\pdfoutput=1\relax\pdfpagewidth=8.26in\pdfpageheight=11.69in\pdfcompresslevel=9 \documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2003(2003), No. 49, pp. 1--16.\newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu (login: ftp)} \thanks{\copyright 2003 Southwest Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE--2003/49\hfil Existence of positive solutions] {Existence of positive solutions for Dirichlet problems of some singular elliptic equations} \author[Zhiren Jin\hfil EJDE--2003/49\hfilneg] {Zhiren Jin} \address{Zhiren Jin\newline Department of Mathematics and Statistics\\ Wichita State University\\ Wichita, Kansas, 67260-0033, USA} \email{zhiren@math.twsu.edu} \date{} \thanks{Submitted December 11, 2002. Published April 29, 2003.} \subjclass[2000]{35J25, 35J60, 35J65} \keywords{Elliptic boundary value problems, positive solutions, \hfill\break\indent singular semilinear equations, unbounded domains, Perron's method, super solutions.} \begin{abstract} When an unbounded domain is inside a slab, existence of a positive solution is proved for the Dirichlet problem of a class of semilinear elliptic equations similar to the singular Emden-Fowler equation. The proof is based on a super and sub-solution method. A super solution is constructed by Perron's method together with a family of auxiliary functions. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \section{Introduction and Main Results} Let $\Omega $ be an unbounded domain in $\mathbb{R}^n$ ($n\geq 3$) with $C^{2,\alpha }$ ($0<\alpha <1$) boundary. We assume that $\Omega $ is inside a slab of width $2M$: $$ \Omega \subset S_M=\{ (\mathbf{x} , y )\in \mathbb{R}^n : |y| 0$ is a constant. The main result of the paper is as follows. \begin{theorem} \label{thm1}Assume \begin{enumerate} \item $p(\mathbf{x}_0,y_0)>0$ for some $(\mathbf{x}_0,y_0)\in \Omega$; \item there is a positive constant $C$ such that \begin{equation} 0\leq p(\mathbf{x},y) \leq C(|\mathbf{x}|+1)^{\gamma} \quad \rm{for} \quad (\mathbf{x},y)\in \Omega ; \label{eq:boundofp} \end{equation} \item $\mathop{\rm Trace}(a_{ij})=1$ and there is a constant $c_1>0$, such that \begin{equation} a_{nn}(\mathbf{x},y)\geq c_1 \quad \mbox{on } \overline{\Omega }. \label{eq:boundofann} \end{equation} \end{enumerate} Then (\ref{eq:theproblem}) has a positive solution $u\in C^{2}(\Omega )\cap C^{0}(\overline{\Omega } )$. \end{theorem} When the principal part in (\ref{eq:theproblem}) is the Laplace operator, (\ref{eq:theproblem}) becomes a boundary value problem for the singular Emden-Fowler equation \begin{equation} -\Delta u = p(\mathbf{x},y)u^{-\gamma } \quad\mbox{on } \Omega; \quad u=0 \quad\mbox{on } \partial \Omega . \label{eq:theproblem1} \end{equation} The singular Emden-Fowler is related to the theory of heat conduction in electrical conduction materials and in the studies of boundary layer phenomena for viscous fluids \cite{Callegari1,Wong}. The existence of positive solutions of the equation on exterior domains (including $\mathbb{R}^n$) has been considered by quite a number of authors (for example, see \cite{Dalmasso,Edelson, Zjin,Kusano,Lair,Shaker}, and references therein). The main approach used to prove existence is to construct super and sub- solutions. To construct super solutions, one needs to assume that $p(\mathbf{x},y)$ decays near infinity in an appropriate rate. A super solution is usually found in the class of radial symmetric functions. If $\Omega $ is an exterior domain (not inside a slab), $\gamma >0$ and there is $C$ such that $p(\mathbf{x},y)\geq \frac{C}{(1+|\mathbf{x}|^{2}+y^{2})}$ for $|\mathbf{x}|^{2}+y^{2}$ large, then (\ref{eq:theproblem1}) has no positive solutions (\cite{Kusano}). On the other hand, if there are constants $\sigma >1$ and $C$, such that $0\leq p(\mathbf{x},y)\leq \frac{C}{(1+|\mathbf{x}|^{2}+y^{2})^{\sigma }}$ for $|\mathbf{x}|^{2}+y^{2}$ large, (\ref{eq:theproblem1}) has a positive solution (\cite{Zjin}). When $\Omega$ is an unbounded domain inside a slab, the situation is quite different. The traditional way to construct a super solution by finding an appropriate radial symmetric function is no longer valid since the domain now is inside a slab (the generality of the coefficient matrix $(a_{ij})$ also makes finding a radial symmetric super solution impossible). In this paper, we combine an idea from \cite{Lopez} and a family of auxiliary functions constructed in \cite{JL4} to construct a super solution which is then used to prove the existence of a positive solution of (\ref{eq:theproblem}). Actually the procedure in the paper can be applied to prove the existence of a positive solution for the Dirichlet problem of more general elliptic equations. A statement for the general case will be given in the last section of the paper. Here we just state a special case of the general result. \begin{theorem} \label{thm2} Assume \begin{enumerate} \item $p(\mathbf{x}_0,y_0)>0$ for some $(\mathbf{x}_0,y_0)\in \Omega$; \item there is a positive constant $C$ such that \begin{equation} 0\leq p(\mathbf{x},y) \leq Ce^{|\mathbf{x}|} \quad \rm{for} \quad (\mathbf{x},y)\in \Omega , \label{eq:boundofp1} \end{equation} \item $\mathop{\rm Trace}(a_{ij})=1$, and there is a constant $c_1>0$, such that \begin{equation} a_{nn}(\mathbf{x},y)\geq c_1 \quad {\rm{on}} \quad \overline{\Omega }. \label{eq:boundofann2} \end{equation} \end{enumerate} Then the problem \begin{equation} -\sum_{i,j=1}^n a_{ij} (\mathbf{x},y)D_{ij} u = p(\mathbf{x},y)e^{-u} \quad\mbox{on } \Omega; \quad u=0 \quad\mbox{on } \partial \Omega \label{eq:theproblem3} \end{equation} has a positive solution $u\in C^{2}(\Omega )\cap C^{0}(\overline{\Omega } )$. \end{theorem} This paper is organized as follows. In Section 2, we construct a family of auxiliary functions that are defined on a family of subdomains of $\Omega$. In Section 3, we combine the family of auxiliary functions constructed in Section 2 and an idea from \cite{Lopez} to prove that (\ref{eq:theproblem}) has a positive supper solution. In Section 4, we prove that (\ref{eq:theproblem}) has a positive solution by the procedure used in \cite{Zjin}. In Section 5, we discuss the general case. \section{A Family of Auxiliary Functions} In this section, we will construct families of sub-domains $\Omega_{\mathbf{x}_0}$ of $\Omega $ and functions $T_{\mathbf{x}_0} +z$ (see definitions below) so that \begin{equation} -\sum_{i,j=1}^n a_{ij}(\mathbf{x}, y)D_{ij}(T_{\mathbf{x}_0} +z) \geq p(\mathbf{x},y) (T_{\mathbf{x}_0} +z)^{-\gamma} \quad \mbox{on } \Omega_{\mathbf{x}_0} \label{eq:auxiliary} \end{equation} and the graphs of the functions $T_{\mathbf{x}_0} +z$ have special relative positions (see below). Our construction is based on the construction of a family of auxiliary functions used in \cite{JL4} (the construction in \cite{JL4} was adapted from \cite{JL1} which in turn was inspired from \cite{Finn} and \cite{Serrin}). We consider the operator $$ Qu =\sum_{i,j=1}^n a_{ij}(\mathbf{x}, y)D_{ij}u . $$ We first extend $a_{ij}$ ($1\leq i,j \leq n$) to be continuous functions on $\overline{S_M}$ in such a way that we still have $\mathop{\rm Trace}(a_{ij})=1$ and \begin{equation} a_{nn}({\bf x},y) \ge c_1 \quad \mbox{on } S_M. \label{eq:coeffbound} \end{equation} In the rest of the paper, we will use $c_{m}$ (for some integer $m\geq 2$) to denote a constant depending only on $c_1$ and $M$. Once a constant $c_{m}$ is used in a formula, it will represent the same constant if the same notation appears again in the paper. It was proved in \cite{JL4} (also see Appendix I) that there are positive decreasing functions $\chi (t)$, $h_a(t)$ and a positive increasing function $A(t)$ ($\chi (t)$ depending on $c_1$ only, $h_a(t)$ and $A(t)$ depending on $c_1$ and $M$ only), such that for any number $K$, there is a number $H_0$, depending only on $K$, $M$ and $c_1$, such that for $H\geq H_0$, we have (for $0T_{\mathbf{x}_1} +10-\frac{\delta_0\sqrt{195A(H)e^{-\chi (H)}}}{A(H)e^{\chi(H)}}. \end{align*} Thus there is a $\delta_0$ small such that for all $|\mathbf{x}-\mathbf{x}_0|\leq \delta_0$ with $(\mathbf{x},y)\in \Omega_{\mathbf{x}_1}$, if $\mathbf{x}_1$ and $\mathbf{x}_0$ satisfy (\ref{eq:overlapping}), we have \begin{equation} T_{\mathbf{x}_1} + z_{\mathbf{x}_1}(\mathbf{x}, y) \geq T_{\mathbf{x}_1} +8. \label{eq:test1} \end{equation} Further for all $\mathbf{x}_0$ and $\mathbf{x}_1$ satisfying (\ref{eq:overlapping}), \begin{align*} T_{\mathbf{x}_0} +2 &\leq T_{\mathbf{x}_1} + T_{\mathbf{x}_0} - T_{\mathbf{x}_1} +2 \\ &\leq T_{\mathbf{x}_1} + (\frac{C}{c_2})^{\frac{1}{\gamma}} (|\mathbf{x}_0| - |\mathbf{x}_1|)+2 \\ &\leq T_{\mathbf{x}_1} + (\frac{C}{c_2})^{\frac{1}{\gamma}} |\mathbf{x}_1-\mathbf{x}_0|+2 \\ &\leq T_{\mathbf{x}_1} + (\frac{C}{c_2})^{\frac{1}{\gamma}} \sqrt{205A(H)e^{\chi (H)}} +2\\ &\leq T_{\mathbf{x}_1} + (\frac{C}{c_2})^{\frac{1}{\gamma}}c_{5}+2 \end{align*} where $c_{5} = \sqrt{205A(H)e^{\chi (H)}}$. Thus if we assume that $C$ in (\ref{eq:boundofp}) satisfies \begin{equation} C\leq 6^{\gamma}c_{5}^{-\gamma }c_2 , \label{eq:smallofp} \end{equation} we have that for all $\mathbf{x}_0$ and $\mathbf{x}_1$ satisfying (\ref{eq:overlapping}), \begin{equation} T_{\mathbf{x}_0} +2 \leq T_{\mathbf{x}_1} +8\,. \label{eq:test2} \end{equation} From (\ref{eq:domain1}) and (\ref{eq:supersolution4}), we can choose a number $\delta_2({\mathbf{x}_0})>0$ such that for all $\mathbf{x}\in R^{n-1}$ with $|\mathbf{x}_0 - \mathbf{x}|\leq \delta_2({\mathbf{x}_0})$, we have $(\mathbf{x},y)\in \Omega_{\mathbf{x}_0}$ for all $|y|< M$, and \begin{equation} T_{\mathbf{x}_0} + z_{\mathbf{x}_0}(\mathbf{x}, y) \leq T_{\mathbf{x}_0} +2\,. \label{eq:test3} \end{equation} Now if we set $\delta_{\mathbf{x}_0} =\min \{ \delta_0, \delta_2({\mathbf{x}_0})\}$, from (\ref{eq:test1}), (\ref{eq:test2}) and (\ref{eq:test3}), we have \begin{equation} T_{\mathbf{x}_0}+z_{\mathbf{x}_0}(\mathbf{x},y) \leq T_{\mathbf{x}_1} +z_{\mathbf{x}_1}(\mathbf{x},y) \label{eq:definitionofradius} \end{equation} for all $\mathbf{x}_0$ and $\mathbf{x}_1$ satisfying (\ref{eq:overlapping}), $|\mathbf{x}_0 - \mathbf{x}|\leq \delta_{\mathbf{x}_0}$ and $(\mathbf{x},y)\in \Omega_{\mathbf{x}_1}$. Finally we define a family of open subsets of $\Omega $ that will be needed in next section. For each point $(\mathbf{x}_0,y_0)\in \overline{\Omega} $, we define an open set $O(\mathbf{x}_0,y_0)$ as follows: \begin{enumerate} \item If $(\mathbf{x}_0,y_0)\in \Omega$, we choose a ball $B$ with center $(\mathbf{x}_0,y_0)$ and a radius less than $\delta_{\mathbf{x}_0}$ so that $B\subset \Omega$. We then set $O(\mathbf{x}_0,y_0)=B$; \item If $(\mathbf{x}_0,y_0)\in \partial \Omega$, since $\Omega $ has $C^{2,\alpha }$ boundary, there is a ball $B$ with center $(\mathbf{x}_0,y_0)$ and a radius less than $\delta_{\mathbf{x}_0}$, such that there is a $C^{2,\alpha }$ diffeomorphism $\Phi $ satisfying $$ \Phi (B\cap \Omega ) \subset \mathbb{R}^n_{+}, \quad \Phi (B\cap \partial \Omega ) \subset \partial \mathbb{R}^n_{+}; \quad \Phi (\mathbf{x}_0,y_0)={\bf{0}}. $$ \end{enumerate} Now we choose a domain $J$ with $C^{3}$ boundary with following properties: (a) $J\subset \Phi (B\cap \Omega )$; (b) $\partial J\cap \partial \mathbb{R}^n_{+}$ is a neighborhood of ${\bf{0}}$ in $\partial \mathbb{R}^n_{+}$. Certainly there are many different $J$'s having those properties. One example is given in the Appendix II at the end of paper to illustrate how to construct such a domain $J$. Now we set $O(\mathbf{x}_0,y_0)=\Phi^{-1}(J)$. It is easy to see that $O(\mathbf{x}_0,y_0)\subset B\cap \Omega $, $O(\mathbf{x}_0,y_0)$ has a $C^{2,\alpha }$ boundary and $\partial O(\mathbf{x}_0,y_0)\cap \partial \Omega$ is a neighborhood of $(\mathbf{x}_0,y_0)$ in $\partial \Omega$. Let $\Pi$ be the collection of all such open sets $O(\mathbf{x}_0,y_0)$ defined in (1) and (2). \section{A Super Solution of (\ref{eq:theproblem})} In this section, using the family of auxiliary functions $T_{\mathbf{x}_0}+z$ constructed in Section 2 and an idea from \cite{Lopez} (that basically says that the Perron's method still works if we can find a family of appropriate auxiliary functions that works like a super solution), we will show that there is a positive function $u\in C^{2}(\Omega )\cap C^{0}(\overline{\Omega})$, satisfies $$ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}u = p(\mathbf{x},y)u^{-\gamma } \quad \mbox{on } \Omega , \quad u=\tau \quad \mbox{on } \partial \Omega . $$ for some constant $\tau >0$. Then $u$ will be a super solution of (\ref{eq:theproblem}). If $u=c_0v$ for some constant $c_0$, $v$ will satisfy $$ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}v = c_0^{-\gamma -1} p(\mathbf{x},y)v^{-\gamma } \quad\mbox{on } \Omega , \quad v=\tau /c_0 \quad \mbox{on } \partial \Omega \,. $$ Thus without loss of generality, we may assume $C$ in (\ref{eq:boundofp}) satisfying (\ref{eq:smallofp}). Then all constructions in Section 2 are valid. Let $v>0$ be a function on $\overline{\Omega}$, for a point $(\mathbf{x}_0,y_0)\in \overline{\Omega } $, we define a new function $M_{(\mathbf{x}_0,y_0)}(v)$, called the lift of $v$ over $O(\mathbf{x}_0,y_0)$ as follows: \begin{gather*} M_{(\mathbf{x}_0,y_0)}(v)(\mathbf{x},y)=v(\mathbf{x},y) \quad {\rm{if}} \quad (\mathbf{x},y)\in \Omega\setminus O(\mathbf{x}_0,y_0) \\ M_{(\mathbf{x}_0,y_0)}(v)(\mathbf{x},y)=w(\mathbf{x},y) \quad {\rm{if}} \quad (\mathbf{x},y)\in O(\mathbf{x}_0,y_0) \end{gather*} where $w(\mathbf{x},y)$ is the positive solution of the boundary-value problem \begin{equation} -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}w = p(\mathbf{x},y) w^{-\gamma } \quad \mbox{in } O(\mathbf{x}_0,y_0), \quad w=v \quad \mbox{on } \partial O(\mathbf{x}_0,y_0) \,. \label{eq:lift} \end{equation} It is easy to see (\ref{eq:lift}) has a unique positive solution in $C^{2}(O(\mathbf{x}_0,y_0))\cap C^{0}(\overline{O(\mathbf{x}_0,y_0)})$. Indeed $m_1=\min \{ v (\mathbf{x},y): (\mathbf{x},y)\in \partial O(\mathbf{x}_0,y_0)\}$ is a sub-solution since $p(\mathbf{x},y)$ is non-negative, $m_2+T_{\mathbf{x}_0} + z_{\mathbf{x}_0}$ is a super solution by (\ref{eq:auxiliary}), where $m_2=\max \{ v (\mathbf{x},y): (\mathbf{x},y)\in \partial O(\mathbf{x}_0,y_0)\}$. Then we can conclude the existence of a desired solution (for example, see \cite{Amann} or \cite{Crandall}). Uniqueness of positive solutions of (\ref{eq:lift}) follows from a standard argument. Set $\tau = (C/c_2)^{1/\gamma} c_{4}$ (see (\ref{eq:definitionoft}) for the source of the constants). We define a class $\Xi $ of functions as follows: a function $v$ is in $\Xi $ if \begin{enumerate} \item $v\in C^{0}(\overline{\Omega })$, $v>0$ on $\overline{\Omega}$ and $v\leq \tau $ on $\partial \Omega$; \item For any $(\mathbf{x}_0,y_0)\in \overline{\Omega}$, $v\leq M_{(\mathbf{x}_0,y_0)}(v)$; \item $v\leq T_{\mathbf{x}_0} +z_{\mathbf{x}_0}$ on $\Omega_{\mathbf{x}_0} \cap \Omega$ for any $(\mathbf{x}_0,y_0)\in \overline{\Omega}$. \end{enumerate} By the following well-known lemma, it is easy to check the function $v=\tau $ is in $\Xi$. Thus $\Xi$ is not empty. \begin{lemma} \label{lm1} Let $D$ be a bounded domain, $f(\mathbf{x},y,t)$ be a $C^{1}$ function that is decreasing in $t$. If $w_1$, $w_2$ are in $C^{2}(D)\cap C^{0}(\overline{D})$, $w_1\leq w_2$ on $\partial D$, and \begin{gather*} -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}w_1 \leq f(\mathbf{x},y,w_1) \quad \mbox{in } D, \\ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}w_2 \geq f(\mathbf{x},y, w_2) \quad\mbox{in }D \end{gather*} then $w_1\leq w_2$ on $D$. \end{lemma} Now we set $$ u(\mathbf{x},y)=\sup_{v\in {\Xi }} v(\mathbf{x},y), \quad (\mathbf{x},y)\in \overline{\Omega} \,. $$ We will show that $u$ is in $C^{2}(\Omega )\cap C^{0}(\overline{\Omega } )$ and satisfies $$ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}u = p(\mathbf{x},y) u^{-\gamma } \quad \mbox{on } \Omega ;\quad u=\tau \quad \mbox{on } \partial \Omega . $$ First we need some lemmas. \begin{lemma} \label{lm2} If $00$ on $\overline{\Omega}$ and $\max \{ v_1, v_2 \}\leq \tau $ on $\partial \Omega$. It is also clear that $\max \{ v_1, v_2 \}\leq T_{\mathbf{x}_0} +z_{\mathbf{x}_0}$ on $\Omega_{\mathbf{x}_0} \cap \Omega$ for any $(\mathbf{x}_0,y_0)\in \overline{\Omega}$. Since $$ v_1\leq \max \{ v_1, v_2 \}, \quad v_2 \leq \max \{ v_1, v_2\} $$ we have (by lemma 2) that for any $(\mathbf{x}_0,y_0)\in \overline{\Omega} $, $$ M_{(\mathbf{x}_0,y_0)}(v_1) \leq M_{(\mathbf{x}_0,y_0)}(\max \{ v_1, v_2\} ),\quad M_{(\mathbf{x}_0,y_0)}(v_2) \leq M_{(\mathbf{x}_0,y_0)}(\max \{ v_1, v_2\} ). $$ Since $v_1\in \Xi $ and $v_2\in \Xi $ imply $$ v_1 \leq M_{(\mathbf{x}_0,y_0)}(v_1),\quad v_2 \leq M_{(\mathbf{x}_0,y_0)}(v_2), $$ we have $$ \max \{ v_1, v_2 \}\leq M_{(\mathbf{x}_0,y_0)}(\max \{ v_1, v_2\}) . $$ Thus $\max \{ v_1, v_2 \}\in \Xi$. \end{proof} \begin{lemma} \label{lm4} If $v\in \Xi $, then $M_{(\mathbf{x}_0,y_0)}(v)\in \Xi$ for any $(\mathbf{x}_0,y_0)\in \overline{\Omega} $. \end{lemma} \begin{proof} By the definition of $M_{(\mathbf{x}_0,y_0)}( v)$, it is clear that $M_{(\mathbf{x}_0,y_0)}( v)>0$ on $\overline{\Omega }$, $M_{(\mathbf{x}_0,y_0)}( v)\in C^{0}(\overline{\Omega} )$ and $M_{(\mathbf{x}_0,y_0)}(v) \leq \tau $ on $\partial \Omega $. For any $(\mathbf{x}^{*},y^{*})\in \overline{\Omega} $, we first show that \begin{equation} M_{(\mathbf{x}_0,y_0)}( v)(\mathbf{x},y)\leq M_{(\mathbf{x}^{*},y^{*})}(M_{(\mathbf{x}_0,y_0)}( v))(\mathbf{x},y). \label{eq:mv} \end{equation} We only need to prove that (\ref{eq:mv}) is true for $(\mathbf{x},y)\in O(\mathbf{x}^{*},y^{*})$. Since $$ v\leq M_{(\mathbf{x}_0,y_0)}( v), $$ we have (by lemma 2) $$ M_{(\mathbf{x}^{*},y^{*})}( v)\leq M_{(\mathbf{x}^{*},y^{*})}(M_{(\mathbf{x}_0,y_0)}( v)). $$ Then from $v\leq M_{(\mathbf{x}^{*},y^{*})}( v)$ (by lemma 2 again), we have $$ v\leq M_{(\mathbf{x}^{*},y^{*})}(M_{(\mathbf{x}_0,y_0)}( v)). $$ Thus for $(\mathbf{x},y)\in O(\mathbf{x}^{*},y^{*}) \setminus O(\mathbf{x}_0,y_0)$, \begin{equation} M_{(\mathbf{x}_0,y_0)}( v)(\mathbf{x},y)=v(\mathbf{x},y) \leq M_{(\mathbf{x}^{*},y^{*})}(M_{(\mathbf{x}_0,y_0)}( v))(\mathbf{x},y). \label{eq:boundary1} \end{equation} That is, (\ref{eq:mv}) is true on $O(\mathbf{x}^{*},y^{*}) \setminus O(\mathbf{x}_0,y_0)$, Now for $\Omega_1=O(\mathbf{x}^{*},y^{*}) \cap O(\mathbf{x}_0,y_0)$, if we set $$ M_{(\mathbf{x}_0,y_0)}( v)=w_1, \quad M_{(\mathbf{x}^{*},y^{*})}(M_{(\mathbf{x}_0,y_0)}( v))=w_2 $$ we have \begin{gather*} -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij} w_1 = p(\mathbf{x},y) w_1^{-\gamma } \quad\mbox{on } \Omega_1, \\ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij} w_2 = p(\mathbf{x},y) w_2^{-\gamma } \quad \mbox{on }\Omega_1 . \end{gather*} On $\partial \Omega_1$, $w_1\leq w_2$ on $O(\mathbf{x}^{*},y^{*})\cap \partial O(\mathbf{x}_0,y_0)$ by (\ref{eq:boundary1}) and $w_1\leq w_2$ on $\partial O(\mathbf{x}^{*},y^{*})\cap O(\mathbf{x}_0,y_0)$ since (\ref{eq:mv}) is true on $\Omega\setminus O(\mathbf{x}^{*},y^{*})$. Then lemma 1 implies $w_1\leq w_2$ on $\Omega_1$. Thus (\ref{eq:mv}) is true on $O(\mathbf{x}^{*},y^{*}) \cap O(\mathbf{x}_0,y_0)$ and on $O(\mathbf{x}^{*},y^{*})$. \end{proof} Now we prove that $M_{(\mathbf{x}_0,y_0)}(v)\leq T_{\mathbf{x}_1}+z_{\mathbf{x}_1}$ on $\Omega_{\mathbf{x}_1}\cap \Omega$ for all $(\mathbf{x}_1,y_1)\in \overline{\Omega }$. By the definition of $M_{(\mathbf{x}_0,y_0)}(v)$, we only need to consider the graph of the function $M_{(\mathbf{x}_0,y_0)}(v)$ over $O(\mathbf{x}_0,y_0)$. If $O(\mathbf{x}_0,y_0)$ is covered completely by $\Omega_{\mathbf{x}_1}$, since $v\leq T_{\mathbf{x}_1}+z_{\mathbf{x}_1}$ and $T_{\mathbf{x}_1}+z_{\mathbf{x}_1}$ satisfies (\ref{eq:auxiliary}), $T_{\mathbf{x}_1}+z_{\mathbf{x}_1}$ is a super solution of (\ref{eq:lift}) on $O(\mathbf{x}_0,y_0)$. Then Lemma \ref{lm1} implies $M_{(\mathbf{x}_0,y_0)}(v)\leq T_{\mathbf{x}_1}+z_{\mathbf{x}_1}$ on $O(\mathbf{x}_0,y_0)$. In the case that $O(\mathbf{x}_0,y_0)$ does not intersect with $\Omega_{\mathbf{x}_1}$, the conclusion is trivial. Now we consider the case that $O(\mathbf{x}_0,y_0)$ is partially covered by $\Omega_{\mathbf{x}_1}$. Since $O(\mathbf{x}_0,y_0)$ is covered by $\Omega_{\mathbf{x}_0}$, we always have \begin{equation} M_{(\mathbf{x}_0,y_0)}(v)\leq T_{\mathbf{x}_0}+z_{\mathbf{x}_0} \quad \mbox{on } O(\mathbf{x}_0,y_0). \label{eq:control} \end{equation} Then by the choice of $\delta_{\mathbf{x}_0}$, $O(\mathbf{x}_0,y_0)$, and the fact that $O(\mathbf{x}_0,y_0) \cap T_{\mathbf{x}_1}$ is not empty, we have that $\mathbf{x}_0$ and $\mathbf{x}_1$ satisfy (\ref{eq:overlapping}), and for all $(\mathbf{x},y)\in O(\mathbf{x}_0,y_0)\cap \Omega_{\mathbf{x}_1}$, $|\mathbf{x}_0 - \mathbf{x}|\leq \delta_{\mathbf{x}_0}$. Then by (\ref{eq:definitionofradius}), the graph of $T_{\mathbf{x}_0}+z_{\mathbf{x}_0}$ over $O(\mathbf{x}_0,y_0)\cap \Omega_{\mathbf{x}_1}$ is under the graph of $T_{\mathbf{x}_1}+z_{\mathbf{x}_1}$. Thus the conclusion follows from (\ref{eq:control}). Now we are ready to prove that $u$ has the desired properties. Let $(\mathbf{x}_0,y_0)\in \overline{\Omega }$. By the definition of $u(\mathbf{x}_0,y_0)$, there is a sequence of functions $v_{k}$ in $\Xi $ such that $$ u(\mathbf{x}_0,y_0)=\lim_{k\to \infty } v_{k}(\mathbf{x}_0,y_0). $$ By lemma 3 and the fact that $v=\tau $ is in $\Xi$, replacing $v_{k}$ by $\max \{ v_{k}, \tau \}$ if it is necessary, we may assume that $v_{k}\geq \tau $ on $\Omega$. We replace $v_{k}$ by $M_{(\mathbf{x}_0,y_0)}(v_{k})$. Then we have a sequence of functions $w_{k}$ satisfying \begin{gather*} u(\mathbf{x}_0,y_0)=\lim_{k\to \infty } w_{k}(\mathbf{x}_0,y_0) ,\\ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij} w_{k} =p(\mathbf{x},y) w_{k}^{-\gamma } \quad on \quad O(\mathbf{x}_0,y_0), \\ w_{k}=v_{k} \quad on \quad \partial O(\mathbf{x}_0,y_0). \end{gather*} Since for all $k$, $$ \tau \leq v_{k}\leq w_{k} \leq T_{\mathbf{x}_0} +z_{\mathbf{x}_0} \quad \mbox{on } O(\mathbf{x}_0,y_0). $$ By \cite[Theorem 9.11]{Trudinger} and an approximation of the boundary value by smooth functions, we see that there is a subsequence of $w_{k}$, for convenience still denoted by $w_{k}$, converges to a $C^{2}(O(\mathbf{x}_0,y_0))\cap C^{0}(\overline{O(\mathbf{x}_0,y_0)})$ function $w(x)$ in $C^{2}(O(\mathbf{x}_0,y_0)) \cap C^{0}(\overline{O(\mathbf{x}_0,y_0)})$. Thus $w(x)$ satisfies $$ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij} w = p(\mathbf{x},y)w^{-\gamma} \quad on \quad O(\mathbf{x}_0,y_0) $$ and $u(\mathbf{x}_0,y_0)=w(\mathbf{x}_0,y_0)$. We claim that $u=w$ on $O(\mathbf{x}_0,y_0)$. Indeed, if there is another point $(\mathbf{x}_2,y_2) \in O(\mathbf{x}_0,y_0)$ such that $u(\mathbf{x}_2,y_2)$ is not equal to $w(\mathbf{x}_2,y_2)$, then $u(\mathbf{x}_2,y_2)>w(\mathbf{x}_2,y_2)$. Then there is a function $u_0\in \Xi $, such that $$ w(\mathbf{x}_2,y_2)< u_0(\mathbf{x}_2,y_2)\leq u(\mathbf{x}_2,y_2). $$ Now the sequence $\max \{ u_0, M_{(\mathbf{x}_0,y_0)}(v_{k}) \}$ satisfying $$ v_{k} \leq \max \{ u_0, M_{(\mathbf{x}_0,y_0)}(v_{k}) \} \leq u . $$ Then similar to the way we obtain $w$, $M_{(\mathbf{x}_0,y_0)}(\max \{ u_0, M_{(\mathbf{x}_0,y_0)}(v_{k}) \})$ will produce a function $w_1$ satisfying \begin{gather*} -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij} w_1 =p(\mathbf{x},y) w_1^{-\gamma } \quad\mbox{on } O(\mathbf{x}_0,y_0), \\ w\leq w_1 \quad on \quad O(\mathbf{x}_0,y_0), \quad w(\mathbf{x}_2,y_2)s$. Finally we take limit of $w_{m}$ to get a desired solution. \section{The General Case} Now we consider the boundary-value problem \begin{equation} - \sum_{i,j=1}^n a_{ij}(\mathbf{x},y)D_{ij}u =g(\mathbf{x},y,u) \quad\mbox{on }\Omega , \quad u=0 \quad\mbox{on }\partial \Omega . \label{eq:general} \end{equation} In addition to the assumptions on $(a_{ij})$ and $\Omega$ given at the beginning of the paper, we assume the following conditions. \begin{enumerate} \item $\mathop{\rm Trace}a_{ij})=1$; \item There is a constant $c_1>0$ such that $a_{nn} \geq c_1$ on $\overline{\Omega}$; \item There is a family of increasing positive functions $T=T(t)$ satisfying (with $T_{\mathbf{x}}=T(|\mathbf{x}|)$) \begin{itemize} \item[(a)] $|T_{\mathbf{x}_0}-T_{\mathbf{x}}|\leq |\mathbf{x}_0-\mathbf{x}|/c_{5}$; \item[(b)] $g(\mathbf{x},y,T_{\mathbf{x}_0}+z_{\mathbf{x}_0} )\leq c_2 $ on $\Omega_{\mathbf{x}_0}$ ($\Omega_{\mathbf{x}_0}$, $z_{\mathbf{x}_0}$ and $c_2$ are defined in Section 2); \end{itemize} \item $g(\mathbf{x},y,t)$ is non-negative, in $C^{1}(\overline{\Omega}\times \mathbb{R}^n_{+})$ and decreasing on $t$. \item ${\lim }_{t\longrightarrow 0^{+}} \frac{g(\mathbf{x},y,t)}{t} \geq v_0(\mathbf{x},y)$ uniformly for $(\mathbf{x},y)$ in any bounded subset on $\overline{\Omega}$, where $v_0(\mathbf{x},y)$ is a non-negative function satisfying that when $m$ is large, the eigenvalue problem $$ -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}w = \lambda v_0(\mathbf{x},y) w \ \ on \ \ {\Omega }_{m}, \quad u =0 \ \ on \ \ \partial {\Omega }_{m} . $$ has a first eigenvalue $\lambda_1<1$. \end{enumerate} Then we have the following conclusion. \begin{theorem} \label{thm3} Under the assumptions (1)-(5), (\ref{eq:general}) has a positive solution. \end{theorem} \begin{proof} We just sketch the proof here. Assumptions (1)--(3) assure that $T_{\mathbf{x}_0}+z_{\mathbf{x}_0} $ is a family of auxiliary functions satisfying (\ref{eq:auxiliary}) on $\Omega_{\mathbf{x}_0}$ and the graphs of these function have the desired relative positions as discussed in Section 2. Assumption (4) assures that lemma 1 can be applied and the boundary value problem \begin{equation} -\sum_{i,j=1}^n a_{ij}(\mathbf{x},y) D_{ij}w = g(\mathbf{x},y, w) \quad \mbox{in } O(\mathbf{x}_0,y_0), \quad w=v \mbox{on } \partial O(\mathbf{x}_0,y_0) \label{eq:lift2} \end{equation} has a unique positive solution for each positive function $v$ on $\overline{\Omega }$. Thus the lift $M_{(\mathbf{x}_0, y_0)}$ and the class $\Xi$ of functions are well defined. The proofs of lemmas 2-4 and the existence of the super solution $u$ are the same. Finally the assumption (5) assures that the proof in Section 4 still works out like that in \cite{Zjin}. \end{proof} Now we apply theorem 3 to the case that $g(\mathbf{x},y,u)= p(\mathbf{x},y)e^{-u}$. We consider a modified problem: \begin{equation} - \sum_{i,j=1}^n a_{ij}(\mathbf{x},y)D_{ij}u =\frac{p(\mathbf{x},y)e^{-c_{5}u}}{c_{5}} \quad\mbox{on }\Omega ,\quad u=0 \quad\mbox{on }\partial \Omega . \label{eq:general2} \end{equation} If we can find a positive solution $u$ of (\ref{eq:general2}), then $c_{5}u$ is a positive solution of (\ref{eq:theproblem3}). For (\ref{eq:general2}), we set $$ T(t) =\frac{1}{c_{5}} (t+c_{4}) + \frac{1}{c_{5}} \ln \frac{C}{c_2c_{5}} +A $$ where $A$ is a positive constant such that $\frac{1}{c_{5}} \ln \frac{C}{c_{5}} +A>1$, $C$ is defined in (\ref{eq:boundofp1}) and $c_2$, $c_{4}$, $c_{5}$ are defined in Section 2. Then $T(t)$ is increasing and the assumption (3)(a) is obviously satisfied for $T_{\mathbf{x}}=T(|\mathbf{x}|)$. For (3)(b), on $\Omega_{\mathbf{x}_0}$, \begin{align*} \frac{1}{c_{5}}p(\mathbf{x},y)e^{-c_{5}(T_{\mathbf{x}_0} +z_{\mathbf{x}_0})} &\leq \frac{C}{c_{5}} e^{|\mathbf{x}|}e^{-c_{5}T_{\mathbf{x}_0}} \\ &\leq \frac{C}{c_{5}} e^{|\mathbf{x}_0|+c_{4}}e^{-c_{5}T_{\mathbf{x}_0}}\\ &=\frac{C}{c_{5}} e^{|\mathbf{x}_0|+c_{4}}e^{-|\mathbf{x}_0| -c_{4} - \ln \frac{C}{c_2c_{5}} -c_{5}A}\\ &= c_2e^{-c_{5}A} 0. $$ It is clear that $\chi (\alpha )$ is a decreasing function with range $(0,\infty).$ Let $\eta$ be the inverse of $\chi.$ Then $\eta$ is a positive, decreasing function with range $(0,\infty)$. Let $c^{*} =11/c_1$. For $\alpha >1$, we have \begin{equation} \chi(\alpha)=\int_{\alpha}^{\infty}\ \frac{d\rho}{\rho^{3}\Phi_1(\rho)} =\int_{\alpha}^{\infty}\ \frac{d\rho}{c^{*}\rho^{3}} = \frac{1}{2c^{*}} \alpha^{-2} . \label{eq:chi2} \end{equation} Thus \begin{equation} \eta (\beta ) = (2c^{*}\beta)^{-\frac{1}{2}}\quad for \quad 0<\beta <(2c^{*})^{-1}. \label{eq:beta1} \end{equation} Let $H\ge 2$. Since $\eta(\chi(H))=H$ and $\eta $ is decreasing, we have $\eta(\beta)> H$ for $0<\beta< \chi(H)$. We define a function $A(H)$ by \begin{equation} A(H) = 2M (\int_1^{e^{\chi(H)}}\ \eta (\ln t) dt)^{-1} . \label{eq:ah} \end{equation} For the rest of this article, we set $a=A(H)$ and define \begin{equation} h_a(r)=\int_{r}^{ae^{\chi(H)}} \eta (\ln \frac{t}{a} )\ dt \quad \mbox{for } a\le r\le ae^{\chi(H)}. \end{equation} Then \begin{equation} h_a(ae^{\chi (H)})=0, \quad h_a(a) =h_{A(H)}(A(H))= 2M. \label{eq:chi3} \end{equation} For $aH, \quad h_a''(r)=\frac{1}{r}(\eta(\ln \frac{r}{a} ))^{3}\Phi_1 (\eta(\ln \frac{r}{a} )). \label{eq:derivativelarge} \end{equation} Thus for $a2$ such that for $H\geq H_0$, \begin{equation} H_0>\frac{1}{\sqrt{2c^{*}}} +3M+4+\frac{24nc_1K}{M}, \quad \sqrt{\frac{4K}{A(H)e^{\chi (H)}}}\leq \frac{1}{\sqrt{2}}. \label{eq:Hlarge} \end{equation} Then we have (\ref{eq:boundofae1}). For $H>H_0$, by (\ref{eq:chi2}), (\ref{eq:beta1}), we have \begin{align*} A(H)^{-1} &= (2M)^{-1} \int_1^{e^{\chi(H)}}\ \eta (\ln t) dt \\ &=(2M)^{-1} \int_0^{\chi(H)}\ \eta (m)e^{m} dm \\ &=(2M)^{-1} \int_0^{\chi(H)}\ \frac{e^{m}}{\sqrt{2c^{*}m}} dm\,. \end{align*} From $$ \frac{1}{\sqrt{2c^{*}}}\int_0^{\chi(H)}\ \frac{1}{\sqrt{m}} dm \leq \int_0^{\chi(H)}\ \frac{e^{m}}{\sqrt{2c^{*}m}} dm \leq \frac{e^{\chi (H)}}{\sqrt{2c^{*}}} \int_0^{\chi(H)}\ \frac{1}{\sqrt{m}} dm\,, $$ we have $$ \frac{1}{c^{*}H} = \frac{2\sqrt{\chi (H)}}{\sqrt{2c^{*}}} \leq \int_0^{\chi(H)}\ \frac{e^{m}}{\sqrt{2c^{*}m}} dm \leq \frac{2e^{\chi (H)}\sqrt{\chi (H)}}{\sqrt{2c^{*}}} =\frac{e^{\frac{1}{2c^{*}H^{2}}}}{c^{*} H} . $$ Thus \begin{equation} 2Mc^{*}H \geq A(H)\geq 2Mc^{*}He^{-\chi (H)} = 2Mc^{*}H e^{-\frac{1}{2c^{*}H^{2}}} . \label{eq:ah2} \end{equation} Thus we have the second half of (\ref{eq:boundofae}) since $c^{*}=11/c_1$. For $\mathbf{x}_0\in \mathbb{R}^{n-1}$, and a fixed constant $K$, we define a domain $\Omega_{\mathbf{x}_0,H,K}$ in $(\mathbf{x},y)$ space by (\ref{eq:domain1}) and define a function $z=z(\mathbf{x}, y)$ by (\ref{eq:solution111}). Since $h_a^{-1}(y+M)\geq 0$ for $|y|\leq M$, $(\mathbf{x}_0, y)\in \Omega_{\mathbf{x}_0,H,K}$ for $|y|< M$. Further it is clear that the function $z=z(\mathbf{x},y)$ is well defined on $\Omega_{\mathbf{x}_0,H,K}$. Now we verify the first half of (\ref{eq:supersolution2}), on $\partial \Omega_{\mathbf{x}_0,H,K} \cap \{ (\mathbf{x},y): |y|0$)}\\ &\leq \frac{1}{S}\big\{ 1+\sum_{i,j=1}^n a_{ij} \frac{\partial z}{\partial x_{i}} \frac{\partial z}{\partial x_{j}} -a_{nn} h_a^{-1} (h_a^{-1})'' \big\}\,. \end{align*} By (\ref{eq:coeffbound}), (\ref{eq:equationofinverse})), (\ref{eq:ah2}) and (\ref{eq:normofgradient})) the above expression is bounded by $$ \frac{-9}{S} \leq \frac{-9} {h_a^{-1}(y+M)} \leq \frac{-9} {A(H)e^{\chi (H)}} \leq \frac{-9}{2Mc^{*}He^{\frac{1}{2c^{*}H^{2}}}} \leq \frac{-3c_1}{22eMH}. $$ This shows (\ref{eq:supersolution1}). \section{Appendix II: A Construction of the Domain $J$} In this part, we give a construction of the domain $J$ used at the end of Section 2 in the definition of $\Pi$. Let \begin{gather*} \mathbb{R}^n_{+}=\{ (y_1, y_2, \dots , y_n)|y_n>0\}\,,\\ J_1=\big\{ (y_1, y_n) : y_1=\pm 1, \; |y_n|\leq 1 \; {\rm{or}} \; y_n=\pm 1, \; |y_1|\leq 1 \big\} \end{gather*} That is, $J_1$ is a square with side length 2 and center $(0,0)$ in $(y_1,y_n)$ plane. In polar coordinate we can write $\partial J_1$ as $$ (y_1, y_n) =(k(\theta )\cos \theta , k(\theta ) \sin \theta ), \quad 0\leq \theta \leq 2\pi, $$ where $k(\theta )$ is a positive, continuous, periodic function of period $2\pi$, $k(\theta )$ is $C^{\infty }$ except at $\theta =\pm \frac{\pi}{4}$, $\pm \frac{3\pi}{4}$. Then we can smooth out $k(\theta )$ near those points to get a function $k_1(\theta )$ such that $k_1(\theta )$ is a positive, $C^{\infty }$, periodic function of period $2\pi$, $k_1(\theta )=k(\theta )$ except in some small neighborhoods of $\theta =\pm \frac{\pi}{4}$, $\pm \frac{3\pi}{4}$, and $k_1(\theta )\leq k(\theta )$ for all $\theta $. Indeed we can modify $k(\theta )$ as follows: Let $s(t)$ be a $C^{\infty }$ function satisfying \begin{enumerate} \item $s(t)=0$ if $t\leq 1$; \item $0