\documentclass[twoside]{article} \usepackage{amssymb} % font used for R in Real numbers \pagestyle{myheadings} \markboth{\hfil $C^{1,\alpha}$ convergence of minimizers \hfil EJDE--2000/14} {EJDE--2000/14\hfil Yutian Lei \& Zhuoqun Wu \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol.~{\bf 2000}(2000), No.~14, pp.~1--20. \newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu \quad ftp ejde.math.unt.edu (login: ftp)} \vspace{\bigskipamount} \\ % $C^{1,\alpha}$ convergence of minimizers of a Ginzburg-Landau functional \thanks{ {\em 1991 Mathematics Subject Classifications:} 35J70. \hfil\break\indent {\em Key words and phrases:} Ginzburg-Landau functional, regularizable minimizer. \hfil\break\indent \copyright 2000 Southwest Texas State University and University of North Texas. \hfil\break\indent Submitted September 2, 1999. Published February 21, 2000.} } \date{} % \author{Yutian Lei \& Zhuoqun Wu} \maketitle \begin{abstract} In this article we study the minimizers of the functional $$ E_\varepsilon(u,G)=\frac{1}{p}\int_G|\nabla u|^p+\frac{1}{4\varepsilon^p} \int_G(1-|u|^2)^2, $$ on the class $W_g=\{v \in W^{1,p}(G,{\mathbb R}^2);v|_{\partial G}=g\}$, where $g:\partial G \to S^1$ is a smooth map with Brouwer degree zero, and $p$ is greater than 2. In particular, we show that the minimizer converges to the $p$-harmonic map in $C_{\mbox{\scriptsize\rm loc}}^{1,\alpha}(G,{\mathbb R}^2)$ as $\varepsilon$ approaches zero. \end{abstract} \renewcommand{\theequation}{\thesection.\arabic{equation}} \newtheorem{theorem}{Theorem}[section] \newtheorem{proposition}[theorem]{Proposition} \section{Introduction } Let $G \subset {\mathbb R}^2$ be a bounded and simply connected domain with smooth boundary $\partial G$ and $g$ be a smooth map from $\partial G$ into $S^1=\{x \in {\mathbb R}^2;|x|=1\}$. Consider the Ginzburg-Landau-type functional $$ E_\varepsilon(u,G)=\frac{1}{p}\int_G|\nabla u|^p +\frac{1}{4\varepsilon^p}\int_G(1-|u|^2)^2 $$ with a small parameter $\varepsilon>0$. This functional has been studied in [1] for $p=2$, $d=\mathop{\rm deg}(g,\partial G)=0$, and in [2] for $p=2$, $d=\mathop{\rm deg}(g,\partial G) \neq 0$. Here $d=\mathop{\rm deg}(g,\partial G)$ denotes the Brouwer degree of the map $g$. For other related papers, we refer to [3]--[11]. In this paper we are concerned with the case $p>2$, $d=\mathop{\rm deg}(g,\partial G)=0$. It is easy to see that the functional $E_\varepsilon(u,G)$ achieves its minimum on $$ W_g=\{v \in W^{1,p}(G,{\mathbb R}^2):v|_{\partial G}=g\} $$ at a function $u_\varepsilon$ and that \begin{equation} \lim_{\varepsilon \to 0}u_\varepsilon= u_p\quad \mbox{in }W^{1,p}(G,{\mathbb R}^2) \label{(1.1)} \end{equation} where $u_p$ is a $p$-harmonic map from $G$ into $S^1$ with boundary value $g$ [9]. Recall that $u \in W^{1,p}(G,S^1)$ is said to be a $p$-harmonic map on $G$, if $u$ is a weak solution of the equation $$ -\mathop{\rm div}(|\nabla u|^{p-2}\nabla u)=u|\nabla u|^p. $$ Under the condition $d=0$, there exists exactly one $p$-harmonic map on $G$ with the given boundary value $g$. However, there may be several minimizers of the functional. Let $\tilde{u}_\varepsilon$ be a minimizer that can be obtained as the limit of a subsequence of the minimizers $u_\varepsilon^{\tau}$ of the regularized functionals $$ E_\varepsilon^{\tau}(u,G)= \frac{1}{p}\int_G(|\nabla u|^2+\tau)^{p/2} +\frac{1}{4\varepsilon^p} \int_G(1-|u|^2)^2,\quad (\tau>0) $$ on $W_g$ as $\tau_k \to 0$, namely \begin{equation} \lim_{\tau_k \to 0}u_\varepsilon^{\tau_k}=\tilde{u}_\varepsilon \quad \mbox{in }W^{1,p}(G,{\mathbb R}^2). \label{(1.2)} \end{equation} $\tilde{u}_\varepsilon$ is called the regularizable minimizer of $E_\varepsilon(u,G)$. Our main result reads as follows. \begin{theorem} \label{th1.1} Assume that $p>2$, $d=\mathop{\rm deg}(g,\partial G)=0$. Let $\tilde{u}_\varepsilon$ be a regularizable minimizer of $E_\varepsilon(u,G)$. Then for some $\alpha \in (0,1)$ we have $$ \lim_{\varepsilon \to 0}\tilde{u}_\varepsilon = u_p\quad \mbox{in } C_{\mbox{\scriptsize\rm loc}}^{1,\alpha}(G,{\mathbb R}^2). $$ \end{theorem} We shall prove a series of preliminary propositions in Sections 2, 3, and 4. Then we complete the proof of the main theorem in $\S5$. In $\S6$, we indicate how to extend our result to the higher dimensional case. \section{Convergence of $|u_\varepsilon^{\tau}|$ } \setcounter{equation}{0} We start our argument with the following proposition. \begin{proposition} \label{prop2.1} \begin{equation} \lim_{\varepsilon,\tau \to 0}|u_\varepsilon^{\tau}|=1 \quad \mbox{in }C(\overline{G},{\mathbb R}^2). \label{(2.1)} \end{equation} \end{proposition} \paragraph{Proof.} For $\tau \in (0,1)$, we have \begin{eqnarray*} E_\varepsilon^{\tau}(u_\varepsilon,G) &\leq& E_\varepsilon^{\tau}(u_p,G) =\frac{1}{p}\int_G(|\nabla u_p|^2+\tau)^{p/2}\\ & \leq& \frac{1}{p}\int_G(|\nabla u_p|^2+1)^{p/2}=C\,. \end{eqnarray*} Hence \begin{eqnarray} &\int_G|\nabla u_\varepsilon^{\tau}|^p \leq \int_G(|\nabla u_\varepsilon^{\tau}|^2+\tau)^{p/2} \leq C,& \label{(2.2)} \\ &\int_G(1-|u_\varepsilon^{\tau}|^2)^2 \leq C\varepsilon^p\,.& \label{(2.3)} \end{eqnarray} Here and below, we denote by $C$ a universal constant which may take different values on different occasions. If necessary, we indicate explicitly its dependence. >From (2.3) it follows that there exists a subsequence $u_{\varepsilon_k} ^{\tau_k}$ of $u_\varepsilon^{\tau}$ with $\varepsilon_k \to 0,\tau_k \to 0$ as $k \to \infty$, such that \begin{equation} \lim_{k \to \infty}|u_{\varepsilon_k}^{\tau_k}| =1,\quad \mbox{ a. e. in }G\,.\label{(2.4)} \end{equation} Inequality (2.2) combined with $|u_\varepsilon^{\tau}| \leq 1$ (which follows from the maximum principle) means that $\|u_\varepsilon^{\tau}\|_{W^{1,p} (G,{\mathbb R}^2)} \leq C$ which implies that there exist a function $u_* \in W^{1,p}(G,{\mathbb R}^2)$ and a subsequence of $u_{\varepsilon_k}^{\tau_k}$, supposed to be $u_{\varepsilon_k}^{\tau_k}$ itself without loss of generality, such that \begin{equation} \lim_{k \to \infty}u_{\varepsilon_k}^{\tau_k}=u_* \quad \mbox{in }C(\overline{G},{\mathbb R}^2)\,. \label{(2.5)} \end{equation} Combining (2.5) with (2.4) yields $|u_*|=1$ in $G$ and hence $$ \lim_{k \to \infty}|u_{\varepsilon_k}^{\tau_k}|=1 \quad \mbox{in }C(\overline{G},{\mathbb R}^2)\,. $$ Since any subsequence of $|u_\varepsilon^{\tau}|$ contains a uniformly convergent subsequence and the limit is the same number 1, we may assert (2.1) which completes the proof. Next, we prove some related facts about the asymptotic behaviour of $|u_\varepsilon^{\tau}|$, although Proposition~\ref{prop2.1} is enough for proving the next steps. \begin{proposition} \label{prop2.2} For all $q \in (1,p)$, there exist constants $C$, $\lambda>0$, independent of $\varepsilon$ such that \begin{equation} \int_G|\nabla |u_\varepsilon^{\tau}||^q \leq C\varepsilon^{\lambda} \label{(2.6)} \end{equation} for $\tau \in (0,1)$ and $\varepsilon \in (0,\eta)$ for some small $\eta>0$. \end{proposition} \paragraph{Proof.} As a minimizer of $E_\varepsilon^{\tau}(u,G)$, $u=u_\varepsilon^{\tau}$ satisfies the corresponding Euler equation \begin{eqnarray} &-\mathop{\rm div}(v^{(p-2)/2}\nabla u) =\frac{1}{\varepsilon^p}u(1-|u|^2) \quad \mbox{in }G &\label{(2.7)}\\ &u|_{\partial G}=g\,,& \end{eqnarray} where $v=|\nabla u|^2+\tau$. Set $u=h(\cos \phi,\sin \phi)$ and $h=|u|$. Then \begin{eqnarray} &-\mathop{\rm div}(v^{(p-2)/2}h^2 \nabla \phi)=0 & \\ &-\mathop{\rm div}(v^{(p-2)/2}\nabla h)+h|\nabla \phi|^2v^{(p-2)/2}= \frac{1}{\varepsilon^p}h(1-h^2)\,. &\label{(2.8)} \end{eqnarray} Fix $\beta \in (0,p/2)$ and set $$ S=\{x \in G;|h(x)|>1-\varepsilon^{\beta}\}, \tilde{h}=\max(h,1-\varepsilon^{\beta}). $$ Multiplying (2.8) with $h(1-\tilde{h})$, integrating over $G$ and noticing that $\tilde{h}|_{\partial G}=1$, we have $$ \displaylines{ -\int_Gv^{(p-2)/2}h \nabla h \nabla \tilde{h} +\int_Gv^{(p-2)/2}|\nabla h|^2(1-\tilde{h}) +\int_Gv^{(p-2)/2}h^2|\nabla \phi|^2(1-\tilde{h}) \cr =\frac{1}{\varepsilon^p} \int_Gh^2(1-h^2)(1-\tilde{h}) \cr} $$ and thus we obtain \begin{equation} \int_Gv^{(p-2)/2}h \nabla h \nabla \tilde{h} \leq C\varepsilon^{\beta} \label{(2.9)} \end{equation} by using (2.2) and the facts $|\nabla u|^2=|\nabla h|^2+h^2|\nabla \phi|^2$ and $h=|u| \leq 1$. Since $\tilde{h}=1-\varepsilon^{\beta}$ on $G \setminus S$, $\tilde{h}=h$ on $S$ and $h>1/2$ for $\varepsilon>0$ small enough, (2.9) implies $$ \int_Sv^{(p-2)/2}|\nabla h|^2 \leq C\varepsilon^{\beta} $$ and hence \begin{equation} \int_S|\nabla h|^p \leq C\varepsilon^{\beta}. \label{(2.10)} \end{equation} On the other hand, from the definition of $S$ and (2.3), we have $$ C\mathop{\rm meas}(G \setminus S)\varepsilon^{2\beta} \leq \int_{G \setminus S}(1-|u|^2)^2 \leq C\varepsilon^p, $$ namely $$ \mathop{\rm meas}(G \setminus S) \leq C\varepsilon^{p-2\beta} $$ and hence using (2.2) again we see that for any $q \in (1,p)$ $$ \int_{G \setminus S}|\nabla h|^q \leq \mathop{\rm meas}(G-S)^{1-q/p} (\int_G|\nabla h|^p)^{q/p} \leq C\varepsilon^{(p-2\beta)(1-q/p)} $$ which and (2.10) imply the conclusion of Proposition~\ref{prop2.2}. \begin{proposition} \label{prop2.3} There exists a constant $C$ independent of $\varepsilon,\tau\in (0,1)$, such that \begin{equation} \frac{1}{\varepsilon^p} \int_G(1-|u_\varepsilon^{\tau}|^2) \leq C\,.\label{(2.11)} \end{equation} \end{proposition} \paragraph{Proof.} First we take the inner product of the both sides of (2.7) with $u$ and then integrate over $G$ $$ -\int_G\mathop{\rm div}(v^{(p-2)/2}\nabla u)u =\frac{1}{\varepsilon^p}\int_G|u|^2(1-|u|^2). $$ Integrating by parts and using (2.2) and the Holder inequality we obtain \begin{eqnarray} \frac{1}{\varepsilon^p}\int_G|u|^2(1-|u|^2) &\leq&\int_Gv^{(p-2)/2}|\nabla u|^2 +\int_{\partial G}v^{(p-2)/2}|u_n||u| \nonumber\\ &\leq& C+\int_{\partial G}v^{(p-2)/2}|u_n| \label{(2.12)}\\ &\leq& C+C\int_{\partial G}v^{(p-2)/2} +C\int_{\partial G}v^{(p-2)/2}|u_n|^2 \nonumber\\ &\leq& C+C\int_{\partial G}v^{p/2} \nonumber \end{eqnarray} where $n$ denotes the unit outward normal to $\partial G$ and $u_n$ the derivative with respect to $n$. To estimate $\int_{\partial G}v^{p/2}$, we choose a smooth vector field $\nu=(\nu_1,\nu_2)$ such that $\nu|_{\partial G}=n$, take the inner product of the both sides of (2.7) with $\nu \cdot \nabla u$ and integrate over $G$. Then we have $$ -\int_G\mathop{\rm div}(v^{(p-2)/2}\nabla u)(\nu \cdot \nabla u) =\frac{1}{2\varepsilon^p} \int_G(1-|u|^2)(\nu \cdot \nabla |u|^2). $$ Integrating by parts and noticing $|u|_{\partial G}=|g|=1$ and $$ \int_G(1-|u|^2)(\nu \cdot \nabla |u|^2) =-\frac{1}{2}\int_G\nabla (1-|u|^2)^2 \cdot \nu =\frac{1}{2}\int_G(1-|u|^2)^2\mathop{\rm div} \nu $$ yield \begin{equation} -\int_{\partial G}v^{(p-2)/2}|u_n|^2 +\int_Gv^{(p-2)/2}\nabla u \cdot \nabla (\nu \cdot \nabla u) =\frac{1}{4\varepsilon^p} \int_G(1-|u|^2)^2\mathop{\rm div} \nu\,.\label{(2.13)} \end{equation} >From the smoothness of $\nu$ and (2.2), (2.3) we have \begin{equation} \frac{1}{\varepsilon^p}\int_G (1-|u|^2)^2|\mathop{\rm div} \nu| \leq C \label{(2.14)} \end{equation} \begin{eqnarray} \int_Gv^{(p-2)/2} \nabla u \nabla (\nu \cdot \nabla u) &\leq& C\int_Gv^{(p-2)/2}|\nabla u|^2 +\frac{1}{2}\int_G v^{(p-2)/2} \nu \cdot \nabla v \nonumber\\ &\leq& C+\frac{1}{p} \int_G\nu \cdot \nabla (v^{p/2}) \label{(2.15)} \\ &\leq& C+\frac{1}{p}\int_G\mathop{\rm div}(\nu v^{p/2}) -\frac{1}{p}\int_Gv^{p/2}\mathop{\rm div}\nu \nonumber\\ &\leq& C+\frac{1}{p}\int_{\partial G}v^{p/2} \nonumber \end{eqnarray} and \begin{eqnarray} \int_{\partial G}v^{p/2} &=&\int_{\partial G}v^{(p-2)/2}(|u_n|^2+|g_t|^2+\tau) \nonumber\\ &\leq& \int_{\partial G}v^{(p-2)/2} |u_n|^2+C\int_{\partial G}v^{(p-2)/2}\label{(2.16)} \end{eqnarray} where $g_t$ denotes the derivative of $g$ with respect to the tangent vector $t$ to $\partial G$. Combining (2.13)-(2.16) we obtain $$ \int_{\partial G}v^{p/2} \leq C\int_{\partial G}v^{(p-2)/2}+C +\frac{1}{p}\int_{\partial G}v^{p/2} $$ and derive \begin{equation} \int_{\partial G}v^{p/2} \leq C \label{(2.17)} \end{equation} by using the Young inequality. Substituting (2.17) into (2.12) yields $$ \frac{1}{\varepsilon^p}\int_G|u|^2(1-|u|^2) \leq C $$ which together with (2.3) implies (2.11). Using Proposition~\ref{prop2.2} and Proposition~\ref{prop2.3}, we may obtain the following result which is similar to but stronger than the result in Proposition~\ref{prop2.1}. \begin{proposition} \label{prop2.4} Uniformly for $\tau \in (0,1)$, $$ \lim_{\varepsilon \to 0}|u_\varepsilon^{\tau}| =1\quad \mbox{in }C(\overline{G},{\mathbb R}^2)\,. \label{(2.18)} $$ \end{proposition} \paragraph{Proof.} From (2.6) and (2.11), we have $$ \int_G|\nabla |u_\varepsilon^{\tau}||^{(p+2)/2} \leq C\varepsilon^{\lambda},\quad ~ \forall \varepsilon \in (0,\eta),\tau \in (0,1) $$ $$ \int_G(1-|u_\varepsilon^{\tau}|)^{(p+2)/2} \leq \int_G(1-|u_\varepsilon^{\tau}|) \leq \int_G(1-|u_\varepsilon^{\tau}|^2) \leq C\varepsilon^p,\quad ~\forall \varepsilon,\tau \in (0,1) $$ Thus $$ \|1-|u_\varepsilon^{\tau}|\|_{W^{1,(p+2)/2}(G,{\mathbb R}^2)} \leq C\varepsilon^{\lambda}, \quad ~\forall \varepsilon \in (0,\eta),\tau \in (0,1) $$ and hence by the embedding inequality, we obtain $$ \|1-|u_\varepsilon^{\tau}|\|_{C(G,{\mathbb R}^2)} \leq C\varepsilon^{\lambda} $$ which is a conclusion stronger than (2.18). \section{Estimate for $\|\nabla u_\varepsilon^{\tau}\|_ {L_{\mbox{\scriptsize\rm loc}}^l}$ } \setcounter{equation}{0} The main goal of this section is to establish uniform estimates for $\|\nabla u_\varepsilon^{\tau}\|_{L_{loc}^l}$. \begin{proposition} \label{prop3.1} There exists a constant $C$ independent of $\varepsilon,\tau \in (0,\eta)$ for small $\eta>0$, such that \begin{equation} \|\nabla u_\varepsilon^{\tau}\|_{L^l(K,{\mathbb R}^2)} \leq C=C(K,l) \label{(3.1)} \end{equation} where $K \subset G$ is an arbitrary compact subset and $l>1$. \end{proposition} \paragraph{Proof.} Differentiate (2.7) with respect to $x_j$ $$ -(v^{(p-2)/2}u_{x_i})_{x_ix_j} =\frac{1}{\varepsilon^p}(u(1-|u|^2))_{x_j}\,. $$ Here and in the sequel, double indices indicate summation. Let $\zeta \in C_0^{\infty}(G,R)$ be a function such that $\zeta=1$ on $K$, $\zeta =0$ on $G \setminus \overline{G}_1, 0 \leq \zeta \leq 1, |\nabla \zeta| \leq C$ on $G$, where $K \subset G_1$ and $G_1 \subset \subset G$ be a sub-domain. Taking the the inner product of the both sides of (2.7) with $u_{x_j}\zeta^2$ and integrating over $G$, we obtain $$ \int_G(v^{(p-2)/2}u_{x_i})_{x_j}(\zeta^2 u_{x_j})_{x_i} =\frac{1}{\varepsilon^p} \int_G(1-|u|^2)\zeta^2|\nabla u|^2 -\frac{1}{2\varepsilon^p} \int_G\zeta^2(|u|^2)_{x_j}^2. $$ Summing up for $j=1,2$ and computing the term of the left side yield \begin{eqnarray} \lefteqn{ \int_G \zeta^2v^{(p-2)/2} \sum_{j=1}^2|\nabla u_{x_j}|^2 +\frac{p-2}{4} \int_G\zeta^2v^{(p-4)/2}|\nabla v|^2 }\nonumber \\ &\leq& \frac{1}{\varepsilon^p} \int_G(1-|u|^2)\zeta^2|\nabla u|^2 +2|\int_G(v^{(p-2)/2}u_{x_i})_{x_j}u_{x_j}\zeta\zeta_{x_i}|\,.\label{(3.2)} \end{eqnarray} Applying Proposition~\ref{prop2.1} and the Young inequality, we derive from (2.7) that for any $\delta \in (0,1)$ \begin{eqnarray}\lefteqn{ \frac{1}{\varepsilon^p} \int_G(1-|u|^2)\zeta^2|\nabla u|^2 }\nonumber\\ &\leq& \int_G|u|^{-1}|\nabla u|^2\zeta^2 |\mathop{\rm div}(v^{(p-2)/2}\nabla u)| \label{(3.3)}\\ &\leq& C\int_G\zeta^2v(v^{(p-2)/2}|\Delta u| +\frac{p-2}{2}v^{(p-3)/2}|\nabla v|) \nonumber \\ &\leq& C(\delta)\int_G\zeta^2v^{(p+2)/2} +\delta \int_G \zeta^2|\Delta u|^2v^{(p-2)/2} +\delta\int_G\zeta^2v^{(p-4)/2}|\nabla v|^2 \nonumber\\ &\leq& C(\delta)\int_G\zeta^2v^{(p+2)/2} +\delta \int_G \zeta^2 \sum_{j=2}^2|\nabla u_{x_j}|^2v^{(p-2)/2} +\delta\int_G\zeta^2v^{(p-4)/2}|\nabla v|^2 \,,\nonumber \end{eqnarray} where $\varepsilon,\tau \in (0,\eta)$ with $\eta>0$ small enough. Since \begin{eqnarray*} |(v^{(p-2)/2}u_{x_i})_{x_j}u_{x_j}\zeta\zeta_{x_i}| &=&|\frac{1}{2}v^{(p-2)/2}v_{x_i} +\frac{p-2}{2}v^{(p-4)/2}v_{x_j} u_{x_i}u_{x_j}\zeta\zeta_{x_i}|\\ &\leq& Cv^{(p-2)/2}\zeta|\nabla \zeta||\nabla v|\,, \end{eqnarray*} using the Young inequality again, for $\delta \in (0,1)$ we have \begin{eqnarray} |\int_G(v^{(p-2)/2}u_{x_i})_{x_j}u_{x_j}\zeta\zeta_{x_i}| &\leq& C\int_Gv^{(p-2)/2}\zeta|\nabla \zeta||\nabla v| \label{(3.4)}\\ &\leq& C(\int_Gv^{(p-4)/2}|\nabla v|^2\zeta^2)^{1/2} (\int_Gv^{p/2}|\nabla \zeta|^2)^{1/2} \nonumber\\ &\leq& \delta\int_Gv^{(p-4)/2}|\nabla v|^2\zeta^2 +C(\delta)\int_Gv^{p/2}|\nabla \zeta|^2\,. \nonumber \end{eqnarray} Substituting (3.3) and (3.4) into (3.2) and choosing $\delta$ small enough,yield \begin{eqnarray} \lefteqn{ \int_G\zeta^2v^{(p-2)/2}\sum_{j=1}^2|\nabla u_{x_j}|^2 +\int_G\zeta^2v^{(p-4)/2}|\nabla v|^2 }\nonumber\\ &\leq& C\int_G\zeta^2v^{(p+2)/2} +C\int_Gv^{p/2}|\nabla \zeta|^2.\label{(3.5)} \end{eqnarray} Hence, by applying (2.2) to the last term, we obtain \begin{equation} \int_G\zeta^2v^{(p-4)/2}|\nabla v|^2 \leq C\int_G\zeta v^{(p+2)/2}+C. \label{(3.6)} \end{equation} Now we estimate $\int_G\zeta v^{(p+2)/2}$. To do this, we take $\phi=\zeta^{2/q}v^{(p+2)/2q}$ in the interpolation inequality $$ \|\phi\|_{L^q}\leq C\|\nabla \phi\|_{L^1}^{\alpha} \|\phi\|_{L^1}^{1-\alpha}, \quad ~q \in (1,2),\alpha=2(1-1/q). $$ Noticing that $$ |\nabla \phi| \leq C\zeta^{2/q-1}|\nabla \zeta|v^{(p+2)/2q} +C\zeta^{2/q}v^{(p+2)/2q-1}|\nabla v|, $$ we have \begin{eqnarray} \int_G\zeta^2v^{(p+2)/2} &\leq& C(\int_G\zeta^{2/q}v^{(p+2)/2q})^{q(1-\alpha)} \label{(3.7)}\\ &&\times(\int_G\zeta^{2/q-1}|\nabla \zeta|v^{(p+2)/2q} +\int_G\zeta^{2/q}v^{(p+2)/2q-1}|\nabla v|)^{q\alpha}\,.\nonumber \end{eqnarray} Since $p>2$, we can choose $q \in (1+2/p,2)$ and hence $\frac{p+2}{2q}<\frac{p}{2}$. Thus using (2.2) again,we derive that $$ \int_G\zeta^{2/q}v^{(p+2)/2q}\mbox{ and } \int_G\zeta^{2/q-1}|\nabla \zeta|v^{(p+2)/2q} $$ are bounded by $$ C\int_Gv^{(p+2)/2q} \leq C(\int_Gv^{p/2})^{(p+2)/pq}\leq C\,. $$ Substituting these inequalities into (3.7) gives \begin{eqnarray} \int_G\zeta^2v^{(p+2)/2} &\leq& C(1+\int_G\zeta^{2/q} v^{(p+2)/2q-1}|\nabla v|)^{q\alpha} \label{(3.8)}\\ &\leq& C[1+(\int_G\zeta^2v^{(p-4)/2}|\nabla v|^2)^{1/2} (\int_G\zeta^{4/q-2}v^{(p+2)/q-p/2})^{1/2}]^{q\alpha} \nonumber\\ &\leq& C+C(\int_G\zeta^2v^{(p-4)/2}|\nabla v|^2)^{q\alpha/2} (\int_G\zeta^{4/q-2}v^{(p+2)/q-p/2})^{q\alpha/2}\,. \nonumber \end{eqnarray} Here we have used the inequality $$ (a+b)^{\lambda} \leq C(a^{\lambda}+b^{\lambda}),\quad (a,b \geq 0). $$ Since $q \in (1+\frac{2}{p},2)$, we have $\frac{q\alpha}{2}<1,\frac{p+2}{q}-\frac{p}{2}<\frac{p}{2}$. Thus, using the Holder inequality and (2.2), we obtain $$ \int_G\zeta^{4/q-2}v^{(p+2)/q-p/2} \leq C(\int_Gv^{p/2})^{2(p+2)/pq-1} \leq C\,. $$ Hence from (3.8),we have for any $\delta \in (0,1)$ \begin{eqnarray*} \int_G\zeta^2v^{(p+2)/2} &\leq& C+C(\int_G\zeta^2v^{(p-4)/2} |\nabla v|^2)^{q\alpha/2}\\ &\leq& C(\delta)+\delta\int_G \zeta^2v^{(p-4)/2}|\nabla v|^2 \end{eqnarray*} since $\frac{q\alpha}{2}<1$. Combining the last inequality with (3.6) we derive $$ \int_G\zeta^2v^{(p-4)/2}|\nabla v|^2 \leq C $$ or $$ \int_G\zeta^2|\nabla w|^2 \leq C $$ where $w=v^{p/4}$. Since (2.2) implies $\int_G\zeta^2|w|^2 \leq C$, we have $\zeta w \in W^{1,2}(G,R)$, and thus the embedding inequality gives $$ \int_G(\zeta w)^l \leq C(l) $$ for any $l>1$, which implies (3.1) since $\zeta = 1$ on $K$. \section{Estimate for $\|\nabla u_\varepsilon^{\tau}\|_{L_{\mbox{\scriptsize\rm loc}}^{\infty}}$ } \setcounter{equation}{0} By means of the Moser iteration, from the estimate (3.1) we can further prove \begin{proposition} \label{prop4.1} There exists a constant $C$ independent of $\varepsilon,\tau\in (0.\eta)$ for small $\eta>0$ such that \begin{equation} \|\nabla u_\varepsilon^{\tau}\|_{L^{\infty}(K,{\mathbb R}^2)} \leq C=C(K) \label{(4.1)} \end{equation} where $K \subset G$ is an arbitrary compact subset. \end{proposition} \paragraph{Proof.} Given any $x_0 \in G$. Let $r$ be small enough and positive so that $B(x_0, 2r) \subset G$. Denote $Q_m=B(x_0,r_m),r_m=r+\frac{r}{2^m}$. Choose $\zeta_m \in C_0^{\infty}(Q_m,R)$ such that $\zeta_m=1$ on $Q_{m+1},|\nabla\zeta_m| \leq Cr^{-1}2^m,(m=1,2,\dots)$. Differentiate both sides of (2.7) with respect $x_j$, multiply by $\zeta_m^2v^bu_{x_j}$ for $b \geq 1$ and integrate over $Q_m$. Then \begin{eqnarray*} \lefteqn{ \int_{Q_m}(v^{(p-2)/2}u_{x_i})_{x_j}(\zeta_m^2v^bu_{x_j})_{x_i} }\\ &=&\frac{1}{\varepsilon^p} \int_{Q_m}(1-|u|^2)\zeta_m^2v^b|\nabla u|^2 -\frac{1}{2\varepsilon^p}\int_{Q_m}\zeta_m^2v^b(|u|^2)_{x_j}^2\,. \end{eqnarray*} Similar to the derivation of (3.2) we can obtain \begin{eqnarray}\lefteqn{ \int_{Q_m}\zeta_m^2v^{(p+2b-2)/2}\sum_j|\nabla u_{x_j}|^2 +\frac{p+2b-2}{4} \int_{Q_m}\zeta_m^2v^{(p+2b-4)/2}|\nabla v|^2 } \nonumber \\ &\leq& \frac{1}{\varepsilon^p} \int_{Q_m}(1-|u|^2)\zeta_m^2v^{b+1} +|\int_{Q_m}(v^{(p-2)/2}u_{x_i})_{x_j}v^bu_{x_j}\zeta_m\zeta_{mx_i}|\,. \label{(4.2)} \end{eqnarray} Also for any $\delta \in (0,1)$, we have \begin{eqnarray}\lefteqn{ \frac{1}{\varepsilon^p}\int_{Q_m}(1-|u|^2)\zeta_m^2v^{b+1} }\nonumber\\ &\leq& \int_{Q_m}|u|^{-1} \zeta_m^2v^{b+1}|\mathop{\rm div}(v^{(p-2)/2}\nabla u)| \nonumber\\ &\leq& C\int_{Q_m}\zeta_m^2v^{(p+2b)/2}|\Delta u|+\frac{C(p+2b-2)}{2} \int_{Q_m}\zeta_m^2v^{(p+2b-2)/2}|\nabla v| \label{(4.3)}\\ &\leq& C(\delta)\int_{Q_m}\zeta_m^2v^{(p+2b+2)/2} +\delta\int_{Q_m}\zeta_m^2v^{(p+2b-2)/2}|\Delta u|^2 \nonumber\\ &&+\frac{C(\delta)(p+2b-2)}{2} \int_{Q_m}\zeta_m^2v^{(p+2b+2)/2} \nonumber\\ &&+\frac{\delta(p+2b-2)}{2} \int_{Q_m}\zeta_m^2v^{(p+2b-4)/2}|\nabla v|^2\nonumber \end{eqnarray} and \begin{eqnarray}\lefteqn{ \bigg|\int_{Q_m}(v^{(p-2)/2} u_{x_i})_{x_j}v^bu_{x_j}\zeta_m\zeta_{mx_i}\bigg| }\nonumber \\ &\leq& C\int_{Q_m}v^{(p+2b-2)/2} |\nabla v|\zeta_m|\nabla \zeta_m| \label{(4.4)}\\ &\leq& \delta\int_{Q_m}v^{(p+2b-1)/2}|\nabla v|^2\zeta_m^2 +C(\delta)\int_{Q_m}v^{(p+2b)/2}|\nabla \zeta_m|^2\,,\nonumber \end{eqnarray} where the constants $C$ and $C(\delta)$ are independent of $b,m,\varepsilon,\tau$. Combining (4.2) with (4.3)(4.4) and choosing $\delta$ small enough yield \begin{equation} \int_{Q_m}\zeta_m^2v^{(p+2b-4)/2}|\nabla v|^2 \leq C\int_{Q_m}v^{(p+2b)/2}|\zeta_m|^2 +C\int_{Q_m}\zeta_m^2v^{(p+2b+2)/2}. \label{(4.5)} \end{equation} To estimate $\int_{Q_m}\zeta_m^2v^{(p+2b+2)/2}$, we take $$ \phi =\zeta_m^{2/q}v^{(p+2b+2)/2q} $$ in the interpolation inequality (3.6) and observe that $$ |\nabla \phi| \leq \frac{2}{q} \zeta_m^{2/q-1}|\nabla \zeta_m|v^{(p+2b+2)/2q} +\frac{p+2b+2}{2q}\zeta_m^{2/q}v^{(p+2b+2)/2q-1}|\nabla v|. $$ Then we obtain \begin{eqnarray} \lefteqn{ \int_{Q_m}\zeta_m^2v^{(p+2b+2)/2} }\nonumber\\ &\leq& C(\int_{Q_m}\zeta_m^{2/q} v^{(p+2b+2)/2q})^{q(1-\alpha)} \bigg(\frac{2}{q}\int_{Q_m}\zeta_m^{2/q-1}|\nabla \zeta_m|v^{(p+2b+2)/2q} \nonumber\\ &&+\frac{p+2b+2}{2q} \int_{Q_m}\zeta_m^{2/q}v^{(p+2b+2)/2q-1}|\nabla v|)^{q\alpha} \nonumber\\ &\leq& C(\int_{Q_m}\zeta_m^{2/q} v^{(p+2b+2)/2q})^{q(1-\alpha)} (\frac{2}{q})^{q\alpha} (\int_{Q_m}\zeta_m^{2/q-1}|\nabla \zeta_m|v^{(p+2b+2)/2q}\bigg)^{q\alpha} \label{(4.6)}\\ &&+(\frac{p+2b+2}{2q})^{q\alpha} (\int_{Q_m}\zeta_m^{2/q}v^{(p+2b+2)/2q-1}|\nabla v|)^{q\alpha}\,. \nonumber \end{eqnarray} Now we estimate all integrals on the right-hand side of (4.6). Choose $r$ small enough so that $\mathop{\rm meas}(Q_m) \leq 1$. In computing we need to notice that $q\in (1+\frac{2}{p},2)$, which implies $q>1+\frac{2}{p+2b}$ or $\frac{p+2b}{2q}<\frac{p+2b}{2}$. We have \begin{eqnarray*} \lefteqn{\int_{Q_m}\zeta_m^{2/q}v^{(p+2b+2)/2q} }\\ &\leq& \int_{Q_m}v^{(p+2b+2)/2q} \\ &\leq& (\mathop{\rm meas}(Q_m))^{1-(p+2b+2)/(q(p+2b))} (\int_{Q_m}v^{(p+2b)/2})^{(p+2b+2)/(q(p+2b))}\\ &\leq& (\int_{Q_m}v^{(p+2b)/2})^{(p+2b+2)/(q(p+2b))}, \end{eqnarray*} \begin{eqnarray*} \int_{Q_m}\zeta_m^{2/q-1}|\nabla \zeta_m|v^{(p+2b+2)/2q} &\leq& \frac{2^m}{r} \int_{Q_m}v^{(p+2b+2)/2q}\\ &\leq& \frac{2^m}{r}(\int_{Q_m}v^{(p+2b)/2})^{(p+2b+2)/(q(p+2b))} \end{eqnarray*} and \begin{eqnarray*} \lefteqn{\int_{Q_m}\zeta_m^{2/q}v^{(p+2b+2)/2q-1}|\nabla v|}\\ &\leq& (\int_{Q_m}\zeta_m^2v^{(p+2b-4)/2}|\nabla v|^2)^{1/2} (\int_{Q_m}\zeta_m^{4/q-2}v^{(p+2b+2)/q-(p+2b)/2})^{1/2}\\ &\leq& (\int_{Q_m}\zeta_m^2v^{(p+2b-4)/2}|\nabla v|^2)^{1/2} (\int_{Q_m}v^{(p+2b)/2})^{(p+2b+2)/(q(p+2b))-1/2}\,. \end{eqnarray*} Combining these inequalities with (4.5) and (4.6) yields \begin{equation} I_1 \leq C[(\frac{2^m}{r})^2I_2 +(\frac{2^m}{r})^{q\alpha}I_2^{1+2/(p+2b)} +(\frac{p+2b+2}{2q})^{q\alpha} I_1^{q\alpha/2}I_2^{1+2/(p+2b)-q\alpha/2}]\,, \label{(4.7)} \end{equation} where $$ I_1=\int_{Q_m}\zeta_m^2v^{(p+2b-4)/2}|\nabla v|^2, \mbox{ and } I_2=\int_{Q_m}v^{(p+2b)/2}. $$ Let $$ p+2b=s^m,\quad w=v^{(p+2b)/4}=v^{s^m/4} $$ with $s>2$ to be determined later. Then (4.7) becomes $$ I_1 \leq C[(\frac{2^m}{r} I_2+(\frac{2^m}{r})^{q\alpha}I_2^{1+2/{S^m}} +(\frac{s^m+2}{2q})^{q\alpha} I_1^{q\alpha/2}I_2^{1+2/{s^m}-q\alpha/2}]\,. $$ The Young inequality applied to the last term on the right-hand side yields \begin{eqnarray*} \lefteqn{C(\frac{s^m+2}{2q})^{q\alpha} I_1^{q\alpha/2}I_2^{1+2/(s^m)-q\alpha/2} }\\ &\leq& \delta I_1+C(\delta)[(\frac{s^m+2}{2q})^{q\alpha} I_2^{1+2/(s^m)-q\alpha/2}]^{2/(2-q\alpha)}\\ &=&\delta I_1+C(\delta)(\frac{s^m+2}{2q})^{2q\alpha/(2-q\alpha)} I_2^{2(1+2/(s^m)-q\alpha/2)/(2-q\alpha)}\,. \end{eqnarray*} Thus we obtain \begin{eqnarray} I_1 &\leq& C(\delta)[(\frac{2^m}{r})^2I_2 +(\frac{2^m}{r})^{q\alpha}I_2^{1+2/(s^m)} \label{(4.8)}\\ &&+(\frac{s^m+2}{2q})^{2q\alpha/(2-q\alpha)} I_2^{2(1+2/(s^m)-q\alpha/2)/(2-q\alpha)}]\,. \nonumber \end{eqnarray} By the embedding theorem, we have for any $s>1$ \begin{eqnarray*} \int_{Q_m}(\zeta_mw)^{2s} &\leq& C(s)[\int_{Q_m}(\zeta_mw)^2 +\int_{Q_m}|\nabla(\zeta_mw)|^2]^s\\ &\leq& C(s)[\int_{Q_m}(\zeta_mw)^2 +\int_{Q_m}|\nabla \zeta_m|^2w^2 +\int_{Q_m}\zeta_m^2|\nabla w|^2]^s\\ &\leq& C(s)[(1+(\frac{2^m}{r})^2)I_2 +(\frac{s^m}{4})^2I_1]^s\,, \end{eqnarray*} which by using (4.8), turns out to be \begin{eqnarray} \int_{Q_m}(\zeta_mw)^{2s} &\leq& C(s)[(1+(\frac{2^m}{r})^2 +(\frac{s^m}{4})^2(\frac{2^m}{r})^2)I_2 +(\frac{s^m}{4})^2 (\frac{2^m}{r})^{q\alpha}I_2^{1+\frac{2}{s^m}} \label{(4.9)}\\ &&+(\frac{s^m}{4})^2 (\frac{s^m+2}{2q})^{\frac{2q\alpha}{2-q\alpha}} I_2^{(1+\frac{2}{s^m}-\frac{q\alpha}{2})\frac{2}{2-q\alpha}}]^s\,.\nonumber \end{eqnarray} If there exists a subsequence of positive integers $\{m_i\}$ with $m_i \to \infty$ such that $$ I_2=\int_{Q_{m_i}}v^{s^{m_i}/2} < 1\,, $$ then as $m_i \to \infty$, \begin{equation} \|v\|_{L^{\infty}(Q_{\infty},R)} \leq C(r)\,.\label{(4.10)} \end{equation} Otherwise, there must be a positive integer $m_0$ such that $$ I_2=\int_{Q_m}v^{s^m/2} \geq 1\,,\ \forall m \geq m_0\,. $$ Since $$ (1+\frac{2}{s^m}-\frac{q\alpha}{2})\frac{2}{2-q\alpha} =1+\frac{2}{s^m}\frac{1}{2-q}>1+\frac{2}{s^m}>1\,, $$ the exponent of the last term in (4.9) is higher than those of the other terms. Now comparing the coefficients of the terms in (4.9), we have $$ (\frac{s^m}{r})^2\geq 1\,, \quad (\frac{2^m}{r})^2 \geq (\frac{2^m}{r})^{q\alpha} $$ and, if we choose $s>2q(\frac{2}{r})^{\frac{2(q-1)}{2-q}}$, $r \leq 1$, then $$ (\frac{s^m+2}{2q})^{\frac{2q\alpha}{2-q\alpha}} =(\frac{s^m+2}{2q})^{\frac{2(q-1)}{2-q}} \geq (\frac{2^m}{r})^2. $$ Therefore, the coefficient of the last term in (4.9) is greater than those of the other terms. Hence we have $$ \int_{Q_m}(\zeta_mw)^{2s}\leq C[(\frac{s^m}{4})^2 (\frac{s^m+2}{2q})^{\frac{2(q-1)}{2-q}} I_2^{1+\frac{2}{s^m}\frac{1}{2-q}}]^s $$ or \begin{equation} \int_{Q_{m+1}}v^{s^{m+1}/2} \leq (C_0C_1^m)^s(\int_{Q_m}v^{s^m/2})^{(1+C_2 /S^m)s} \label{(4.11)} \end{equation} with some constant $C_0>0$, $C_2=\frac{2}{2-q}$, $C_1=s^{(2+\frac{2(q-1)}{2-q})s}$. Using an iteration proposition which will be stated and proved later, we also reach estimate (4.10). Thus the proof of Proposition~\ref{prop4.1} is complete. \begin{proposition} \label{prop4.2} Let $Q_m(m=1,2,\dots) \subset G$ be a sequence of bounded, open subsets such that $Q_{m+1} \subset Q_m$. If for any $l \geq 1$, $v \in L^l(Q_1,R)$ and there exist constants $\lambda, C_0, C_1, C_2>0$, $s>1$, $\lambda s \geq 1$, such that $$ \int_{Q_{m+1}}|v|^{\lambda s^{m+1}}\,dx \leq (C_0C_1^m)^s(\int_{Q_m} |v|^{\lambda s^m}\,dx)^{(1+C_2/s^m)s}, \label{(4.12)} $$ for $m=1,2,\dots$, then $$ \|v\|_{L^{\infty}(Q_{\infty},R)} \leq C_0^{A_1}C_1^{A_2}(\int_{Q_{n_0}} |v|^{\lambda s^{n_0}}\,dx)^{A_3/(\lambda s^{n_0})}, \label{(4.13)} $$ where $A_1, A_2, A_3$ are constants depending only on $s$, $C_2$, and $n_0$ is an arbitrary nonnegative integer. \end{proposition} \paragraph{Proof.} From (4.12) by iteration, we obtain \begin{equation} \int_{Q_{m+1}}|v|^{\lambda s^{m+1}}\,dx \leq C_0^{X_m}C_1^{Y_m}(\int_{Q_{n_0}} |v|^{\lambda s^{n_0}}\,dx)^{s^{m-n_0+1}Z_m} \label{(4.14)} \end{equation} where \begin{eqnarray*} X_m&=&s+s^2\lambda_m+\dots+s^{m-n_0+1} \lambda_m\lambda_{m-1}\dots\lambda_{n_0+1}\\ Y_m&=&ms+(m-1)s^2\lambda_m+\dots+{n_0}s^{m-n_0+1} \lambda_m\lambda_{m-1}\dots\lambda_{n_0+1}\\ Z_m&=&\lambda_{n_0}\dots\lambda_{m-1}\lambda_m \end{eqnarray*} with $\lambda_m=1+C_2/s^m$. Since $\lambda_j \geq 1$ for $j=n_0-1,\dots,m-1,m,\dots$, $Z_m$ is an increasing sequence. Noticing that $\ln(1+x) \leq x$ for $x>0$, we have \begin{eqnarray*} lnZ_m &=&ln\lambda_{n_0}+\dots+ln\lambda_{m-1}+ln\lambda_m\\ &\leq& C_2[(\frac{1}{s})^{n_0}+\dots+ (\frac{1}{s})^{m-1}+(\frac{1}{s})^m]\\ &\leq& C_2\frac{(1/s)^{n_0}}{1-1/s}=\gamma \end{eqnarray*} or $Z_m \leq e^{\gamma}$. Hence $\lim_{m \to \infty}Z_m=A_3$ exists. Clearly, we also have \begin{eqnarray*} X_m &\leq& e^{\gamma}[s+s^2+\dots+s^{m-n_0+1}]\\ Y_m &\leq& e^{\gamma}[ms+(m-1)s^2+\dots+n_0s^{m-n_0+1}]\,. \end{eqnarray*} From which it is easily seen that the following two limits exist: \\ $\lim_{m\to\infty}s^{-(m+1)}X_m=A_1$ and $\lim_{m\to \infty}s^{-(m+1)}Y_m=A_2$. Taking the $1/\lambda s^{m+1}$ power on the both sides of (4.14), letting $m \to \infty$, and noticing that $$ \|v\|_{L^{\infty}(Q_{\infty},R)} =\lim_{m \to \infty} (\int_{Q_{m+1}}|v|^{\lambda s^{m+1}}\,dx)^{1/\lambda s^{m+1}}, $$ we obtain (4.13). \section{Completion of the proof } \setcounter{equation}{0} Once the locally uniform estimate $\|\nabla u_\varepsilon^{\tau}\|_{L^{\infty}(K)}$ is established, it is not difficult to prove the following proposition. \begin{proposition} \label{prop5.1} Let $\psi_\varepsilon^{\tau} =\frac{1}{\varepsilon^p}(1-|u_\varepsilon^{\tau}|^2)$. Then there exists a constant $C$ independent of $\varepsilon,\tau \in (0,\eta)$ with $\eta>0$ small enough, such that $$ \|\psi_\varepsilon^{\tau}\|_{L^{\infty}(K,R)} \leq C=C(K), \label{(5.1)} $$ where $K \subset G$ is an arbitrary compact subset. \end{proposition} \paragraph{Proof.} Take the inner product of both sides of (2.7) with $u$, $$ -\mathop{\rm div}(v^{(p-2)/2}\nabla u)u =\frac{1}{\varepsilon^p}|u|^2(1-|u|^2)=|u|^2\psi $$ where $u=u_\varepsilon^{\tau}, \psi=\psi_\varepsilon^{\tau}$. This and $$\displaylines{ \nabla \psi=-\frac{2}{\varepsilon^p}u \cdot \nabla u \cr -\mathop{\rm div}(v^{(p-2)/2}\nabla u)u =-\mathop{\rm div}(v^{(p-2)/2}u \cdot \nabla u) +v^{(p-2)/2}|\nabla u|^2 \cr} $$ give $$ |u|^2\psi=v^{(p-2)/2}|\nabla u|^2+\frac{\varepsilon^p}{2} \mathop{\rm div}(v^{(p-2)/2}\nabla \psi). $$ Using Proposition~\ref{prop2.1} we obtain $$ \frac{1}{2}\psi \leq v^{(p-2)/2}|\nabla u|^2 +\frac{\varepsilon^p}{2}\mathop{\rm div}(v^{(p-2)/2}\nabla \psi), ~\forall \varepsilon,\tau \in (0,\eta). \label{(5.2)} $$ Since at the point where $\psi$ achieves its maximum, $\nabla \psi=0$, $\Delta \psi \leq 0$, and $$ \mathop{\rm div}(v^{(p-2)/2}\nabla \psi) =v^{(p-2)/2}\Delta \psi +\frac{p-2}{2}v^{(p-4)/2}\nabla v \nabla \psi \leq 0\,, $$ we derive (5.1) from (5.2) by using Proposition~\ref{prop4.1}. To complete the proof of Theorem 1.1, we apply a theorem in [12] (Page 244 Line 19--23). Now according to Proposition~\ref{prop5.1} the right hand side of (2.7) is bounded on every compact subset $K \subset G$ uniformly in $\varepsilon, \tau \in (0,\eta)$. Thus applying the theorem in [12] (Page 244) yields \begin{equation} \|u_\varepsilon^{\tau}\|_{C^{1,\beta}(K)} \leq C=C(K) \label{(5.3)} \end{equation} for some $\beta \in (0,1)$, where the constant $C$ does not depend on $\varepsilon,\tau \in (0,\eta)$. From this it follows that there exist a function $u_*$ and a subsequence $u_{\varepsilon_k}^{\tau_k}(\varepsilon_k, \tau_k \to 0$, as $k \to \infty$) of $u_\varepsilon^{\tau}$, such that $$ \lim_{k \to \infty}u_{\varepsilon_k}^{\tau_k}=u_*, \quad \mbox{ in }C^{1,\alpha}(K,{\mathbb R}^2),\alpha \in (0,\beta)\,. $$ By an argument similar to that in the proof of (1.1) and (1.2), we obtain $$ \lim_{\varepsilon,\tau \to 0} u_\varepsilon^{\tau}=u_p,\quad\mbox{in }W^{1,p}(K,{\mathbb R}^2). $$ Certainly $u_*=u_p$. From the fact that any subsequence of $u_\varepsilon^{\tau}$ contains a subsequence convergent in $C^{1,\alpha}(K,{\mathbb R}^2)$ and the limit is the same function $u_p$, we may assert \begin{equation} \lim_{\varepsilon,\tau \to 0} u_\varepsilon^{\tau}=u_p,\quad \mbox{ in }C^{1,\alpha}(K,{\mathbb R}^2). \label{(5.4)} \end{equation} On the other hand, for any $\varepsilon \in (0,\eta)$, as a regularizable minimizer of $E_\varepsilon(u,G)$, $\tilde{u}_\varepsilon$ is the limit of some subsequence $u_\varepsilon^{\tau_k}$ of $u_\varepsilon^{\tau}$ in $W^{1,p}(G,{\mathbb R}^2)$. For large $k$, $u_\varepsilon^{\tau_k}$ satisfies (5.3) and hence it contains a subsequence, for simplicity we suppose it is $u_\varepsilon^{\tau_k}$ itself, such that $$ \lim_{k \to \infty}u_\varepsilon^{\tau_k} =w,\quad \mbox{ in }C^{1,\alpha}(K,{\mathbb R}^2) $$ where the function $w$ must be $\tilde{u}_\varepsilon$. Combining this with (5.4) we finally obtain $$ \lim_{\varepsilon \to 0}\tilde{u}_\varepsilon =u_p,\quad \mbox{ in }C^{1,\alpha}(K,{\mathbb R}^2) $$ and complete the proof of Theorem 1.1. \paragraph{Remark} Using Proposition~\ref{prop2.4} instead of Proposition~\ref{prop2.1} we may also prove our theorem. In this way, we may obtain (3.1), (4.1) and (5.1) for $\varepsilon \in (0,\eta),\tau \in (0,1)$ instead of those for $\varepsilon,\tau \in (0,\eta)$. The remainder of the proof is just the same as above. \section{Extension of the argument} \setcounter{equation}{0} Our argument can be extended to the higher dimensional case. Let $n>2$ $G \subset {\mathbb R}^n$ be a bounded and simply connected domain with smooth boundary $\partial G$, and $g:\partial G \to S^{n-1}=\{x \in {\mathbb R}^n;|x|=1\}$ be a smooth map with $d=\mathop{\rm deg}(g,\partial G)=0$. Consider the functional $$ E_\varepsilon(u,G)=\frac{1}{p}\int_G|\nabla u|^p +\frac{1}{4\varepsilon^p} \int_G(1-|u|^2)^2,\quad (\varepsilon>0) $$ and its regularized functional $$ E_\varepsilon^{\tau}(u,G) =\frac{1}{p}\int_G(|\nabla u|^2+\tau)^{p/2} +\frac{1}{4\varepsilon^p} \int_G(1-|u|^2)^2,\quad (\varepsilon,\tau>0) $$ on $$ W_g=\{v \in W^{1,p}(G,{\mathbb R}^n);v|_{\partial G}=g\}. $$ Similar to the case $n=2$, we may prove that if $p>1$, then $E_\varepsilon(u,G)$ and $E_\varepsilon^{\tau}(u,G)$ achieve their minimum on $W_g$ by some $u_\varepsilon$ and $u_\varepsilon^{\tau}$; $u_\varepsilon$ and $u_\varepsilon^{\tau}$ satisfy $$ -\mathop{\rm div}(|\nabla u|^{p-2} \nabla u) =\frac{1}{\varepsilon^p}u(1-|u|^2),\quad \mbox{in }G $$ and $$ -\mathop{\rm div}(v^{(p-2)/2} \nabla u) =\frac{1}{\varepsilon^p}u(1-|u|^2),\quad \mbox{in }G \label{(6.1)} $$ respectively where $v=|\nabla u|^2+\tau$, and $$ |u_\varepsilon|,|u_\varepsilon^{\tau}| \leq 1,\quad\mbox{ in }G. $$ It can also be proved that if $p>1$, then there exists a subsequence $u_{\varepsilon_k}$ of $u_\varepsilon$ with $\varepsilon_k \to 0$ such that $$ \lim_{\varepsilon_k \to 0} u_{\varepsilon_k}=u_p,\quad \mbox{in }W^{1,p}(G,{\mathbb R}^n) $$ where $u_p$ is a $p$-harmonic map with boundary value $g$. However, differ to the case $n=2$, here we can only prove the convergence for a subsequence because of the lack of uniqueness result for $p$-harmonic map with given boundary value. Similarly, for some subsequence we have $u_\varepsilon^{\tau_k}(\tau_k \to 0)$ of $u_\varepsilon^{\tau}$ $$ \lim_{\tau_k \to 0}u_\varepsilon^{\tau_k} =\tilde{u}_\varepsilon,\quad \quad{in } W^{1,p}(G,{\mathbb R}^n) $$ and the limit $\tilde{u}_\varepsilon$ is a minimizer of $E_\varepsilon(u,G)$, called regularizable minimizer. The main result is the following \begin{theorem} \label{th6.1} Assume that $p>2n-2$ and $d=\mathop{\rm deg}(g,\partial G)=0$. Let $\tilde{u}_\varepsilon$ be a regularizable minimizer of $E_\varepsilon(u,G)$. Then there exists a subsequence $\tilde{u}_{\varepsilon_k}$ of $\tilde{u}_\varepsilon$ with $\varepsilon_k \to 0$ such that for some $\alpha \in (0,1)$ $$ \lim_{\varepsilon_k \to 0} \tilde{u}_{\varepsilon_k}=u_p,\quad \mbox{in } C_{\mbox{\scriptsize\rm loc}}^{1,\alpha}(G,{\mathbb R}^n). $$ \end{theorem} The proof is similar to the case $n=2$. First we have $$ \lim_{\varepsilon,\tau \to 0} |u_\varepsilon^{\tau}|=1\quad \mbox{in }C(\overline{G},{\mathbb R}^n) $$ and also $$ \lim_{\varepsilon \to 0} |u_\varepsilon^{\tau}|=1\quad \mbox{in }C(\overline{G},{\mathbb R}^n) $$ uniformly for $\tau \in (0,1)$. Next we prove $$ \|\nabla u_\varepsilon^{\tau}\|_{L^l(K,{\mathbb R}^n)} \leq C=C(K,l) \label{(6.2)} $$ where $K \subset G$ is an arbitrary compact subset and $l>1$. For this purpose, we proceed as in section 3: first differentiate (6.1) with respect to $x_j$, take the inner product of the both sides with $\zeta^2v^bu_{x_j}(b \geq 0)$, where $\zeta \in C_{0}^{\infty}(G,R)$ with $0 \leq \zeta \leq 1$, and integrate over $G$. Then as in (3.5) and in (4.5), we obtain \begin{equation} \int_G\zeta^2v^{(p+2b-4)/2}|\nabla v|^2 \leq C\int_G\zeta^2v^{(p+2b+2)/2} +C\int_Gv^{(p+2b)/2}|\nabla \zeta|^2\,. \label{(6.3)} \end{equation} Using the interpolation inequality $$ \|\phi\|_{L^q} \leq C\|\nabla \phi\| _{L^1}^{\alpha}\|\phi\|_{L^1}^{1-\alpha}, \quad q \in (1,n/(n-1)),\ \alpha=n(q-1)/q $$ for $\phi =\zeta^{2/q}v^{(p+2b+2)/2q}$ to estimate the last term of (6.3) yields \begin{eqnarray*} \lefteqn{\int_G\zeta^2v^{(p+2b-4)/2}|\nabla v|^2 }\\ &\leq& C\int_G|\nabla \zeta|^2v^{(p+2b)/2}\\ && +C(\int_G\zeta^{2/q}v^{(p+2b+2)/2q})^{q(1-\beta)} (\int_G \zeta^{2/q-1} |\nabla \zeta|v^{(p+2b+2)/2q})^{q\beta}\\ && +C(\int_G\zeta^{2/q}v^{(p+2b+2)/2q})^{\lambda_1} (\int_G \zeta^{4/q-2}v^{(p+2b+2)/q-(p+2b)/2})^{\lambda_2}\,, \end{eqnarray*} where the constants $\lambda_1$, $\lambda_2>0$ depend on $p$, $q$, $b$, $\alpha$ only. Set $w=v^{(p+2b)/4}$. Since $p>2n-2$, we may choose $q \in (1-2/p,n/(n-1))$ such that $$ \frac{p+2b+2}{2q}<\frac{p+2b}{2} \mbox{ and } \frac{p+2b+2}{q}-\frac{p+2b}{2} <\frac{p+2b}{2}\,. $$ Then use the Holder inequality to obtain $$ \int_G\zeta^2|\nabla w|^2 \leq C\int_G|\nabla \zeta|^2w^2+C(\int_Gw^2)^{\lambda} [1+(\int_G|\nabla \zeta|^{\lambda})^{\lambda}]\,, \label{(6.4)} $$ where the constants $C$ and $\lambda>0$ are independent of $\varepsilon$ and $\tau$. Now we choose $\zeta$ such that $\zeta=1$ on $G_1$, where $G_1, G_0$ are sub-domains of $G$ satisfying $K \subset G_1 \subset \subset G_0 \subset \subset G$, $\zeta=0$ on $G \setminus \overline{G}_0$, $|\nabla \zeta| \leq C$ on $G$ and $b=0$. From $$ E_\varepsilon^{\tau}(u_\varepsilon^{\tau},G) \leq E_\varepsilon^{\tau}(u_p,G) \leq C $$ we have $$ \int_G \zeta^2|\nabla w|^2 \leq C $$ and hence $\|\zeta w\|_{L^2(G,{\mathbb R}^n)} \leq C$. By the embedding theorem, \begin{equation} \|\zeta w\|_{L^r(G,{\mathbb R}^n)}=(\int_G(\zeta w)^r)^{1/r} \leq C\|\zeta w\|_{L^2(G,{\mathbb R}^n)} \leq C\,, \label{(6.5)} \end{equation} where $r \leq \frac{2n}{n-2}$. Clearly $r=2+\frac{8}{np} \leq \frac{2n}{n-2}$. Choosing $r=2+\frac{8}{np}$ in (6.5) and noticing that $\zeta=1$ on $G_1$ we see that $\nabla u \in L^{s_1}(G_1)$and \begin{equation} \int_{G_1}|\nabla u|^{s_1} \leq C\,, \label{(6.6)} \end{equation} where $s_1=p+\frac{4}{n}$. In the present case $n>2$, we can not derive (6.2) directly by using the embedding theorem once. To prove (6.2) we choose $G_2$, a sub-domain of $G_1$, such that $K \subset G_2 \subset \subset G_1$ and $\zeta=1$ on $G_2,\zeta=0$ on $G \setminus \overline{G}_1,|\nabla \zeta| \leq C$ on $G$. Set $b=\frac{2}{n},w=v^{(p+4/n)/4}$. Then from (6.6) $$ \int_{G_1}w^2 =\int_{G_1}v^{(p+4/n)/2} =\int_{G_1}|\nabla u|^{s_1} \leq C\,, $$ and from (6.4), $$ \int_{G_1}\zeta^2 |\nabla w|^2 \leq C $$ Thus $\|\zeta w\|_{L^2(G,{\mathbb R}^n)} \leq C$. Applying the embedding theorem to $\zeta w$, we obtain $$ \int_{G_2}|\nabla u|^{s_2} \leq C\, $$ where $$ s_2=s_1+\frac{4(n+2)}{n^2}=p+\frac{4}{n}+\frac{4(n+2)}{n^2} =p+\frac{8}{n}+\frac{8}{n^2} $$ For a given $l>1$, proceeding inductively, we find an $s_i$ for some $i$ such that $s_i>l$ and $$ \int_{G_i}|\nabla u|^{s_i} \leq C\,, $$ where $G_i$ is a sub-domain of $G_{i-1}$ such that $K \subset G_i \subset \subset G_{i-1} \subset \subset \dots\subset \subset G$. Thus (6.2) is proved. The remainder of the proof is just the same as in sections 4 and 5. However, we are not able to establish the convergence for the full $\tilde{u}_\varepsilon$ because of the lack of uniqueness result for the $p$-harmonic map with the given boundary value. \begin{thebibliography}{20} \bibitem{BBH1;93} F. Bethuel, H. Brezis, F.Helein: {\it Asymptotics for the minimization of a Ginzburg-Landau functional,} Calc. Var. PDE., {\bf 1} (1993).123-148. \bibitem{BBH2} F. Bethuel, H. Brezis, F. Helein: {\it Ginzburg-Landau Vortices,} Birkhauser. 1994. \bibitem{Ding;96} S.J. Ding, Z.H. Liu: {\it On the zeroes and asymptotic behaviour of minimizers to the Ginzburg-Landau functional with variable coefficient}, J. Partial Diff. Eqns., {\bf 10} No.1. (1997).45-64. \bibitem{HL2;95} R. Hardt, F.H. Lin: {\it Singularities for p-energy minimizing unit vector fields on planner domains,} Cal. Var. PDE., {\bf 3} (1995), 311-341. \bibitem{Ho1;96} M.C. Hong: {\it Asymptotic behavior for minimizers of a Ginzburg-Landau type functional in higher dimensions associated with n-harmonic maps,} Adv. in Diff. Eqns., {\bf 1} (1996), 611-634. \bibitem{Ho2;96} M.C. Hong: {\it On a problem of Bethuel, Brezis and Helein concerning the Ginzburg-Landau functional,} C. R. Acad. Sic. Paris, {\bf 320} (1995), 679-684. \bibitem{Ho3;96} M.C. Hong: {\it Two estimates concerning asymptotics of the minimizations of a Ginzburg-Landau functional, } J. Austral. Math. Soc. Ser. A, {\bf 62} No.1, (1997), 128-140. \bibitem{Ho4;96} F.H. Lin: {\it Solutions of Ginzburg-Landau equations and critical points of the renormalized energy,} Ann. Inst. P. Poincare Anal. Nonlineare, {\bf 12} No.5, (1995), 599-622. \bibitem{L;99} Y.T. Lei: {\it Asymptotic behavior of minimizers for a functional} Acta. Scien. Natur. Univer. Jilin., {\bf 1} (1999), 1-6. \bibitem{St1;94} M. Struwe: {\it On the asymptotic behavior of minimizers of the Ginzburg-Landau model in 2-dimensions,} Diff. and Int. Eqns., {\bf 7} (1994), 1613-1624. \bibitem{St2;93} M. Struwe: {\it An aysmptotic estimate for the Ginzburg-Landau model,} C.R.Acad. Sci. Paris., {\bf 317} (1993), 677-680. \bibitem{To2;83} P. Tolksdorf: {\it Everywhere regularity for some quasilinear systems with a lack of ellipticity,} Anna. Math. Pura. Appl., {\bf 134} (1983), 241-266. \end{thebibliography} \medskip \noindent {\sc Yutian Lei} \\ {\sc Zhuoqun Wu} (e-mail: wzq@mail.jlu.edu.cn )\\ Institute of Mathematics, Jilin University \\ 130023 Changchun, P. R. China \end{document}