., South China Normal University, Guangzhou 510631, China} \email{yangzhou@scnu.edu.cn} \address{Di Geng \newline School of Math. Sci., South China Normal University, Guangzhou 510631, China} \email{gengdi@scnu.edu.cn} \address{Huiwen Yan \newline School of Math. Sci., South China Normal University, Guangzhou 510631, China} \email{hwyan10@yahoo.com.cn} \thanks{Submitted January 14, 2008. Published April 22, 2008.} \thanks{Supported by grants 10671075 from the National Natural Science Foundation of China, \hfill\break\indent 5005930 from the National Natural Science Foundation of Guangdong, and \hfill\break\indent 20060574002 from the University Special Research Fund for Ph. D. Program.} \subjclass[2000]{35J60} \keywords{$p$-Laplacian operator; singularity; multiple solutions} \begin{abstract} In this paper, we consider the singular $p$-Laplacian type equation \begin{gather*} -\mathop{\rm div}(|x|^{-\beta} a(x,\nabla u)) =\lambda f(x,u),\quad \mbox{in }\Omega,\\ u=0,\quad \mbox{on }\partial\Omega, \end{gather*} where $0\leq\beta1$. In this case \eqref{eq1.1} reduces to an equation involving the $p$-Laplacian operator. Under the assumptions that the nonlinear term $f(u):\mathbb{R}\to\mathbb{R}$ is continuous, $(p-1)$-sublinear at infinity and $(p-1)$-superlinear at the origin, Krist\'aly applied Bonanno's variational principle to \eqref{eq1.1} and obtain the existence of three weak solutions. In the present paper, we investigate the existence and multiplicity of solutions to the singular $p$-Laplacian type equation \begin{equation} \label{eq1.2} \begin{gathered} -\mathop{\rm div}(|x|^{-\beta} a(x,\nabla u)) =\lambda f(x,u),\quad \mbox{in }\Omega,\\ u=0,\quad \mbox{on }\partial\Omega, \end{gathered} \end{equation} where $0\leq\beta1$ and a positive constant $a_1$ such that $$ |a(x,\xi)|\leq a_1(1+|\xi|^{p-1})\quad\mbox{for a.e. }x\in\Omega \mbox{ and all } \xi\in \mathbb{R}^N; $$ \item[(A3)] $A(x,\xi)$ is strictly convex in $\xi$, that is, for $\xi,\eta\in\mathbb{R}^N$ with $\xi\neq\eta$ $$ 2A\left(x,{\xi+\eta\over2}\right)0. \end{equation} If $F(x,u)$ admits the asymptotic property at the origin: \begin{equation} \label{C2} \mathcal{F}(u)|u|^{-p}\to0\quad\mbox{as }u\to0, \end{equation} then, there exists an open interval $\Lambda\subset[0,+\infty)$ and a number $R>0$ such that for every $\lambda\in \Lambda$, equation \eqref{eq1.2} has at least three distinct solutions in $X$, whose $X$-norms are less than $R$. \end{theorem} Note that when $\beta=\alpha=0$ and $f(x,u)=f(u)$, Theorem \ref{thm1.1} implies the conclusion in \cite[Theorem 2.1]{Alexandru}. The conclusion in Theorem \ref{thm1.1} still holds if the asymptotic property of $f(x,u)$ at the origin is replaced by some other properties. To state the next result, we introduce the following notation: \begin{equation} \label{C3} c_2(s)=\inf_{x\in B(x_0,r/2)} {F(x,s)\over1+|s|^p},\quad c_3(s)=\sup_{|u|\geq s} \mathcal{F}(u)|u|^{-p},\quad c_4(s)=\sup_{|u|\leq s}\mathcal{F}(u), \end{equation} where $B(x_0,r)\subset\Omega$ and $s\geq0$. \begin{theorem} \label{thm1.2} Assume {\rm (A1)--(A4), (B1)--(B2)} are satisfied. Let $E=B(x_0,r)$ be a ball contained in $\Omega$, such that \begin{equation} \label{C3b} F(x,u)\geq0,\quad\mbox{ for a.e. }x\in E\mbox{ and all }u\in I, \end{equation} where $I$ is either $\mathbb{R}^+$ or $\mathbb{R}^-$. If there exist $L>0$ and $K\in I$ such that \begin{equation} \label{C4} c_2(K)|K|^p\geq C c_4(L),\quad c_2(K)>C(c_3(L))^{p\over q}(c_4(L))^{q-p\over q}L^{p(p-q)\over q}, \end{equation} where $q=p^*(\beta,\alpha)$ and $C$ is a certain positive constant only dependent on $p$, $\beta$, $\alpha$, $N$, $E$, $a_1$ and $a_2$. Then the conclusion in Theorem \ref{thm1.1} remains valid. \end{theorem} \begin{remark} \label{rmk1.3}\rm The above result is new even in the case of $\beta=\alpha=0$. Moreover, by the method similar to \cite{Yang}, we can show a more general result. \end{remark} \begin{remark} \label{rmk1.4} \rm If we fix some $L$ and keep $c_2(K)/c_3(L)$ less than a fixed constant, then assumption \eqref{C4} holds when $K>L$ and $c_2(K)$ is large enough. \end{remark} \section{Preliminaries} Firstly, we recall the generalized Hardy-Sobolev imbedding theorem, which can be deduced from Caffarelli-Kohn-Nirenberg inequality (see \cite{Caffarelli,Xuan}). \begin{lemma} \label{lem2.1} Suppose that $\beta^*_1\leq\widetilde{\alpha}\leq\beta^*_2$ and $\beta^*_1\leq\widehat{\alpha}<\beta^*_3$. Let $U$ be an arbitrary smooth bounded domain in ${\mathbb{R}^N}$ containing the origin. We have \begin{itemize} \item[(i)] There exists a constant $S_{\widetilde{\alpha}}>0$, such that for any $u\in\mathcal{D}^{1,p}({\mathbb{R}^N}, |x|^{-\beta}dx)$, there holds $$ S_{\widetilde{\alpha}}\|u\|_{L^{p^*(\beta,\widetilde{\alpha})} ({\mathbb{R}^N},|x|^{-\widetilde{\alpha}}dx)}^p\leq\|u\|_ {\mathcal{D}^{1,p}({\mathbb{R}^N},|x|^{-\beta}dx)}^p\;, $$ where $L^p(U,|x|^{-\alpha}dx)$ is $L^p$ space with $|x|^{-\alpha}$ as weight. \item[(ii)] For $1\leq\widetilde q\leq p^*(\beta,\widetilde{\alpha})$, there exists a constant $ S_{\widetilde{q},\widetilde{\alpha}}>0$ such that for any $u\in\mathcal{D}$$^{1,p}(U,|x|^{-\beta}dx)$, there holds $$ S_{\widetilde{q},\widetilde{\alpha}}\|u\|_{L^{\widetilde q} (U,|x|^{-\widetilde{\alpha}}dx)}^p \leq\|u\|_{\mathcal{D}^{1,p}(U,|x|^{-\beta}dx)}^p\;, $$ Moreover, $S_{\widetilde{\alpha}}=S_{\widetilde{q},\widetilde{\alpha}}$ is independent of the domain $U$ provided $\widetilde{q}= p^*(\beta,\widetilde{\alpha})$. \item[(iii)] $\mathcal{D}$$^{1,p}(U,|x|^{-\beta}dx)$ compactly imbeds into $L^{\widehat{q}}(U,|x|^{-\widehat{\alpha}})$ provided $1\leq\widehat q\widehat{q}$. Combining the above two inequalities, we obtain $$ 0\leq\int_U|u_n-u|^{\widehat{q}}|x|^{-\widehat{\alpha}}dx \leq C_m\|u_n-u\|^{\widehat{q}}_{L^{\widehat{q}}(U)} +C\Big(\int_{\overline{B}_{\rho_m}(0)}|x| ^{-\widehat{\alpha}}dx\Big)^{(\tau-\widehat{q}\;)/\tau}. $$ First let $n\to\infty$, then $m\to\infty$, and we derive that $u_n$ strongly converges to $u$ in $L^{\widehat{q}}(U,|x|^{-\widehat{\alpha}})$. \end{proof} Secondly, we review Bonanno's three critical points theorem (see \cite{Bonanno}), which is the main variational tool in this paper. \begin{lemma} \label{lem2.2} Let $\mathcal{X}$ be a separable and reflexive real Banach space, and let $\phi,\psi:{\mathcal{X}}\to \mathbb{R}$ be two continuously G\^ateaux differentiable functionals. Assume that \begin{itemize} \item[(D1)] There exists a function $u_0\in{\mathcal{X}}$ such that $\phi(u_0)=\psi(u_0)=0$ and $\phi(u)\geq0$ for every $u\in{\mathcal{X}}$. \item[(D2)] There exists a function $u_1\in{\mathcal{X}}$ and a positive number $\rho$ such that \begin{equation} \label{eq2.1} \rho<\phi(u_1),\quad \sup_{\phi(u)<\rho}\psi(u)<\rho {\psi(u_1)\over\phi(u_1)}. \end{equation} \item[(D3)] Further, put $$ \gamma=\xi\rho\Big[\rho {\psi(u_1)\over\phi(u_1)}- \sup_{\phi(u)<\rho}\psi(u)\Big]^{-1}, $$ with $\xi>1$, and suppose that for every $\lambda\in[0,\gamma]$, the functional $\phi(u)-\lambda\psi(u)$ is sequentially weakly lower semicontinuous, satisfies the P.-S. condition and \begin{equation} \label{eq2.2} \lim_{\|u\|\to+\infty}\Big[\phi(u)-\lambda\psi(u)\Big]=+\infty. \end{equation} \end{itemize} Then, there exists an open interval $\Lambda\subset[0,\gamma]$ and a number $R>0$ such that, for any $\lambda\in \Lambda$, the equation $\phi'(u)-\lambda\psi'(u)=0$ admits at least three solutions in ${\mathcal{X}}$ whose norms are less than $R$. \end{lemma} In the sequel, by setting ${\mathcal{X}}=X=\mathcal{D}^{1,p}(\Omega,|x|^{-\beta}dx)$, $\phi(u)=\Phi(u)$, $\psi(u)=\Psi(u)$ and $\xi=+\infty$ we show that the variational functional $I(u)$ satisfies all assumptions in Lemma \ref{lem2.2}. \begin{lemma} \label{lem2.3} Suppose that the assumptions {\rm (B1), (B2)} are satisfied. Then $\Psi(u)$ is weakly continuous on $X$, i.e., if $u_n$ weakly converges to $u$ in $X$, $\Psi(u_n)$ converges to $\Psi(u)$. \end{lemma} \begin{proof} According to assumptions {\rm (B1), (B2)}, it is not difficult to deduce that, for each $\epsilon>0$, there exists some positive number $M_\epsilon$ such that \begin{gather} \label{eq2.3} |f(x,u)u|+|F(x,u)|\leq \epsilon|u|^p|x|^{-\beta_2^*},\quad \mbox{a.e. }x\in \Omega \mbox{ and all }|u|\in[M_\epsilon,+\infty);\\ \label{eq2.4} |f(x,u)u|+|F(x,u)|\leq \epsilon|u|^p|x|^{-\beta_2^*} +C_\epsilon|u||x|^{-\alpha},\quad \mbox{a.e. }x\in \Omega\mbox{ and all }u\in \mathbb{R}, \end{gather} here $C_\epsilon$ is a positive number dependent only on $\epsilon$. Assume that $u_n$ converges weakly to $u$ in $X$, then for any $\epsilon\geq0$, we conclude \begin{align*} |F(x,u_n)-F(x,u)|&\leq |f(x,\theta u+(1-\theta)u_n)||u_n-u|\\ &\leq (\epsilon|u|^{p-1}|x|^{-\beta_2^*}+\epsilon|u_n|^{p-1}|x|^{-\beta_2^*} +\overline{C}_\epsilon|x|^{-\alpha})|u_n-u|, \end{align*} where $0<\theta<1$. The definition of $\Psi(u)$ thus implies that \begin{align*} |\Psi(u_n)-\Psi(u)| &\leq\int_\Omega|F(x,u_n)-F(x,u)|dx \\ &\leq \int_\Omega\Big(\epsilon {|u|^{p-1}+|u_n|^{p-1} \over|x|^{\beta_2^*}}+ {\overline{C}_\epsilon\over |x|^\alpha}\Big)|u_n-u|\,dx\\ &\leq C\epsilon(\|u_n\|_X^p+\|u\|_X^p) +\overline{C}_\epsilon\|u_n-u\|_{L^1(\Omega;|x|^{-\alpha}dx)}. \end{align*} Since $X$ compactly imbeds into $L^1(\Omega;|x|^{-\alpha}dx)$, taking $n\to\infty$, we obtain $$ \limsup_{n\to\infty}|\Psi(u_n)-\Psi(u)|\leq C\epsilon\|u\|_X^p. $$ Let $\epsilon\to0^+$ in the above inequality, and the conclusion in the lemma follows. \end{proof} \begin{lemma} \label{lem2.4} Suppose that the assumptions {\rm (A1)--(A4), (B1)--(B2)} are satisfied. Then $I(u)$ is weakly lower semicontinuous on $X$. \end{lemma} \begin{proof} Owing to previous lemma, it suffice to show weakly lower semicontinuity of $\Phi(u)$ on $X$. We argue by contradiction, assume that $\{u_n\}$ is a function sequence weakly converging to $u$ in $X$, but there is a subsequence $u_{n_k}$ such that $\lim_{k\to\infty}\Phi(u_{n_k})>\Phi(u)$. Without loss of generalization, one can assume that $$ \Phi(u_{n_k})>\Phi(u)+\delta,\quad\mbox{for }k=1,2,\dots, $$ where $\delta$ is a positive number. In view of Mazur theorem, there exists a sequence $\{v_m\}$ strongly converging to $u$ in $X$, where $v_m$ is a convex combination of finitely many $u_{n_k}$; i.e., for any $m\in{\mathbb Z}^+$, $$ v_m=\sum_{i=1}^m \alpha_{m i} u_{n_{k_i}}, \quad\mbox{with } \alpha_{m i}>0,\;\; \sum_{i=1}^m \alpha_{m i}=1. $$ Since $A(x,\xi)$ is convex with respect to $\xi$, we then derive \begin{align*} \Phi(v_m)&\geq \sum_{i=1}^m \alpha_{m i} \int_\Omega A(x,\nabla u_{n_{k_i}}) |x|^{-\beta}dx \\ &= \sum_{i=1}^m \alpha_{m i} \Phi(u_{n_{k_i}})>\Phi(u)+\delta,\quad \mbox{for }m=1,2,\dots, \end{align*} which contradicts that $\{v_m\}$ strongly converges to $u$ in $X$. \end{proof} \begin{lemma} \label{lem2.5} Suppose that the assumptions {\rm (A1)--(A4), (B1)--(B2)} are satisfied. Then $I(u)$ satisfies the P.-S. condition. \end{lemma} \begin{proof} Suppose that $\{u_n\}\subset X$ is a P.-S. sequence for $I(u)$, that is, $\{I(u_n)\}$ is bounded, and $\|I'(u_n)\|_{X^*}\to 0$ as $n\to0$, where $X^*$ is the dual space of $X$. We claim that $\{u_n\}$ admits a strongly convergent subsequence. Firstly, we show that $\{u_n\}$ is bounded in $X$. In fact, combining assumption {\rm (A4)}, (\ref{eq2.4}) and Lemma \ref{lem2.1}, we calculate \begin{align*} C\geq I(u_n) &=\int_\Omega A(x,\nabla u_n)|x|^{-\beta}dx- \lambda \int_\Omega F(x, u_n)dx\\ &\geq a_2\int_\Omega |\nabla u_n|^p|x|^{-\beta}dx -\lambda \int_\Omega ( \epsilon|u|^p|x|^{-\beta_2^*} +C_\epsilon|x|^{-\alpha})dx\\ &\geq (a_2-\lambda S^{-1}_{\beta_2^*}\epsilon)\|u_n\|_X^p -\overline C_\epsilon. \end{align*} Fix $\epsilon>0$ small enough that $a_2-\lambda S_{\beta_2^*}^{-1}\epsilon \geq a_2/2$, then we discover that $\{u_n\}$ is bounded in $X$. There thus exists a subsequence of $\{u_n\}$, still denoted by itself, such that $\{u_n\}$ weakly converges to $u$ in $X$. Moreover, without loss of generalization, one can assume that $f(x,u_n)$ weakly converges to $f(x,u)$ in $X^*$. We next demonstrate that there exists a subsequence of $\{u_n\}$, still denoted by itself, such that \begin{equation} \label{eq2.5} \lim_{n\to\infty}\nabla u_n=\nabla u \quad \mbox{a.e. in }\Omega. \end{equation} Indeed, the facts that $\{u_n\}$ is bounded in $X$ and $\|I'(u_n)\|_{X^*}\to 0$ as $n\to\infty$ implies that \begin{equation} \label{eq2.6} \langle I'(u_n)-I'(u), u_n-u\rangle=\langle I'(u_n), u_n-u\rangle -\langle I'(u), u_n-u\rangle=o(1),\quad\mbox{as }n\to\infty. \end{equation} Furthermore, repeat the argument in the proof of Lemma \ref{lem2.3}, and it is easy to deduce \begin{equation} \label{eq2.7} \begin{aligned} J(u,u_n)&:=\int_\Omega[f(x,u_n)-f(x,u)][u_n-u]dx\\ &= \int_\Omega f(x,u_n)(u_n-u)\, dx- \int_\Omega f(x,u)(u_n-u)\,dx=o(1), \end{aligned} \end{equation} as $n\to\infty$. On the other hand, \begin{equation} \label{eq2.8} \langle I'(u_n)-I'(u), u_n-u\rangle=\int_\Omega H(x,u,u_n)|x|^{-\beta}dx -\lambda J(u,u_n), \end{equation} where $$ H(x,u,u_n):=[ a(x,\nabla u_n) -a(x,\nabla u)]\cdot[\nabla u_n-\nabla u]. $$ Combining (\ref{eq2.6}), (\ref{eq2.7}) and (\ref{eq2.8}), we obtain \begin{equation} \label{eq2.9} \lim_{n\to\infty}\int_\Omega H(x,u,u_n)|x|^{-\beta}dx=0. \end{equation} Notice that $H(x,u,u_n)\geq0$ since $A(x,\xi)$ is convex in $\xi$. So, (\ref{eq2.9}) implies that there exists a subsequence of $\{u_n\}$, still denoted by itself, such that $H(x,u,u_n)\to0$ a.e. in $\Omega$ as $n\to\infty$. Hence, (\ref{eq2.5}) follows from the strict convexity of $A(x,\xi)$. Then, we prove that there exists a subsequence of $\{u_n\}$, still denoted by itself, such that \begin{equation} \label{eq2.10} \lim_{n\to\infty} \int_\Omega |x|^{-\beta} a(x,\nabla u_n)\cdot \nabla u_n\, dx= \int_\Omega |x|^{-\beta} a(x,\nabla u)\cdot \nabla u \,dx. \end{equation} According to the growth condition (A2) and (\ref{eq2.5}), we can assume that $a(x,\nabla u_n)$ weakly converges to $a(x,\nabla u)$ in $X^*$, maybe a subsequence of $\{u_n\}$. Recalling that $f(x,u_n)$ weakly converges to $f(x,u)$ in $X^*$, we infer that $I'(u_n)$ weakly converges to $I'(u)$ in $X^*$. Hence, as $n\to\infty$, we deduce \begin{align*} o(1) &=\langle I'(u_n), u_n-u\rangle-\langle I'(u_n)-I'(u),u\rangle\\ &=\langle I'(u_n), u_n\rangle-\langle I'(u),u\rangle \\ &=\int_\Omega |x|^{-\beta} [ a(x,\nabla u_n)\cdot \nabla u_n -a(x,\nabla u)\cdot \nabla u]dx\\ &\quad -\lambda\int_\Omega [f(x,u_n)u_n-f(x,u)u]\,dx. \end{align*} Repeating the procedure as in the proof of (\ref{eq2.7}), we can achieve (\ref{eq2.10}). On the other hand, since $A(x,\xi)$ is convex with $A(x,0)=0$ and satisfies elliptic condition, we observe \[ a(x,\xi)\cdot\xi\geq A(x,\xi)\geq a_2|\xi|^p,\quad \mbox{for all }\xi\in \mathbb{R}^N, \] which implies $a_2|\nabla u_n|^p$ and $a_2|\nabla u|^p$ being dominated by $a(x,\nabla u_n)\cdot \nabla u_n,\;a(x,\nabla u)\cdot \nabla u$, respectively. Combining (\ref{eq2.5}), (\ref{eq2.10}) and the dominated convergence theorem, we conclude that $\nabla u_n$ converges to $\nabla u$ in $L^p(\Omega,|x|^{-\beta}dx)$, that is $u_n$ strongly converges to $u$ in $X$. \end{proof} \section{Proof of the main results} To prove Theorems \ref{thm1.1} and \ref{thm1.2}, we set notation as follows: \begin{equation} \label{eq3.1} \Pi(F;M)=M^{p-q}\Big({\rho\over a_2 S_\alpha}\Big)^{q/p} \sup_{|u|\geq M}\mathcal{F}(u)|u|^{-p} + \mu(\Omega)\sup_{|u|\leq M}\mathcal{F}(u), \end{equation} where $\mu(\Omega):=\int_\Omega |x|^{-\alpha}\,dx$ and $q=p^*(\beta,\alpha)$ as defined in (\ref{betastar}). One can establish the next result. \begin{lemma} \label{lem3.6} Suppose that the hypothesis {\rm (B2)} and {\rm (A4)} are satisfied. For every $u\in X$ with $\Phi(u)\leq\rho$, we have $$ \Psi(u)\leq\Pi(F;M). $$ \end{lemma} \begin{proof} According to assumption (A4) and Lemma \ref{lem2.1}, for every $u\in \Phi^{-1}(-\infty,\rho]$, we have \begin{equation} \label{eq3.2} \|u\|^p_X\leq {\Phi(u)\over a_2}\leq {\rho\over a_2}\,,\quad \|u\|_\alpha^q\leq {\|u\|_X^q\over S_\alpha^{q/p}}\leq \Big({\rho\over a_2 S_\alpha}\Big)^{q/p}, \end{equation} where $\|u\|_\alpha^q:=\int_\Omega|u|^q|x|^{-\alpha}dx$. By setting $\Omega_M:=\{x\in\Omega:|u(x)|\geq M\}$, we can deduce \begin{equation} \label{eq3.3} \mu(\Omega_{M})\leq M^{-q}\int_{\Omega_{M}} |u|^q|x|^{-\alpha}\,dx \leq M^{-q} \|u\|_\alpha^q\,. \end{equation} By assumption (B2), for every $u\in\Phi^{-1}(-\infty,\rho]$, we have the following estimate: \begin{align*} \Psi(u)&=\int_\Omega F(x,u)\,dx\leq \sup_{|u|\geq M}\mathcal{F}(u)|u|^{-p} \int_{\Omega_{M}} |u|^p|x|^{-\alpha} +\int_{\Omega\setminus \Omega_{M}}F(x,u)\,dx\\ &\leq \sup_{|u|\geq M} \mathcal{F}(u)|u|^{-p}\|u\|_\alpha^p\mu(\Omega_{M})^{1-p/q} +\sup_{|u|\leq M} \mathcal{F}(u)\mu(\Omega). \end{align*} Combining (\ref{eq3.2}) and (\ref{eq3.3}), we obtain $\Psi(u)\leq\Pi(F;M)$ for every $u\in \Phi^{-1}(-\infty,\rho]$. \end{proof} \begin{proof}[Proof of Theorem \ref{thm1.1}] To apply Bonanno's three critical points theorem, we have to verify all conditions in Lemma \ref{lem2.2}. Recalling the definition of $\Phi(u),\Psi(u)$, we conclude that $\Phi(0)=\Psi(0)=0$ and $\Phi(u)\geq0$ for all $u\in X$, which is the condition {\bf (D1)} in Lemma \ref{lem2.2}. Put $\gamma=+\infty$, then Lemma \ref{lem2.4} and Lemma \ref{lem2.5} imply that the functional $I(u)=\Phi(u)-\lambda\Psi(u)$ is sequentially weakly lower semicontinuous and satisfies the P.-S. condition. Moreover, using {\rm (A4)}, (\ref{eq2.4}) and Lemma \ref{lem2.1}, we compute \begin{align*} \Phi(u)-\lambda\Psi(u) &\geq a_2\|u\|^p_X- \lambda\int_\Omega(\epsilon|u|^p|x|^{-\beta_2^*} +C_\epsilon|u||x|^{-\alpha})\,dx\\ &\geq a_2\|u\|^p_X-\epsilon \lambda S_{\beta^*_2}^{-1}\|u\|^p_X -C_\epsilon \lambda S_{1,\alpha}^{-1/p}\|u\|_X, \end{align*} fix a positive $\epsilon$ less than $a_2\lambda^{-1}S_{\beta^*_2}/2$, then (\ref{eq2.2}) is obvious and we manifest assumption (D3). In the following, we verify the condition (D2), or equivalently, (\ref{eq2.1}). In fact, we can define a function the same as in \cite{Alexandru}: \begin{equation} \label{eq3.4} u_\sigma(x)= \begin{cases} 0, &x\in \mathbb{R}^N\setminus E;\\ K, &x\in B(x_0,\sigma r);\\ { K\over r(1-\sigma)}(r-|x-x_0|), &x\in E\setminus B(x_0,\sigma r), \end{cases} \end{equation} where $0<\sigma<1$ to be determined later. Owing to assumption (\ref{C1}) and (B2), we observe that \begin{align*} \Psi(u_\sigma )&=\int_E F(x,u_\sigma )dx\\ &\geq \int_{E\cap\{u_\sigma (x)=K\}}F(x,u_\sigma )dx -\max_{|u|\leq |K|}\mathcal{F}(u)\int_{E\cap\{|u_\sigma (x)|<|K|\}} |x|^{-\alpha}\,dx \\ &\geq \inf_{x\in E}F(x,K)\int_{B(x_0,\sigma r)}\,dx -\max_{|u|\leq |K|}\mathcal{F}(u)\int_{E\setminus B(x_0,\sigma r)} |x|^{-\alpha}\,dx. \end{align*} As $\sigma\to1^-$, the first term on the right hand side of the above inequality tends to the positive constant $\omega r^N\inf_EF(x,K)$, here $\omega$ is the volume of the unit ball, and the second term goes to zero. We thus pick up some $\sigma$ and $u_\sigma $ such that $\Psi(u_\sigma )>0$. Furthermore, from assumption {\rm (A4)}, we see that $\Phi(u_\sigma )\geq a_2\|u_\sigma \|_X^p>0$. According to the Lemma \ref{lem3.6}, to verify (\ref{eq2.1}), it suffice to turn up two positive numbers $M$ and $\rho$, such that \begin{equation} \label{eq3.5} 0<\rho<\Phi(u_\sigma )\quad\mbox{and}\quad {\Pi(F;M)\over\rho}<{\Psi(u_\sigma )\over\Phi(u_\sigma )}. \end{equation} Indeed, in view of assumption (\ref{C2}) and (B2), we see that, for any $\varepsilon>0$, there exist some positive constant $M$ such that $ \mathcal{F}(u)\leq\varepsilon |u|^p$, for all $u\in[-M,M]$ and $\mathcal{F}(u)|u|^{-p}\leq C$ for all $u\in \mathbb{R}$, where $C$ is independent of $M$. Put $\rho=\delta^p M^p$ with $\delta$ is a positive number to be determined later, then we deduce $$ {\Pi(F;M)\over\rho}\leq C\delta^{q-p} \Big({1\over a_2 S_\alpha}\Big)^{q/p} +\varepsilon\delta^{-p}\mu(\Omega) $$ One can first fix $\delta>0$ small enough, then choose $\varepsilon>0$ so small that the right hand side of the above inequality is less than $\Psi(u_\sigma)/\Phi(u_\sigma)$, finally choose $M$ and $\rho$ satisfy (\ref{eq3.5}), which yields condition (\ref{eq2.1}). Hence, we testify all the conditions in Lemma \ref{lem2.2} and the desired conclusion follows. \end{proof} \begin{proof}[Proof of Theorem \ref{thm1.2}] Similar to the proof of Theorem \ref{thm1.1}, denote $u_\sigma$ as (\ref{eq3.4}) and fix $\sigma =1/2$. Owing to assumptions (\ref{C3b}) and (\ref{C4}), it is clear that \[ \Psi(u_\sigma) \geq\int_{E\cap\{u_\sigma(x)=K\}}F(x,u_\sigma)dx \geq c_2(K)(1+|K|^p)\int_{B(x_0,r/2)}\,dx. \] Moreover, recalling assumptions (A4) and (A2), we have \begin{equation} \label{eq3.11} \begin{gathered} \Phi(u_\sigma) \geq a_2\int_E|\nabla u_\sigma|^p|x|^{-\beta}\,dx \geq a_2\Big({2|K|\over r}\Big)^p \int_{E\setminus B(x_0,r/2)} |x|^{-\beta}\,dx, \\ \Phi(u_\sigma)\leq a_1\int_E(|\nabla u_\sigma| +|\nabla u_\sigma|^p)|x|^{-\beta}\,dx \leq a_1\Big({2|K|\over r}+\Big({2|K|\over r}\Big)^p\Big) \int_E|x|^{-\beta}\,dx. \end{gathered} \end{equation} We thus get \begin{equation} \label{eq3.6} \rho {\Psi(u_\sigma)\over\Phi(u_\sigma)}\geq\delta c_2(K)\rho, \end{equation} where $\delta$ is a positive constant dependent only on $p,\beta,N,E$ and $a_1$. On the other hand, let $M=L$ in \eqref{eq3.1}, according to the definition in \eqref{C3}, we obtain \begin{equation} \label{eq3.7} \Pi(F;L)\leq c_3(L){L}^{p-q} \Big({\rho\over a_2 S_\alpha}\Big)^{q/p}+ c_4(L) \mu(\Omega). \end{equation} Denote by \begin{align*} \rho_1=\Big({\delta c_2(K)L^{q-p}(a_2S_\alpha)^{q/p} \over 2c_3(L)}\Big)^{p\over q-p},\quad \rho_2={\Phi(u_\sigma)\over2}. \end{align*} Let $\rho=\min\{\rho_1,\rho_2\}$. When $\rho=\rho_1$, in view of (\ref{eq3.6}), (\ref{eq3.7}) and assumption (\ref{C4}), we compute \begin{align*} \rho {\Psi(u_\sigma)\over\Phi(u_\sigma)}-\Pi (F;L) &\geq \delta c_2(K) \rho_1-\Pi (F;L) \\ &\geq {\delta\over2} c_2(K) \rho_1 -c_4(L)\mu(\Omega)\\ &= \delta^* (c_2(K))^{q\over q-p}L^p (c_3(L))^{p\over p-q}-c_4(L)\mu(\Omega) \\ &\geq \delta^* C^{q\over q-p}c_4(L)-c_4(L)\mu(\Omega) >0, \end{align*} where $\delta^*$ and $C$ are constants dependent only on $p,\beta,\alpha,N,E, \Omega, a_1$ and $a_2$. In the other case of $\rho=\rho_2$, owing to (\ref{eq3.11}), (\ref{eq3.6}), (\ref{eq3.7}) and assumption (\ref{C4}), we deduce \begin{align*} \rho{\Psi(u_\sigma)\over\Phi(u_\sigma)}-\Pi(F;L) &\geq {\delta\over2}c_2(K)\rho_2 -c_4(L)\mu(\Omega)\\ &\geq \delta^{**} c_2(K)|K|^p-c_4(L)\mu(\Omega)\\ &\geq \delta^{**} Cc_4(L)-c_4(L)\mu(\Omega) >0, \end{align*} where $\delta^{**},C$ are constants dependent only on $p,\beta,\alpha,N,E, \Omega, a_1$ and $a_2$. So, we achieve assumption (\ref{eq2.1}) in any cases and the conclusion in the theorem is derived from Lemma \ref{lem2.2}. \end{proof} In the following, we give two simple examples: \begin{example} \label{exa3.1} \rm Consider the mean curvature equation \begin{equation} \label{eq3.8} \begin{gathered} -\mathop{\rm div}(|x|^{-\beta}(1+|\nabla u|^2)^{p-2\over2}\nabla u) =\lambda |u|^{m+{p-m\over |u|+1}}|x|^{-\alpha}, \quad x\in\Omega,\\ u=0,\quad x\in\partial\Omega\,. \end{gathered} \end{equation} Employing Theorem \ref{thm1.1}, we can get the following result: If $2\leq p