%\pdfoutput=1\relax\pdfpagewidth=8.26in\pdfpageheight=11.69in\pdfcompresslevel=9 \documentclass[twoside]{article} \usepackage{amssymb} % font used for R in Real numbers \pagestyle{myheadings} \markboth{\hfil Asymptotic behavior of regularizable minimizers \hfil EJDE--2001/15} {EJDE--2001/15\hfil Yutian Lei \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol. {\bf 2001}(2001), No. 15, pp. 1--28. \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} \\ % Asymptotic behavior of regularizable minimizers of a Ginzburg-Landau functional in higher dimensions % \thanks{ {\em Mathematics Subject Classifications:} 35J70. \hfil\break\indent {\em Key words:} Ginzburg-Landau functional, module and zeroes of regularizable minimizers. \hfil\break\indent \copyright 2001 Southwest Texas State University. \hfil\break\indent Submitted December 13, 2000. Published February 23, 2001.} } \date{} % \author{Yutian Lei } \maketitle \begin{abstract} We study the asymptotic behavior of the regularizable minimizers of a Ginzburg-Landau type functional. We also dicuss the location of the zeroes of the minimizers. \end{abstract} \newtheorem{theorem}{Theorem}[section] \newtheorem{proposition}[theorem]{Proposition} \renewcommand{\theequation}{\thesection.\arabic{equation}} \catcode`@=11 \@addtoreset{equation}{section} \catcode`@=12 \section{Introduction } Let $G \subset \mathbb{R}^n$ ($n\geq 2$) be a bounded and simply connected domain with smooth boundary $\partial G$. Let $g$ be a smooth map from $\partial G$ into $S^{n-1}$ satisfying $d=\deg(g,\partial G) \neq 0$. Consider the Ginzburg-Landau-type functional $$ E_\varepsilon(u,G)= \frac{1}{p}\int_G|\nabla u|^p +\frac{1}{4\varepsilon^p}\int_G(1-|u|^2)^2 ,\quad (p>1) $$ with a small parameter $\varepsilon >0$. It is known that this functional achieves its minimum on $$ W_p=\{v \in W^{1,p}(G,\mathbb{R}^n) :v|_{\partial G}=g\} $$ at a function $u_{\varepsilon}$. We are concerned with the asymptotic behavior of $u_{\varepsilon}$ and the location of the zeroes of $u_{\varepsilon}$ as $\varepsilon \to 0$. The functional $E_{\varepsilon}(u,G)$ was introduced in the study of the Ginzburg-Landau vortices by F. Bethue, H. Brezis and F. Helein [1] in the case $p=n=2$. Similar models are also used in many other theories of phase transition. The minimizer $u_{\varepsilon}$ of $E_{\varepsilon}(u,G)$ represents a complex order parameter. The zeroes of $u_{\varepsilon}$ and the module $|u_{\varepsilon}| $ both have physics senses, for example, in superconductivity $|u_{\varepsilon}|^2$ is proportional to the density of supercoducting electrons, and the zeroes of $u_{\varepsilon}$ are the vortices, which were introduced in the type-II superconductors. In the case $1n$, $W_g^{1,p}(G,S^{n-1})=\emptyset$. Thus there is no map of least p-energy on $G$ with the boundary value $g$. It seems to be very difficult to study the convergence for minimizers of $E_{\varepsilon}(u,G)$ in $W_p$. Some results on the asymptotic behavior of the radial minimizers of $E_{\varepsilon}(u,G)$ were presented in [7]. When $p=n$, this problem was introduced in [1] (the open problem 17). M. C. Hong studied the asymptotic behavior for the regularizable minimizers of $E_{\varepsilon}(u,G)$ in $W_n$ [6]. He proved that there exist $\{a_1,a_2,\dots,a_J\}\subset \overline{G}$, $J\in N$ and a subsequence $u_{\varepsilon_k}$ of the regularizable minimizers $u_{\varepsilon}$ such that $$ u_{\varepsilon_k} \stackrel{w}{\to} u_n, \quad\mbox{in } W_{{\rm loc}}^{1,n}(G\setminus \{a_1,a_2,\dots,a_J\},\mathbb{R}^n) \eqno{(1.1)} $$ as $\varepsilon_k \to 0$, where $u_n$ is an n-harmonic map. In this paper we shall discuss the asymptotic behavior for the regularizable minimizers of $E_{\varepsilon}(u,G)$ on $W_n$ in the case $p=n$. Without loss of generality, we may assume $d>0$. Recalling a minimizer of $E_{\varepsilon}(u,G)$ on $W_n$ be called the regularizable minimizer, if it is the limit of the minimizer of the regularized functional $$ E_{\varepsilon}^{\tau}(u,G) =\frac{1}{p}\int_G(|\nabla u|^2+\tau)^{p/2} +\frac{1}{4\varepsilon^p}\int_G(1-|u|^2)^2 ,\quad (\tau \in (0,1)) $$ on $W_n$ in $W^{1,p}$. It is not difficult to prove that the regularizable minimizer is also a minimizer of $E_{\varepsilon}(u,G)$. In order to find the zeroes of the minimizers, we should first locate the singularities of the n-harmonic map $u_n$. \begin{theorem} \label{th 1.1} If $a_j \in \overline{G}, j=1,2,\dots,J$ are the singularities of $n$-harmonic map $u_n$, then $J=d$, the degree $\deg(u_n, a_j)=1$, and $\{a_j\}_{j=1}^d \subset G$. Moreover, for every $j$, there exists at least one zero of the regularizable minimizer $u_{\varepsilon}$ near to $a_j$. \end{theorem} Because the module of the minimizer has the physics sense, we have also studied its asymptotic behavior. \begin{theorem} \label{th 1.2} Let $u_{\varepsilon}$ be a regularizable minimizer of $E_{\varepsilon}(u,G)$, $\rho=|u_{\varepsilon}|$, then there exists a constant $C$ independent of $\varepsilon$ such that $$ \int_G|\nabla \rho|^n \leq C, \quad\mbox{and} \quad \frac{1}{\varepsilon^n}\int_G(1-\rho^2) \leq C(1+|\ln \varepsilon|). $$ For any given $\eta>0$, denote $G_{\eta}=G\setminus \cup_{j=1}^dB(a_j,\eta)$, then as $\varepsilon \to 0$, $$ \frac{1}{\varepsilon^n} \int_{G_{\eta}}(1-\rho^2)^2 \to 0, $$ $$ \rho \to 1,\quad\mbox{ in } C_{{\rm loc}}(G_{\eta},R). $$ \end{theorem} At last, we develop the conclusion of (1.1) into following \begin{theorem} \label{th 1.3} There exists a subsequence $u_{\varepsilon_k}$ of $u_{\varepsilon}$ such that as $\varepsilon \to 0$, $$ u_{\varepsilon_k} \to u_n, \quad\mbox{ in } W_{{\rm loc}}^{1,n}(G\setminus \cup_{j=1}^d\{a_j\},\mathbb{R}^n). $$ \end{theorem} We shall prove Theorems 1.2 and 1.3 in $\S5$ and $\S7$ respectively, and the proof of Theorem 1.1 will be given in $\S6$. \section{Basic properties of the regularizable minimizers } First we recall the minimizer of the regularized functional $$ E_{\varepsilon}^{\tau}(u,G) =\frac{1}{n}\int_G (|\nabla u|^2+\tau)^{n/2} +\frac{1}{4\varepsilon^n} \int_G(1-|u|^2)^2, \quad \tau \in (0,1) $$ on $W_n$, denoted by $u_{\varepsilon}^{\tau}$. As $\tau \to 0$, there exists a subsequence $u_{\varepsilon}^{\tau_k}$ of $u_{\varepsilon}^{\tau}$ such that $$ \lim_{\tau_k \to 0}u_{\varepsilon}^{\tau_k}=u_{\varepsilon}, \quad\mbox{in } W^{1,n}(G,\mathbb{R}^n), \eqno{(2.1)} $$ and the limit $u_{\varepsilon}$ is one minimizer of $E_{\varepsilon}(u,G)$ on $W_n$, which is named the regularizable minimizer. It is not difficult to prove that $u_{\varepsilon}^{\tau}$ solves the problem $$\displaylines{ \hfill -\mathop{\rm div}[(|\nabla u|^2+\tau)^{(n-2)/2}\nabla u] =\frac{1}{\varepsilon^n}u(1-|u|^2),\quad\mbox{on }G, \hfill \llap{(2.2)}\cr u|_{\partial G}=g }$$ and satisfies the maximum principle: $|u_{\varepsilon}^{\tau}|\leq 1$ on $\overline{G}$. Moreover \begin{proposition}[{Theorem 2.2 in [6]}] \label{prop2.1} For any $\delta >0$, there exists a constant $C$ independent of $\varepsilon$ such that $$ \overline{\lim}_{\tau \to 0} |\nabla u_{\varepsilon}^{\tau}|\leq C\varepsilon^{-1},\quad\mbox{on }G^{\delta \varepsilon}, \eqno{(2.3)} $$ where $G^{\delta \varepsilon}=\{x \in G: \mathop{\rm dist}(x,\partial G)\geq \delta \varepsilon \}$. \end{proposition} In this section we shall present some basic properties of the regularizable minimizer $u_{\varepsilon}$. Clearly it is a weak solution of the equation $$ -\mathop{\rm div}(|\nabla u|^{n-2}\nabla u)= \frac{1}{\varepsilon^n}u(1-|u|^2),\quad\mbox{on }G, \eqno{(2.4)}$$ and it is known that $|u_{\varepsilon}|\leq 1$ a.e. on $\overline{G}$ [6]. We also have \begin{proposition} \label{prop2.2} For any $\delta>0$, there exists a constant $C$ independent of $\varepsilon$ such that $$ \|\nabla u_{\varepsilon}\|_{L^{\infty} (B(x,\delta \varepsilon/8,\mathbb{R}^n)} \leq C\varepsilon^{-1}, \quad\mbox{if }x \in G^{\delta \varepsilon}. $$ \end{proposition} \paragraph{Proof.} Let $y=x\varepsilon^{-1}$ in (2.4) and denote $v(y)=u(x)$, $G_{\varepsilon}=\{y=x\varepsilon^{-1} :x \in G\}, \quad G^{\delta}=\{y \in G_{\varepsilon}:\mathop{\rm dist}(y,\partial G_{\varepsilon}) >\delta\}$. Since that $u$ is a weak solution of (2.4), we have $$ \int_{G_{\varepsilon}}|\nabla v|^{n-2}\nabla v\nabla \phi =\int_{G_{\varepsilon}}v (1-|v|^2)\phi, \quad \phi \in W_0^{1,n}(G_{\varepsilon},\mathbb{R}^n). $$ Taking $\phi=v \zeta^n, \zeta \in C_0^{\infty} (G_{\varepsilon},R)$, we obtain $$ \int_{G_{\varepsilon}}|\nabla v|^n\zeta^n \leq n\int_{G_{\varepsilon}}|\nabla v|^{n-1}\zeta^{n-1} |\nabla \zeta||v| +\int_{G_{\varepsilon}}|v|^2(1-|v|^2)\zeta^n. $$ Setting $y \in G^{\delta}, B(y,\delta/2) \subset G_{\varepsilon}$, and $\zeta=1\quad\mbox{in } B(y,\delta/4), \zeta=0 \quad\mbox{in } G_{\varepsilon} \setminus B(y,\delta/2), |\nabla \zeta| \leq C(\delta)$, we have $$ \int_{B(y,\delta/2)}|\nabla v|^n\zeta^n \leq C(\delta)\int_{B(y,\delta/2)} |\nabla v|^{n-1}\zeta^{n-1}+C(\delta). $$ Using Holder inequality we can derive $\int_{B(y,\delta/4)}|\nabla v|^n \leq C(\delta)$. Combining this with the theorem of [9] yields $$ \|\nabla v \|_{L^{\infty}(B(y,\delta/8))}^n \leq C(\delta)\int_{B(y,\delta/4)}(1+|\nabla v|)^n \leq C(\delta) $$ which implies $$ \|\nabla u\|_{L^{\infty}(B(x,\varepsilon \delta /8))} \leq C(\delta)\varepsilon^{-1}. $$ \begin{proposition}[{Lemma 2.1 in [6]}] \label{prop2.3} There exists a constant $C$ independent of $\varepsilon$ such that for $\varepsilon \in (0,1)$, $$ E_{\varepsilon}(u_{\varepsilon},G)\leq d \frac{(n-1)^{n/2}}{n}|S^{n-1}||\ln \varepsilon|+C. \eqno{(2.5)} $$ \end{proposition} \begin{proposition} \label{prop2.4} There exists a constant $C$ independent of $\varepsilon$ such that $$ \frac{1}{\varepsilon^n} \int_G(1-|u_{e}|^2)^2 \leq C. \eqno{(2.6)} $$ \end{proposition} \paragraph{Proof.} By (3.6) in [6], $$ \int_G|\nabla u_{\varepsilon}|^n \geq d(n-1)^{n/2}|S^{n-1}||\ln \varepsilon|-C. $$ Applying Proposition 2.3 we may obtain (2.6). \section{A class of bad balls} Fix $\rho >0$. For the regularizable minimizer $u_{\varepsilon}$, from Theorem 2.2 in [6] we know $$ |u_{\varepsilon}| \geq \frac{1}{2}, \quad\mbox{on }G\setminus G^{\rho \varepsilon}, \eqno{(3.1)} $$ where $G^{\rho \varepsilon}=\{x \in G : \mathop{\rm dist}(x,\partial G)\geq \rho \varepsilon \}$. Thus there exists no zero of $u_{\varepsilon}$ on $G\setminus G^{\rho \varepsilon}$. \begin{proposition} \label{prop3.1} Let $u_{\varepsilon}$ be a regularizable minimizer of $E_{\varepsilon}(u,G)$, There exist positive constants $\lambda, \mu$ which are independent of $\varepsilon \in (0,1)$ such that if $$ \frac{1}{\varepsilon^n} \int_{G^{\rho \varepsilon}\cap B^{2l\varepsilon}} (1-|u_{\varepsilon}|^2)^2 \leq \mu, \eqno{(3.2)} $$ where $B^{2l\varepsilon}$ is some ball of radius $2l\varepsilon$ with $l\geq \lambda$, then $$ |u_{\varepsilon}| \geq \frac{1}{2}, \quad \forall x \in G^{\rho \varepsilon}\cap B^{l\varepsilon}. \eqno{(3.3)} $$ \end{proposition} \paragraph{Proof.} First it is known that there exists a constant $\beta>0$ such that for any $x \in G^{\rho \varepsilon}$ and $0 \frac{1}{16}, \quad \forall x \in B(x_0,\lambda \varepsilon)\cap G^{\rho \varepsilon}, $$ $$\int_{B(x_0,\lambda \varepsilon) \cap G^{\rho \varepsilon}}(1-|u_\varepsilon|^2)^2 > \frac{1}{16} |G^{\rho \varepsilon} \cap B(x_0,\lambda \varepsilon)| \geq \beta \frac{1}{16}(\lambda \varepsilon)^n =\mu \varepsilon^n. \eqno{(3.4)} $$ Since $x_0 \in B^{l\varepsilon} \cap G^{\rho \varepsilon}$, we have $(B(x_0,\lambda \varepsilon) \cap G^{\rho \varepsilon}) \subset (B^{2l\varepsilon} \cap G^{\rho \varepsilon})$, thus (3.4) implies $$ \int_{B^{2l\varepsilon} \cap G^{\rho \varepsilon}}(1-|u_\varepsilon|^2)^2 > \mu \varepsilon^n $$ which contradicts (3.2) and thus the proposition is proved. To find the zeroes of the regularizable minimizer $u_{\varepsilon}$ based on Proposition 3.1, we may take (3.2) as the ruler to distinguish the ball of radius $\lambda \varepsilon$ which contain the zeroes. Let $\lambda,\mu$ be constants in Proposition 3.1. If $$ \frac{1}{\varepsilon^n} \int_{G^{\rho \varepsilon}\cap B(x^{\varepsilon},2\lambda \varepsilon)} (1-|u_{\varepsilon}|^2)^2 \leq \mu, $$ then $B(x^{\varepsilon},\lambda \varepsilon)$ is called good ball. Otherwise $B(x^{\varepsilon},\lambda \varepsilon)$ is called bad ball. From Proposition 3.1 we are led to $$ |u_{\varepsilon}| \geq \frac{1}{2}, \quad\mbox{on }G^{\rho \varepsilon}\setminus \cup_{ x^{\varepsilon} \in \Lambda} B(x^{\varepsilon},\lambda \varepsilon), \eqno{(3.5)} $$ where $\Lambda$ is the set of the centres of all bad balls. (3.5) and (3.1) imply that the zeroes of $u_{\varepsilon}$ are contained in these bad balls. Now suppose that $\{B(x_i^{\varepsilon},\lambda \varepsilon), i \in I\}$ is a family of balls satisfying \begin{description} \item{(i)} $x_i^{\varepsilon} \in G^{\rho \varepsilon},i \in I$ \item{(ii)} $G^{\rho \varepsilon} \subset \cup_{i \in I}B(x_i^{\varepsilon},\lambda \varepsilon)$ \item{(iii)} $$B(x_i^{\varepsilon},\lambda \varepsilon /4) \cap B(x_j^{\varepsilon},\lambda \varepsilon /4)=\emptyset,i \neq j\,. \eqno{(3.6)}$$ \end{description} Let $J_\varepsilon=\{i \in I : B(x_i^{\varepsilon}, \lambda \varepsilon)\mbox{ is a bad ball}\}$. \begin{proposition} \label{prop3.2} There exists a positive integer $N$ which is independent of $\varepsilon$ such that the number of bad balls $\mathop{\rm card}J_\varepsilon \leq N$. \end{proposition} \paragraph{Proof.} Since (3.6) implies that every point in $G^{\rho \varepsilon}$ can be covered by finite, say m (independent of $\varepsilon$) balls, from (2.6) and the definition of bad balls,we have \begin{eqnarray*} \mu \varepsilon^n \mathop{\rm card} J_\varepsilon &\leq& \sum_{i \in J_\varepsilon} \int_{B(x_i^{\varepsilon},2\lambda \varepsilon) \cap G^{\rho \varepsilon}}(1-|u_\varepsilon|^2)^2\\ &\leq& m\int_{\cup_{i \in J_\varepsilon} B(x_i^{\varepsilon},2\lambda \varepsilon) \cap G^{\rho \varepsilon}}(1-|u_\varepsilon|^2)^2\\ &\leq& m\int_G(1-|u_\varepsilon|^2)^2 \leq mC\varepsilon^n \end{eqnarray*} and hence $\mathop{\rm card} J_\varepsilon\leq \frac{mC}{\mu} \leq N$. Similar to the argument of Theorem IV.1 in [1], we have \begin{proposition} \label{prop3.3} There exist a subset $J \subset J_{\varepsilon}$ and a constant $h\geq \lambda$ such that $$\displaylines{ \cup_{i \in J_{\varepsilon}}B(x_i^{\varepsilon},\lambda \varepsilon) \subset \cup_{i \in J}B(x_j^{\varepsilon},h \varepsilon), \cr \hfill |x_i^{\varepsilon}-x_j^{\varepsilon}|> 8h\varepsilon,\quad i,j \in J,\quad i \neq j. \hfill\llap{(3.7)} }$$ \end{proposition} \paragraph{Proof.} If there are two points $x_1, x_2$ such that (3.7) is not true with $h=\lambda$, we take $h_1=9\lambda$ and $J_1=J_{\varepsilon}\setminus\{1\}$. In this case, if (3.7) holds we are done. Otherwise we continue to choose a pair points $x_3, x_4$ which does not satisfy (3.7) and take $h_2=9h_1$ and $J_2=J_{\varepsilon}\setminus\{1,3\}$. After at most $N$ steps we may conclude this proposition. Applying Proposition 3.3 we may modify the family of bad balls such that the new one, denoted by $\{B(x_i^{\varepsilon},h\varepsilon) :i \in J\}$, satisfies $$\displaylines{ \cup_{i \in J_\varepsilon}B(x_i^{\varepsilon},\lambda \varepsilon) \subset \cup_{i \in J}B(x_i^{\varepsilon},h \varepsilon), \cr \hfill \lambda \leq h; \quad \mathop{\rm card} J \leq \mathop{\rm card} J_\varepsilon, \hfill\llap{(3.8)} \cr |x_i^{\varepsilon}-x_j^{\varepsilon}|>8h \varepsilon,i,j \in J,i \neq j. }$$ The last condition implies that every two balls in the new family do not intersect. As $\varepsilon \to 0$, there exist a subsequence $x_i^{\varepsilon_k}$ of $x_i^{\varepsilon}$ and $a_i \in \overline{G}$ such that $$ x_i^{\varepsilon_k} \to a_i,\quad i=1,2,\dots,N_1=\mathop{\rm card} J. $$ Perhaps there may be at least two subsequences converge to the same point, we denote by $$ a_1,a_2,\dots,a_{N_2},\quad N_2 \leq N_1 $$ the collection of distinct points in $\{a_i\}_1^{N_1}$. To prove $a_j \overline{\in}\partial G$, it is convenient to enlarge a little $G$. Assume $G'\subset \mathbb{R}^n$ is a bounded, simply connected domain with smooth boundary such that $\overline{G} \subset G'$, and take a smooth map $\bar {g}:(G'\setminus G) \to S^{n-1}$ such that $\bar{g}=g$ on $\partial G$. We extend the definition domain of every element in $\{u:G\to \mathbb{R}^n : u|_{\partial G}=g\}$ to $G'$ such that $u=\overline{g}$ on $G'\setminus G $. In particular, the regularizable minimizer $u_{\varepsilon}$ can be defined on $G'$. Fix a small constant $\sigma >0$ such that $$\displaylines{ \overline{B(a_j,\sigma)} \subset G',\quad j=1,2,\dots,N_2;\cr 4\sigma < |a_j -a_i|,\quad i \neq j;\quad 4\sigma < \mathop{\rm dist}(G,\partial G'). }$$ Writing $\Lambda_j=\{i \in J :x_i^{\varepsilon_k} \to a_j\},j=1,2,\dots,N_2$, we have $$\displaylines{ \cup _{i \in \Lambda_j} \overline{B(x_i^{\varepsilon_k}, h \varepsilon_k)} \subset B(a_j,\sigma), \quad j=1,2,\dots,N_2 \cr \cup _{j \in J}B(x_j^{\varepsilon_k},h \varepsilon_k) \subset \cup _{j=1}^{N_2}B(a_j,\sigma /4) \cr B(x_i^{\varepsilon_k},h \varepsilon_k) \cap B(x_j^{\varepsilon_k},h \varepsilon_k)=\emptyset,\quad i,j \in J,i \neq j }$$ as long as $\varepsilon_k$ is small enough. Let $u_{\varepsilon}$ is the regularizable minimizer of $E_{\varepsilon}(u,G)$ and denote $d_i^k=deg({u_{\varepsilon_k}},\partial B(x_i^{\varepsilon_k},h \varepsilon_k)), l_j^k=deg({u_{\varepsilon_k}},\partial B(a_j,\sigma))$, thus $$l_j^k=\sum_{i \in \Lambda_j}d_i^k,\quad d=\sum_{j=1}^{N_2} l_j^k. \eqno{(3.9)} $$ To prove that the degrees $d_i^k$ and $l_j^k$ are independent of $\varepsilon_k$, we recall a proposition stated in [6] (Lemma 3.3) or [2] (Theorem 8.2). \begin{proposition} \label{prop3.4} Let $\phi :S^{n-1} \to S^{n-1}$ be a $C^0$-map with $\deg \phi=d$. Then $$ \int_{S^{n-1}}|\nabla _{\tau}\phi |^{n-1}dx \geq |d|(n-1)^{(n-1)/2}|S^{n-1}|. $$ \end{proposition} \begin{proposition} \label{prop3.5} There exists a constant $C$ which is independent of $\varepsilon_k$ such that $$ |d_i^k| \leq C,i \in J;\quad |l_j^k| \leq C,j=1,2,\dots,N_2. $$ \end{proposition} \paragraph{Proof.} Since $u=u_{\varepsilon}$ is a weak solution of (2.4), applying the theory of the local regularity in [9], we know $u \in C(\partial B(x_i^{\varepsilon_k},h \varepsilon_k))$. Since (3.5) implies $|u|\geq 1/2$ on $\partial B(x_i^{\varepsilon_k},h \varepsilon_k)$, thus $\phi=\frac{u}{|u|} \in C(\partial B(x_i^{\varepsilon_k},h \varepsilon_k),S^{n-1})$. From Proposition 3.4, we have $$ |d_i^k| \leq |S^{n-1}|^{-1}(n-1)^{(1-n)/2} \int_{\partial B(x_i^{\varepsilon_k},h \varepsilon_k) }|(\frac{u}{|u|})_{\tau}|^{n-1}. $$ Since $|u|\geq \frac{1}{2}$ on $G'\setminus G^{\rho \varepsilon}$, there is no zero of $u_{\varepsilon}$ in it. Thus $$ \deg(u_{\varepsilon_k},\partial B(x_i^{\varepsilon_k},h \varepsilon_k)) =\deg(u_{\varepsilon_k},\partial (B(x_i^{\varepsilon_k}, h \varepsilon_k)\cap G^{\rho \varepsilon_k})) $$ and $$ |d_i^k| \leq |S^{n-1}|^{-1}(n-1)^{(1-n)/2} \int_{\partial[B(x_i^{\varepsilon_k},h \varepsilon_k) \cap G^{\rho \varepsilon}]}|(\frac{u}{|u|})_{\tau}|^{n-1}. \eqno{(3.10)} $$ Substituting (2.3) and the fact $|u_{\varepsilon_k}| \geq \frac{1}{2}$ on $\partial [B(x_i^{\varepsilon_k},h \varepsilon_k)\cap G^{\rho \varepsilon}]$ into (3.10), we obtain $$ |d_i^k| \leq C\varepsilon_k^{1-n}|S^{n-1}|^{-1}(n-1)^{(1-n)/2} (h \varepsilon_k)^{n-1} \leq C, $$ where $C$ is a constant which is independent of $\varepsilon_k$. Combining this with (3.9) we can complete the proof of the proposition. Proposition 3.5 implies that there exist a number $k_j$ which is independent of $\varepsilon_k$ and a subsequence of $l_j^k$ denoted itself such that $$ l_j^k \to k_j,\quad as \quad k \to \infty. $$ Since $l_j^k, k_j \in N, \{l_j^k\}$ must be constant sequence for any fixed $j$, namely $l_j^k=k_j$. The same reason shows $d_i^k$ can be writen as $d_i$ which is also a number independent of $\varepsilon_k$ later. \section{An estimate for the lower bound } Write $\Omega '=G'\setminus \cup_{j=1}^{N_2}B(a_j,\sigma)$. Fixing $j \in \{1,2,\dots,N_2\}$ and taking $i_0 \in \Lambda_j$, we have $x_{i_0} \to a_j$ as $\varepsilon \to 0$. Thus $$ \cup_{i \in \Lambda_j}\overline{B(x_i^{\varepsilon}, h \varepsilon)}\subset B(x_{i_0},\sigma /4) \subset B(a_j,\sigma) \eqno{(4.1)} $$ holds with $\varepsilon$ small enough. Denote $\Omega_j=B(a_j,\sigma)\setminus\cup_{i \in \Lambda_j} B(x_i^{\varepsilon},h \varepsilon), \Omega_{j \sigma}=B(x_{i_0},\sigma /4) \setminus\cup_{i \in \Lambda_j}B(x_i^{\varepsilon},h \varepsilon)$. To estimate the lower bound of $\|\nabla u_{\varepsilon}\| _{L^n(\Omega_j)}$, the following proposition is necessary that was given by Theorem 3.9 in [6]. \begin{proposition} \label{prop4.1} Let $A_{s,t}(x_i)=(B(x_i,s)\setminus B(x_i,t))\cap G$ with $\varepsilon \leq t0$ such that $$ R_0^l \leq M \varepsilon,\quad \mathbb{R}^L \geq \sigma /M,\quad R_0^{l+1}\leq MR^l \eqno{(4.3)} $$ for all $l=1,2,\dots,L-1$. Finally, observe that for all $R \in R_{\varepsilon}^{\sigma}$ and $J \in J_R$, $$ |k_j|=|\sum_{i \in J_R}d_{i,R}|\leq \sum_{i \in J_R}|d_{i,R}|^{n/(n-1)}. \eqno{(4.4)} $$ Applying (4.3)(4.4) and proposition 4.1 we have \begin{eqnarray*} \int_{\Omega_{j,\sigma}}|\nabla {u_{\varepsilon}}|^n &\geq& \sum_{l=1}^L \sum_{i \in J^l} |\int_{A_{\mathbb{R}^l,R_0^l}(x_i)} \nabla u_{\varepsilon}|^n\\ &\geq& \sum_{l=1}^L \sum_{i \in J^l} |S^{n-1}|(n-1)^{n/2} |d_{i,\mathbb{R}^l}|\ln (\mathbb{R}^l/R_0^l)-C\\ &\geq& |S^{n-1}|(n-1)^{n/2} |k_j|\sum_{l}(\ln \mathbb{R}^l-\ln R_0^l)-C\\ &\geq& (n-1)^{n/2}|S^{n-1}||k_j| \ln\frac{\sigma}{\varepsilon}-C. \end{eqnarray*} This and (4.1) imply that (4.2) holds. \paragraph{Remark} In fact the following results $$ \int_{\Omega_j}|\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n \geq (n-1)^{n/2}|S^{n-1}||k_j|^{n/(n-1)}\ln {\frac{\sigma}{\varepsilon}}, $$ and $$ \int_{\Omega_j}(1-|u_{\varepsilon}|^n) |\nabla \frac{u_{\varepsilon}} {|u_{\varepsilon}|}|^n\leq C $$ had been presented in the proof of Theorem 3.9 in [6], where $C$ which is independent of $\varepsilon$. Noticing $$ \int_{\Omega_j}|u_{\varepsilon}|^n |\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n =\int_{\Omega_j}|\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n -\int_{\Omega_j}(1-|u_{\varepsilon}|^n) |\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n, $$ we have $$ \int_{\Omega_j}|u_{\varepsilon}|^n |\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n \geq (n-1)^{n/2}|k_j|^{n/(n-1)}|S^{n-1}|\ln \frac{\sigma}{\varepsilon}-C. $$ \begin{theorem} \label{th4.3} There exists a constant $C$ which is independent of $\varepsilon, \sigma \in (0,1)$ such that $$ \int_{\cup_{j=1}^{N_2}\Omega_j}| \nabla {u_{\varepsilon}}|^n \geq (n-1)^{n/2}|S^{n-1}|d\ln\frac{\sigma}{\varepsilon}-C, \eqno{(4.5)} $$ $$ \frac{1}{n}\int_{G_{\sigma}} |\nabla {u_{\varepsilon}}|^n +\frac{1}{4\varepsilon^n}\int_G (1-|{u_{\varepsilon}}|^2)^2 \leq \frac{1}{n}(n-1)^{n/2} |S^{n-1}|d\ln\frac{1}{\sigma}+C \eqno{(4.6)} $$ where $G_{\sigma}=G\setminus \cup_{j=1}^{N_2}B(a_j,\sigma)$. \end{theorem} \paragraph{Proof.} From (4.2) and Proposition 2.3 we have $$ (n-1)^{n/2}|S^{n-1}|(\sum_{j=1}^{N_2}|k_j|) \ln\frac{\sigma}{\varepsilon} \leq (n-1)^{n/2}|S^{n-1}|d\ln\frac{1}{\varepsilon}+C $$ or $(\sum_{j=1}^{N_2}|k_j|-d)\ln\frac{1}{\varepsilon} \leq C$. It is seen as $\varepsilon$ small enough $$ \sum_{j=1}^{N_2}|k_j| \leq d=\sum_{j=1}^{N_2}k_j $$ which implies $$ k_j \geq 0. \eqno{(4.7)} $$ This and (3.9) imply $$\sum_{j=1}^{N_2}|k_j|=\sum_{j=1}^{N_2}k_j=d. \eqno{(4.8)} $$ Substituting (4.8) into (4.2) yields (4.5), and (4.6) may be concluded from (4.5) and Proposition 2.3. From (4.6) and the fact $|u_{\varepsilon}| \leq 1$ a.e. on $G$, we may conclude that there exists a subsequence $u_{\varepsilon_k}$ of $u_{\varepsilon}$ such that $$ u_{\varepsilon_k} \stackrel{w}{\to} u_*, \quad W^{1,n}(G_{\sigma},\mathbb{R}^n) \eqno{(4.9)} $$ as $\varepsilon_k \to 0$. Compare (4.9) with (1.1) we known $u_*=u_n$ on $G_{\sigma}$, and $$ \{a_j\}_{j=1}^{N_2}=\{a_j\}_{j=1}^J. \eqno{(4.10)} $$ These points were called the singularities of $u_n$. To show these singularities $a_j \overline{\in}{\partial G}$, the following conclussion is necessary. \begin{proposition} \label{prop4.4} Assume $a \in \partial G$ and $\sigma \in (0,R)$ with a small constant $R$. If $$ u\in W^{1,n}(A_{R,\sigma}(a),S^{n-1})\cap C^0, \quad u=\overline{g} $$ on $(G'\setminus G)\cap B(a,R)$ and $\deg(u,\partial B(a,R))=1$, then there exists a constant $C$ which is independent of $\sigma$ such that $$ \int_{A_{R,\sigma}(a)}|\nabla u|^n \geq 2^{\frac{1}{n}}(n-1)^{n/2}|S^{n-1}| \ln\frac{1}{\sigma}-C\,. \eqno{(4.11)} $$ \end{proposition} \paragraph{Proof.} Similar to the proof of Lemma VI.1 in [1], we may write $G$ as the half space $$\{(x_1,x_2,\dots,x_n) :x_n >0\}$$ locally and $a$ as $0$ by a conformal change. Denote $S_t=\partial B(0,t), t \in (\sigma,R)$. Noticing that $\overline{g}$ is smooth on $G'\setminus G$, we have $$ \sup_{\overline{G'}\setminus G}|\overline{g}_{\tau}| \leq C_1. $$ Taking $t$ sufficiently small such that $$ t \leq (n-1)^{1/2}\frac{(2^{n-1}-1)^{1/(n-1)}}{2C_1}, $$ then $$\int_{S_t^{-}}|\bar{g}_{\tau}|^{n-1} \leq |S_t^-|C_1^{n-1} \leq |S^{n-1}|t^{n-1}C_1^{n-1} \leq (n-1)^{(n-1)/2}|S^{n-1}|(1-2^{1-n}) \eqno{(4.12)} $$ with $R<1$ small enough, where $S_t^{-}=S_t \cap \{x_n <0\}$. On the other hand we can be led to $$ (n-1)^{(n-1)/2}|S^{n-1}| \leq \int_{S_t}|u_{\tau}|^{n-1} =\int_{S_t^{+}}|u_{\tau}|^{n-1} +\int_{S_t^{-}}|\bar{g}_{\tau}|^{n-1} $$ from Proposition 3.4. Here $S_t^+=S_t\setminus S_t^-$. Combining this with (4.12) yields \begin{eqnarray} \int_{S_t^{+}}|u_{\tau}|^n &\geq& |S_t^+|^{-1/(n-1)}(\int_{S_t^+} |u_{\tau}|^{n-1})^{n/(n-1)}\\ &\geq& 2^{\frac{1}{n}}|S^{n-1}|(n-1)^{n/2}t^{-1}. \end{eqnarray} Integrating this over $(\sigma,R)$, we obtain $$ \int_{A_{R,\sigma}}|\nabla u|^n \geq 2^{\frac{1}{n}}|S^{n-1}|(n-1)^{n/2} \ln\frac{R}{\sigma} $$ which implies (4.11). To prove $k_j=1$ for any $j$, we suppose $R>2\sigma$ is a small constant such that $$ \overline{B(a_j,R)} \subset G';\quad B(a_j,R)\cap B(a_i,R)=\emptyset, i \neq j. \eqno{(4.13)} $$ Denote $\Pi=\{v \in W^{1,n}(\Omega',S^{n-1})\cap C^0: \deg(v,\partial B(a_j,r))=k_j,r \in (\sigma,R), j=1,2,\dots,N_2\}$. \begin{proposition} \label{prop4.5} For any $v \in \Pi$, if $k_j\geq 0, j=1,2,\dots,N_2$, then there exists a constant $C=C(R)$ which is independent of $\sigma$ such that $$ \int_{\Omega'}|\nabla v|^n \geq (n-1)^{n/2}|S^{n-1}|(\sum_{j=1}^{N_2}k_j^{\frac {n}{n-1}})\ln\frac{1}{\sigma}-C. \eqno{(4.14)} $$ \end{proposition} \paragraph{Proof.} Write $A_{R,\sigma}(a_j)=B(a_j,R)\setminus B(a_j,\sigma)$, thus $\cup_{j=1}^{N_2}A_{R,\sigma}(a_j) \subset \Omega'$. From Proposition 3.4 we have \begin{eqnarray*} k_j=|k_j| &\leq& (n-1)^{(1-n)/2}|S^{n-1}|^{-1} \int_{S^{n-1}}|v_{\tau}|^{n-1}\\ &\leq& (n-1)^{(1-n)/2}|S^{n-1}|^{(n-1)/n} (\int_{S^{n-1}}|v_{\tau}|^n)^{(n-1)/n} \end{eqnarray*} namely $$ \int_{S^{n-1}}|v_{\tau}|^n \geq (n-1)^{n/2}|S^{n-1}|k_j^{n/(n-1)}. $$ On the other hand, we may obtain \begin{eqnarray*} \int_{\Omega'}|\nabla v|^n &\geq& \sum_{j=1}^{N_2} \int_{A_{R,\sigma}(a_j)}|\nabla v|^n\\ &\geq& \sum_{j=1}^{N_2} \int_{\sigma}^R \int_{S^{n-1}}r^{-n}|\nabla _{\tau}v| ^nr^{n-1}d\zeta dr\\ &\geq& (n-1)^{n/2}|S^{n-1}|\sum_{j=1}^{N_2}k_j^{n/(n-1)} \int_{\sigma}^Rr^{-1}dr\\ &=&(n-1)^{n/2}|S^{n-1}|(\sum_{j=1}^{N_2}k_j^ {n/(n-1)})\ln\frac{R}{\sigma} \end{eqnarray*} which implies (4.14). \section{The proof of Theorem 1.2 } Let $u_{\varepsilon}$be a regularizable minimizer of $E_{\varepsilon}(u,G)$. Proposition 2.4 has given one estimate of convergence rate of $|u_{\varepsilon}|$. Moreover, we also have \begin{theorem} \label{th5.1} There exists a constant $C$ which is independent of $\varepsilon \in (0,1)$ such that \begin{equation} \label{5.1} \frac{1}{\varepsilon^n} \int_G(1-|{u_{\varepsilon}}|^2) \leq C(1+\ln\frac{1}{\varepsilon}). \end{equation} \end{theorem} \paragraph{Proof.} The minimizer $u=u_{\varepsilon}^{\tau}$ of the regularized functional $E_{\varepsilon}^{\tau}(u,G)$ solves (2.2). Taking the inner product of the both sides of (2.2) with $u$ and integrating over $G$ we have \begin{eqnarray} \frac{1}{\varepsilon^n}\int_G|u|^2(1-|u|^2) &=&-\int_Gdiv(v^{(n-2)/2}\nabla u)u \nonumber \\ &=&\int_Gv^{(n-2)/2}|\nabla u|^2 -\int_{\partial G}v^{(n-2)/2}uu_n \\ %\eqno(5.22) &\leq& \int_Gv^{(n-2)/2}|\nabla u|^2 +C\int_{\partial G}v^{n/2}+C \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^{n/2}$, we choose a smooth vector field $\nu$ such that $\nu |_{\partial G}=n$. Multiplying (2.2) by $(\nu\cdot \nabla u)$ and integrating over $G$, we obtain \begin{eqnarray*} \frac{1}{\varepsilon^n} \int_Gu(1-|u|^2)(\nu \cdot \nabla u) &=&-\int_Gdiv(v^{(n-2)/2}\nabla u)(\nu \cdot \nabla u)\\ &=&\int_Gv^{(n-2)/2}\nabla u \cdot (\nu \cdot \nabla u) -\int_{\partial G}v^{(n-2)/2}|u_n|^2. \end{eqnarray*} Combining this with \begin{eqnarray*} \frac{1}{\varepsilon^n} \int_Gu(1-|u|^2)(\nu \cdot \nabla u) &=&\frac{1}{2\varepsilon^n} \int_G(1-|u|^2)(\nu \cdot \nabla (|u|^2))\\ &=&-\frac{1}{4\varepsilon^n} \int_G(1-|u|^2)^2\mathop{\rm div}\nu \end{eqnarray*} and \begin{eqnarray*} \lefteqn{ \int_Gv^{(n-2)/2}\nabla u \cdot \nabla(\nu \cdot \nabla u) }\\ &=&\int_Gv^{(n-2)/2}|\nabla u|^2 \mathop{\rm div}\nu+\frac{1}{n}\int_G\nu \cdot \nabla (v^{n/2})\\ &=&\int_Gv^{(n-2)/2}|\nabla u|^2 \mathop{\rm div}\nu+\frac{1}{n}\int_{\partial G}v^{n/2} -\frac{1}{n}\int_Gv^{n/2}\mathop{\rm div}\nu \end{eqnarray*} we obtain $$ \int_{\partial G}v^{(n-2)/2}|u_n|^2 \leq \frac{C}{4\varepsilon^n}\int_G(1-|u|^2)^2 +C\int_Gv^{n/2} +\frac{1}{n}\int_{\partial G}v^{n/2}. $$ Thus \begin{eqnarray*} \int_{\partial G}v^{n/2} &=&\int_{\partial G} v^{(n-2)/2}(|u_n|^2+|g_t|^2+\tau)\\ &\leq& C\int_{\partial G} v^{(n-2)/2}+\frac{1}{n} \int_{\partial G}v^{n/2} +CE_{\varepsilon}^{\tau}(u_{\varepsilon}^{\tau},G). \end{eqnarray*} Substituting this into (5.2) yields $$ \frac{1}{\varepsilon^n}\int_G |u|^2(1-|u|^2) \leq CE_{\varepsilon}^{\tau}(u_{\varepsilon}^{\tau}, G). $$ Let $\tau \to 0$, applying (2.1) and Proposition 2.3 we have $$ \frac{1}{\varepsilon^n}\int_G |u_{\varepsilon}|^2(1-|u_{\varepsilon}|^2) \leq CE_{\varepsilon}(u_{\varepsilon},G) \leq C(1+|\ln\varepsilon|) $$ which and (2.6) imply (5.1). \begin{theorem} \label{th5.2} Denote $\rho=|u_{\varepsilon}|$. There exists a constant $C$ which is independent of $\varepsilon \in (0,1)$ such that $$ \|\nabla \rho \|_{L^n(G)} \leq C. \eqno{(5.3)} $$ \end{theorem} \paragraph{Proof.} Denote $u=u_{\varepsilon}$. From the Remark in $\S4$ we know $$ \int_{\Omega_j}|u|^n|\nabla \frac{u}{|u|}|^n dx \geq (n-1)^{n/2}|k_j|^{\frac{n}{n-1}}|S^{n-1}| \ln{\frac{\sigma}{\varepsilon}} -C. $$ Thus we may modify (4.5) as $$ \int_{\cup_{j=1}^{N_2}\Omega_j}\rho^n|\nabla \frac{u}{|u|}|^n \geq (n-1)^{n/2}|S^{n-1}|d\ln\frac{\sigma}{\varepsilon}-C. $$ Combining this with $$ \int_{\cup_{j=1}^{N_2}\Omega_j}|\nabla u|^n \geq \int_{\cup_{j=1}^{N_2} \Omega_j}\rho^n|\nabla \frac{u}{|u|}|^n + \int_{\cup_{j=1}^{N_2}\Omega_j}|\nabla \rho|^n-C $$ and Proposition 2.3, we derive $$ \int_{\cup_{j=1}^{N_2}\Omega_j}|\nabla \rho|^n \leq C. \eqno{(5.4)} $$ On the other hand, from (2.1) and Proposition 2.1 we are led to $$ \int_{G^{\rho \varepsilon} \cap B(x_i,h \varepsilon)}|\nabla {u_{\varepsilon}}|^n =\lim_{\tau_k \to 0} \int_{G^{\rho \varepsilon} \cap B(x_i,h \varepsilon)}|\nabla {u_{\varepsilon}}^{\tau_k}|^n \leq C(\lambda \varepsilon)^n (C/{\varepsilon})^n \leq C, $$ for $i \in \Lambda_j$. Summarizing for $i$ and using (5.4) we can obtain (5.3). \begin{theorem} \label{th5.3} For the $\sigma >0$ in Theorem 4.4, then as $\varepsilon \to 0$, $$ \frac{1}{\varepsilon^n} \int_{G_{3\sigma}}(1-\rho^2)^2 \to 0, \eqno{(5.5)} $$ where $G_{3\sigma}=G\setminus \cup_{j=1}^{N_2}B(a_j,3\sigma)$. \end{theorem} \paragraph{Proof.} The regularizable minimizer $u_{\varepsilon}$ satisfies $$ \int_{G_{\sigma}}|\nabla u|^{n-2}\nabla u \nabla \phi =\frac{1}{\varepsilon^n} \int_{G_{\sigma}}u \phi (1-|u|^2), \eqno{(5.6)} $$ where $\phi \in W_0^{1,n}(G_{\sigma},\mathbb{R}^n)$ since $u_{\varepsilon}$ is a weak solution of (2.4). Denoting $u=u_{\varepsilon}^{\tau}=\rho w, \rho=|u|,w=\frac{u}{|u|}$ in $G_{\sigma}$ and taking $\phi =\rho w \zeta, \zeta \in W_0^{1,n}(G_{\sigma}, \mathbb{R}^n)$, we have $$ \int_{G_{\sigma}}|\nabla u|^{n-2}(w\nabla \rho+\rho \nabla w) (\rho \zeta \nabla w+\rho w \nabla \zeta +w\zeta \nabla \rho) =\frac{1}{\varepsilon^n} \int_{G_{\sigma}}\rho^2 \zeta (1-\rho^2). \eqno{(5.7)} $$ Substituting $2w\nabla w=\nabla (|w|^2)=0$ into (5.7), we obtain $$ \int_{G_{\sigma}}|\nabla u|^{n-2} (\rho \nabla \rho \nabla \zeta +|\nabla u|^2 \zeta) =\frac{1}{\varepsilon^n} \int_{G_{\sigma}}\rho^2 \zeta (1-\rho^2). \eqno{(5.8)} $$ Set $S=\{x \in G_{\sigma} :\rho(x)>1 -\varepsilon^{\beta}\}$ for some fixed $\beta \in (0,n/2)$ and $\overline{\rho}=\max(\rho, 1-\varepsilon^{\beta})$, thus $\rho=\overline{\rho}$ on $S$. In (5.8) taking $\zeta=(1-\overline{\rho})\psi$, where $\psi \in C^{\infty}(G_{\sigma},R), \psi=0$ on $G_{\sigma} \setminus G_{2\sigma}, 0<\psi<1$ on $G_{2\sigma} \setminus G_{3\sigma}, \psi=1$ on $G_{3\sigma}$, we have \setcounter{equation}{8} \begin{eqnarray} \lefteqn{\int_{G_{\sigma}}|\nabla u|^{n-2} \rho \nabla \rho \cdot \nabla \bar{\rho}\psi +\frac{1}{\varepsilon^n} \int_{G_{\sigma}}l^2(1-\rho^2)(1-\bar{\rho})\psi }\\ &=&\int_{G_{\sigma}}|\nabla u|^{n-2}\rho \nabla \rho \nabla \psi(1-\bar{\rho}) +\int_{G_{\sigma}}|\nabla u|^n\psi(1-\overline{\rho}) \nonumber \end{eqnarray} Noticing $1/2 \leq l\leq 1$ in $G_{\sigma}$ and applying (4.6) we obtain $$ \frac{1}{\varepsilon^n} \int_{G_{3\sigma}}(1-\overline{\rho})(1-\rho^2) +\int_{S \cap G_{3\sigma}} |\nabla u|^{n-2}|\nabla \rho|^2 \leq C \varepsilon^{\beta}. \eqno{(5.10)} $$ On the other hand, (2.6) implies $$ \varepsilon^{2\beta}|G_{\sigma}\setminus S| \leq \int_{G_{\sigma}\setminus S} (1-l^2)^2 \leq C\varepsilon^n, $$ namely $|G_{\sigma}\setminus S|\leq C\varepsilon^{n-2\beta}$. Then there exists a small constant $\varepsilon_0>0$ such that $$ G_{3\sigma} \subset S \cup E $$ as $\varepsilon \in (0,\varepsilon_0)$ where $E$ is a set, the measure of which converges to zero. Thus $$ \lim_{\varepsilon \to 0} \int_{G_{3\sigma}}(1-\rho^2)(1-\overline{\rho}) =\lim_{\varepsilon \to 0} \int_{G_{3\sigma}}(1+\rho)(1-\rho)^2. $$ By (5.10), \begin{eqnarray*} \lefteqn{ \lim_{\varepsilon \to 0} \frac{1}{\varepsilon^n} \int_{G_{3\sigma}}(1+\rho)^2(1-\rho)^2}\\ &\leq& \lim_{\varepsilon \to 0} \frac{2}{\varepsilon^n} \int_{G_{3\sigma}}(1-\overline{\rho})(1-\rho^2) =0 \end{eqnarray*} This is our conclusion. \begin{theorem} \label{th5.4} Assume $B(x,2\sigma) \subset G_{\sigma}$ satisfies $$ \frac{1}{\varepsilon^n} \int_{B(x,\sigma)}(1-|u_{\varepsilon}|^2)^2 \to 0, \mbox{ as }\varepsilon \to 0, \eqno{(5.11)} $$ then $|u_{\varepsilon}| \to 1$ in $C(B(x,\sigma),R)$. \end{theorem} \paragraph{Proof.} Since $B(x,2\sigma) \subset G_{\sigma}$, there exists $\varepsilon_0$ sufficiently small so that $B(x,\sigma) \subset G^{2\delta \varepsilon_0}$. We always assume $\varepsilon <\varepsilon_0$. For $x_0 \in B(x,\sigma)$, set $\alpha =|u_{\varepsilon}(x_0)|$. Proposition 2.2 implies $$ |u_{\varepsilon}(x)-u_{\varepsilon}(x_0)| 0$, we can see that the zeroes of $u_{\varepsilon}$ are also in $G$ . Moreover, the zeroes are contained in finite bad balls, i.e. $B(x_i^{\varepsilon},h \varepsilon), i \in J$. As $\varepsilon \to 0, B(x_i^{\varepsilon}, h \varepsilon) \to a_j, i\in \Lambda_j$. This implies that the zeroes of $u_{\varepsilon}$ distribute near these singularities of $u_n$ as $\varepsilon \to 0$. Thus it is necessary to describe these singularities $\{a_j\}, j=1,2,\dots,N_2$. \begin{proposition} \label{prop6.1} $k_j=\deg(u_n,a_j)$. \end{proposition} \paragraph{Proof.} Denote $\Omega'=G'\setminus \cup_{j=1}^{N_2}B(a_j,\sigma)$. Combining (4.6) and $$ \int_{G'\setminus G}|\nabla {u_{\varepsilon}}|^n =\int_{G'\setminus G}|\nabla \bar{g}|^n \leq C, $$ we have $$ \int_{\Omega'}|\nabla {u_{\varepsilon}}|^n \leq C+(n-1)^{n/2}|S^{n-1}|d|\ln \sigma|, \eqno{(6.1)} $$ where $C$ is a constant which is independent of $\varepsilon$. For $R$ in (4.13), from (6.1) we have $$ \int_{A_{R,\sigma}(a_j)} |\nabla {u_{\varepsilon}}|^n \leq C. $$ Then we know that there exists a constant $r \in (\sigma,R)$ such that $$ \int_{\partial B(a_j,r)} |\nabla {u_{\varepsilon}}|^n \leq C(r) $$ by using integral mean value theorem. Thus there exists a subsequence $u_{\varepsilon_k}$ of $u_{\varepsilon}$ such that $$ {u_{\varepsilon_k}} \to u_n,\quad \mbox{in } C(\partial B(a_j,r)) $$ as $\varepsilon_k \to 0$, which implies $$ k_j=\deg({u_{\varepsilon}},\partial B(a_j,\sigma))=\deg(u_n,a_j). $$ \begin{proposition} \label{prop6.2} $k_j=0$ or $k_j=1$. \end{proposition} \paragraph{Proof.} From the regularity results on n-harmonic maps (see [3][5] or [9]), we know $u_n \in C^0(G_{\sigma},\mathbb{R}^n)$. Set $$ w=\left\{\begin{array}{ll} \bar{g} &\mbox{on }G'\setminus G,\\ u_n &\mbox{on } G_{\sigma}, \end{array}\right. $$ then $w \in \Pi$. Using Proposition 4.5 and (4.7) we have $$ \int_{\Omega'}|\nabla w|^n \geq (n-1)^{n/2}|S^{n-1}|(\sum_{j=1}^{N_2}k_j^{\frac {n}{n-1}})\ln\frac{1}{\sigma}-C(R). \eqno{(6.2)} $$ On the other hand, (6.1) and (4.9) imply $$ {u_{\varepsilon_k}} \stackrel{w}{\to} w, \quad\mbox{in } W^{1,n}(\Omega',\mathbb{R}^n). $$ Noting this and the weak lower semicontinuity of $\int_{\Omega'}|\nabla u|^n$, applying (6.1) we have $$ \int_{\Omega'}|\nabla w|^n \leq \underline{\lim}_{\varepsilon_k \to 0} \int_{\Omega'}|\nabla u_{\varepsilon_k}|^n \leq (n-1)^{n/2}|S^{n-1}|d\ln\frac{1}{\sigma}+C. \eqno{(6.3)} $$ Combining this with (6.2), we obtain $$ (\sum_{j=1}^{N_2}k_j^{\frac{n}{n-1}}-d)\ln \frac{1}{\sigma} \leq C \quad \mbox{or} \quad \sum_{j=1}^{N_2}k_j^{\frac{n}{n-1}} \leq d=\sum_{j=1}^{N_2}k_j $$ for $\sigma$ small enough. Thus $(k_j^{1/(n-1)}-1)k_j \leq 0$ which implies that the Proposition holds. \begin{proposition} \label{prop6.3} $k_j >0$, $j=1,2,\dots,N_2$. \end{proposition} \paragraph{Proof.} Suppose $k_1=0$ and $k_2,k_3,\dots,k_{N_2}>0$. Similar to the proof of Theorem 4.3 we have $$ \int_{\cup_{j=2}^{N_2}\Omega_j} |\nabla {u_{\varepsilon}}|^n \geq (n-1)^{n/2}|S^{n-1}|d\ln\frac{\sigma}{\varepsilon}-C. $$ By this we can rewrite (4.6) as $$ \int_{G\setminus \cup_{j=2}^{N_2}B(a_j,\sigma)} |\nabla {u_{\varepsilon}}|^n +\frac{1}{4\varepsilon^n}\int_G (1-|{u_{\varepsilon}}|^2)^2 \leq C(\sigma). $$ Thus similar to the proof of Theorem 5.3 we may modify (5.5) as $$ \frac{1}{\varepsilon^n} \int_{G\setminus \cup_{j=2}^{N_2}B(a_j,3\sigma)} (1-|{u_{\varepsilon}}|^2)^2 \to 0 \eqno{(6.4)} $$ as $\varepsilon \to 0$. Noticing $$ G \cap B(a_1,\sigma) \subset G \cap B(a_1,R) \subset G\setminus \cup_{j=2}^{N_2}B(a_j,R) \subset G\setminus \cup_{j=2}^{N_2}B(a_j,3\sigma) $$ we have $$ \frac{1}{\varepsilon^n} \int_{G \cap B(a_1,\sigma)} (1-|{u_{\varepsilon}}|^2)^2 \to 0. \eqno{(6.5)} $$ On the other hand, the definition of $a_1$ implies that there exists at least one bad ball $B(x_0^{\varepsilon},h \varepsilon)$ such that $$ G \cap B(x_0^{\varepsilon},h \varepsilon)\subset G \cap B(a_1,\sigma). $$ Applying the definition of bad ball we obtain $$ \frac{1}{\varepsilon^n} \int_{G \cap B(a_1,\sigma)} (1-|{u_{\varepsilon}}|^2)^2 \geq \frac{1}{\varepsilon^n} \int_{G \cap B(x_0^{\varepsilon},h \varepsilon)} (1-|{u_{\varepsilon}}|^2)^2 \geq \mu >0 $$ which is contrary to (6.5). This contradiction shows $k_1 >0$. \paragraph{Remark} We may conclude $k_j=1,j=1,2,\dots,N_2$ from Proposition 6.2 and Proposition 6.3. Noticing $d=\sum_{j=1}^{N_2}k_j$, we obtain $$ N_2=d,\quad 1=k_j=\sum_{i \in \Lambda_j}d_i. $$ Thus on one hand, although the number of the singularities of $n-$ harmonic maps is indefinite (see Theorem A and Theorem C in [3]), we can say that for this $n-$ harmonic map $u_n$, the limit of the regularizable minimizer $u_{\varepsilon_k}$ in $W^{1,n}$ as $k \to \infty$, the number of its singularities is just the degree $d$ by applying (4.10). On the other hand, there exists at least one $i_0 \in \Lambda_j$ such that $d_{i_0} \neq 0$. Then we know that there exists at least one zero of $u_{\varepsilon}$ in $B(x_{i_0}^{\varepsilon},h \varepsilon)$ by using Kronecker's theorem. \begin{theorem} \label{th6.4} $a_j \in G,\quad j=1,2,\dots,d$. \end{theorem} \paragraph{Proof.} Suppose $a_1 \in \partial G$, $a_2,a_3,\dots,a_d \in G$. Set $$ \Omega_{\sigma}=(G'\setminus B(a_1,R))-\cup_{j=2}^dB(a_j,\sigma), \quad w=\left\{ \begin{array}{ll} u_n &\mbox{on } G_{\sigma},\\ \bar{g} &\mbox{on }G'\setminus G.\end{array}\right. $$ Using Proposition 4.5 on $\Omega_{\sigma}$ we have $$ \int_{\Omega_{\sigma}}|\nabla w|^n \geq (n-1)^{n/2}|S^{n-1}|(d-1) \ln\frac{1}{\sigma}-C(R). \eqno{(6.6)} $$ Taking $u=w,a=a_1$ in Proposition 4.4 we have $$ \int_{A_{R,\sigma}(a_1)}|\nabla w|^n \geq 2^{\frac{1}{n}}(n-1)^{n/2}|S^{n-1}| \ln\frac{1}{\sigma}-C. $$ Combining this with (6.6) yields $$ \int_{\Omega'}|\nabla w|^n \geq (d+2^{\frac{1}{n}}-1)(n-1)^{n/2}|S^{n-1}| \ln\frac{1}{\sigma}-C. $$ Compare this to (6.3) we obtain $$ (d+2^{\frac{1}{n}}-1-d)\ln\frac{1}{\sigma}\leq C $$ where $C$ is a constant which is independent of $\sigma$. It is impossible as $\sigma$ small enough, so $a_1 \in G$. \section{The proof of Theorem 1.3 } \begin{theorem} \label{th 7.1} Let $u_{\varepsilon}$ be the regularizable minimizer of $E_{\varepsilon}(u,G)$. Then there exists a subsequence $u_{\varepsilon_k}$ of $u_{\varepsilon}$ such that $$ u_{\varepsilon_k} \to u_n, \quad\mbox{ in } W_{{\rm loc}}^{1,n}(G\setminus \cup _{j=1}^d \{a_j\},\mathbb{R}^n). $$ \end{theorem} \paragraph{Proof.} Step 1: Suppose the ball $B(x_0,2\sigma) \subset G\setminus \cup_{j=1}^d \{a_j\}$, where the constant $\sigma$ may be sufficiently small but independent of $\varepsilon$. Since (4.6) implies $$ E_{\varepsilon}(u_{\varepsilon},B(x_0,2\sigma) \setminus B(x_0,\sigma)) \leq C, $$ we know there is a constant $r \in (\sigma, 2\sigma)$ such that $$ \int_{\partial B(x_0,r)}|\nabla u_{\varepsilon}|^n +\frac{1}{\varepsilon^n}\int _{\partial B(x_0,r)}(1-|u_{\varepsilon}|^2)^2 \leq C(r), \eqno{(7.1)} $$ by applying the integral mean value theorem. Thus, there exists a subsequence $u_{\varepsilon_k}$ of $u_{\varepsilon}$ such that $$ u_{\varepsilon_k} \to u_n, \quad\mbox{in } C(\partial B(x_0,r),\mathbb{R}^n), $$ which leads to $$ \frac{u_{\varepsilon_k}}{|u_{\varepsilon_k}|} \to u_n, \quad\mbox{in } C(\partial B(x_0,r),\mathbb{R}^n). \eqno{(7.2)} $$ Step 2: Denote $\rho=|u_{\varepsilon}|$ on $B=B(x_0,r)$. It is not difficult to prove that the minimizer $w$ of the problem $$ \min\{\int_B|\nabla u|^n : u \in W_{\frac{u_{\varepsilon}}{|u_{\varepsilon}|}}^{1,n}(B,S^{n-1})\} \eqno{(7.3)} $$ exists. Noting $u_{\varepsilon}$ be a minimizer of $E_{\varepsilon}(u,G)$, we have $$ E_{\varepsilon}(u_{\varepsilon},B) \leq \frac{1}{n}\int_B|\nabla (\rho w)|^n +\frac{1}{4\varepsilon^n}\int_B(1-\rho^2)^2. $$ Obviously (4.6) and $|u_{\varepsilon}|\geq 1/2$ on $B$ imply $$ \frac{1}{2^n}\int_B|\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n \leq \int_B|\nabla u_{\varepsilon}|^n \leq C, $$ thus $$ \int_B|\nabla w|^n \leq \int_B|\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^n\leq C. \eqno{(7.4)} $$ Applying this we may claim that $$ \int_B|\nabla u_{\varepsilon}|^n \leq C\varepsilon^{\lambda}+\int_B|\nabla w|^n, \eqno{(7.5)} $$ for some $\lambda >0$. Its proof can be seen in $\S8$. Step 3: Let $w^{\tau}$ is a solution of $$ \min\{\int_B(|\nabla w|^2+\tau)^{n/2} : w \in W_{\frac{u_{\varepsilon}}{|u_{\varepsilon}|}}^{1,n} (B,S^{n-1})\},\quad \tau \in (0,1). \eqno{(7.6)} $$ It is easy to see that $w^{\tau}$ solves $$ -\mathop{\rm div}(v_{\varepsilon}^{(n-2)/2}\nabla w) =w|\nabla w|^2v_{\varepsilon}^{(n-2)/2}, \quad v_{\varepsilon}=|\nabla w|^2+\tau. \eqno{(7.7)} $$ as $\tau \to 0$. Noticing $\frac{u_{\varepsilon}}{|u_{\varepsilon}|} \in W_{\frac{u_{\varepsilon}}{|u_{\varepsilon}|}}^{1,n} (B,S^{n-1})$ we have \setcounter{equation}{7} \begin{eqnarray} \int_B|\nabla w^{\tau}|^n &\leq& \int_B(|\nabla w^{\tau}|^2+\tau)^{n/2}\\ &\leq& \int_B(|\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^2+\tau)^{n/2} \leq \int_B(|\nabla \frac{u_{\varepsilon}}{|u_{\varepsilon}|}|^2+1)^{n/2} \leq C \nonumber \end{eqnarray} by using (7.4), where $C$ is a constant which is independent of $\varepsilon, \tau$. Then there exist $w^* \in W_{\frac{u_{\varepsilon}}{|u_{\varepsilon}|}}^{1,n} (B,S^{n-1})$ and a subsequence of $w^{\tau}$ such that \begin{equation} w^{\tau} \stackrel{w}{\longrightarrow} w^*,\quad\mbox{in } W^{1,n}(B,\mathbb{R}^n). \end{equation} Noting the weak lower semicontinuity of $\int_B|\nabla w|^n$, we have \begin{eqnarray} \int_B|\nabla w^*|^n &\leq& \underline{\lim}_{\tau \to 0} \int_B|\nabla w^{\tau}|^n \\ &\leq& \overline{\lim}_{\tau \to 0} \int_B|\nabla w^{\tau}|^n \leq \overline{\lim}_{\tau \to 0} \int_B(|\nabla w^{\tau}|^2+\tau)^{n/2}. \nonumber \end{eqnarray} The fact that $w^{\tau}$ solves (7.6) implies $$ \overline{\lim}_{\tau \to 0} \int_B(|\nabla w^{\tau}|^2+\tau)^{n/2} \leq \lim_{\tau \to 0} \int_B(|\nabla w_*|^2+\tau)^{n/2} =\int_B|\nabla w_*|^n, $$ where $w_*$ is a solution of (7.3). This and (7.10) lead to $$ \int_B|\nabla w^*|^n \leq \underline{\lim}_{\tau \to 0} \int_B|\nabla w^{\tau}|^n \leq \overline{\lim}_{\tau \to 0} \int_B|\nabla w^{\tau}|^n \leq \int_B|\nabla w_*|^n. \eqno{(7.11)} $$ Since $w^* \in W_{\frac{u_{\varepsilon}}{|u_{\varepsilon}|}}^{1,n} (B,S^{n-1})$, we know $w^*$ also solves (7.3), namely $$ \int_B|\nabla w_*|^n=\int_B|\nabla w^*|^n. $$ Combining this with (7.11) yields $$ \lim_{\tau \to 0}\int_B|\nabla w^{\tau}|^n =\int_B|\nabla w^*|^n, $$ which and (7.9) imply $$ \nabla w^{\tau} \to \nabla w^*, \quad \mbox{in } L^n(B,\mathbb{R}^n). \eqno{(7.12)} $$ Step 4: Similar to the discussion of Step 3, we may derive the following conclusion: Let $u^{\tau}$ be a solution of $$ \min\{\int_B(|\nabla u|^2+\tau)^{n/2} : u \in W_{u_n}^{1,n}(B,S^{n-1})\},\quad \tau \in (0,1). \eqno{(7.13)} $$ Then $u^{\tau}$ satisfies $$ \int_B|\nabla u^{\tau}|^n \leq C, \eqno{(7.14)} $$ where $C$ is which is independent of $\tau$, and $u^{\tau}$ solves $$ -\mathop{\rm div}(v^{(n-2)/2}\nabla u)=u|\nabla u|^2v^{(n-2)/2}, \quad v=|\nabla u|^2+\tau. \eqno{(7.15)} $$ As $\tau \to 0$, there exists a subsequence of $u^{\tau}$ denoted itself such that $$ \nabla u^{\tau} \to \nabla u^*, \quad\mbox{in } L^n(B,\mathbb{R}^n), \eqno{(7.16)} $$ where $u^*$ is a minimizer of $\int_B|\nabla u|^n$ in $W_{u_n}^{1,n}(B,S^{n-1})$. It is well-known that $u^*$ is a map of the least n-energy, and also an n-harmonic map. Fix $R>2\sigma$ such that $B(x_0,R) \subset G\setminus \cup_{j=1}^d \{a_j\}$. Applying the regularity results on the map of the least n-energy (for example, Theorem 3.1 in [5]), we have $$ \sup_{B(x_0,r)}|\nabla u^*|^n \leq \sup_{\overline{B(x_0,R)}}|\nabla u^*|^n:=C_0. \eqno{(7.17)} $$ It is obvious that $C_0$ is a constant which is independent of $r$. Step 5: From (7.7) subtracts (7.15). Then $$ -\mathop{\rm div}(v_{\varepsilon}^{(n-2)/2}\nabla w-v^{(n-2)/2}\nabla u) =w|\nabla w|^2v_{\varepsilon}^{(n-2)/2}-u|\nabla u|^2v^{(n-2)/2}. \eqno{(7.18)} $$ Multiplying both sides of (7.18) by $w-u$ and integrating over $B$ we obtain \begin{eqnarray*} \lefteqn{-\int_{\partial B}(v_{\varepsilon}^{(n-2)/2}w_{\nu} -v^{(n-2)/2}u_{\nu})(w-u)}\\ \lefteqn{+\int_B(v_{\varepsilon}^{(n-2)/2}\nabla w -v^{(n-2)/2}\nabla u)\nabla (w-u)}\\ &=&\int_B(w|\nabla w|^2v_{\varepsilon}^{(n-2)/2} -u|\nabla u|^2v^{(n-2)/2})(w-u), \end{eqnarray*} where $\nu$ denotes the unit outside-norm vector of $\partial B$. Thus \setcounter{equation}{18} \begin{eqnarray} \lefteqn{|\int_B(v_{\varepsilon}^{(n-2)/2}\nabla w -v^{(n-2)/2}\nabla u)\nabla (w-u)| }\nonumber\\ &\leq& |\int_{\partial B}(v_{\varepsilon}^{(n-2)/2}w_{\nu} -v^{(n-2)/2}u_{\nu})(w-u)|\\ &&+|\int_B(w|\nabla u|^2v^{(n-2)/2} -u|\nabla u|^2v^{(n-2)/2})(w-u)| \nonumber\\ &&+|\int_B(w|\nabla w|^2v_{\varepsilon}^{(n-2)/2} -w|\nabla u|^2v^{(n-2)/2})(w-u)| \nonumber\\ &=&I_1+I_2+I_3. \nonumber \end{eqnarray} First we give an estimate for $I_1$. Let $w=w^{\tau}$ is a solution of (7.6). Integrating both sides of (7.7) over $B$, we have $$ -\int_{\partial B}v_{\varepsilon}^{(n-2)/2}w_{\nu} =\int_Bw|\nabla w|^2v_{\varepsilon}^{(n-2)/2}, $$ which and (7.8) imply $$ |\int_{\partial B}v_{\varepsilon}^{(n-2)/2}w_{\nu}| \leq \int_B v_{\varepsilon}^{n/2} \leq C. \eqno{(7.20)} $$ An analogous discussion shows that for the solution $u=u^{\tau}$ of (7.13) which equips with (7.14), we may also obtain $$ |\int_{\partial B}v^{(n-2)/2}u_{\nu}| \leq \int_B|\nabla u|^n \leq C. \eqno{(7.21)} $$ Applying (7.20)(7.21) we derive \setcounter{equation}{21} \begin{eqnarray} I_1 &\leq& \sup_{\partial B}|w-u| (|\int_{\partial B}v_{\varepsilon}^{(n-2)/2}w_{\nu}| +|\int_{\partial B}v^{(n-2)/2}u_{\nu}|)\\ &\leq& C\sup_{\partial B}|w-u| =C\sup_{\partial B}|\frac{u_{\varepsilon}} {|u_{\varepsilon}|}-u_n|, \nonumber \end{eqnarray} where $C$ is independent of $\varepsilon, \tau$. For the estimate of $I_3$, we have \begin{eqnarray} I_3 &\leq& \int_B|u-w|||\nabla u|^2v^{(n-2)/2}-|\nabla w|^2v_{\varepsilon}^{(n-2)/2}|\\ &\leq& 2\int_B||\nabla u|^2v^{(n-2)/2} -|\nabla w|^2v_{\varepsilon}^{(n-2)/2}|. \nonumber \end{eqnarray} For estimating $I_2$, we multiply both sides of (7.15) by $(u-w)$ and integrate over $B$, then \begin{eqnarray*} \lefteqn{-\int_{\partial B}v^{(n-2)/2}u_{\nu}(u-w) +\int_Bv^{(n-2)/2} \nabla u \nabla (u-w)}\\ &=&\int_B|\nabla u|^2v^{(n-2)/2}u(u-w) =\int_B|\nabla u|^2v^{(n-2)/2}(1-uw). \end{eqnarray*} Thus, we have \begin{eqnarray*} I_2 &\leq& \int_B|\nabla u|^2v^{(n-2)/2}|u-w|^2 =2\int_B|\nabla u|^2v^{(n-2)/2}(1-uw)\\ &\leq& 2|\int_{\partial B}v^{(n-2)/2}u_{\nu}(u-w)| +2|\int_Bv^{(n-2)/2}\nabla u \nabla (u-w)|. \end{eqnarray*} Noting (7.21) we may derive $$ I_2 \leq C\sup_{\partial B}| \frac{u_{\varepsilon}}{|u_{\varepsilon}|}-u_n| +2|\int_Bv^{(n-2)/2}\nabla u \nabla (u-w)|. \eqno{(7.24)} $$ Step 6: Substituting (7.22)-(7.24) into (7.19) yields \begin{eqnarray*} \lefteqn{|\int_B(v_{\varepsilon}^{(n-2)/2}\nabla w -v^{(n-2)/2}\nabla u)\nabla (w-u)|}\\ &\leq& C\sup_{\partial B}|\frac {u_{\varepsilon}}{|u_{\varepsilon}|}-u_n| +2|\int_Bv^{(n-2)/2}\nabla u \nabla (u-w)| \\ && +2\int_B|v_{\varepsilon}^{(n-2)/2}|\nabla w|^2 -v^{(n-2)/2}|\nabla u|^2|. \end{eqnarray*} Letting $\tau \to 0$ and applying (7.12)(7.16) we obtain \begin{eqnarray*} \lefteqn{|\int_B(|\nabla w^*|^{(n-2)/2}\nabla w^* -|\nabla u^*|^{(n-2)/2}\nabla u^*)\nabla (w^*-u^*)|}\\ &\leq& C\sup_{\partial B}|\frac {u_{\varepsilon}}{|u_{\varepsilon}|}-u_n| +2|\int_B|\nabla u^*|^{n-1} \nabla (u^*-w^*)| +2\int_B||\nabla w^*|^n -|\nabla u^*|^n|. \end{eqnarray*} Using Lemma 1.2 in [4], we have $$ 2^{n-1}\int_B|\nabla w^*-\nabla u^*|^n \leq |\int_B(|\nabla w^*|^{(n-2)/2}\nabla w^* -|\nabla u^*|^{(n-2)/2}\nabla u^*)\nabla (w^*-u^*)|. $$ Thus $$ (2^{n-1}-2)\int_B|\nabla w^*-\nabla u^*|^n \leq C\sup_{\partial B}|\frac {u_{\varepsilon}}{|u_{\varepsilon}|}-u_n| +2|\int_B|\nabla u^*|^{n-1} \nabla (u^*-w^*)|. $$ Denote $\psi (\varepsilon)=\int_B |\nabla w^*-\nabla u^*|^n$ and let $\varepsilon \to 0$, then $$ (2^{n-1}-2)\psi(\varepsilon) \leq o(1)+2(C_0|B|)^{(n-1)/n}(\psi(\varepsilon))^{1/n} \eqno{(7.25)} $$ holds by using (7.2), where $C_0$ is the constant in (7.17). We claim that for some small constant $\sigma >0$, the following holds: $$ \psi(\varepsilon) \to 0, \quad\mbox{ as } \varepsilon \to 0. \eqno{(7.26)} $$ Suppose (7.26) is not true, then there exists $\tau>0$, for any $\varepsilon_0>0$, such that as $\varepsilon<\varepsilon_0$ we have $\psi(\varepsilon) \geq 2\tau >\tau$ or $$ (\psi(\varepsilon))^{(n-1)/n}>\tau^{(n-1)/n}, \quad \forall \varepsilon<\varepsilon_0. \eqno{(7.27)} $$ Taking $\sigma$ small enough so that $$ 2(C_0|B(x_0,r)|)^{(n-1)/n}=(2^{n-2}-1)\tau^{(n-1)/n}, $$ we obtain from (7.25) \setcounter{equation}{27} \begin{eqnarray} \lefteqn{(\psi(\varepsilon))^{1/n}[(\psi(\varepsilon))^{(n-1)/n} -\frac{2(C_0|B|)^{(n-1)/n}}{2^{n-1}-2}]}\\ &=&(\psi(\varepsilon))^{1/n}[(\psi(\varepsilon))^{(n-1)/n} -\frac{1}{2}\tau^{(n-1)/n}] = o(1). \nonumber \end{eqnarray} Substituting (7.27) into (7.28) we derive $(\psi(\varepsilon))^{1/n} = o(1)$, which is contrary to (7.27). Step 7: Noting the weak lower semicontinuity of the functional $\int_B|\nabla u|^n$, from (4.9) we are led to $$ \int_B|\nabla u_n|^n \leq \underline{\lim}_{\varepsilon_k \to 0} \int_B|\nabla u_{\varepsilon_k}|^n. $$ Combining this with (7.5) and (7.26) we obtain \begin{eqnarray*} \int_B|\nabla u_n|^n \leq \underline{\lim}_{\varepsilon_k \to 0}\int_B|\nabla u_{\varepsilon_k}|^n &\leq& \overline{\lim}_{\varepsilon_k \to 0}\int_B|\nabla u_{\varepsilon_k}|^n \\ &\leq& \lim_{\varepsilon_k \to 0}\int_B|\nabla w^*|^n =\int_B|\nabla u^*|^n. \end{eqnarray*} Recalling the definition of $u^*$ in Step 4, and noticing $u_n \in W_{u_n}^{1,n}(B,S^{n-1})$, we know that $u_n$ is also a minimizer of $\int_B|\nabla u|^n$ and $$ \lim_{\varepsilon_k \to 0}\int_B|\nabla u_{\varepsilon_k}|^n=\int_B|\nabla u_n|^n=\int_B|\nabla u^*|^n, \eqno{(7.29)}$$ which and (4.9) imply $$ \nabla u_{\varepsilon_k} \to \nabla u_n, \quad\mbox{ in } L^n(B,\mathbb{R}^n). $$ Combining this with the fact $$ u_{\varepsilon_k} \to u_n, \quad\mbox{ in } L^n(B,\mathbb{R}^n), $$ which can be deduced from (4.6), we derive $$ u_{\varepsilon_k} \to u_n, \quad\mbox{ in } W^{1,n}(B,\mathbb{R}^n). $$ Then it is not difficult to complete the proof of this theorem. \section{The proof of (7.5) } To prove (7.5), we will introduce a comparison function first. Consider the functional $$ E(\rho,B)=\frac{1}{n}\int_B(|\nabla \rho|^2+1)^{n/2} +\frac{1}{2\varepsilon^n}\int_B(1-\rho)^2. $$ It is easy to prove that the minimizer $\rho_1$ of $E(\rho,B)$ on $W_{|u_{\varepsilon}|}^{1,n}(B,R^+)$ exists and satisfies $$ -div(v^{(n-2)/2}\nabla \rho) =\frac{1}{\varepsilon^n}(1-\rho) \quad on \quad B, \eqno{(8.2)} $$ $$ \rho|_{\partial B}=|u_{\varepsilon}|, \eqno{(8.3)} $$ where $v=|\nabla \rho|^2+1$. Since $1/2 \leq |u_{\varepsilon}| \leq 1$ on $B$, it follows from the maximum principle that $$ 1/2 \leq |u_{\varepsilon}| \leq \rho_1 \leq 1 \eqno{(8.4)} $$ on $\overline{B}$. Applying (4.6) we see easily that $$ E(\rho_1,B) \leq E(|u_{\varepsilon}|,B) \leq CE_{\varepsilon}(u_{\varepsilon},B) \leq C. \eqno{(8.5)} $$ Multiplying (8.2) by $(\nu \cdot \nabla \rho)$, where $\rho=\rho_1$, and integrating over $B$, we obtain $$ -\int_{\partial B}v^{(n-2)/2}(\nu \cdot \nabla \rho)^2 +\int_Bv^{(n-2)/2}\nabla \rho \cdot \nabla (\nu \cdot \nabla \rho) =\frac{1}{\varepsilon^n}\int_B(1-\rho)(\nu \cdot \nabla \rho), \eqno{(8.6)} $$ where $\nu$ denotes the unit outside norm vector on $\partial B$. Using (8.5) we have $$\begin{array}{ll} &~~|\int_Bv^{(n-2)/2}\nabla \rho \nabla(\nu \cdot \nabla \rho)| \leq C\int_Bv^{(n-2)/2}|\nabla \rho|^2 +\frac{1}{2}|\int_Bv^{(n-2)/2}\nu \cdot \nabla v|\\[3mm] &\leq C+\frac{1}{n}|\int_B\nu \cdot \nabla(v^{n/2})| \leq C+\frac{1}{n}\int_B|div(\nu v^{n/2})-v^{n/2}div \nu|\\[3mm] &C+\frac{1}{n}\int_{\partial B}v^{n/2}. \end{array} \eqno{(8.7)} $$ Combining (8.3)(7.1) and (8.5) we also have $$\begin{array}{ll} |\frac{1}{\varepsilon^n}\int_B(1-\rho)(\nu \cdot \nabla \rho)| \leq \frac{1}{2\varepsilon^n} |\int_B(1-\rho)^2div\nu-\int_{\partial B}(1-\rho)^2|\\[3mm] \leq \frac{1}{2\varepsilon^n} \int_B(1-\rho)^2|div\nu| +\frac{1}{2\varepsilon^n}\int_{\partial B}(1-\rho)^2 \leq C. \end{array} $$ Substituting this and (8.7) into (8.6) yields $$ |\int_{\partial B}v^{(n-2)/2} (\nu \cdot \nabla \rho)^2| \leq C+\frac{1}{n}\int_{\partial B}v^{n/2}. \eqno{(8.8)} $$ Applying (8.3)(7.1) and (8.8), we obtain for any $\delta \in (0,1)$, $$\begin{array}{ll} &~~\int_{\partial B}v^{n/2}= \int_{\partial B}v^{(n-2)/2}[1+ (\tau \cdot \nabla \rho)^2 +(\nu \cdot \nabla \rho)^2]\\[3mm] &=\int_{\partial B}v^{(n-2)/2}[1+ (\tau \cdot \nabla |u_{\varepsilon}|)^2 +(\nu \cdot \nabla \rho)^2]\\[3mm] &\leq \int_{\partial B}v^{(n-2)/2} +\int_{\partial B}v^{(n-2)/2}(\nu \cdot \nabla \rho)^2\\[3mm] &~~+(\int_{\partial B}v^{n-2})^{(n-2)/n} (\int_{\partial B}(\tau \cdot \nabla |u_{\varepsilon}|)^n)^{2/n}\\[3mm] &\leq C(\delta)+(\frac{1}{n}+2\delta)\int_{\partial B}v^{n/2}, \end{array} $$ where $\tau$ denotes the unit tangent vector on $\partial B$. Hence it follows by choosing $\delta >0$ so small that $$ \int_{\partial B}v^{n/2} \leq C. \eqno{(8.9)} $$ Now we multiply both sides of (8.2) by $(1-\rho)$ and integrate over $B$. Then $$ \int_Bv^{(n-2)/2}|\nabla \rho|^2 +\frac{1}{\varepsilon^n}\int_B(1-\rho)^2 =-\int_{\partial B}v^{(n-2)/2} (\nu \cdot \nabla \rho)(1-\rho). $$ From this, using (7.1)(8.3)(8.4) and (8.9) we obtain $$\begin{array}{ll} &~~E(\rho_1,B) \leq C|(\nu \cdot \nabla \rho)(1-\rho)|\\[3mm] &\leq C|\int_{\partial B}v^{n/2}|^{(n-1)/n} |\int_{\partial B}(1-\rho)^2|^{1/n}\\[3mm] &\leq C|\int_{\partial B}(1-|u_{\varepsilon}|)^2|^{1/n} \leq C\varepsilon \end{array} \eqno{(8.10)} $$ Since $u_{\varepsilon}$ is a minimizer of $E_{\varepsilon}(u,B)$, we have $$\begin{array}{ll} &~~E_{\varepsilon}(u_{\varepsilon},B) \leq E_{\varepsilon}(\rho_1 w,B)\\[3mm] &=\frac{1}{n}\int_B(|\nabla \rho_1|^2 +\rho_1^2|\nabla w|^2)^{n/2} +\frac{1}{4\varepsilon^n}\int_B(1-\rho_1^2)^2, \end{array} \eqno{(8.11)} $$ where $w$ is a solution of (7.3). On on hand, $$\begin{array}{ll} &~~\int_B(|\nabla \rho_1|^2 +\rho_1^2|\nabla w|^2)^{n/2}dx -\int_B(\rho_1^2|\nabla w|^2)^{n/2}dx\\[3mm] &=\frac{n}{2}\int_B\int_0^1[(|\nabla \rho_1|^2 +\rho_1^2|\nabla w|^2)^{(n-2)/2}s +(\rho_1^2|\nabla w|^2)^{(n-2)/2}(1-s)]ds|\nabla \rho_1|^2dx\\[3mm] &\leq C\int_B(|\nabla \rho_1|^n +|\nabla \rho_1|^2|\nabla w|^{(n-2)/2})dx. \end{array} \eqno{(8.12)} $$ On the other hand, by using (8.10) and (7.4) we have $$ \int_B|\nabla \rho_1|^2|\nabla w|^{(n-2)/2} \leq (\int_B|\nabla \rho_1|^{4n/(n+2)})^{(n+2)/2n} (\int_B|\nabla w|^n)^{(n-2)/2n} \leq C\varepsilon^{\lambda}. \eqno{(8.13)} $$ Combining (8.11)-(8.13), we can derive $$ E_{\varepsilon}(u_{\varepsilon},B) \leq \frac{1}{n}\int_B\rho_1^n|\nabla w|^n +C\varepsilon^{\lambda}, $$ where $\lambda$ is a constant only depending on $n$. Thus (7.5) can be seen by (8.4). \begin{thebibliography}{00} \bibitem{BBH2} F. Bethuel, H. Brezis, F. Helein: {\it Ginzburg-Landau Vortices, } Birkhauser. 1994. \bibitem{BCL;86} H. Brezis, J. Coron, E. Leib: {\it Harmonic maps with defects, } Comm. Math. Phys. ,{\bf 107} (1996).649-705. \bibitem{ChH;95} B. Chen, R.Hardt: {\it Prescribing singularities for p-harmonic maps, } Ind. Univ. Math. J. ,{\bf 44} (1995).575-602. \bibitem{ChHH;94} Y. Chen, M. Hong, N. Hungerbuhler: {\it Heat flow of p-harmonic maps with values into spheres,} Math. Z., {\bf 215} (1994), 25-35. \bibitem{HL;87} R. Hardt, F. Lin: {\it Mappings minimizing the $L^p$ norm of the gradient,} Comm. Pure Appl. Math., {\bf 40} (1987), 555-588. \bibitem{Ho2;96} M. 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{LWY;99} Y. Lei, Z. Wu, H. Yuan: {\it Radial minimizers of a Ginzburg-Landau type functional}, Electron. J. Diff Eqns., {\bf 1999} (1999), No. 30, 1-21. \bibitem{Str;94} M. Struwe: {\it On the asymptotic behaviour of minimizers of the Ginzburg-Landau model in 2 dimensions,} Diff. and Inter. Eqns., {\bf 7} (1994), 1613-1624. \bibitem{T;83} P. Tolksdorf: {\it Everywhere regularity for some quasilinear systems with a lack of ellipticity, } Anna. Math. Pura. Appl., {\bf 134} (1983), No. 30, 241-266. \end{thebibliography} \noindent{\sc Yutian Lei} \\ Mathematics Department, SuZhou University\\ 1 Shizi street \\ Suzhou, Jiangsu, 215006\\ P. R. China \\ e-mail: lythxl@163.com \end{document}