\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2006(2006), No. 71, pp. 1--17.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2006 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2006/71\hfil Energy quantization] {Energy quantization for Yamabe's problem in conformal dimension} \author[F. Mahmoudi\hfil EJDE-2006/71\hfilneg] {Fethi Mahmoudi} \address{Fethi Mahmoudi \newline Scuola Internazionale Superiore Di Studi Avanzati (Sissa)\\ Via Beirut 2-4, 34014 Trieste, Italy} \email{mahmoudi@sissa.it} \date{} \thanks{Submitted February 20, 2006. Published July 7, 2006.} \subjclass[2000]{35B33, 46E30, 46L65} \keywords{Critical exponents; Lorentz spaces; quantization phenomena} \begin{abstract} Rivi\`ere \cite{T.R} proved an energy quantization for Yang-Mills fields defined on $n$-dimensional Riemannian manifolds, when $n$ is larger than the critical dimension 4. More precisely, he proved that the defect measure of a weakly converging sequence of Yang-Mills fields is quantized, provided the $W^{2,1}$ norm of their curvature is uniformly bounded. In the present paper, we prove a similar quantization phenomenon for the nonlinear elliptic equation \[ - \Delta{u}= u |u|^{4/(n-2)}, \] in a subset $\Omega$ of $\mathbb{R}^n$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \section{Introduction} Let $ \Omega $ be an open subset of $\mathbb{R}^n$ with $n\geq 3$. We consider the equation \begin{equation} -\Delta{u}=u|{u}|^{4/(n-2)} \qquad \hbox{ in}\quad \Omega \label{eq1.1} \end{equation} We will say that $u$ is a weak solution of \eqref{eq1.1} in $\Omega$, if, for all $ \Phi \in C^{\infty}({\Omega})$ with compact support in $ \Omega $, we have \begin{equation} -\int_{\Omega}{\Delta{\Phi}(x) u(x)} dx = \int_{\Omega}{\Phi}(x) u(x)|{u(x)}|^{4/(n-2)}dx \label{eq1.2} \end{equation} If in addition $u$ satisfies \begin{equation} \int_{\Omega}\Big[{\frac{\partial{u}}{\partial{x_{i}}}} {\frac{\partial{u}}{\partial{x_{j}}}} {\frac{\partial{\Phi^{j}}}{\partial{x_{i}}}}-\frac{1}{2}|\nabla{u}|^{2} {\frac{\partial{\Phi^{i}}}{\partial{x_{i}}}} +\frac{n-2}{2n}|u|^{2n/(n-2)} {\frac{\partial{\Phi^{i}}}{\partial{x_{i}}}}\Big]dx = 0 \label{eq1.3} \end{equation} for any $\Phi=(\Phi^{1},\Phi^{2}\dots ,\Phi^n)$ $\in C^{\infty}(\Omega)$ with compact support in $\Omega$, we say that u is stationary. In other words, a weak solution $u$ in $\textbf{H}^{1}(\Omega)\cap \textbf{L}^{2n/(n-2)}(\Omega)$ of \eqref{eq1.1} is stationary if the functional $E$ defined by \[ E(u)=\frac{1}{2}\int_{\Omega}|\nabla{u}|^{2} +\frac{n-2}{2n}\int_{\Omega}|u|^{2n/(n-2)} \] is stationary with respect to domain variations, i.e. \[ \frac{d}{dt}(E(u_{t}))|_{t=0}=0 \] where $u_{t}(x)=u(x+t\Phi)$. It is easy to verify that a smooth solution is stationary. In this paper we prove a monotonicity formula for stationary weak solution $u$ in $\textbf{H}^{1}(\Omega)\cap \textbf{L}^{2n/(n-2)}(\Omega)$ of \eqref{eq1.1} by a similar idea as in \cite{F.P}. More precisely we have the following result. \begin{lemma} Suppose that $u \in \mathbf{L}^{2n/(n-2)}(\Omega ) \cap \mathbf H^1 (\Omega)$ is a stationary weak solution of \eqref{eq1.1}. Consider the function \[ E_u (x,r) =\int_{B(x,r)}|u|^{2n/(n-2)}dy + \frac{d}{dr}\int_{\partial B(x,r)}u^2 ds + r^{-1}\int_{B(x,r)}u^2 ds. \] Then $r \mapsto E_u(x,r) $ is positive, nondecreasing and continuous. \label{lem1.1} \end{lemma} This monotonicity formula together with ideas which go back to the work of Schoen \cite{Sch}, allowed to prove the following result. \begin{theorem} There exists $ \varepsilon>0 $ and $r_{0}>0 $ depend only on $n$ such that, for any smooth solution $u\in \textbf{H}^{1}(\Omega)\cap \textbf{L}^{2n/(n-2)}(\Omega)$ of \eqref{eq1.1}, we have: For any $x_{0}\in\Omega$, if \[\int_{B(x_{0},r_{0})}|\nabla{u}|^{2}+|u|^{2n/(n-2)} \leq \varepsilon, \] then \[\|u\|_{\textbf{L}^{\infty}(B_{\frac{r}{2}}(x_{0}))}\leq\frac{C(\varepsilon)}{r^{(n-2)/2}} \quad\hbox{for any} \quad r0$ and $\varepsilon_0 >0$ such that \[ \Sigma = \cap_{0 < r < r_0 } \big\{ x \in \Omega: \liminf_{i \to \infty}E_{u_i}(x,r) \geq \varepsilon _0 \big\}. \] \label{thm1.2} \end{theorem} We define the sequence of Radon measures \[ {\eta}_i:= (\frac{1}{2} { \vert{\nabla u_{i}} \vert}^2 + \frac{n-2}{2n} {\vert{u_i}\vert}^{2n/(n-2)}) \,dx \] Assumption that the sequence $( { \Vert {\nabla u_i} \Vert}_{\mathbf H^{1}(\Omega)} +{\Vert{u_i}\Vert}_{\mathbf{L} ^{2n/(n-2)}(\Omega)})_i$ is bounded, and up to a subsequences, we can assume that $\eta_i \rightharpoonup \eta$ in the sense of measures as $i \to \infty$. Namely, for any continuous function $\phi$ with compact support in $\Omega $ \[ \lim_{i \to \infty} \int_\Omega \phi \,d\eta_i = \int_ \Omega \phi \,d\eta. \] Fatou's Lemma then implies that we can decompose \[ \eta = (\frac{1}{2} { \vert{\nabla u} \vert}^2 + \frac{n-2}{2n} {\vert{u}\vert}^{2n \over n-2}) \,dx+\nu \] where $\nu $ is a nonnegative Radon measure. Moreover, we prove that $\nu$ satisfies the following lemma. \begin{lemma}\label{lem1.2} Let $\delta>0$ such that $B_\delta\subset \Omega$. Then we have \begin{itemize} \item[(i)] $\Sigma\subset spt(\nu)$ \item[(ii)] There exists a measurable, upper-semi-continuous function $\Theta$ such that \[ \nu (x)= \Theta (x) \mathcal{H}^{0} \lfloor \Sigma ,\quad \text{for } x \in \Sigma. \] \end{itemize} Moreover, there exists some constants $c$ and $C >0$ (only depending on $n$ and $\Omega$) such that \[ c \varepsilon_0 < \Theta (x) < C \quad \mathcal{H}^{0}-\mbox{a.e. in } \Sigma \] where ${H}^{0}\lfloor \Sigma$ is the restriction to $\Sigma$ of the Hausdorff measure and $\Theta$ is a measurable function on $\Sigma$. \end{lemma} The main question we would like to address in the present paper concerns the multiplicity $\Theta$ of the defect measure which has been defined above. More precisely, we have proved the following theorem. \begin{theorem} Let $\nu$ be the defect measure of the sequence $(|\nabla{u_{i}}|^{2}+|u_{i}|^{2n/(n-2)})dx$ defined above. Then $\nu$ is quantized. That is, for a.e $x\in \Sigma$, \begin{equation} \Theta(x)=\sum_{j=1}^{j=N_{x}}{{\|\nabla{v_{x,j}}\|^2_{{L}^{2}(\Omega)}+ \|{v_{x,j}}\|^{2n/(n-2)}_{{{L}^{2n/(n-2)}}(\Omega)}}} \end{equation} where $N_{x}$ is a positive integer and where the functions $v_{x,j}$ are solutions of $\Delta v + v^{\frac{n+2}{n-2}} =0$ which are defined on $\mathbb{R}^n$, issued from $(u_{i'})$ and that concentrate at $x$ as $i\to \infty$. \label{thm1.3} \end{theorem} The sentence ``issued from $(u_{i'})$ and that concentrate at $x$ as $i \to \infty $'' means that there are sequences of conformal maps ${\psi}_{j}^{i}$ , a finite family of balls $(B^{l}_{i,j})_{l}$ such that the pulled back function \[ {\tilde{u}_{i,j}} = ({\psi}_{j}^{i})^* u_{i'} \] satisfies \begin{gather*} {\tilde{u}_{i,j}} \to v_j \quad \mbox{strongly in} \quad \mathbf{L}^{2}({\mathbb R^n} \setminus {\cup}_{l}{B^{l}_{i,j})}),\\ \nabla{{\tilde{u}_{i,j}}} \to \nabla{v_j} \quad \mbox{strongly in} \quad \mathbf{L}^{2}({\mathbb R^n} \setminus {\cup}_{l} {B^{l}_{i,j})} ) \end{gather*} In the context of Yang-Mills fields in dimension $n \geq 4$ a similar concentration result has been proven by Rivi\`ere \cite{T.R}. More precisely, Rivi\`ere has shown that, if $(A_i)_i$ is a sequence of Yang-Mills connections such that $( \Vert {\nabla}_{A}{\nabla}_{A} F_{A} \Vert_{\mathbf{L}^{1}(B_{1}^n) } )_i$ is bounded, then the corresponding defect measure $\nu = \Theta \mathcal{H}^{n-4} \lfloor{\Sigma}$ of a sequence of smooth Yang-Mills connections is quantized. The proof of Theorem \ref{thm1.3} uses technics introduced by Lin and Rivi\`ere in their study of Ginzburg-Landau vortices \cite{L.R} and also the technics developed by Rivi\`ere in \cite{Z.G}. These technics use as an essential tool the Lorentz spaces, more specifically the $\mathbf{L}^{2,\infty}$-$\mathbf{L}^{2,1}$ duality \cite{L.T}. \ This paper is organized in the following way: In Section 2 we establish first a monotonicity formula for smooth solutions of problem \eqref{eq1.1} which allows us to prove an $\varepsilon$-regularity Theorem. Then, we prove Theorem \ref{thm1.1} and Lemma \ref{lem1.2}. While Section 3 is devoted to the proof of our main result, Theorem \ref{thm1.3}. \section{A monotonicity Inequality} In this section, we establish a monotonicity formula for smooth solutions of problem \eqref{eq1.1}. Using Pohozaev identity: Multiplying \eqref{eq1.1} by $x_{i} \frac{\partial{u}}{\partial{x_{i}}}$ (summation over i is understood) and integrating over B(x,r), the ball centered at $x$ of radius r, we obtain \[ -\int_{B(x,r)}{x_{i}\frac{\partial{u}}{\partial{x_{i}}}}{\Delta{u}}\,dy =-\int_{B(x,r)}{x_{i}\frac{\partial{u}}{\partial{x_{i}}}}u |{u}|^{4/(n-2)}\,dy \] By Green formula, we get \begin{equation} \begin{aligned} &\frac{n-2}{2} \int_{B(x,r)}{|u|^{2n/(n-2)}}dy -\frac{n-2}{2} \int_{B(x,r)}{|\nabla{u}|^{2}}dy\\ &-\frac{n-2}{2n} \int_{\partial{B(x,r)}}{|u|^{2n/(n-2)}}\,ds + \frac{1}{2} r \int_{\partial{B(x,r)}}{|\nabla{u}|^{2}}ds \\ &= r \int_{\partial{B(x,r)}}{|\frac{\partial{u}}{\partial{r}}|^{2}}dy \end{aligned} \label{eq2.1} \end{equation} On the other hand, multiplying \eqref{eq1.1} by $u$ and integrating over $B(x,r)$, we get \begin{equation} \int_{B(x,r)}{|\nabla{u}|^{2}}dy -\int_{\partial{B(x,r)}}{u\frac{\partial{u}}{\partial{r}}}ds =\int_{B(x,r)}{|u|^{2n/(n-2)}}dy \label{eq2.2} \end{equation} Deriving (\ref{eq2.2}) with respect to $r$, we obtain \begin{equation}\int_{\partial{B(x,r)}}{|\nabla{u}|^{2}}dy -\frac{d}{dr} \int_{\partial{B(x,r)}}{u\frac{\partial{u}}{\partial{r}}}ds =\int_{\partial{B(x,r)}}{|u|^{2n/(n-2)}}dy \label{eq2.3} \end{equation} Combining \eqref{eq2.1}, \eqref{eq2.2} and \eqref{eq2.3}, we get \begin{align*} &-\frac{r}{n} \int_{\partial B(x,r)}|u|^{2n/(n-2)}\,ds \\ &= \frac{1}{2}r \frac{d}{dr}\int_{{\partial}B(x,r)} u\frac{{\partial}u}{{\partial}r}\,ds - r\int_{{\partial}B(x,r)} |\frac{{\partial}u}{{\partial}r}|^{2}\,dy + r^{-1} u \frac{{\partial}u}{{\partial}r}\,ds . \end{align*} Moreover, we have that \begin{align*} \frac{d^2}{dr^2} (\int_{{\partial}B(x,r)}u^2 \,ds ) &=\frac{d}{dr}( 2\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}\,ds + \frac{n-1}{r}\int_{{\partial}B(x,r)}u^2\,ds ) \\ &=(n-1)\Big[\frac{2}{r}\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}\,ds +(\frac{n-1}{r^2}-\frac{1}{r^2})\int_{{\partial}B(x,r)}u^2\,ds \Big]\\ &\quad +2\frac{d}{dr}\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}\,ds\\ &=\frac{n-1}{r}\Big[2\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}\,ds +\frac{n-2}{r}\int_{{\partial}B(x,r)}u^2 \,ds\Big]\\ &\quad + 2\frac{d}{dr}\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}\,ds. \end{align*} Hence \begin{align*} &\frac{1}{n} \frac{d}{dr}\int_{B(x,r)}|u|^{2n/(n-2)}\,dy +\frac{1}{n} \frac{d^2}{dr^2} \int_{{\partial}B(x,r)}u^2 \,ds\\ &= \int_{{\partial}B(x,r)}( |\frac{{\partial}u}{{\partial}r}|^2 + \frac{2n-3}{2r} u \frac{{\partial}u}{{\partial}r} + \frac{(n-1)(n-2)}{4}r^{-2}u^{2})\,ds. \end{align*} Moreover \begin{align*} &\frac{d}{dr}( \frac{1}{r}\int_{{\partial}B(x,r)}u^2 ds )\\ &=-\frac{1}{r^2}\int_{{\partial}B(x,r)}u^2 ds + \frac{2}{r}\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}ds+ \frac{n-1}{r^2}\int_{{\partial}B(x,r)}u^2 ds \\ &= \frac{n-2}{r^2}\int_{{\partial}B(x,r)}u^2 ds + \frac{2}{r}\int_{{\partial}B(x,r)}u\frac{{\partial}u}{{\partial}r}ds. \end{align*} We obtain \begin{align*} &\frac{d}{dr}\Big[ \frac{1}{n}\int_{B(x,r)}|u|^{2n/(n-2)}dy +\frac{1}{n}\frac{d}{dr}\int_{{\partial}B(x,r)}u^2 ds -\frac{1}{n}\frac{1}{r}\int_{{\partial}B(x,r)}u^2 ds \Big]\\ &=\int_{{\partial}B(x,r)}(|\frac{{\partial}u}{{\partial}r}|^2+ (n-2) r^{-1}u|\frac{{\partial}u}{{\partial}r}| +\frac{(n-2)^2}{4}r^{-2}u^{2})ds\\ &=\int_{{\partial}B(x,r)}(\frac{{\partial}u}{{\partial}r} +\frac{n-2}{2}r^{-1}u)^2 ds \geq 0 \end{align*} We conclude that \begin{equation} E_u (x,r) = \frac{1}{n}\int_{B(x,r)}|u|^{2n/(n-2)}dy + \frac{1}{n}\frac{d}{dr}\int_{B(x,r)}u^2 ds + \frac{1}{n}r^{-1}\int_{B(x,r)}u^2 ds \label{eq2.4} \end{equation} is a nondecreasing function of $r$. Using the fact that \[ \int_{B(x,r)}|u|^{2n/(n-2)}dy - \int_{{\partial}B(x,r)}|\nabla u|^2 dy = - \int_{{\partial}B(x,r)} u\frac{{\partial}u}{{\partial}r}ds, \] one can easily get \begin{align*} E_u(x,r) &= \frac{1}{n}\int_{B(x,r)}|u|^{2n/(n-2)}dy + \frac{1}{4}\frac{d}{dr}\int_{{\partial}B(x,r)}u^2 ds - \frac{1}{4}r^{-1}\int_{{\partial}B(x,r)}u^2 ds\\ &= \frac{\frac{n}{2}}{n}\int_{B(x,r)}|u|^{2n/(n-2)}dy +\frac{1-\frac{n}{2}}{n}\int_{B(x,r)}|u|^{2n/(n-2)}dy\\ &\quad +\frac{1}{4}\frac{d}{dr}\int_{{\partial}B(x,r)}u^2 ds -\frac{1}{4}r^{-1}\int_{{\partial}B(x,r)}u^2 ds \\ &= \frac{1}{2}\int_{B(x,r)}|{\nabla}u|^2 dy - \frac{1}{2}\int_{{\partial}B(x,r)} u\frac{{\partial}u}{{\partial}r}ds - \frac{n-2}{2n}\int_{B(x,r)}|u|^{2n/(n-2)}dy \\ &\quad +\frac{1}{4}\frac{d}{dr}\int_{{\partial}B(x,r)}u^2 ds -\frac{1}{4}r^{-1}\int_{{\partial}B(x,r)}u^2 ds \\ &= \frac{1}{2}\int_{B(x,r)} ( |{\nabla}u|^2 - \frac{n-2}{2n}|u|^{2n/(n-2)} ) dy + \frac{1}{4} \frac{d}{dr}\int_{{\partial}B(x,r)}u^2 \,ds \\ &\quad -\frac{1}{4}r^{-1}\int_{{\partial}B(x,r)}u^2 ds - \frac{1}{2}\int_{{\partial}B(x,r)} u\frac{{\partial}u}{{\partial}r} ds . \end{align*} We obtain an equivalent formulation of $ E_u(x,r)$ \begin{equation} E_{u}(x,r)=\frac{1}{2}\int_{B(x,r)} ( |{\nabla}u|^2 -\frac{n-2}{2n}|u|^{2n/(n-2)} dy +\frac{n-2}{4}r^{-1}\int_{{\partial}{B(x,r)}}u^2 ds \label{eq2.5} \end{equation} Moreover, using the fact that \[ \frac{d}{dr}\int_{\partial{B(x,r)}}{u^{2}}ds =2\int_{\partial{B(x,r)}}{u\frac{\partial{u}}{\partial{r}}}ds +\frac{n-1}{r}\int_{\partial{B(x,r)}}{u^2} \] we obtain \begin{align*} \frac{1}{r}\int_{\partial{B(x,r)}}{u^2}\,ds &=\frac{1}{n-1}\frac{d}{dr}\int_{\partial{B(x,r)}}{u^{2}}\,ds -\frac{2}{n-1}\int_{\partial{B(x,r)}}{u\frac{\partial{u}}{\partial{r}}}\,ds\\ &=\frac{1}{n-1}\frac{d}{dr}\int_{\partial{B(x,r)}}{u^{2}}\,ds\\ &\quad +\frac{2}{n-1}\Big[\int_{B(x,r)}{|u|^{2n/(n-2)}}\,dy -\int_{B(x,r)}{|\nabla{u}|^{2}}\,dy\Big] \end{align*} Then $E_{u}(x,r)$ can also be written \begin{align*} &E_{u}(x,r)\\ &=\frac{1}{2(n-1)}\int_{B(x,r)}(|\nabla{u}|^{2} +\frac{n-2}{n}|u|^{2n/(n-2)})\,dy +\frac{n-2}{4(n-1)}\frac{d}{dr}\int_{\partial{B(x,r)}}{u^{2}}\,ds. \end{align*} \begin{proof}[Proof of Lemma \ref{lem1.1}] To prove that $(x,r) \mapsto E_{u}(x,r)$ is continuous it suffices to prove that \[ (x,r) \mapsto\int_{\partial{B(x,r)}}{u^{2}}ds \] is continuous with respect to $x$ and $r$. We have \[ \int_{\partial{B(x,r)}}{u{\frac{\partial{u}}{\partial{r}}}}ds = \int_{B(x,r)}{|\nabla{u}|^{2}}- \int_{B(x,r)}{|u|^{2n/(n-2)}}dy \] Thus $(x,r)\mapsto{\int_{\partial{B(x,r)}}}{u{\frac{\partial{u}}{\partial{r}}}}$ is continuous, and this allows to get the conclusion. Now, to prove that $E_u$ is positive, we proceed by contradiction. If the result is not true, then there would exists $x \in \Omega $ and $R > 0 $ such that $E_{u}(x,R)< 0$. For almost every $y$ in some neighborhood of $x$, we have \[ \lim_{r\to 0}{\int_{\partial{B(x,r)}}u{\frac{\partial{u}}{\partial{r}}}}\,ds =0 \] integrating $E_{u}(x,r)$ over the interval $[0,R]$ and using the fact that r $\mapsto{E_{u}(x,r)} $ is increasing, we obtain \begin{align*} {\int_{0}}^{R}{E_{u}(y,r)}dr &=\frac{1}{2(n-1)}{\int_{0}}^{R}{dr \int_{B(y,r)}}(|\nabla{u}|^{2}+\frac{n-2}{2n}{|u|^{2n/(n-2)}})dx\\ &\quad + \frac{n-2}{4(n-1)}\int_{\partial{B(y,R)}}{u^{2}}\,ds \\ &\leq {R E_{u}(y,R)} < 0 \end{align*} which is not possible. This proves Lemma \ref{lem1.1}. \end{proof} \begin{lemma} There exist $r_{0}>0$ and some constant $c>0 $, depending only on $n$, such that \[ \int_{B(x,r)}(|\nabla{u}|^{2}+{|u|^{2n/(n-2)}})\,dy < c E_{u}(x,r) \] for any r $ 0 $ such that $E_{u}(x_{0},r_{0})\leq \varepsilon $ then \[ \int_{B(x,r)}(|\nabla{u}|^{2}+\frac{n-2}{n}|u|^{2n/(n-2)})\,dy\leq C \varepsilon \quad \forall \quad 0 0$ such that \[ \liminf_{i \to \infty} E_{u_i}(x_0,r_1) < \varepsilon_0. \] Then, we may find a sequence $n_j\to \infty$ as $j \to \infty$ such that \[ \sup_{n_j}E_{u_{n_j}}(x_0,r_1) < \varepsilon_0. \] We deduce from the $\varepsilon$-regularity Theorem (Theorem \ref{thm1.1}) that \[ \sup_{n_j}\sup_{x\in B_{\frac{r_1}{16}}(x_0)}|u_{n_j}|\leq \frac{C}{r_1^{(n-2)/2}}. \] for some constant $C$ depending only on $n$. Then \[ u_{n_j}\to u \quad\text{in } C^1(B_{\frac{r_1}{16}}(x_0)) \] a similar argument allows to show that \[ \nabla u_{n_j}\to \nabla u \quad\text{in } C^1(B_{\frac{r_1}{16}}(x_0)) \] Then \[ \mu_{n_j}:=\left(\frac{1}{2} |\nabla u_{n_j}|^2+ \frac{n-2}{2n} u_{n_j}^{2n/(n-2)}\right)\,dx\to \left(\frac{1}{2} |\nabla u|^2+\frac{n-2}{2n} u^{2n/(n-2)}\right)\,dx \] as radon measure. Hence $\nu = 0$ on $B_{\frac{r_1}{16}}(x_0)$ i.e $x_0 \notin \mathop{\rm supp}(\nu)$ and then we deduce that $\mathop{\rm supp}(\nu)\subset \Sigma$. To show (ii), let us first recall some properties of the function $E_u(x,r)$ that has been defined above: \noindent $\bullet$ For all $x \in \Omega$, there exists $r_0 > 0$ and a constant $ C>0$ such that \[\int_{B(x,r)}(\frac{1}{2} |\nabla u|^2+ \frac{n-2}{2n}|u|^{2n/(n-2)}) < C E_u(x,r_0) \quad \forall r<\frac{r_0}{2}. \] This is explained in the proof of Lemma \ref{lem1.1}. \noindent$\bullet$ Using the fact that $E_u(x,.)$ is increasing on $r$ together with the fact that \[ \lim_{r\searrow 0} E_u(x,r)=0 \quad \mathcal{H}^{0}-\text{a.e. } x\in\Omega \] we deduce that for $\mathcal{H}^{0}$-a.e. $x\in\Sigma$, $\lim_{r \searrow 0} \int_{B(x,r)} \nu$ exists. and the density $\Theta (\eta, .)$ defined by \begin{equation} \Theta (\eta ,x):= \lim_{r\searrow 0} \eta(B_r(x)) \end{equation} exists for every $x\in \Omega$. Moreover, for $\mathcal{H}^{0}$-a.e. $x\in\Omega$, $\Theta_u(x)=0$, where \begin{equation} \Theta_u(x):= \lim_{r \searrow 0} \int_{B(x,r)}(\frac{1}{2} {\vert \nabla u \vert}^2 + \frac{n-2}{2n} {\vert u \vert}^{2n \over {n-2}}) \,dy. \end{equation} Now, for $r$ sufficiently small and $i$ sufficiently large \begin{equation} \int_{B(x,r)} \frac{1}{2}|\nabla u_i |^2+\frac{n-2}{2n} u_i^{2n/(n-2)} \leq C E_{u_i}(x,r) \leq C(\Lambda,\Omega) \label{eq4.13} \end{equation} where $\Lambda$ is given above and $ C(\Lambda,\Omega)$ is a constant depending only on $\Lambda $ and $\Omega$. Hence \begin{equation} \eta(B(x,r)) \leq C(\Lambda,\Omega) \quad\text{for } x \in B_1^n \label{eq4.14} \end{equation} In particular, this implies that $\eta \lfloor \Sigma $ is absolutely continuous with respect to $\mathcal{H}^{0}\lfloor \Sigma $. Applying Radon-Nikodym's Theorem \cite{E.G}, we conclude that \begin{equation} \eta \lfloor \Sigma = \Theta (x) \mathcal{H}^{0} \lfloor \Sigma \quad \text{for } \mathcal{H}^{0}\text{-a.e. } x \in \Sigma \end{equation} Using \ref{eq4.13} we conclude that \begin{equation} \nu (x) = \Theta (x) \mathcal{H}^{0}\lfloor \Sigma \end{equation} for a $\mathcal{H}^{0}$-a.e. $x \in \Sigma$ (recall that $\eta =(\frac{1}{2}{\vert \nabla u \vert }^2 +\frac{n-2}{2n} {\vert u \vert}^{2n \over {n-2}})\,dx + \nu $ and $\mathop{\rm supp}( \nu) \subset \Sigma $). The estimate on $\Theta$ follows from \ref{eq4.14}. \end{proof} For any $ y\in B^n_1$ and any sufficiently small $\lambda > 0$, we define the scaled measure $\eta_{y,\lambda}$ by \begin{equation} \eta_{y,\lambda}(x) := \eta (y + \lambda x ) \end{equation} We have the following lemma. \begin{lemma} \label{lem4.1} Assume that $(\lambda_j)_j$ satisfies $\lim_{j\to \infty}\lambda_j=0$. Then, there exist a subsequence $({{\lambda}_{j'}} )_{j'}$ and a Radon measure $\chi$ defined on $\Omega$, such that ${\eta_{y,\lambda_{j'}}} \rightharpoonup \chi $ in the sense of measures. \end{lemma} \begin{proof} For each $i \in {\mathbb N}$, we define the scaled function $u_{i,y,\lambda}$ by \begin{equation} u_{i,y,\lambda}(x):={\lambda}^{{n-2 \over 2}} u_i(\lambda x + y)\quad\text{for } y \in B^n_1. \end{equation} Then $u_{i,y,\lambda}$ is a solution of \[ -\Delta{u}= u|u|^{4/(n-2)} \quad \text{on } B^n_1. \] In addition, for any $r > 0$ sufficiently small, we have \begin{equation} \begin{aligned} &\int_{B_r(0)}\left(\frac{1}{2}{\vert \nabla u_{i,y,\lambda} \vert}^2 +\frac{n-2}{2n} {\vert u_{i,y,\lambda}\vert}^{{2k \over {k-2}}}\right) \,dx \\ &= \int_{B_{\lambda r}(y)} \left(\frac{1}{2} {\vert \nabla u_i \vert}^2 + \frac{n-2}{2n} {\vert u_i\vert}^{{2n \over {n-2}}}\right) \,dx \leq C(\Lambda , \Omega). \end{aligned} \label{eq419} \end{equation} Finally for fixed $\lambda$, \begin{align*} &\left(\frac{1}{2} {\vert \nabla u_{i,y,\lambda} \vert}^2 + \frac{n-2}{2n} {\vert u_{i,y,\lambda}\vert}^{2n/(n-2)} \right) (x) \,dx \\ &={\lambda}^n \left(\frac{1}{2} {\vert \nabla u_i \vert}^2 - \frac{n-2}{2n} {\vert u_i \vert}^{2n/(n-2)}\right)(\lambda x +y)\,dx \\ & \rightharpoonup \eta(\lambda x +y) = \eta_{y,\lambda}(x) \end{align*} in the sense of measures as $i\to \infty$. On the other hand letting $i$ tends to infinity in (\ref{eq419}), we conclude that for any $r>0$ \begin{equation} \eta_{y , \lambda} (B_r(0)) \leq C(\Omega,\Lambda). \end{equation} Hence, we may find a subsequence $\{\lambda'_j \}$ of $\{ \lambda_j\}$ and a Radon measure $\chi$ such that $\eta_{y , \lambda'_j}$ converge weakly to $\chi$ as Radon measure on $\Omega$. Then \[ \lim_{j\to \infty} \lim_{i\to \infty} \left( \frac{1}{2} {\vert \nabla u_{i,y,\lambda'_j} \vert}^2 +\frac{n-2}{2n} {\vert u_{i,y,\lambda'_j} \vert}^{2n \over {n-2}}\right)\,dx = \lim_{j\to \infty} \eta_{y,\lambda'_j}(x) = \chi \] Using a diagonal subsequence argument, we may find a subsequence $i_j \to\infty$, such that \[ \lim_{j\to \infty} \left(\frac{1}{2} {\vert \nabla u_{i_j,y,\lambda'_j} \vert}^2 +\frac{n-2}{2n} {\vert u_{i_j,y,\lambda'_j} \vert}^{2n \over {n-2}}\right)dx = \chi \] This proves the Lemma. \end{proof} \begin{remark} \label{rmk1}\rm Observe that \[ \chi(B_r(0))= \lim_{j\to \infty } \eta_{y , \lambda'_j}(B_r(0))=\lim_{j\to \infty }\eta (B_{\lambda'_j r}(y))=\Theta(\eta, y) \] In particular, we deduce that $\chi(B_r(0))$ is independent of r. \end{remark} \section{Proof of Theorem \ref{thm1.3}} The idea of the proof comes from Rivi\`ere \cite{T.R} in the context of Yang-Mills Fields. To simplify notation and since the result is local, we assume that $\Omega$ is the unit ball $B^n$ of $\mathbb{R}^n$. Let $(u_{k}) $ be a sequence of smooth solutions of \eqref{eq1.1} such that \[ \Big( \|u_{k}\|_{\textbf{H}^{1}(\Omega)} +\|{u_{k}}\|_{\textbf{L}^{2n/(n-2)}(\Omega)} \Big) \] is bounded and let $\nu$ be the defect measure defined above. We claim that for $\delta>0$, we have \begin{equation} \lim_{k\to\infty} \sup_{y\in{B_{1}}(x_{0})}\int_{B_\delta(y_{0})} \left(|u_{k}|^{2n/(n-2)}+|\nabla{u_{k}}|^{2}\right)\geq{\varepsilon(n)} \label{eq3.1} \end{equation} where $\varepsilon (n)$ is given by Theorem \ref{thm1.3}. Indeed if (\ref{eq3.1}) would not hold, we have for $\delta>0$ and $k\in\mathbb{N}$ large enough \[ \sup_{y\in{B_{1}}(x_{0})}\int_{B_\delta(y_{0})}\left(|u_{k}|^{2n/(n-2)}+|\nabla{u_{k}}|^{2}\right)\leq{\varepsilon(n)} \] and by Theorem \ref{thm1.1} we have \[ \|\nabla{u_{k}}\|_{\textbf{L}^{\infty}(B_\frac{\delta}{2}(y))} \leq C(\epsilon)/ r^{n/2} \] This contradict the concentration phenomenon and the claim is proved. We then conclude that there exists sequences $\delta_{k}\to 0$ as $k\to\infty$ and $ (y_{k})\subset{B_{1}(x_{0})}$ such that \begin{equation} \begin{aligned} \int_{B_{\delta_{k}}(y_{0})}\left(|u_{k}|^{2n/(n-2)}+|\nabla{u_{k}}|^{2}\right)dx &= \sup_{y\in{B_{1}}(x_{0})}\int_{B_{\delta_{k}(y_{0})}}\left(|u_{k}|^{2n/(n-2)} +|\nabla{u_{k}}|^{2}\right)\,dx\\ &= \frac{\varepsilon(n)}{2}. \label{eq3.2} \end{aligned} \end{equation} In other words, $y_{k}$ is located at a bubble of characteristic size $\delta_{k}$. More precisely, if one introduces the function \[ {\widetilde{u}}_{k}(x)=\delta_{k}^{(n-2)/2} u_{k}(\delta_{k}x+y_{k}); \] we have, up to a subsequence, that \begin{gather*} {\widetilde{u}}_{k} \to {u_{\infty}}\quad \text{in }{\textbf{C}}^{\infty}_{\rm loc}(\mathbb{R}^n) \quad \text{as }k \to \infty ,\\ \nabla{{\widetilde{u}}_{k}}\to{\nabla{u_{\infty}}} \quad\text{in } {\textbf{C}}^{\infty}_{\rm loc}(\mathbb{R}^n)\quad \text{as } k \to \infty\,. \end{gather*} Therefore, \[-\Delta{u_{\infty}}=u_{\infty} |{u_{\infty}}|^{4/(n-2)} \quad \text{in } \mathbb{R}^n. \] This is the first bubble we detect. On the other hand, we have clearly that \begin{equation} \int_{{\mathbb{R}}^n}\left(|u_{\infty}|^{2n/(n-2)}+|\nabla{u_{\infty}}|^{2}\right)\,dx = \lim_{R\to\infty} \lim_{k\to\infty}\int_{B_{R{\delta_{k}}}(y_{k})} \left(|u_{k}|^{2n/(n-2)}+|\nabla{u_{k}}|^{2}\right)\,dx. \label{eq3.3} \end{equation} Indeed: \begin{align*} &\lim_{R\to\infty} \lim_{k\to\infty}\int_{B_{R{\delta_{k}}}(y_{k})} \left(|u_{k}|^{2n/(n-2)}+|\nabla{u_{k}}|^{2}\right) \,dx \\ &= \lim_{R\to\infty} \lim_{k\to\infty}\int_{B_{R}(0)} \left(|u_{k}|^{2n/(n-2)}+|\nabla({u_{k})}|^{2}\right)(\delta_{k}x+y_{k})\, \delta_{k}^n\,dx \\ &= \lim_{R\to\infty} \lim_{k\to\infty}\int_{B_{R}(0)} \left(|{\delta_{k}}^{\frac{2-n}{2}}{\widetilde{u}_{k}(x)}|^{2n/(n-2)}+ |{\delta_{k}}^{\frac{2-n}{2}}{\delta_{k}}^{-1}\nabla{\widetilde{u}_{k}(x)}|^{2}\right) \delta_{k}^n\,dx \\ &= \lim_{R\to\infty} \lim_{k\to\infty}\int_{B_{R}(0)} \left(|{\widetilde{u}_{k}(x)}|^{2n/(n-2)}+|\nabla{\widetilde{u}_{k}(x)}|^{2}\right)\,dx \\ &= \lim_{R\to\infty}\int_{B_{R}(0)}\left(|u_{\infty}(x)|^{2n/(n-2)}+|\nabla{u}_{\infty}(x)|^{2}\right)\,dx \\ &=\int_{{\mathbb{R}}^n}\left(|u_{\infty}(x)|^{2n/(n-2)} +|\nabla{u}_{\infty}(x)|^{2}\right)\,dx\,. \end{align*} Assume first that we have only one bubble of characteristic $\delta_{k}$. We have shown that \begin{equation} \Theta=\lim_{k\to\infty}\int_{B_{1}^n(0)}\left(|\nabla{u_{k}}|^{2} +|u_{k}|^{2n/(n-2)}\right)\,dx =\int_{\mathbb{R}^n}\left(|\nabla{u_{\infty}}|^{2}+|u_{\infty}|^{2n/(n-2)}\right)\,dx, \label{eq5.24} \end{equation} where $\Theta$ is defined above. It suffices to prove that \begin{equation} \lim_{R\to\infty} \lim_{k\to\infty}\int_{B_{1}^n(0) \setminus{B_{R\delta_{k}(y_{k})}}} \left(|{u}_{k}(x)|^{2n/(n-2)}+|\nabla{{u}_{k}(x)}|^{2}\right) \,dx = 0\,. \label{eq3.4} \end{equation} In other words there is no ``neck'' of energy which is quantized. To simplify notation, we assume that $y_{k} = 0$. We claim that for any $\varepsilon > 0$ small enough, there exists $R> 0$ and $k_{0}\in \mathbb{N}$ such that for any $k\geq{k_{0}}$ and $R\delta_{k}\leq{r}\leq{\frac{1}{2}}$, we have \begin{equation} \int_{B_{2r}^n(0)\setminus{B_{r(0)}}}\left(|{u}_{k}(x)|^{2n/(n-2)} +|\nabla{{u}_{k}(x)}|^{2}\right)\,dx \leq\varepsilon \label{eq3.5} \end{equation} Indeed, if is not the case, we may find $\varepsilon_{0}>0$, a subsequence $k'\to\infty$ (Still denoted $k$ ) and a sequence $r_{k}$ such that \begin{equation} \begin{gathered} \int_{B_{2r}^n(0)\setminus{B_{r}(0)}}\left(|{u}_{k}(x)|^{2n/(n-2)} +|\nabla{{u}_{k}(x)}|^{2}\right)\,dx >\varepsilon_0, \\ \frac{r_k}{\delta_k}\to\infty \quad \text{as }\quad k\to\infty \end{gathered} \label{eq3.6} \end{equation} Let $\alpha_k\to0$ such that $r_k/\alpha_k=o(1)$ and $\alpha_k r_k/\delta_k \to\infty$ and let \[ v_k(x)=r_k^{(n-2)/2} u_k(r_k x) \] clearly $v_k $ satisfies \[ -\Delta{v_k}=v_k |v_k|^{4/(n-2)} \quad\text{in } B_{2\alpha_k} \setminus{B_{\alpha_k}} \] Therefore, \[ \int_{B_{2}^n(0)\setminus{B_{1}(0)}}\left(|{v}_{k}(x)|^{2n/(n-2)} +|\nabla{{v}_{k}(x)}|^{2}\right)\,dx >\varepsilon(n) \] and then we have a second bubble. This contradict our assumption. We deduce from (\ref{eq3.6}) and Theorem \ref{thm1.1} that for any $\varepsilon <\varepsilon(n)$, there exist $R>0$ and $ k_0 \in \mathbb{N}$ such that for all $k\geq{k_0}$ and $|x|\geq{R\delta_k}$ \[ |\nabla{u_k}|(x)\leq C(\epsilon) /|x|^{n/2} \] where $C(\varepsilon)\to 0$ as $\varepsilon\to0$. Then \begin{equation} |\nabla{u_k}|^{2}(x)\leq C(\varepsilon)/|x|^n . \label{eq3.7} \end{equation} We define $E_\lambda^k$ by \[ E_\lambda^k = \mathop{\rm meas} \left\{x \in{\mathbb{R}^n} : |\nabla{u_k}|(x)\geq\lambda\right\} \] We have $ E_\lambda^k \leq C(\varepsilon)/ \lambda ^2$; indeed \[ \left\{x\in{\mathbb{R}^n} : |\nabla{u}_k|(x)\geq\lambda\right\}\subset\{x\in{\mathbb{R}^n} : |x|^n\leq{\frac{C(\varepsilon)}{{\lambda}^{2}}}\} \] and \[ \mathop{\rm meas} \left\{x\in{\mathbb{R}^n} : |x|^n\leq{\frac{C(\varepsilon)}{{\lambda}^{2}}}\right\} \leq{\frac{C(\varepsilon)}{{\lambda}^2}} \] We deduce from (\ref{eq3.7}) that \begin{equation} \|\nabla{u_k}\|_{{\textbf{L}}^{2,\infty}(C_{B_{R{\delta}_k}})}\leq C(\varepsilon) \label{eq3.8} \end{equation} where ${{\textbf{L}}^{2,\infty}}$ is the Lorentz space defined in \cite{L.T}, the weak ${\textbf{L}}^{2}$ space, and $\|\cdot\|_{{\textbf{L}}^{2,\infty}}$ is the weak norm defined by \[ \|f\|_{{\textbf{L}}^{2,\infty}}= \sup_{0< t <\infty} {t^{1/2}f^*(t)} \] where $f^*$ is the nonincreasing rearrangement of $|f|$. Indeed \[ \|\nabla{u_k}\|_{{\textbf{L}}^{2,\infty}(C_{B_{R{\delta}_k}})} = \sup_{0< t <\infty} {t^{1/2}(\nabla{u_k})^*(t)} \] by definition, \[ (\nabla{u_k})^*(t)=inf\{\lambda>0 / E^k_\lambda \leq t \} \] For all $t >0$ such that $\frac{C(\varepsilon)}{\lambda^2}\leq t$, we have $E^k_\lambda \leq t$. Then \begin{align*} \inf \left\{\lambda>0: E^k_\lambda \leq t \right\} &\leq \inf\left\{\lambda>0 : \frac{C(\varepsilon)}{\lambda^2} \leq t \right\}\\ &\leq \inf\left\{\lambda>0 : \lambda\geq \frac{(C(\varepsilon))^{1/2}}{t^{1/2}} \right\}\\ &=\frac{(C(\varepsilon))^{1/2}}{t^{1/2}} \end{align*} Hence $t^{1/2}(\nabla{u_k})^*(t) \leq C(\varepsilon)$ and so \begin{equation} \|\nabla{u_k}\|_{{\textbf{L}}^{2,\infty}(C_{B_{R{\delta}_k}})}\leq C(\varepsilon) \label{eq4.19} \end{equation} We claim that the sequence $(\nabla{u_k})$ is uniformly bounded in the Lorentz space ${\textbf{L}}^{2,1}(B^n_1)$ (see \cite{L.T} for the definition). We prove this claim using an iteration proceeding; Indeed, the sequence $(u_k)$ is bounded in ${\textbf{L}}^\frac{2n}{n-2}(B_1^n)$. Then \[ \Delta{u_k}= - u_k |u_k|^{4/(n-2)} \] is bounded in ${\textbf{L}}^\frac{2n}{n+2}(B_1^n)$ which implies by the elliptic regularity Theorem that the sequence $(u_k)$ is bounded in ${\textbf{W}}^{2,\frac{2n}{n+2}}(B_1^n)$. Using the imbedding Theorem for Sobolev spaces \[ {\textbf{W}}^{m,p}(B_1^n)\subset {\textbf{W}}^{r,s}(B_1^n) \quad \text{if } m \geq r,\; p \geq s \text{ and } m-\frac{n}{p}=r-\frac{n}{s}. \] In particular, ${\textbf{W}}^{2,\frac{2n}{n+2}}(B_1^n)$ is continuously imbedded in ${\textbf{W}}^{1,2}(B_1^n)$. On the other hand by Proposition 4 in \cite{L.T}, we have \[ {\textbf{W}}^{1,2}(B_1^n)\hookrightarrow {\textbf{L}}^{2^*,2}(B_1^n) = {\textbf{L}}^{\frac{2n}{n-2},2}(B_1^n) \] continuously. We then deduce that \[ \Delta{u_k}= - u_k |u_k|^{4/(n-2)} \] is bounded in ${\textbf{L}}^{\frac{2n}{n+2},\frac{2(n-2)}{n+2}}(B_1^n)$. Here, we have used the following lemma. \begin{lemma} If $f \in{\textbf{L}}^{p,q}(B_1^n)$ and $\alpha\in\mathbb{Q^+}$, then $f^\alpha\in{\textbf{L}}^{\frac{p}{\alpha},\frac{q}{\alpha}}(B_1^n)$. \label{lem5.1} \end{lemma} \begin{proof} In the case where $\alpha\in\mathbb{N}$, the result follows from the fact that \[ f \in{\textbf{L}}^{a,b}(B_1^n) \text{ and } g \in{\textbf{L}}^{c,d}(B_1^n)\Rightarrow f.g \in{\textbf{L}}^{q,r}(B_1^n), \] where $\frac{1}{q}=\frac{1}{a}+\frac{1}{b}$ and $\frac{1}{r}=\frac{1}{c}+\frac{1}{b}$ (see \cite{B.W}). The general case is a consequence of the fact that the increasing rearrangement of the function $|f|^\beta$ is equal to the puissance $\beta$ of the increasing rearrangement of $|f|$ since $(f^\beta)^*$ is the only one function verifying \[ \mathop{\rm meas} \{x\in{\mathbb{R}^n} : f^{\beta}(x) \geq\lambda \} = \mathop{\rm meas} \{t>0 : (f^{\beta})^*(x) \geq\lambda \} \] This in turns proves Lemma \ref{lem5.1}. \end{proof} Now, using in \cite[Theorem 8]{L.T}, we deduce from (\ref{eq3.6}) that $(\nabla{u_k})$ is uniformly bounded in the space ${\textbf{L}}^{(\frac{2n}{n+2})^*,\frac{2(n-2)}{n+2}}(B_1^n)$ = ${\textbf{L}}^{2,\frac{2(n-2)}{n+2}}(B_1^n)$. Hence $(u_k)$ is bounded in ${\textbf{L}}^{2^*,\frac{2(n-2)}{n+2}}(B_1^n)$. Then \[ \Delta{u_k}= - u_k |u_k|^{4/(n-2)} \] is bounded in ${\textbf{L}}^{\frac{2n}{n+2},\frac{2(n-2)^2}{(n+2)^2}}(B_1^n)$. Hence, again by \cite[Theorem 8]{L.T}, the sequence $(\nabla{u_k})$ is bounded in ${\textbf{L}}^{2,\frac{2(n-2)^2}{(n+2)^2}}(B_1^n)$ and by elliptic regularity Theorem \[ \Delta{u_k}= - u_k |u_k|^{4/(n-2)} \] is bounded in ${\textbf{L}}^{\frac{2n}{n+2},\frac{2(n-2)^3}{(n+2)^3}}(B_1^n)$. We obtain after $p$ iterations that \[ \Delta{u_k}= - u_k |u_k|^{4/(n-2)} \] is bounded in ${\textbf{L}}^{\frac{2n}{n+2},\frac{2(n-2)^p}{(n+2)^p}}(B_1^n)$. We choose $p >0$ such that $6p >n$, we have in particular $\frac{2(n-2)^p}{(n+2)^p}<1 $ which gives \[ \Delta{u_k}= - u_k |u_k|^{4/(n-2)} \] is bounded in ${\textbf{L}}^{\frac{2n}{n+2},1}(B_1^n)$. Here we have used the fact that \[ {\textbf{L}}^{p, q_1}(B_1^n)\subset{\textbf{L}}^{p, q_2}(B_1^n) \quad\text{if }q_1 < q_2 \] We use also \cite[Theorem 8]{L.T} to deduce that $(\nabla{u_k})$ is bounded in ${\textbf{L}}^{(\frac{2n}{n+2})^*,1}(B_1^n)={\textbf{L}}^{2,1}(B_1^n)$. In particular, there exist a constant $C > 0 $ depending only on $n$ such that \begin{equation} \| \nabla {u_k}\|_{\textbf{L}^{2,1} (B_1^n)}\leq C \label{eq4.20} \end{equation} We deduce from (\ref{eq4.19}), (\ref{eq4.20}) together with the ${\textbf{L}}^{2,1}-{\textbf{L}}^{2,\infty}$ duality that \[ \|\nabla{u_k}\|_{{\textbf{L}}^{2}(B_1^n\setminus{B_{R\delta_k}})} \leq \|\nabla{u_k}\|_{{\textbf{L}}^{2,1}(B_1^n\setminus{B_{R\delta_k}})} \|\nabla{u_k}\|_{{\textbf{L}}^{2,\infty}(B_1^n\setminus{B_{R\delta_k}})}\\ \leq C(\epsilon) \] for a constant $C(\varepsilon)\to 0$ as $\varepsilon\to 0$. Now, we use the embedding $\textbf{H}^1\hookrightarrow{\textbf{L}}^{2n/(n-2)}$ continuously, we obtain \begin{align*} \|u_k\|_{{\textbf{L}}^{2n/(n-2)}(B_1^n\setminus{B_{R\delta_k}})} &\leq C \|\nabla{u_k}\|_{{\textbf{L}}^{2}(B_1^n\setminus{B_{R\delta_k}})}\\ &\leq C(\varepsilon)\to 0 \quad\text{as } \varepsilon\to 0 . \end{align*} We deduce that \[ \lim_{R\to\infty} \lim_{k\to\infty} \int_{B_{1}^n(0)\setminus{B_{R\delta_{k}(y_{k})}}}(|{u}_{k}|^{2n/(n-2)} +|\nabla{{u}_{k}}|^{2})(x)\,dx = 0 \] This proves Theorem \ref{thm1.3} in the case of one bubble. The case of more than one bubble can be handled in a very similar way and we just give few details for $m = 2$. The proof starts the same until (\ref{eq5.24}) which cannot hold any more otherwise we would have had one bubble only as it is (\ref{eq5.24}) holds. It remains to show that: for any $\varepsilon \ge 0 $, there are sufficiently large R $> 0 $ and a sequence $r_i \to 0 $ such that for any $ R {\delta}_i \le r_i \le 1/2$, \begin{equation} \begin{gathered} \lim_{R \to \infty} \lim_{i \to \infty} {\int_{\{0\} \times{ B^n_{r_i} \setminus {B^n_{R {\delta}_i}}(0)}}}({ { \frac{1}{2} \vert{\nabla{v_i}} \vert}^{2} +\frac{n-2}{2n} {\vert {v_i} \vert }^{2n/(n-2)} })\,dx = 0\,, \\ \lim_{i \to \infty} {\int_{\{0\}\times{ B^n_{1/2}\setminus{B^n_{r_i}(0)}}} ( \frac{1}{2} {\vert {\nabla{v_i}} \vert}^{2} +\frac{n-2}{2n} {\vert {v_i} \vert } ^{ 2n/(n-2)} )\,dx= 0} \end{gathered} \label{eq5.31} \end{equation} where $v_i$ is defined by $v_{i}(y) = {r_i}^{(n-2)/2} $ $u_{i} (r_i y)$ , $y \in {\mathbb{R}}^n $. The proof of (\ref{eq5.31}) can be done exactly as the proof of (\ref{eq5.24}), the case of 2 bubbles is then proved. To prove the general case, for any number $m\ge 2$, one can follow exactly the same strategy. \begin{thebibliography}{00} \bibitem{A} W. Allard, \emph{An integrity Theorem a regularity Theorem for surfaces whose first variation with respect to a parametric elliptic integrand is controlled}, Proc. Symp. Pure Math., 44, (1986), 1-28. \bibitem{B.W} H. Brezis and S. Wainger, \emph{A note on limiting cases of Sobolev embedding and convolution inequalities}, Comm. P.D.E, 5, (1980), 773-789. \bibitem{D.T} S. K. Donaldson and R. P. Thomas, \emph{Gauge theory in higher dimensions} in The geometric Universe (Oxford, 1996), Oxford Univ. Press, 1998, 31-47. \bibitem{E.G} L.C. Evans and R.F. Gariepy, \emph{Measure Theory and Fine Properties of Functions}, Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1992. \bibitem{Z.G} Z. Guo and Jiayu-Li, \emph{The blow-up locus of semilinear elliptic equations with subcritical exponent}, Calc. Var. Partial Differential Equations 15 (2002), no. 2, 133-153. \bibitem{F.P} F. Pacard, \emph{Partial regulatity for weak solutions of a nonlinear elliptic equation}, Manuscripta Math. 79 (1993), 161-172. \bibitem{T.P} T. Parker, \emph{Bubble tree convergence for harmonic maps}, J. Diff. Geom. , 44, (1996), 545-633. \bibitem{J.P} J. Peetre, \emph{Espaces d'interpolations et th\'eor\`eme de Sobolev}, Ann. Instit. Fourier , Grenoble, 16, (1966), 279-317. \bibitem{Lin} F. G. Lin, \emph{Gradient estimates and blow-up analysis for stationary harmonic maps}, Ann. Math. , 149, (1999), 785-829. \bibitem{L.R} F. G. Lin and T. Rivi\`ere, \emph{A Quantization Property for Static Ginzburg-Landau vortices}, Comm. Pure Appl. Math. 54 (2001), no. 2, 206--228. \bibitem{T.R} T. Rivi\`ere, \emph{Interpolation Spaces and Energy Quantization for Yang-Mills Fields}, Comm. Anal. Geom. 10 (2002), no. 4, 683--708. \bibitem{Sch} R. Schoen, \emph{Analytic aspects for the harmonic map problem}, Math. Sci. Res. Insti. Publi. 2, Springer, Berlin (1984), 312-358. \bibitem{Sim} L. Simon, \emph{Lectures on Geometric Measure Theory}, Proc. of Math. Anal.3, Australian National Univ. (1983). \bibitem{L.T} L. Tartar, \emph{Imbedding Theorems of Sobolev Spaces into Lorentz Spaces}, Boll. U.M.I. 1, B, (1998) 479-500. \end{thebibliography} \end{document}