\documentclass[reqno]{amsart} \usepackage{amssymb} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2004(2004), No. 90, pp. 1--24.\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 2004 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2004/90\hfil Partial compactness] {Partial compactness for the 2-D Landau-Lifshitz flow} \author[P. Harpes\hfil EJDE-2004/90\hfilneg] {Paul Harpes} \address{Paul Harpes \hfill\break ETH Z\"urich \\ R\"amistrasse 101, 8092 Z\"urich, Switzerland} \email{pharpes@math.ethz.ch} \date{} \thanks{Submitted September 11, 2003. Published July 5, 2004.} \subjclass[2000]{35B65, 35B45, 35D05, 35D10, 35K45, 35K50, 35K55} \keywords{Partial compactness; partial regularity; Landau-Lifshitz flow; \hfill\break\indent a priori estimates; harmonic map flow; non-linear parabolic; Struwe-solution; approximations} \begin{abstract} Uniform local $C^\infty$-bounds for Ginzburg-Landau type approximations for the Landau-Lifshitz flow on planar domains are proven. They hold outside an energy-concentration set of locally finite parabolic Hausdorff-dimension 2, which has finite times-slices. The approximations subconverge to a global weak solution of the Landau-Lifshitz flow, which is smooth away from the energy concentration set. The same results hold for sequences of global smooth solutions of the 2-d Landau-Lifshitz flow. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemm}[theorem]{Lemma} \newtheorem{prop}[theorem]{Proposition} \newtheorem{coro}[theorem]{Corollary} \newtheorem{remark}[theorem]{Remark} \def\abs#1{|#1|} \def\norm#1{\|#1\|} \section{Introduction} The Ginzburg-Landau approximations $u_\epsilon: \overline{\Omega} \times \mathbb{R}_+ \to \mathbb{R}^3$ to the Landau-Lifshitz flow are solutions of \begin{gather} \label{Eq_eps_LL} \gamma_1 {\partial}_t u_\epsilon - \gamma_2 u_\epsilon \times {\partial}_t u_\epsilon - \Delta u_\epsilon = - \frac{1}{\epsilon^2} f(u_\epsilon) \quad \mbox{ in } \Omega \times \mathbb{R}_+ \\ \label{Eq_bdry_eps_LL} u_\epsilon = u_0 \quad \mbox{ on } \bigl( \Omega \times \{ 0 \} \bigr) \cup \bigl( \partial \Omega \times \mathbb{R}_+ \bigr) \,. \end{gather} where $ \gamma_1 > 0 $ and $\gamma_2 \in \mathbb{R}$. "$\times$" denotes the usual vector product in $\mathbb{R}^3$. The domain $ \Omega \subset \mathbb{R}^2$ is open, bounded and smooth. The initial and boundary data $u_0$ is always assumed to map a.e. into the standard sphere $S^2 \subset \mathbb{R}^3$ or an embedded manifold $N \subset \mathbb{R}^n$ (see below). For the definition of $f_\epsilon$ we distinguish two cases: \noindent{\bf Case (I):} If $\gamma_2 \neq 0$, the target is $S^2 \hookrightarrow \mathbb{R}^3$ and the right hand side is given by $$ f(u_\epsilon) := - (1-\abs{u_\epsilon}^2) u_\epsilon = \frac{1}{4} \frac{d}{du} \bigl( 1-|u_\epsilon|^2\bigr)^2 \,. $$ For small $\epsilon > 0$ the maps $u_\epsilon$ then approximate the Landau-Lifshitz flow \begin{equation} \label{LL} \gamma_1 \partial_t u - \gamma_2 u \times \partial_tu - \Delta u = \abs{\nabla u}^2 u \quad \mbox{ in } \Omega \times \mathbb{R}_+ \,. \end{equation} For sufficiently regular solutions $u:\overline{\Omega}\times\mathbb{R} \to S^2$ equation \eqref{LL} is equivalent to \begin{equation} \label{ LL-original} \partial_tu = - \alpha u \times (u\times\Delta u) + \beta u \times \Delta u \mbox{ in } \Omega \times \mathbb{R}_+ \,, \end{equation} where $ \alpha := \frac{\gamma_1}{\gamma_1^2+\gamma_2^2} > 0$ and $ \beta := \frac{\gamma_2}{\gamma_1^2+\gamma_2^2} \in \mathbb{R}$. This is the usual form of the Landau-Lifshitz equations known in physics. (Compare \cite{LL} and \cite{Guo-Hong1}, \cite{Harpes0}, \cite{Harpes1}.) \noindent{\bf Case (II):} If $\gamma_2 = 0$, the target is a smooth, closed, isometrically embedded manifold $N \hookrightarrow \mathbb{R}^n$. For small $\epsilon >0$, the map $u_\epsilon: \overline{\Omega} \times\mathbb{R}_+ \to \mathbb{R}^n$ will then be an approximation of a harmonic map flow (compare \cite{Struwe-Chen}) and is defined to be a solution of \begin{gather}\label{Eq_eps_HMF} {\partial}_t u_\epsilon - \Delta u_\epsilon = - \frac{1}{2\epsilon^2} \frac{d}{du}\chi \bigl(\operatorname{dist}^2(u_\epsilon,N)\bigr) \quad \mbox{ in } \Omega \times \mathbb{R}_+ \\ \label{Eq_bdry_eps_HMF} u_\epsilon = u_0 \quad \mbox{ on } \bigl( \Omega \times \{ 0 \} \bigr) \cup \bigl( \partial\Omega \times \mathbb{R}_+ \bigr) \,, \end{gather} That is, for the function $f(u_\epsilon)$ in \eqref{Eq_eps_LL}, we choose $$ f(u_\epsilon):= \frac{1}{2} \frac{d}{du}\chi \bigl(\operatorname{dist}^2(u_\epsilon,N)\bigr) \,, $$ The cut-off function $ \chi:\mathbb{R}_+ \to \mathbb{R}_+$ is smooth, non decreasing and satisfies $ \chi (t) = t $ for $ 0 \leq t \leq \delta_N^2 $ and $ \chi (t) \equiv 2 \delta_N^2 $ for $ t \geq 4 \delta_N^2$. The parameter $\delta_N >0$ is chosen in such a way that the nearest neighbour projection $ U \ni x \mapsto \pi_N(x) \in N$ is defined and smooth on a tubular neighborhood $ U \subset \mathbb{R}^n $ of $N$ with uniform radius radius $ 2 \delta_N >0 $. (Such a $\delta_N > 0$ always exists if $N$ is closed. Compare \cite{Struwe-Chen} and \cite[Section 2.12.3 p.42]{Noebi}.) For fixed $\epsilon >0$, smooth solutions of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL} or (\ref{Eq_eps_HMF})-(\ref{Eq_bdry_eps_HMF}) on $ \Omega \times \mathbb{R}_+ $ exist and if $u_0 \in H^{1,2}(\Omega;N) \cap H^{3/2,2}(\partial\Omega;N)$, they are unique in $$ H^{1,2}_{\rm loc}\cap L^\infty(H^{1,2}) := H^{1,2}_{\rm loc}( \overline{\Omega} \times \mathbb{R}_+;\mathbb{R}^n)\cap L^\infty(\mathbb{R}_+;H^{1,2}(\Omega;\mathbb{R}^n))\,. $$ (Compare \cite{Chen1},\cite{Struwe-Chen}.) Existence is obtained by Galerkin's method, regularity ($C^\infty$) follows from a standard bootstrap argument and uniqueness may be proven as for the two dimensional harmonic map flow (see \cite{Struwe1} or \cite{Struwe2} (5${}^\circ$) p.234 in the proof of Theorem 6.6). The total energy of the flow at time $t\geq 0$ is defined by \begin{equation} \label{def-en} G_\epsilon \bigl(u_\epsilon (t)\bigr) := \int_\Omega g_\epsilon (u_\epsilon)(x,t) dx \end{equation} where \begin{gather*} \label{g_epsilon} g_\epsilon (u_\epsilon) := \frac{1}{2} \abs{\nabla u_\epsilon}^2 +\frac{1}{4 \epsilon^2} (1-\abs{u_\epsilon}^2)^2 \quad \mbox{if } \gamma_2 \neq 0 \,,\\ g_\epsilon (u_\epsilon) := \frac{1}{2} \abs{\nabla u_\epsilon}^2 +\frac{1}{2\epsilon^2} \chi\bigl( \operatorname{dist}^2(u_\epsilon,N) \bigr) \quad \mbox{if } \gamma_2 = 0 \,. \end{gather*} While the total energy of the "$\epsilon$-approximations" always decreases (see Lemma \ref{en-est-LL} below), the local energy given by \begin{equation} \label{def-loc-en } G_\epsilon \bigl( u_\epsilon(t)\,, B^\Omega_R(x_0) \bigr) := \int_{B_R(x_0)\cap \Omega} g_\epsilon (u_\epsilon)(x,t) \, dx \,. \end{equation} may concentrate at space-time points $(x_0,t_0)$ as $\epsilon \searrow 0 $ either for fixed \mbox{$t=t_0$} or for variable $t \nearrow t_0$ or $t \searrow t_0$. It characterizes the local "asymptotic regularity behaviour" of the flow. Here asymptotic refers to the limit $\epsilon \searrow 0$. We will show that all the derivatives of the family of maps $\{ u_\epsilon \}_{\epsilon>0}$ are locally uniformly bounded on a regular set $\mathop{\rm Reg}\bigl(\{u_\epsilon\}_{\epsilon>0} \bigr)$ consisting of all points $$ z_0 = (x_0,t_0) \in\overline{\Omega} \times ]0,\infty[ $$ for which there is $R_0 = R_0(z_0) >0 $, such that \begin{equation}\label{reg-cond} \limsup_{\epsilon \searrow 0} \sup_{t_0 - R_0^2 < t < t_0 } G_\epsilon \bigl( u_\epsilon(t)\,, B^\Omega_{R_0}(x_0) \bigr) < \epsilon_0 \,, \end{equation} for a constant $\epsilon_0 >0$ that will be determined later. The complement $$ \mathbb{S}\bigl(\{u_\epsilon\}_{\epsilon>0} \bigr) := \bigl(\overline{\Omega} \times \mathbb{R}_+\bigr) \smallsetminus \mathop{\rm Reg} \bigr( \{u_\epsilon\}_{\epsilon>0} \bigl) $$ is referred to as the energy-concentration set. % We will show that It is closed, has locally finite parabolic Hausdorff dimension two and finite slices at fixed time. The limits of converging subsequences $ \{u_{\epsilon_j}\}_j$ are distributional solutions of the Landau-Lifshitz flow (or harmonic map flow if $\gamma_2 =0$) on all $ \Omega \times ]0,\infty[$ in $H^{1,2}_{\rm loc}\cap L^\infty(H^{1,2})$. Bubbling phenomena of the $\epsilon$-approximations as $\epsilon \searrow 0 $ either for fixed \mbox{$t=t_0$} or for variable $t \nearrow t_0$ or $t \searrow t_0$ as described in \cite{Harpes0} will be presented in \cite{Harpes3}. Strong subconvergence of the harmonic map flow penalty-approximations in $$ W^{1,0}_{2,\rm loc}\Bigl(\mathop{\rm Reg}(\{u_{\epsilon}\}_{\epsilon>0}); \mathbb{R}^n \Bigr) $$ to a global distributional $ H^{1,2}_{\rm loc}\cap L^\infty(H^{1,2})$-solution of the harmonic map flow was already proved by M. Struwe and Y. Chen in \cite{Struwe-Chen} for the case of a closed domain manifold $\Omega =M$ with dim$M = m \geq 2 $ or $ M = \mathbb{R}^m $. ($W^{1,0}_{2,\rm loc}$ refers to functions $f$, whose restriction to any closed ball (in space-time) lies in $L^2$ as well as the restriction of the space-gradient $\nabla f$.) Struwe and Chen provided uniform local $L^\infty$-bounds for $g_\epsilon (u_\epsilon)$ on $\mathop{\rm Reg}(\{u_{\epsilon}\}_{\epsilon>0})$. Their result was extended to compact domains with boundary by Chen and Lin in \cite {Chen-Lin1}. The energy-concentration set $ \mathbb{S}(\{u_\epsilon\}_{\epsilon >0})$ is known to have locally finite $m (=\dim M)$-dimensional Hausdorff measure in the case of the harmonic map flow (see \cite{Struwe-Chen}). X. Cheng investigates in \cite{Cheng} weak(*) $H^{1,2}_{\rm loc}\cap L^\infty (H^{1,2})$-limits $u_*$ of sequences of smooth solutions of the harmonic map flow on the domain $ M = \mathbb{R}^m $ and shows that the time slice $\mathbb{S}(\{u_k\}_k) \cap \bigl(\mathbb{R}^m \times \{ t \}\bigr)$ has finite $(m-2)$-dimensional Hausdorff measure. Weak(*)-subconvergence in $H^{1,2}_{\rm loc}\cap L^\infty(H^{1,2})$ of the Landau-Lifshitz $\epsilon$-approx\-imations from closed surfaces to a distributional $ H^{1,2}_{\rm loc}\cap L^\infty(H^{1,2})$-solution of the Landau-Lifshitz flow was proven by B. Guo and M.C. Hong in \cite{Guo-Hong1}. Guo and Ding also studied partial convergence of the two dimensional Landau-Lifshitz penalty-approximations in \cite{Ding-Guo1},\cite{Ding-Guo2} and \cite{Ding-Guo3}. Their arguments however contain several gaps and inconsistencies. \section{Energy-estimates} In the case $\gamma_2 = 0$, equation \eqref{Eq_eps_HMF} is the $L^2$-gradient flow of the functional $ u \mapsto G_\epsilon(u)$. \eqref{Eq_eps_LL} is not known to be a gradient flow, but the total energy still decreases along the (smooth) flow \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL}. \begin{lemm} \label{en-est-LL} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL}. Then \begin{gather} \label{en-est} G_\epsilon \bigl(u_\epsilon(T)\bigr) + \gamma_1\int_0^T \int_\Omega \abs{\partial_t u_\epsilon }^2 dx dt = G_\epsilon \bigl(u_\epsilon(0)\bigr) = E(u_0)=: E_0 \,, \\ \label{loc-en} G_\epsilon \bigl(u_\epsilon(T_2), B_R^\Omega(x_0)\bigr) \leq G_\epsilon \bigl(u_\epsilon(T_1), B_{2R}^\Omega(x_0)\bigr) + \frac{C}{\gamma_1 R^2} \int_{T_1}^{T_2} G_\epsilon \bigl(u_\epsilon(t), B_{2R}^\Omega(x_0)\bigr) dt, \end{gather} for $0\leq T_1 0 $, there exist $ T_0 > 0 $ and $ R_0 > 0 $, such that for all $ x_0 \in \Omega $ and all $ \epsilon > 0 $ we have \begin{equation} \label{gs} \sup_{0 \leq t \leq T_0} G_\epsilon \bigl(u_\epsilon(t), B_{R_0}^\Omega(x_0)\bigr) \leq \eta \,. \end{equation} \end{lemm} \begin{proof} Inequality \eqref{en-est} is obtained by multiplying \eqref{Eq_eps_LL} with $ \partial_t u_\epsilon $. Inequality \eqref{loc-en} follows by multiplying \eqref{Eq_eps_LL} with $ \partial_t u_\epsilon \phi^2 $ for an adequate cut-off function $\phi$ and then integrating by parts and absorbing. Note that $ \partial_t u_\epsilon \equiv 0 $ on $ \partial\Omega \times \mathbb{R}_+ $. Inequality (\ref{gs}) follows from \eqref{loc-en}, if we set $T_1 =0$ and $T_2 = T_0 = \frac{\gamma_1 R_0^2}{2 C E_0}$ for sufficiently small $R_0>0$, such that $ G_\epsilon \bigl(u_\epsilon(0), B_{R_0}^\Omega(x_0)\bigr) = E\bigl( u_0, B_{R_0}^\Omega(x_0)\bigr) < \eta / 2$. \end{proof} The energy estimates imply the penalty-approximations subconverge weak(*) in $H^{1,2}_{\rm loc} (\overline{\Omega}\times \mathbb{R}_+;\mathbb{R}^n)$ $\cap L^\infty \bigl(\mathbb{R}_+; H^{1,2}(\Omega; \mathbb{R}^n)\bigr)$. This was already pointed out by B.Guo and M.C.Hong in section 4 of \cite{Guo-Hong1}. \section{Partial compactness} \label{sec-high} In this section we show that, under the uniform smallness condition (\ref{reg-cond}) on the local energy, all higher derivatives of $u_\epsilon$ are locally and uniformly bounded. Here ``uniform'' of course always means uniform in $\epsilon >0$. In Section \ref{subsec-standard}, estimates for linear parabolic systems that can be applied to \eqref{Eq_eps_LL} as soon as $\nabla u_\epsilon$ is locally bounded are recalled. In Section \ref{subsec-sup} we show that $\nabla u_\epsilon$ is necessarily locally uniformly bounded, whenever (\ref{reg-cond}) holds. In Section \ref{subsec-high} we derive estimates that will provide bounds for the right hand side of \eqref{Eq_eps_LL} and allow to combine the previous estimates into a bootstrap argument. \subsection{Some ``standard'' parabolic estimates} \label{subsec-standard} Equation \eqref{Eq_eps_LL} may be written as \begin{equation} \label{para-LL} L_\epsilon (u_\epsilon) := \partial_t u_\epsilon - M(u_\epsilon) \Delta u_\epsilon = - \frac{1}{\epsilon^2}M(u_\epsilon) f (u_\epsilon) = f_\epsilon (u_\epsilon) \,. \end{equation} The coefficient-matrix $ M(u) $ is smooth with respect to $ u $ and also strictly elliptic: $$ \frac{\gamma_1}{\gamma_1^2 + \gamma_1^2} \abs{\xi}^2 < \xi^T M(u) \xi = \frac{1}{\gamma_1 (\gamma_1^2 + \gamma_2^2 \abs{u}^2)} \Bigl( \gamma_1^2 \abs{\xi}^2 + \gamma_2^2 (u\cdot\xi)^2\Bigr) < \frac{1}{\gamma_1} \abs{\xi}^2 , $$ for all $\xi \in \mathbb{R}^3 $ (See \cite[p.12]{Ding-Guo2}, \cite[p.316]{Guo-Hong1}, \cite[p.37]{Ding-Guo1}). Note that for $\gamma_2=0$, we obtain $M(u) = \frac{1}{\gamma_1} Id$. The results of this section are indeed merely interesting in the case $\gamma_2 \neq 0$, where the left hand side of \eqref{Eq_eps_LL} is non-linear. We will therefore restrict ourselves to the case $\gamma_2 \neq 0$. For fixed $\epsilon >0$, the solution $ u_\epsilon $ of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL} is smooth and in particular continuous. $u_\epsilon$ has the same regularity up to the boundary as the boundary data $u_0$. The family of solutions $\{u_\epsilon\}_{\epsilon >0}$ is also uniformly bounded in $\epsilon > 0$, since \mbox{$\abs{u_\epsilon (x,t)} \leq 1$} $ \forall x,t $. This follows from the Maximum Principle applied to the equation obtained by multiplying \eqref{Eq_eps_LL} with $(1-\abs{u_\epsilon})$. $L_\epsilon$ defines a strongly parabolic system in the sense of Petrovskii (Definition 2, p.599 in \cite{LadySolUral}) but satisfies as well all the other (not necessarily equivalent) definitions of strong parabolicity for general linear parabolic systems (Definitions 3-6) in \cite{LadySolUral}. The boundary-data operators also fullfill the required conditions. First we have estimates in the $W^{2,1}_p$-Sobolev spaces with $p > 1$ (see \cite{LadySolUral} Chapter IV, Theorem 9.10 p.342 and (10.12) p.355 but also Chapter VII, Theorem 10.4 p.621, for the generalization to parabolic systems). Let $f_\epsilon \in L^p(\Omega\times [0,T];\mathbb{R}^n)$ and $u_0 \in H^{2,p}(\Omega;\mathbb{R}^n)$. Then for any $ \delta \in ]0,1[ $, $p> 3/2$ and for $t_0 - R^2>0$ a solution of $$ L_\epsilon(v) = f_\epsilon \mbox{ in } \Omega \times ]0,T[ \quad \mbox{and}\quad v = u_0 \mbox{ on } \bigl( \Omega \times \{ 0 \} \bigr) \cup \bigl( \partial \Omega \times ]0,T[ \bigr) $$ satisfies \begin{gather}\label{star1} \norm{v}_{W^{2,1}_p(\Omega \times [0,T])} \leq C_p(\Omega, T, \omega_{u_\epsilon}) \bigr(\norm{f_\epsilon}_{L^p(\Omega \times [0,T])} + \norm{u_0}_{H^{2,p}(\Omega)}\bigl) ,\\ \label{star2} \begin{aligned} \norm{v}_{ W^{2,1}_p (P_{\delta R}^\Omega(z_0)) } & \leq \tilde{C}_p(R, \delta, \Omega, \omega_{u_\epsilon}) \Bigr(\norm{f_\epsilon}_{L^p(P_R^{\Omega}(z_0))} +\norm{v}_{L^q(P_R^{\Omega}(z_0))} \\ &\quad + \delta_{B_R\cap\partial\Omega} \norm{u_0}_{H^{2-(1/p),p}(B_R^\Omega\cap\partial\Omega(z_0))} \Bigl) , \end{aligned} \end{gather} with $1\leq q \leq p$. Here $$ P_R^{\Omega}(z_0) := (B_R(x_0) \times ]t_0 -R^2, t_0[) \cap (\Omega \times ]0,\infty[) $$ and $\delta_{B_R\cap\partial\Omega} = 1 $ if $ B_R\cap\partial\Omega \neq \emptyset $ and $0$ otherwise. The trace theorems of course imply $$\norm{u_0}_{H^{2-(1/p),p}(\partial\Omega)} \leq \norm{u_0}_{H^{2,p}(\Omega)}\,.$$ The constants $ C_p $ and $ \tilde{C_p} $ depend on the indicated quantities and additionally on the uniform lower and upper bounds for the eigenvalues of $M(u_\epsilon)$, which may be chosen independent of $ \epsilon > 0 $. Note that the constants $ C_p, \tilde{C_p} $ also depend on the moduli of continuity of the coefficients of the leading term, i.e. the modulus of continuity $\omega_{u_\epsilon}$ of $ u_\epsilon $. The equation can also be written in divergence form, $$ L_{\epsilon} (v) := \partial_t v - div \bigl( M(u_\epsilon) \nabla v \bigr) + \bigl(DM(u_\epsilon) \partial_k u_\epsilon\bigr) \partial_k v = f_\epsilon \,. $$ If we assume in addition \begin{equation} \label{ass-sup} \limsup_{\epsilon\searrow 0} \sup_{P_R^\Omega} \abs{\nabla u_\epsilon} < \infty \,, \end{equation} then estimates for equations in divergence form imply $ v \in C^{\gamma,(\gamma/2)}(P_{\delta R}^\Omega;\mathbb{R}^n)$ for some $ \gamma \in ]0,1[ $ and any $\delta \in ]0,1[$. (See \cite{LadySolUral} Chapter VII Theorem 3.1 p.582 or Chapter V, Theorem 1.1, p.419.) Indeed if the right hand side $f_\epsilon \in L^p(P_R^\Omega;\mathbb{R}^n)$ with $p>2$, the following estimate for the mixed H\"older-norm of $v$ on $P_{\delta R}^\Omega$ holds \begin{equation} \label{cont-est} \norm{v}_{C^{\gamma,\gamma/2}(P_{\delta R}^\Omega)} \leq C(f_\epsilon) \,. \end{equation} (See \cite{LadySolUral} p.7 for the definition of the mixed H\"older-spaces denoted there by $H^{\gamma,\gamma/2}$) The bound $C(f_\epsilon)$ depends on the parabolicity constants, on $0<\delta<1$, $\sup_{P_R^\Omega} \abs{ u_\epsilon}$, $\norm{f_\epsilon}_{L^p(P_R^\Omega)}$, bounds for the coefficients of the equations depending on $ \sup_{P_R^\Omega} \abs{\nabla u_\epsilon} $ and also on $ \norm{u_0}_{C^{\gamma}(B_R\cap\partial\Omega)}$ if $ B_R\cap\partial\Omega \neq \emptyset $. Therefore, if (\ref{ass-sup}) holds and $\norm{f_\epsilon}_{L^p(P_R^\Omega)}$ or $ \sup_{P_R^\Omega} \abs{f_\epsilon} $ are uniformly bounded with respect to $\epsilon > 0 $, then estimate (\ref{cont-est}) holds for $u_\epsilon$ and is uniform in $\epsilon >0$. Now the modulus of continuity of $ u_\epsilon $ on $P_{\delta R}^\Omega$ is bounded from above (by an increasing function $ h $ with $ \lim_{t\searrow 0} h(t) = 0 $) independently of $ \epsilon > 0 $. We gain uniform bounds for the modulus of continuity of $ u_\epsilon $ with respect to $t\geq 0$ and estimate $(\ref{star2})$ is now uniform in $ \epsilon > 0 $. Further by Lemma 3.3 p.80 in Chapter II of \cite{LadySolUral} for $p> m+2 (= 4)$ ($m$ being the dimension of the spatial domain, in our case $m=2$), we have $$ \norm{\nabla v}_{C^\lambda(P_R^\Omega)} \leq C(m,p, \lambda, \Omega ) \norm{v}_{W^{2,1}_p(P_R^\Omega)} \quad \mbox{for } \lambda = 1 - (m+2)/p \,. $$ Also if (\ref{ass-sup}) holds and $\norm{f_\epsilon}_{L^p(P_R^\Omega)}$ is uniformly bounded, then $(\ref{star2})$ yields $\epsilon$-uniform estimates for $\norm{\nabla u_\epsilon}_{C^\lambda(P_{\delta R}^\Omega)}$. \subsection{The main sup-estimates for the energy-density} \label{subsec-sup} \subsubsection{An interior sup-estimate for the Landau-Lifshitz-flow approximations} \label{subsubsec-sup} In this section we derive an interior sup-estimate for the energy density in the case $\gamma_2 \neq 0$, but the proof also works if $\gamma_2 =0$ and the target is $N$. The proof of the interior estimate is much simpler than in the boundary case and we therefore consider each case separately. The estimate will result from a scaling argument combined to the following higher estimates, that will be proven in the next section. Let $$ P_R(z_0) := B_R(x_0) \times ]t_0 -R^2, t_0[ \quad \mbox{for } z_0=(x_0,t_0)\,. $$ \begin{lemm} \label{Cinfty1} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL} for each $ \epsilon > 0 $. Assume $$ \limsup_{\epsilon \searrow 0} \sup_{P_R(z_0)} g_{\epsilon} (u_\epsilon) \leq C_0 $$ and $ B_R(x_0) \subset \Omega $, $ 0 0 $. If $ \gamma_2 =0$ and the target is $N$, they also depend on the geometry of $N$ (i.e. the metric on $N$ and its derivatives). \end{lemm} We will now prove the following ``$\epsilon_1$-regularity'' result. \begin{theorem} \label{eps-reg-LL} There are constants $ C_1 = C_1(N), \epsilon_1 = \epsilon_1(N) > 0 $, such that if, for some $0 2 $. Since $ B_{R_0}(x_0) \subset \Omega $, all the higher derivatives of $ v_\epsilon$ are then bounded on $P_{1} $ independently of $ \epsilon > 0 $. Indeed if $ \liminf_{\epsilon\searrow 0} \sqrt{e_\epsilon} \epsilon > 0 $, the uniform estimates are immediate and if $ \liminf_{\epsilon\searrow 0} \sqrt{e_\epsilon} \epsilon = 0 $, they follow from Lemma \ref{Cinfty1}. In particular $$ \sqrt{\abs{\partial_t g_{\tilde{\epsilon}}(v_\epsilon)}}, \abs{\nabla g_{\tilde{\epsilon}}(v_\epsilon)} \leq C < \infty \mbox{ on } P_1 \mbox{ (uniformly in $ \epsilon > 0 $)} $$ and therefore, $$ \inf_{ P_{r_0} }g_{\tilde{\epsilon}}(v_\epsilon) \geq \frac{1}{2} \quad \mbox{ for } r_0 := \min \{ \frac{1}{4C},1 \} \,. $$ Note that $ C $ is an absolute constant in the sense that it merely depends on the radius $2$, the factor $ \delta = \frac{1}{2}$, the $L^\infty$-bound $4$ and the parabolicity constants and the geometry of $N$. This lower bound implies \begin{align*} 1 &= g_{\sqrt{e_\epsilon} \epsilon}(v_\epsilon)(0,0) \leq \frac{2}{\pi r_0^2} \sup_{-r_0^2 < s < 0 } \int_{B_{r_0}} g_{\sqrt{e_\epsilon} \epsilon}(v_\epsilon)(y,s) \, dy \\ &\leq C_* \sup_{t_\epsilon - r_0^2 e_\epsilon^{-1} < t < t_\epsilon } \int_{B_{e_\epsilon^{-1/2}r_0}(x_\epsilon)} g_{ \epsilon}(u_\epsilon)(x,t) \, dx \\ &\leq C_*\sup_{-(\frac{r_0^2}{e_\epsilon} + \sigma_\epsilon^2) < t < 0 } \int_{B_{\frac{r_0}{\sqrt{e_\epsilon}} + \sigma_\epsilon }(x_0)} g_{ \epsilon}(u_\epsilon)(x,t) \, dx\,. \end{align*} Set $ \epsilon_1 := \min \{\frac{1}{2}, \frac{1}{2C_*} \}$. Since $ r_\epsilon = \sqrt{e_\epsilon} \rho_\epsilon > 2 > r_0 $, we have $ \frac{r_0}{\sqrt{e_\epsilon}} + \sigma_\epsilon \leq \rho_\epsilon + \sigma_\epsilon \leq R_0 $ and $ (\frac{r_0}{\sqrt{e_\epsilon}})^2 + \sigma_\epsilon^2 \leq (\rho_\epsilon + \sigma_\epsilon)^2 \leq R_0^2 $. Then the last estimate %, which gives an upper bound for 1, yields a contradiction, since the right hand side is smaller than $\epsilon_1 \leq \frac{1}{2}$. Therefore $r_\epsilon = \sqrt{e_\epsilon} \rho_\epsilon \leq 2 $ and $$ (1-\delta)^2 R_0^2 \sup_{P_{\delta R_0}} g_\epsilon \leq 16 \,. $$ \end{proof} \subsubsection{ A local boundary sup-estimate for the energy density} \label{subsec-sup-bd} Local $L^p$-estimates for $\nabla^3 u_\epsilon$ up to the boundary which are uniform in $\epsilon > 0$ cannot be expected, even if $u_0 \in C^\infty(\Omega;S^2)$. Indeed for fixed $ \epsilon > 0 $, $ u_\epsilon $ is smooth up to the boundary and we may thus evaluate \eqref{Eq_eps_LL} at $ x \in \partial\Omega $ for any $t\geq0$. This gives $ \Delta u_\epsilon = 0 $ on $ \partial\Omega \times \mathbb{R}_+ $. As we will see later, uniform estimates imply the existence of a subsequence $ u_{\epsilon_i} $ converging to a map $ u_* $, which is a smooth solution of the Landau-Lifshitz or harmonic map flow in $ \mathop{\rm Reg}(\{u_{\epsilon_i}\}) $ and satisfies $$ - \Delta u_* = |\nabla u_*|^2 u_* \quad \mbox{ on } (\partial\Omega \times \mathbb{R}_+) \cap \mathop{\rm Reg}(\{u_{\epsilon_i}\}) , $$ since $\partial_t u_* = 0$ on $\partial\Omega \times \mathbb{R}_+$. However $L^p_{\rm loc}$-estimates for $ \nabla^3 u_\epsilon $ would imply $$ 0 = \Delta u_{\epsilon_i} \to \Delta u_* \mbox{ in } L^p_{\rm loc}((\partial\Omega \times \mathbb{R}_+) \cap \mathop{\rm Reg}(\{u_{\epsilon_i}\}) ; \mathbb{R}^3) $$ by compactness of the ``projection'' $ H^{1,p}(\Omega) \hookrightarrow L^p(\partial\Omega)$. This is not possible unless $ u_* \equiv const.$ on $ (\partial\Omega\times \mathbb{R}_+) \cap \mathop{\rm Reg}(\{u_{\epsilon_i}\})$. (Compare \cite{BBH1}, Remark 1 p.125 for a similar argument in the time independent case.) The following lemma will be proven in Section \ref{subsec-high}. \begin{lemm} \label{Cinfty2} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL}, with \\ $ u_0 \in H^{1,2}(\Omega;S^2) \cap H^{2,p}(\partial\Omega;S^2)$ and $ p \geq2 $ for each $ \epsilon > 0 $. Assume $$ \sup_{P^\Omega_R(z_0)} g_{\epsilon} \leq C_0 $$ and $B_R(x_0) \cap \partial\Omega \neq \emptyset$, $ 00$. There are constants $ C_0 =C_0(\norm{u_0}_{C^2(\partial\Omega)}, E_0, \Omega)$ and $\epsilon_0$ $= \epsilon_0(\norm{u_0}_{C^2(\partial\Omega)}, E_0, \Omega)> 0 $, such that if for some $z_0 =(x_0,t_0)$ and $R_0 \in ]0,\min\{1,\sqrt{t_0}\}[$ $$ \limsup_{\epsilon\searrow 0} \underset{t_0-R_0^2 < t < t_0}{\sup } \int_{B_{R_0}(x_0)\cap \Omega} g_\epsilon (u_\epsilon ) dx < \epsilon_0 \,, $$ then $$ \limsup_{\epsilon\searrow 0} \sup_{P^\Omega_{\delta R_0}(z_0)} g_\epsilon (u_\epsilon ) \leq \frac{C_0}{(1-\delta)^2 R_0^2} , $$ for any $ \delta \in ]0,1[$. \end{theorem} If the target is $N$, the above constants $C_0$ and $\epsilon_0$ also depend on the geometry of $N$. \begin{proof} Without loss of generality let $ (x_0,t_0) = 0 $. Since $u_0 \in C^2(\partial\Omega)$ admits an extension $w_0\in C^2(\overline{\Omega})$ and since $t_0 -R^2 >0$, we may assume $u_0 \in C^2(\overline{\Omega})$. We have $ u_\epsilon \in C^{1,0}_{\alpha}(\overline{\Omega}\times \mathbb{R}_+;S^2)$ for any $ 0<\alpha <1$ and so there are $\sigma_\epsilon \in [0,R_0[$ and $ z_\epsilon = (x_\epsilon, t_\epsilon) \in \overline{P^\Omega_{\sigma_\epsilon }}$, such that \begin{gather*} (R_0 - \sigma_\epsilon)^2 \sup_{P^\Omega_{\sigma_\epsilon}} g_\epsilon = \max_{0 \leq \sigma \leq R_0} \Bigl( (R_0 - \sigma)^2 \underset{P^\Omega_{\sigma}}{ \sup } g_\epsilon \Bigr),\\ e_\epsilon = g_\epsilon(u_\epsilon (z_\epsilon)) = \underset{P^\Omega_{\sigma_\epsilon}}{ \sup } g_\epsilon \,. \end{gather*} Again for $ \rho_\epsilon := \frac{1}{2} (R_0 - \sigma_\epsilon)$, we have $ \sup_{P^\Omega_{\rho_\epsilon}(z_\epsilon)} g_\epsilon \leq 4 e_\epsilon$. Consider the rescaled map $$ v_\epsilon(y,s) := u(x_\epsilon + e_\epsilon^{-1/2} y, t_\epsilon + e_\epsilon^{-1}s) \,. $$ By construction $ v_\epsilon $ satisfies \begin{equation} \gamma_1 \partial_t v_\epsilon - \gamma_2 v_\epsilon \times \partial_t v_\epsilon - \Delta v_\epsilon = \frac{1}{\tilde{\epsilon}^2} (1-|v_\epsilon|^2) v_\epsilon \mbox{ on } P_{r_\epsilon}^{\Omega_\epsilon}\,,\label{*} \end{equation} with $\tilde{\epsilon} := \sqrt{e_\epsilon} \epsilon$, $ r_\epsilon := \sqrt{e_\epsilon} \rho_\epsilon$, $ \Omega_\epsilon := \sqrt{e_\epsilon} \bigl( \Omega - x_\epsilon \bigr) $ and $$P_{r_\epsilon}^{\Omega_\epsilon} := \bigl( B_{r_\epsilon} \cap \Omega_\epsilon \bigr)\times ]-r_\epsilon^2, 0[\,.$$ \\ Further by construction, \begin{equation} g_{\tilde{\epsilon}}(v_\epsilon)(0,0) = 1 \mbox{ and } \sup_{ P_{r_\epsilon}^{\Omega_\epsilon} } g_{\tilde{\epsilon}}(v_\epsilon) \leq 4 \,. \label{**} \end{equation} The boundary data are also rescaled. Set $v_{\epsilon,0}(y) := u_0(x_\epsilon + e_\epsilon^{-1/2} y)$. Then $$ v_\epsilon (y,s) = v_{\epsilon,0}(y) \mbox{ on } (\partial\Omega_\epsilon \cap B_{r_\epsilon}) \times ]-r_\epsilon^2,0[ $$ and $$ \sup_{P_{r_\epsilon}^{\Omega_\epsilon}} |\nabla v_{\epsilon,0}| \leq e_\epsilon^{-1/2} \sup_{P_{R_0}^{\Omega}} |\nabla u_0| \,, \quad \sup_{P_{r_\epsilon}^{\Omega_\epsilon}} |\nabla^2 v_{\epsilon,0}| \leq e_\epsilon^{-1} \sup_{P_{R_0}^{\Omega}} |\nabla^2 u_0| \,. $$ Now we claim that for sufficiently small $\epsilon> 0$, we have $$ r_\epsilon \leq C_0 := \max\{2,\tilde{C}(\Omega,\norm{u_0}_{C^2(\overline{\Omega})})\} , $$ where $\tilde{C}(\cdot)>0$ will be specified later. Again by definition of $r_\epsilon$, this will prove the theorem. Assume by contradiction $ r_\epsilon > C_0 \geq 2 $ for small $\epsilon >0$. Then $$ e_\epsilon^{-1/2} = \rho_\epsilon/r_\epsilon < R_0/(2C_0) \leq 1/(2C_0) , $$ since $0 0$, the right hand side of \eqref{*} is uniformly bounded in $\tilde{\epsilon} = \sqrt{e_\epsilon} \epsilon >0$ and together with \eqref{**} we obtain uniform bounds in $C^\infty(P_{2}^{\Omega_\epsilon})$. This however leads to a contradiction as in the proof of Theorem \ref{eps-reg-LL}, if $\epsilon_0$ is smaller than $\epsilon_1$. Further if $$ \limsup_{\epsilon\searrow 0} \bigl(\sqrt{e_\epsilon} \mathop{\rm dist}(x_\epsilon,\partial\Omega)\bigr) = \limsup_{\epsilon\searrow 0} \mathop{\rm dist}(0,\partial\Omega_\epsilon) \geq \frac{1}{2}\,, $$ we can also use uniform interior estimates in $C^\infty(P_{1/4}^{\Omega_\epsilon})$ and proceed as in the proof of Theorem \ref{eps-reg-LL} to get a contradiction, if we choose $\epsilon_0$ sufficiently small. So far the required upper bound on $\epsilon_0$ is universal in the sense that it only depends on the geometry of $N$ and the parabolicity constants. We therefore have $$ \limsup_{\epsilon\searrow 0} \mathop{\rm dist}(0,\partial\Omega_\epsilon) < 1/2 \,, $$ and in the sequel we consider sufficiently small $\epsilon >0$, such that $\mathop{\rm dist}(0,\partial\Omega_\epsilon)< 1/2$. Lemma \ref{Cinfty2} combined to the embedding $W^{2,1}_p(P_1) \hookrightarrow C^1(P_1)$ for $p>4$, implies \begin{align*} &\sup_{P_1^{\Omega_\epsilon}}|\nabla v_\epsilon|^2 \\ &\leq C \norm{v_\epsilon}_{W^{2,1}_p(P_1^{\Omega_\epsilon} )}^2\\ &\leq C(p,\Omega_\epsilon \cap B_2) \Bigl( \norm{\frac{1}{\tilde{\epsilon}^2} (1-|v_\epsilon|^2)}_{L^p(P_2^{\Omega_\epsilon} )}^2 +\norm{v_\epsilon}_{L^2(P_2^{\Omega_\epsilon})}^2 +\norm{v_{\epsilon,0}}_{H^{2,p}(P_2^{\Omega_\epsilon})}^2 \Bigr) \end{align*} Note that $\Omega_\epsilon \cap B_2$ has uniformly bounded curvature and so $$ 0< C(p,\Omega_\epsilon \cap B_2) < C(p,\Omega)\,.$$ Since $(1-|v_\epsilon|^2) \leq 1$ and %by construction $\sup_{ P_2^{\Omega_\epsilon} } g_{\tilde{\epsilon}}(v_\epsilon) \leq 4 $, Lemma \ref{Cinfty2} implies $$ \norm{\frac{1}{\tilde{\epsilon}^2} (1-|v_\epsilon|^2)}_{L^p(P_2^{\Omega_\epsilon} )}^2 \leq C_{p} \bigl( o(\epsilon_0) + o(\tilde{\epsilon}) \bigr) \,, $$ where $C_p = C(p,E_0)$ and $o(\tau)$ denotes a generic function that satisfies $$ \lim_{\tau \searrow 0} o(\tau)=0 \,.$$ A Poincar\'e inequality on $P_2^{\Omega_\epsilon}$ leads to \begin{align*} \norm{v_\epsilon}_{L^2(P_2^{\Omega_\epsilon})}^2 &\leq 2 \bigl(\norm{v_{\epsilon,0}}_{L^2(P_2^{\Omega_\epsilon})}^2 + \norm{v_\epsilon - v_{\epsilon,0}}_{L^2(P_2^{\Omega_\epsilon})}^2\bigr) \\ &\leq 2 \norm{v_{\epsilon,0}}_{L^2(P_2^{\Omega_\epsilon})}^2 +C(\Omega)\bigl( \norm{\nabla v_{\epsilon,0}}_{L^2(P_2^{\Omega_\epsilon})}^2 +\norm{\nabla v_{\epsilon}}_{L^2( P_2^{\Omega_\epsilon})}^2 \bigr) \,. \end{align*} Again $\norm{\nabla v_{\epsilon}}_{L^2( P_2^{\Omega_\epsilon})}^2 \leq o(\epsilon_0)$. Of course \begin{align*} \norm{v_{\epsilon,0}}_{H^{1,2}(B_2^{\Omega_\epsilon})} &\leq C(p) \norm{v_{\epsilon,0}}_{H^{1,p}(B_2^{\Omega_\epsilon})}\\ &\leq C(p) \norm{v_{\epsilon,0}}_{H^{2,p}(B_2^{\Omega_\epsilon})}^2\\ &\leq C(p,\Omega) \norm{v_{\epsilon,0}}_{C^2(B_2^{\Omega_\epsilon})}^2 \end{align*} and we still need to estimate $\norm{v_{\epsilon,0}}_{C^2(B_2^{\Omega_\epsilon})}^2$. For each $\epsilon>0$ we may chose coordinates for the target such that $ v_{\epsilon,0} (0) =0$. Then \begin{gather*} \sup_{B_2^{\Omega_\epsilon} }|v_{\epsilon,0}| \leq 4 \sup_{B_2^{\Omega_\epsilon}} |\nabla v_{\epsilon,0}|\,, \\ \norm{v_{\epsilon,0}}_{C^2(B_2^{\Omega_\epsilon})} \leq C e_\epsilon^{-1/2} \sup_{B_{R_0}^\Omega} |\nabla u_0| + e_\epsilon^{-1} \sup_{B_{R_0}^\Omega} |\nabla^2 u_0|\,. \end{gather*} The above estimates combined to the one for $\norm{ \frac{1}{\tilde{\epsilon}^2} (1 - \abs{v_\epsilon} ^2)}_{L^\infty (P^{\Omega_\epsilon}_1)} $ in Lemma \ref{Cinfty2} yield \begin{align*} 1 &\leq \sup_{P_1^{\Omega_\epsilon}} g_{\tilde{\epsilon}}(v_\epsilon)\\ & \leq\sup_{P_1^{\Omega_\epsilon}}\frac{1}{2}|\nabla v_\epsilon|^2 +\tilde{\epsilon}^2 \sup_{P_1^{\Omega_\epsilon}}\bigl( \frac{1}{\tilde{\epsilon}^2} (1 - \abs{v_\epsilon} ^2)\bigr)^2\\ &\leq C_1 \left(o(\epsilon_0) +o(\tilde{\epsilon}) + e_\epsilon^{-1} \norm{\nabla u_0}_{C^1(B_{R_0}^\Omega)}^2\right) \,, \end{align*} where $C_1 = C_1(\Omega , E_0)$. Now if both $ o(\epsilon_0)< (1/4) C_1^{-1}$ and $o(\tilde{\epsilon}) < (1/4) C_1^{-1}$, this leads to $$ e_\epsilon < 2 C_1 \norm{\nabla u_0}^2_{C^1(B_{R_0}^\Omega)} , $$ which is in contradiction with $r_\epsilon > C_0 := \max\{2,2 C_1 \norm{u_0}_{C^2(\overline{\Omega})}\}$ and $\sqrt{e_\epsilon} > 2C_0$. Thus $ r_\epsilon \leq C_0$ and by definition of $r_\epsilon$ also $$ \frac{1}{4} (R_0 - \delta R_0)^2 \sup_{P^\Omega_{ \delta R_0}} g_\epsilon \leq C_0^2 = C(\Omega, E_0, \norm{u_0}_{C^2(\overline{\Omega})}) \,. $$ Since $t_0 -R^2 >0$, we could replace $u_0$ in the above by any $w_0 \in C^2(\overline{\Omega})$ with $w_0 = u_0$ on $\partial\Omega \cap B_{R_0}$. Therefore the above constants merely depend on $\norm{u_0}_{C^2(\partial\Omega)}$. \end{proof} \subsection{Higher estimates} \label{subsec-high} In this section, we prove Lemmata \ref{Cinfty1} and \ref{Cinfty2}, for which the following uniform estimates will be needed. \subsubsection{Uniform estimates in $ \epsilon > 0 $} \label{subsec-comp} The ``distance-to-the-target-function'' $\rho_\epsilon := 1-|u_\epsilon|^2$ satisfies \begin{equation} \label{rho} \gamma_1 \partial_t \rho_\epsilon - \Delta \rho_\epsilon + \frac{2}{\epsilon^2} \rho_\epsilon = 2 |\nabla u_\epsilon|^2 + \frac{2}{\epsilon^2} \rho_\epsilon^2 \,. \end{equation} Since $\gamma_1 >0$, we may assume $\gamma_1 = 1$ without loss of generality. We will now derive uniform a priori estimates for this equation. Lemma \ref{f} extends a comparision argument from \cite{BBH1} (Lemma 2, p.130) to the time dependent case and to non-positive solutions. The parabolic boundary of $ P_R := B_R(0) \times ]-R^2,0[$ is denoted as $$ \tilde{\partial} P_R := \bigl( B_R(0) \times \{ - R^2 \} \bigr) \cup \bigl( \partial B_R(0) \times [-R^2,0 ] \bigr) \,. $$ \begin{lemm} \label{f} Let $ a > 0$, $R \in ]0, \frac{1}{4}[ $, $\epsilon \in ]0, 1[ $ and $ g \in C^0(\overline{P_R})$ with $\epsilon^2 \sup_{P_R} |g| \leq a$. Let $ f \in C^0(\overline{P_R}) \cap C^2(P_R) $ be a solution of \begin{gather*} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f = g \quad \mbox{in } P_R\,,\\ |f| \leq a \quad \mbox{on } \tilde{\partial} P_R \,. \end{gather*} Then for any $ \delta \in ]0,1[ $, we have $$ \frac{1}{\epsilon^2 } |f| \leq \underset{P_R }{ \sup } |g| + \frac{2a}{\epsilon^2 } e^{-\frac{1}{\epsilon } (1-\delta^2)^2 R^4 } \mbox{ on } P_{\delta R} \,. $$ \end{lemm} \begin{proof} Consider $\omega(x,t) = 2 a e^{-\frac{1}{\epsilon}(R^2 - |x|^2)(R^2 + t )}$. Then \begin{gather*} \epsilon^2 \bigl( \partial_t \omega - \Delta \omega \bigr) + \omega > 0\quad \mbox{in }P_R \,,\\ \omega = 2 a \quad \mbox{on } \tilde{\partial} P_R \,. \end{gather*} For $ f_1 := f - \epsilon^2 \underset{P_R }{ \sup } |g| $ and $ f_2 := f + \epsilon^2 \underset{ P_R }{ \sup } |g| $, we have $$ |f_1| \leq 2 a \quad \mbox{and}\quad |f_2| \leq 2 a \mbox{ on } \tilde{\partial}P_R\,, $$ and hence $$ f_1 - \omega \leq 0 , \quad f_2 + \omega \geq 0 \quad \mbox{on } \tilde{\partial} P_R \,. $$ Moreover $$ \epsilon^2 \bigl( \partial_t f_1 - \Delta f_1 \bigr) + f_1 \leq 0 \,, \quad \epsilon^2 \bigl( \partial_t f_2 - \Delta f_2 \bigr) + f_2 \geq 0 \quad \mbox{in } P_R \,. $$ The Maximum Principle now implies $ f_1 - \omega \leq 0 $ and $ f_2 + \omega \geq 0 $ on $ P_R $, that is $$ - \omega - \epsilon^2 \underset{ P_R }{ \sup } |g| \leq f \leq \omega + \epsilon^2 \sup_{P_R } |g| \,. $$ \end{proof} The above lemma will yield interior estimates. If $ B_R \cap \Omega \neq \emptyset $ and $ f \equiv 0 $ on $ B_R \cap \partial\Omega $, we still obtain a local estimate up to the boundary, i.e. on $ P^\Omega_{\delta R} = (B_{\delta R}\cap \Omega) \times ]-(\delta R)^2,0[$. \begin{coro} \label{bd-f} Consider a smooth domain $\Omega \subset \mathbb{R}^2$, $ a > 0$, $R \in ]0, \frac{1}{4}[ $, $\epsilon \in ]0, 1[ $ and $ g \in C^0(\overline{P_R^\Omega})$ with $\epsilon^2 \sup_{P_R} |g| \leq a$. Let $ f \in C^0(\overline{P_R^\Omega}) \cap C^2(P_R^\Omega) $ be a solution of \begin{gather*} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f = g \quad \mbox{in } P^\Omega_R \,, \\ |f| \leq a \quad \mbox{on } \tilde{\partial} P_R \cap \Omega \,, \\ f = 0 \quad \mbox{on } \partial\Omega \cap P_R \,. \end{gather*} Then for any $ \delta \in ]0,1[$, we have $$ \frac{1}{\epsilon^2 } |f| \leq \sup_{ P^\Omega_R } |g| + \frac{2a}{\epsilon^2 } e^{-\frac{1}{\epsilon } (1-\delta^2)^2 R^4 } \quad \mbox{on } P^\Omega_{\delta R} \,. $$ \end{coro} The proof of Lemma \ref{f} also applies in this case. The next interior-estimate-version of Lemma \ref{f} deals with the case $ B_R(x_0) \cap \Omega \neq \emptyset $ and $ f \neq 0 $ on $ \partial B_R(x_0) \cap \Omega $. The estimate then also depends on $\operatorname{dist}(x,\partial\Omega)$. We formulate the following lemma in such a way that it readily extends to the case $\Omega = M$ is a manifold. \begin{coro} \label{fbdry} % subsolution 2 Let $ U \subset \mathbb{R}^2 $ be an open smooth neighborhood of $ 0 $ with $\mathop{\rm diam} U \leq 1 $ and set $P_{R,U} := U \times ]-R^2,0[ $. Consider $ a > 0 $, $R \in ]0, \frac{1}{4}[ $, $\epsilon \in ]0,\frac{1}{4} [ $ and $ g \in C^0(\overline{P_{R,U}})$ with $\epsilon^2 \sup_{P_{R,U}} |g| \leq a$. Let $ f \in C^0(\overline{P_{R,U}}) \cap C^2(P_{R,U}) $ be a solution of \begin{equation} \label{eq_rho_2} \begin{gathered} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f = g \quad \mbox{in } P_{R,U} \,,\\ |f| \leq a \quad \mbox{on } \tilde{\partial} P_{R,U} \,. \\ \end{gathered} \end{equation} Then there is a constant $ C = C(U) > 0 $, such that for any $ \delta \in ]0,1[$ we have $$ \frac{1}{\epsilon^2} |f (x,t)| \leq \underset{P_{R,U}}{\sup} |g| + \frac{2a}{\epsilon^2} e^{-\frac{ R^2}{C \epsilon } (1-\delta^2) \operatorname{dist}^2 (x,\partial U) } \quad \mbox{ on } P_{\delta R,U}\,. $$ \end{coro} Of course $$ \tilde{\partial } P_{R,U} := \bigl( U \times \{ - R^2 \} \bigr) \cup \bigl( \partial U \times [-R^2,0] \bigr) \,. $$ \begin{proof} Set $ d(x) := \operatorname{dist} (x, \partial U ) $, $ C = C(U) := \max \{ 1,\norm{ \Delta d^2}_{ L^\infty (U) }, \norm{\nabla d^2}_{ L^\infty (U) } \} $. Note that $ d(x) \leq 1 $ on $ U $ since $\mathop{\rm diam} U \leq 1 $. We claim that $$ \omega (x,t) := 2 a e^{-\frac{1}{C \epsilon} d^2(x) (R^2 + t )} $$ is a supersolution of equation (\ref{eq_rho_2}), if $ 0 < R < \frac{1}{4} $ and $ 0 < \epsilon < \frac{1}{4} $. Indeed \begin{align*} \epsilon^2 (\partial_t - \Delta ) \omega + \omega &= \omega \Bigl[1 - \frac{\epsilon }{C} d^2 + \frac{\epsilon }{C} (R^2 + t) \Delta d^2 - \frac{\epsilon }{C}\frac{1}{\epsilon C} (R^2 + t)^2 \abs{\nabla d^2}^2 \Bigr] \\ & \geq \omega \bigl[ 1 - \epsilon - \epsilon R^2 - R^4\bigr] \\ &\geq \frac{1}{4} \omega > 0 \quad \mbox{on } P_{R,U}\,. \end{align*} The claim now follows just as in the proof of Lemma \ref{f}. \end{proof} We will also need a priori $L^p$-estimates for the above equation. \begin{lemm} \label{l1} Consider a smooth domain $\Omega \subset \mathbb{R}^2$, $ g \in L^1(\Omega\times ]0,T[)$ and $\epsilon >0$. Let $ f \in C^1(\overline{\Omega}\times [0,T]) \cap C^2(\Omega\times ]0,T[) $ be a solution of \begin{gather*} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f = g \quad \mbox{in } \Omega \times ]0,T[\,,\\ f = 0 \quad \mbox{on } \Omega \times \{0\} \cup \partial\Omega \times ]0,T[ \,. \end{gather*} For $f\geq 0$, we only need to assume \begin{gather*} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f \leq g \quad \mbox{ in } \Omega \times ]0,T[ \,, \\ f = 0 \quad \mbox{on } \Omega \times \{0\} \cup \partial\Omega \times ]0,T[ \,. \end{gather*} Then \begin{equation} \label{global-l1} \norm{ \frac{1}{\epsilon^2} f }_{L^1(\Omega \times ]0,T[)} \leq \norm{g}_{L^1(\Omega \times ]0,T[)} \,. \end{equation} and for any $R,\rho >0$ and $z_0 =(x_0,t_0) \in \Omega \times ]0,T]$ with $R^2+\rho^2 < t_0$, \begin{equation} \label{1/2} \int_{P^\Omega_{R}(z_0)} \frac{1}{\epsilon^2} |f| dz \leq \int_{P^\Omega_{R+\rho}(z_0)}\Bigl( |g| + \frac{C}{\rho^2} |f| %+ \frac{C}{\rho} |\nabla f| \Bigr) dz \,. \end{equation} \end{lemm} \begin{proof} (i) Multiplication of the equation for $f$ by $\frac{f}{\sqrt{f^2+\delta^2}}$ leads to $$ \frac{\partial_t |f| |f|}{\sqrt{f^2+\delta^2}} + \frac{|\nabla f|^2}{\sqrt{f^2+\delta^2}} \big(1 -\frac{f^2}{f^2+\delta^2}\big) + \frac{1}{\epsilon^2} \frac{f^2}{\sqrt{f^2+\delta^2}} = \frac{g f}{\sqrt{f^2+\delta^2}} + \Delta \sqrt{f^2+\delta^2} \,. $$ Now integrate over $\Omega \times ]0,t[$ for any $t\in ]0,T]$ and let $\delta \to 0$ to obtain $$ \sup_{0\leq t \leq T} \int_\Omega |f(x,t)| \, dx + \int_0^T \int_\Omega \frac{1}{\epsilon^2} |f| \leq \int_0^T \int_\Omega |g| \, dx \,dt \,. $$ (ii) We multiply the equation by $f$ with $$ \Bigl( \frac{f}{\sqrt{f^2+\delta^2}}\Bigr) (x,t) \phi(x) \eta(t) \,. $$ The cut-off function $\phi$ satisfies $ 0 \leq \phi \in C^\infty_c(\mathbb{R}^2)$ with spt$\phi \subset B_{R+\rho}(x_0) $ and $ \phi \equiv 1 $ on $ B_{R}(x_0) $, whereas $ \eta \in C^\infty(\mathbb{R}_+) $ with $ 0 \leq \eta (t)\leq 1$, $\eta (t_0 - R^2 - \rho^2) = 0 $ and $ \eta (t) \equiv 1$ if $ t \geq t_0 - R^2$. We may assume $$ |\nabla \phi| \leq \frac{C}{\rho} \,, \quad |\nabla^2 \phi| \leq \frac{C}{\rho^2} \quad \mbox{and} \quad |d_t \eta| \leq \frac{C}{\rho^2} \,. $$ This leads to \begin{align*} &\frac{\partial_t \bigl(|f| \phi^2 \eta\bigr) |f|}{\sqrt{f^2+\delta^2}} + \frac{|\nabla f|^2 \phi^2 \eta}{\sqrt{f^2+\delta^2}} \big(1 -\frac{f^2 }{f^2+\delta^2}\big) + \frac{1}{\epsilon^2} \frac{f^2 \phi^2 \eta}{\sqrt{f^2+\delta^2}}\\ &= \frac{g f \phi^2 \eta}{\sqrt{f^2+\delta^2}} + \operatorname{div} \big( \nabla f \frac{f \phi^2 \eta}{\sqrt{f^2+\delta^2}} \big) + \frac{f^2 \phi^2 \partial_t\eta }{2 \sqrt{f^2+\delta^2}} - 2 \eta \phi \nabla \phi \nabla \sqrt{f^2+\delta^2} \,. \end{align*} Of course $$ \int_\Omega \phi \nabla \phi \eta \nabla \sqrt{f^2+\delta^2}\, dx = - \int_\Omega \eta \sqrt{f^2+\delta^2} \bigl( \phi \nabla^2\phi + |\nabla \phi|^2 \bigr)\, dx \,. $$ After integrating (\ref{troet}) and letting $\delta \to 0$, we obtain $$ \sup_{t_0-(R^2+\rho^2)< t 0$. Let $ f \in C^1(\overline{\Omega}\times [0, T]) \cap C^2(\Omega\times ]0, T[) $ be a solution of \begin{gather*} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f = g\quad \mbox{in } \Omega\times ]0, T[\,,\\ f = 0 \quad \mbox{on } \partial\Omega \times ]0,T[\,. \end{gather*} For $f\geq 0$, we only need to assume \begin{gather*} \bigl( \partial_t f - \Delta f \bigr) + \frac{1}{\epsilon^2} f \leq g \quad \mbox{ in } \Omega\times ]0,T[ \,,\\ f = 0\quad \mbox{ on } \partial\Omega\times ]0,T[ \,. \end{gather*} (i) For any $\delta \in ]0,1[$ and $z_0 =(x_0,t_0) \in \Omega \times ]0,T]$ with $01$ with $\frac{1}{q} + \frac{1}{p} = 1$ and $ a,b,\delta >0$, we have $$ \left| |g| |f|^{2s-1} \right| \leq \frac{1}{2\epsilon^2} \frac{2s-1}{2s} |f|^{2s} + (2\epsilon^2)^{(2s-1)} \frac{1}{2s} |g|^{2s} $$ and \begin{align*} \left| \nabla |f| |f|^{2s-1} 2 \nabla \phi \phi \eta \right| &=2 |(\frac{1}{s} \nabla |f|^s \phi) (|f|^s \nabla\phi) \eta |\\ &\leq \frac{2s-1}{2s^2} \abs{\nabla |f|^s}^2 \phi^2 \eta + \frac{2}{2s-1} |f|^{2s} \abs{\nabla \phi}^2 \eta \,. \end{align*} This leads to \begin{equation} \label{pain} \begin{aligned} &\frac{1}{2s} \partial_t \bigl( |f|^{2s}(x,t) \phi^2(x) \eta(t) \bigr) + \frac{2s-1}{2s^2} \abs{\nabla |f|^s}^2 \phi^2 \eta + \frac{1}{2\epsilon^2} |f|^{2s} \phi^2 \eta \\ &\leq - \operatorname{div}\bigl( \nabla f f |f|^{2s-2} \phi^2 \eta \bigr) + (2\epsilon^2)^{(2s-1)} \frac{1}{2s} |g|^{2s} \phi^2 \eta\\ &\quad + \frac{2}{2s-1}|f|^{2s} \bigl( \abs{\nabla \phi}^2 \eta + \phi^2 |d_t\eta| \bigr) \,. \end{aligned} \end{equation} For notational ease we relabel the domain as $\Omega \times ]-T,0[$ and assume $z_0 = (0,0) \in \Omega \times ]-T,0[$ and $ 0< R^2 + \rho^2 0$ \begin{equation} \label{est2} \norm{\frac{1}{\epsilon^2} f}_{L^p(P^\Omega_{R})}^p \leq C(p) \norm{g}_{L^p(P^\Omega_{(1+\delta)R} )}^p + \epsilon^2 C_2 \,, \end{equation} where $$ C_2 = C_2 \bigl(\norm{g}_{L^p(P^\Omega_{(1+\delta)R})}, \norm{f}_{L^{2p-1}(P^\Omega_{(1+\delta)R})}, p, \delta, R \bigr)\,. $$ Claim (i) follows by setting $R_{new} = (1+\delta)R$ and $\delta_{new} = \frac{1}{1+\delta}$, i.e. $\delta_{new} R_{new} =R$. \noindent (ii) By applying the estimate $$ \Bigl(\int_{[a,b]} \int_{B_R} \abs{u}^4 dx dt \Bigr)^{\frac{1}{2}} \leq C \Bigl(\max_{t \in [a,b]} \int_{B_R} \abs{u}^2 dx dt + \int_{[a,b]} \int_{B_R} \abs{\nabla u}^2 dx dt \Bigr) $$ (see Theorem 6.9 p.110 in \cite{Lieberman}) to $ u := f^{p/2} \phi \sqrt{\eta}$, we find that the expression \begin{equation} \label{s4} \bigl(\frac{1}{\epsilon^2} \bigr)^{p-1} \Bigl( \int_{P^\Omega_R} |f|^{2p}\phi^4 \eta^2 dz \Bigr)^{\frac{1}{2}} + \int_{P^\Omega_R} \bigl(\frac{1}{\epsilon^2} \bigr)^p |f|^p \phi^2 \eta dz \end{equation} admits the same bound as (\ref{main-s}) with a different constant $C(p) >0$. By combining (\ref{s4}) with (\ref{main-s}), we see that the same bounds as in (\ref{est2}) also holds for $$ \bigl(\frac{1}{\epsilon^2} \bigr)^{p-1} \Bigl(\sup_{-R^2 0 $. Then the following assertions are equivalent: \noindent (i) \quad $ z_0 =(x_0,t_0)\in \mathop{\rm Reg}\bigl( \{u_\epsilon\}_{\epsilon>0}\bigr)$. \noindent (ii) \quad $\exists \delta,R>0: \phantom{ppppp|} $ $ \limsup_{\epsilon\searrow 0}\sup_{t_0-\delta < t < t_0} G_\epsilon \bigl(u_\epsilon (t) , B_R^\Omega (x_0)\bigr) < \epsilon_0 $. \noindent (iii) \quad $ \exists \delta>0:\lim_{R\searrow 0} \limsup_{\epsilon\searrow 0} \sup_{t_0-\delta < t < t_0} G_\epsilon \bigl(u_\epsilon (t), B_R^\Omega (x_0)\bigr) = 0 $. \noindent (iv) \quad $\exists R>0: \limsup_{\epsilon\searrow 0} \frac{1}{R^2} \int_{t_0-R^2}^{t_0} \int_{B_R^\Omega (x_0)} g_\epsilon (u_\epsilon)\,dx\,dt < \frac{1}{4} \delta_0 \epsilon_0 $. \noindent (v) \quad $\exists \delta,R>0: \limsup_{\epsilon\searrow 0} \sup_{t_0-\delta < t < t_0 + \delta} G_\epsilon \bigl(u_\epsilon (t) , B_R^\Omega (x_0)\bigr) < \epsilon_0 $. \end{lemm} \begin{proof} ``$(i) \Leftrightarrow (ii)$'' is obvious.\\ ``$(ii) \Rightarrow (iii)$'' follows from Theorem \ref{eps-reg-LL-bd} in Section \ref{subsec-sup-bd}.\\ ``$(iii) \Rightarrow (iv)$'' is obvious.\\ ``$(iv) \Rightarrow (ii)$'': Assume $(iv)$ holds. By (\ref{reminder}) and the above choice of $\delta_0$, we have for sufficiently small $\epsilon >0$, \begin{align*} &\sup_{t_0 - (1/2)\delta_0 R^2 < t < t_0} G_\epsilon \bigl(u_\epsilon(t), B_{(1/2)R}^\Omega(x_0)\bigr)\\ &\leq \inf_{t_0 -\delta_0 R^2 < s < t_0 - (1/2)\delta_0 R^2} G_\epsilon \bigl(u_\epsilon(s), B_R^\Omega(x_0)\bigr) + \frac{C \delta_0 R^2 E_0}{\gamma_1 R^2} \\ &\leq \frac{2}{\delta_0 R^2} \int_{t_0-\delta_0 R^2}^{t_0 - (1/2)\delta_0 R^2 } G_\epsilon \bigl(u_\epsilon(t), B_R^\Omega(x_0)\bigr) \,dt + \frac{1}{2} \epsilon_0 \\ &< \frac{2}{\delta_0} \frac{1}{4} \delta_0 \epsilon_0 + \frac{1}{2} \epsilon_0 < \epsilon_0 \,. \end{align*} ``$(v) \Rightarrow (ii)$'' is obvious.\\ ``$(iii) \Rightarrow (v)$'': Assume (iii) holds. Then there are $R,\delta>0$, such that $$ \limsup_{\epsilon\searrow 0} \sup_{t_0-\delta\leq t\leq t_0} \int_{B_R(x_0)\cap \Omega} g_\epsilon (u_\epsilon(x,t)) dx < \epsilon_0/2 \,. $$ On the other hand by (\ref{reminder}), we have for $\delta_{new} := \frac{\epsilon_0 \gamma_1 R^2}{2C E_0} = \delta_0 R^2 $, $$ \sup_{t_0\leq t\leq t_0 +\delta_{new}} \int_{B_{\frac{1}{2}R}(x_0)\cap \Omega} g_\epsilon (u_\epsilon(x,t)) dx \leq \int_{B_{R}(x_0)\cap \Omega} g_\epsilon (u_\epsilon(x,t_0)) dx + \frac{\delta_{new} C E_0}{\gamma_1 R^2}\,. $$ Now (v) holds for $\frac{1}{2}R$ and $\min \{ \delta, \delta_{new}\}$. \end{proof} \begin{coro} \label{prop-limits} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL} with $ u_0$ in $H^{1,2}(\Omega;S^2)\cap \break C^{2}(\partial\Omega;S^2)$ for each $ \epsilon > 0 $. Let $\{\epsilon_i\}_i$ be a sequence with $\epsilon_i \searrow 0$ as $i\to\infty$. Then the following holds: \noindent(i) $\mathop{\rm Reg}(\{u_\epsilon\}_\epsilon)$ and $\mathop{\rm Reg}(\{u_{\epsilon_i}\}_i)$ are open in $ \overline{\Omega}\times \mathbb{R}_+$. \noindent (ii) There is some $ T_0 > 0 $, such that $ \overline{\Omega} \times [0,T_0[ \subset \mathop{\rm Reg}(\{u_\epsilon\}_\epsilon) $. \end{coro} \begin{proof} (i) follows from Lemma \ref{equiv} (v).\\ (ii) The existence of $T_0$ immediately follows from Lemma \ref{en-est-LL} (\ref{gs}). \end{proof} Set $$ Q_R(z) := B_R(x) \times ]t-R^2,t+R^2[ \quad \mbox{ for } z = (x,t) \,. $$ and let $\mathfrak{H}^2$ denote the 2-dimensional parabolic Hausdorff measure. \begin{prop} \label{Hausdorff} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL} with $ u_0\in H^{1,2}(\Omega;S^2)\cap C^{2}(\partial\Omega;S^2)$ for each $ \epsilon > 0 $. Then the following holds: \noindent (i) $\mathbb{S}(\{u_\epsilon\}_\epsilon)$ has locally finite two dimensional parabolic Hausdorff-measure. More precisely there is a constant $K_1 =K_1(E_0,\epsilon_0)>0$, such that for any compact intervall $I \subset \mathbb{R}_+$ $$ \mathfrak{H}^2 \bigl( \mathbb{S}(\{u_\epsilon\}_\epsilon) \cap (\overline{\Omega} \times I) \bigr) \leq K_1 |I| \,. $$ \noindent (ii) There is a constant $ K_2 = K_2(E_0,\epsilon_0) > 0 $, such that for any $t>0$ the set $\mathbb{S}^t(\{u_\epsilon\}_\epsilon) := $ $\mathbb{S}(\{u_\epsilon\}_\epsilon) \cap (\overline{\Omega} \times \{t\}) $ consists of at most $K_2$ points. \end{prop} \begin{proof} (i) By $(iv)$ of Lemma \ref{equiv}, we have for any $z_0 =(x_0,t_0) \in \mathbb{S}(\{u_\epsilon\}_\epsilon)$, any $R>0$ and sufficiently small $0 <\epsilon \leq \epsilon (z_0)$ \begin{equation} \label{troet} \frac{1}{R^2} \int_{t_0-R^2}^{t_0} \int_{B_R^\Omega (x_0)} g_\epsilon (u_\epsilon) dx dt > \frac{1}{4} \delta_0 \epsilon_0 \,. \end{equation} Fix a compact intervall $I\subset \mathbb{R}_+$ and $\delta>0$. By compactness and Vitali's Covering Theorem any covering of $ \mathbb{S}(\{u_\epsilon\}_\epsilon) \cap \bigl(\overline{\Omega} \times I\bigr) $ by parabolic cylinders $Q_{R}^\Omega (z)$ with $ 0 0 , \exists \epsilon \in ]0,\gamma [ : \quad \underset{T - \delta < t < T}{\sup} \int_{B_R^\Omega (x_l)} g_\epsilon (u_\epsilon (x,t)) dx \geq \frac{\epsilon_0}{2} \mbox{ for } 1 \leq l \leq k \,. $$ We may choose $ R > 0 $, such that the $ B_R^\Omega (x_l) (1 \leq l \leq k ) $ are pairewise disjoint. Choose $ \delta \in ] 0,\frac{\gamma_1 R^2 \epsilon_0}{4 C E_0}[ $, where $C$ is the constant from Lemma \ref{en-est-LL} \eqref{en-est} and $ \epsilon\in ]0,\gamma [ $ as above. Since $ t \mapsto \int_{B_R^\Omega (x_l)} g_\epsilon (u_\epsilon (x,t)) dx $ is continuous, we may find $ t_\delta^l \in ]T-\delta , T[ $ such that $$ \int_{B_R^\Omega (x_l)} g_\epsilon (u_\epsilon (x,t_\delta^l )) dx \geq \frac{\epsilon_0}{2} \quad \mbox{ for } 1 \leq l \leq k \,. $$ The energy estimate and the local energy inequality, Lemma \ref{en-est-LL},\eqref{en-est} and \eqref{loc-en} now imply \begin{align*} E_0 &\geq \sum_{l=1}^k \int_{B_R^\Omega (x_l)} g_\epsilon (u_\epsilon (x,T-\delta )) \, dx\\ &\geq \sum_{l=1}^k \Bigl( \int_{B_R^\Omega (x_l)} g_\epsilon \Bigl(u_\epsilon \bigl(x,t_\delta^l \bigl) \Bigl) dx - \frac{C}{\gamma_1 R^2} \int_{T - \delta}^T \int_{B_R^\Omega (x_l)} \abs{\nabla u_\epsilon (x,t)}^2 dx dt \Bigr). \end{align*} Thus $E_0 \geq k \bigl( \frac{\epsilon_0}{2} - \frac{C E_0}{R^2} \delta \bigr)$. Now since $ \delta < \frac{R^2 \epsilon_0}{4 C E_0} $, this implies $ k \leq \frac{8 E_0}{\epsilon_0} =: K_2 $. (Compare \cite{Struwe3} and \cite{Struwe2} ($1^\circ$) of the proof of Theorem 6.6 p.229 for a similar argument in the case of the harmonic map flow.) \end{proof} \begin{theorem} \label{limits} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL} with $ u_0$ in $H^{1,2}(\Omega;S^2)\cap \break H^{3/2,2}(\partial\Omega;S^2)$ for each $ \epsilon > 0 $. Then the following holds: \\ There is at least one sequence $ \{ \epsilon_i \}_i $, with $ \epsilon_i \to 0 $ as $ i \to \infty $ and $$ u_*\in H^{1,2}_{\rm loc}( \overline{\Omega} \times \mathbb{R}_+; S^2) \cap L^\infty(\mathbb{R}_+;H^{1,2}(\Omega;S^2)) , $$ such that $ u_{\epsilon_i} \rightharpoonup u_* $ weakly in $ H^{1,2}_{\rm loc}(\overline{\Omega} \times \mathbb{R}_+;\mathbb{R}^3) $ and weak* in $L^\infty(\mathbb{R}_+; H^{1,2}(\Omega;\mathbb{R}^3))$. In addition: (i) For any such sequence $ \{ u_{\epsilon_i} \}_i $, we have $$ \lim_{i\to\infty} u_{\epsilon_i } = u_*\quad \mbox{ in } C^\infty (\mathop{\rm Reg}(\{u_{\epsilon_i}\}_i) \cap (\Omega \times \mathbb{R}_+); \mathbb{R}^3) $$ and $\frac{1}{\epsilon^2} (1-|u_\epsilon|^2) \to |\nabla u_*|^2$ in $C^\infty (\mathop{\rm Reg}(\{u_{\epsilon_i}\}_i) \cap (\Omega \times \mathbb{R}_+))$. \\ (ii) $ u_*$ is a smooth solution of \eqref{LL} in $ \mathop{\rm Reg}(\{u_{\epsilon_i}\}_i)\cap (\Omega \times \mathbb{R}_+) $ and a distributional solution in $$ H^{1,2}_{\rm loc}(\overline{\Omega} \times \mathbb{R}_+) \cap L^\infty \bigl(\mathbb{R}_+;H^{1,2}(\Omega;\mathbb{R}^n)\bigr) $$ on all $\Omega \times \mathbb{R}_+$. Further $ \lim_{t\searrow 0} u_*(.,t) = u_0 $ in $H^{1,2}(\Omega;\mathbb{R}^3)$ and $$ {u_*(.,t)}_{|\partial\Omega} = {u_0}_{|\partial\Omega} \mbox{ as a } H^{2,2}(\Omega;\mathbb{R}^3)\mbox{-trace for a.e. } t>0\,. $$ (iii) If $ u_* $ is regular at $ z_0 = (x_0,t_0) \in \overline{\Omega}\times \mathbb{R}_+ $ in the sense that $$ \lim_{R\searrow 0} \sup_{t_0-R^2 \leq t \leq t_0} \int_{B_R(x_0)} |\nabla u_*|^2 dx = 0 $$ and if $z_0$ is parabolically isolated for $\{ u_{\epsilon_i} \}_i$, i.e. $$ B_{R_0}(x_0) \times ] t_0 - R_0^2,t_0[ \subset \mathop{\rm Reg}(\{u_{\epsilon_i}\})\mbox{ for some }R_0 >0\,, $$ then $z_0 \in \mathop{\rm Reg}(\{u_{\epsilon_i}\})$. (In particular $ u_* $ cannot (backwards) concentrate energy and (backwards) bubble at $z_0$ as $ t\searrow t_0 $. Compare \cite{Harpes1}, \cite{Harpes3}) \end{theorem} \begin{proof} (i) The convergence statements follow from the energy estimate (Lemma \ref{en-est-LL}), \mbox{Theorem \ref{eps-reg-LL}} and Lemma \ref{Cinfty1}. \noindent(ii) For the case $ \gamma_2 = 0 $ and $ f(u_\epsilon) = \frac{1}{2} \frac{d}{du}\chi \bigl(\operatorname{dist}^2(u_\epsilon,N)\bigr) $, this is proven in \cite{Struwe-Chen} III p.95. We will prove it in the case $\gamma_2 \neq 0$. If we apply ``$ u_{\epsilon_i} \times\,. $'' from the left to \eqref{Eq_eps_LL} and pass to the limit $\epsilon_i \to 0$ on $ \mathop{\rm Reg}(\{u_{\epsilon_i}\})\cap (\Omega \times \mathbb{R}_+) $, we obtain \begin{equation}\label{bof} \gamma_1 u_* \times {\partial}_t u_* - \gamma_2 u_* \times( u_* \times {\partial}_t u_*) - u_* \times \Delta u_* = 0 \,. \end{equation} Since $(1-|u_{\epsilon_i}|^2) \to 0$ smoothly, we also have $$ |u_*(x,t)| = 1 \quad \mbox{ in } \mathop{\rm Reg}(\{u_{\epsilon_i}\})\cap (\Omega \times \mathbb{R}_+). $$ Now we use $ a \times (b\times c) = (ac) b - (ab) c $ and $ |u_*| \equiv 1 $ while applying ``$ u_* \times $.'' from the left to (\ref{bof}), to obtain \begin{equation}\label{goal} \gamma_1 \partial_t u_* - \gamma_2 u_* \times \partial_t u_* - \Delta u_* = \abs{\nabla u_*}^2 u_* \quad \mbox{ in } \Omega \times \mathbb{R}_+ \,. \end{equation} In particular, since the left side of \eqref{Eq_eps_LL} converges to the left side of (\ref{goal}), we have $$ \frac{1}{\epsilon_i^2} (1-|u_{\epsilon_i}|^2) \to |\nabla u_*|^2 \mbox{ in } C^\infty (\mathop{\rm Reg}(\{u_{\epsilon_i}\})\cap (\Omega \times \mathbb{R}_+))\,. $$ We now prove that $u_*$ is a distributional $H^{1,2}_{\rm loc}\cap L^\infty(H^{1,2})$-solution of \eqref{LL} on all $\Omega\times\mathbb{R}_+ $. Note that the sequence $\{ u_{\epsilon_i} \}_i$ converges weakly in $H^{1,2}(\Omega\times \mathbb{R}_+; S^2)$ and smoothly on $\mathop{\rm Reg}(\{u_{\epsilon_i}\})\cap (\Omega \times \mathbb{R}_+)$. Further since $\mathbb{S}^t (\{u_{\epsilon_i}\}) := $ $\mathbb{S}(\{u_{\epsilon_i}\}) \cap (\overline{\Omega} \times \{t\}) $ is finite for all $t\geq 0$, we have both $u_{\epsilon_i} \to u_*$ pointwise a.e. in $\Omega \times \mathbb{R}_+$ and $u_{\epsilon_i}(\cdot,t) \to u_*(\cdot,t)$ pointwise a.e. in $\Omega$ for all $t \in \mathbb{R}_+$. Since $ \int_0^\infty\int_\Omega |\partial_t u_{\epsilon_i}|^2 dx dt \leq E_0 $, by Fatou's Lemma the complement of $$ A :=\{ t \geq 0 | \liminf_{\epsilon_i\searrow 0} \int_\Omega |\partial_t u_{\epsilon_i}|^2(x,t) dx < \infty \} $$ has measure 0. Pick $t_0 \in A$. Then there is a subsequence still denoted by $u_{\epsilon_i}$, such that $ \partial_t u_{\epsilon_i}(\cdot,t_0) \rightharpoonup \partial_t u_*(\cdot,t_0)$ weakly in $L^2(\Omega ; \mathbb{R}^3)$. By the local energy estimate, we may assume that, for the same subsequence, we also have $u_{\epsilon_i}(\cdot,t_0) \rightharpoonup u_*(\cdot,t_0)$ weakly in $H^{1,2}(\Omega ; S^2)$. By pointwise a.e. uniqueness of the limit, the whole sequence converges. Also $$ u_*(\cdot,t_0) \in H^{1,2}(\Omega;S^2) \quad \mbox{and} \quad \partial_t u_* (\cdot,t_0) \in L^2(\Omega;\mathbb{R}^3) \quad \mbox{for all } t_0 \in A \,. $$ Now $$ -\Delta u_*(\cdot,t_0) = (|\nabla u_*|^2 u_*)(\cdot,t_0) + f \,, $$ where $$ f = - \gamma_1 \partial_t u_*(\cdot,t_0) + \gamma_2 u_* \times \partial_t u_*(\cdot,t_0) \in L^2(\Omega;\mathbb{R}^3) $$ and by a regularity result of T.Rivi\`ere (see \cite{Tristan} Lemma p.3), we have $$ u_*(\cdot,t_0) \in H^{2,2}(\Omega;S^2) \quad \mbox{if} \quad u_0 \in H^{3/2,2}(\partial\Omega;S^2) \cap H^{1,2}(\Omega;S^2)\,. $$ This in particular implies $ {u_*(.,t)}_{|\partial\Omega} = {u_0}_{|\partial\Omega} $ as a $H^{2,2}(\Omega)$-trace for any $t\in A$. Further since $ \mathbb{S}^{t_0}(\{u_{\epsilon_i}\}_i) $ consists of finitely many points, it has vanishing $2$-capacity in $\mathbb{R}^2$, i.e. $$ Cap_2 \bigl(\mathbb{S}^{t_0}(\{u_{\epsilon_i}\})\bigr) = 0 $$ (see \cite{Evans-Gariepy}). Therefore, there is a sequence $\{\eta_k \}_k = \{ \eta_{k,q}\}_k \subset C^\infty_c(\mathbb{R}^2)$ with $$ \eta_k(x) = 1 \forall x \in \mathbb{S}^{t_0}(\{u_{\epsilon_i}\}_i) \quad \mbox{ and }\quad \norm{\eta_k}_{H^{1,2}(\mathbb{R}^2)} \overset{(k\to\infty)}{\to} 0 $$ (see \cite{Evans-Gariepy} 4.7.1). For $ \phi \in C^\infty_c(\Omega)$, we may test equation (\ref{bof}) with the cut-off function $ (1-\eta_k) \phi$, which has support in $\mathop{\rm Reg} (\{u_{\epsilon_i}\}_i)$. After passing to the limit $ k\to\infty $, we find that for %all $ \phi \in C^\infty_c(\Omega)$ and any $t\in A$, \begin{align*} &\int_\Omega \gamma_1 \partial_t u_*(x,t) \phi(x) - \gamma_2 (u_* \times \partial_t u_*)(x,t) \phi(x) + \nabla u_* (x,t) \nabla \phi(x) \, dx \\ &=\int_\Omega (|\nabla u_*|^2 u_*)(x,t) \phi(x) \, dx \,. \end{align*} This equation holds for a.e. $t\geq0$. On the other hand, we have $u_*\in H^{1,2}(\Omega \times [0,T];S^2)$ for any $T>0$ and so both sides of the above equation are locally integrable on $\mathbb{R}_+$. Therefore we may multiply the equation with $ \psi \in C^\infty_c([0,\infty[) $ and integrate over $\mathbb{R}_+$. Moreover linear combinations $\sum_k a_k \phi_k(x) \psi_k(t) $ with $ \phi_k \in C^\infty_c(\Omega) $ and $ \psi_k \in C^\infty_c([0,\infty[) $ are dense in $ C^\infty_c(\Omega \times [0,\infty[) $ and so \begin{align*} &\int_0^\infty \int_\Omega \gamma_1 \partial_t u_*(x,t) \phi(x,t) -\gamma_2 (u_* \times \partial_t u_*)(x,t) \phi(x,t) +\nabla u_* (x,t) \nabla \phi(x,t) dx\, dt \\ &= \int_0^\infty \int_\Omega (|\nabla u_*|^2 u_*)(x,t) \phi(x,t) dx \,dt \,, \end{align*} for any $\phi \in C^\infty_c(\Omega \times [0,\infty[) $. Finally $ \lim_{t\searrow 0} u_*(.,t) = u_0 $ in $H^{1,2}(\Omega;S^2)$ immediately follows from $ E(u_*(t_0)) \leq E(u_0) $, since we have weak convergence as $ t\searrow 0$. \\ (iii): By assumption there is $R>0$, such that $$ \sup_{t_0-R^2\leq t \leq t_0} \int_{B_R(x_0)\cap\Omega} \frac{1}{2} |\nabla u_* (x,t)|^2 dx < \frac{\epsilon_0}{4} \,. $$ Set $\delta := \min\{ R^2,\frac{\gamma_1 R^2\epsilon_0}{2 C E_0}\} $ and $ s_0 := t_0 -\frac{1}{2} \delta $. We may assume we have for the same $R>0$ $ P_{2R}^\Omega(z_0)\setminus \{z_0\} \subset \mathop{\rm Reg}(\{u_{\epsilon_i}\}) $. Then Theorem \ref{eps-reg-LL} and \ref{eps-reg-LL-bd} and Lemma \ref{Cinfty1} and \ref{Cinfty2} imply $$ \lim_{i\to\infty} \int_{B_R(x_0)\cap\Omega} g_{\epsilon_i}(u_{\epsilon_i}(x,s_0)) dx = \int_{B_R(x_0)\cap\Omega} \frac{1}{2} |\nabla u_* (x,s_0)|^2 dx < \frac{\epsilon_0}{4}\,. $$ Now by Lemma \ref{en-est-LL} \eqref{loc-en}, we have \begin{align*} \sup_{s_0\leq t\leq s_0 +\delta} \int_{B_{\frac{1}{2}R}(x_0)\cap \Omega} g_{\epsilon_i} (u_{\epsilon_i}(x,t)) \, dx &\leq \int_{B_{R}(x_0)\cap \Omega} g_{\epsilon_i} (u_{\epsilon_i}(x,s_0)) \,dx + \frac{\delta C E_0}{\gamma_1 R^2} \\ &< \frac{\epsilon_0}{2} + \frac{\epsilon_0}{2}\,, \end{align*} for $i$ sufficiently large. Since by construction $t_0 \in ]s_0,s_0+\delta[$, the claim follows. \end{proof} By Theorem \ref{limits}, Corollary \ref{prop-limits} (ii) and uniqueness of smooth solutions, we obtain the following. \begin{remark} \label{unique-Struwe} \rm There is $T_0 >0$, such that $$ \lim_{\epsilon \searrow 0} u_\epsilon = u_* \quad \mbox{in } C^\infty( \Omega \times ]0,T_0[; S^2) , $$ where $u_*$ is the unique smooth solution of \eqref{LL} with initial and boundary data $u_0$. (Compare \cite{Harpes1}.) \end{remark} If the energy of a (sub-)limit $u_*$ was everywhere decreasing, A.Freire's uniqueness result \cite{Freire1} would imply that $u_*$ is (globally) the Struwe-solution. However all we can say about the energy of sublimits $u_*$ is the following Lemma \ref{en-jump}. In particular extension $u_*$ after the maximal smooth existence time $T_0$ with backward bubbling cannot be excluded. (See \cite{Harpes3}.) \begin{lemm} \label{en-jump} Let $ u_\epsilon $ be a solution of \eqref{Eq_eps_LL}-\eqref{Eq_bdry_eps_LL} for fixed $ \epsilon > 0 $ and assume $ u_* = \mbox{weak-}H^{1,2}\mbox{-}\lim_ {i \to \infty} u_{\epsilon_i}$ for a sequence $0<\epsilon_i \searrow 0$. If $ s < t $ and $ \mathbb{S}^s(\{u_{\epsilon_i}\}) := (\overline{\Omega}\times \{s\}) \cap \mathbb{S}(\{ u_{\epsilon_i}\}_i ) = \emptyset $ and $ \mathbb{S}^t(\{u_{\epsilon_i}\}) \neq \emptyset $, then $$ \int_\Omega \frac{1}{2} |\nabla u_*|^2 (x,s) \, dx \geq \int_\Omega \frac{1}{2} |\nabla u_*|^2 (x,\tau) \, dx \quad \forall \tau > s \,, $$ and $$ \int_\Omega \frac{1}{2} |\nabla u_*|^2 (x,s) \, dx \geq \int_\Omega \frac{1}{2} |\nabla u_*|^2 (x,t) \, dx + \epsilon_0 \,, $$ where $ \epsilon_0>0 $ is the constant from Theorem \ref{eps-reg-LL-bd}. \end{lemm} \begin{proof} Set $ \overline{x} := (x_1, \ldots , x_K) $ if $ \mathbb{S}^t(\{u_{\epsilon_i}\}) = \{ x_1, \ldots , x_K \} $ and\\ $ B_R(\overline{x}) := \bigcup_{j=1}^K B_R(x_j) $. Then \begin{align*} E(u_*(s),\Omega ) &:= \int_\Omega \frac{1}{2} |\nabla u_* |^2 (x,s) \, dx\\ &= \lim_{i \to \infty} \int_{\Omega} g_{\epsilon_i}(u_{\epsilon_i})(x,s)\, dx\\ &\geq \limsup_{i \to \infty} \int_{\Omega} g_{\epsilon_i}(u_{\epsilon_i})(x,\tau) \, dx \quad \forall \tau > s \quad \mbox{ (by Lemma \ref{en-est-LL})} \\ &\geq \int_\Omega \frac{1}{2} |\nabla u_* |^2 (x,\tau) \, dx \quad \forall \tau > s \quad \mbox{(by weak lower semi-continuity)} \end{align*} Also $$ E(u_*(s),\Omega ) \geq \limsup_{i \to \infty} \Bigl(\int_{\Omega \smallsetminus B_R(\overline{x})} g_{\epsilon_i}(u_{\epsilon_i})(x,\tau) \, dx +\int_{B_R(\overline{x})} g_{\epsilon_i}(u_{\epsilon_i})(x,\tau) \, dx \Bigr) , $$ for all $\tau \in ]s,t]$. Now for any $\delta \in ]0,1[$ and $R>0$, there are sequences $s 0\,,\; \delta \in ]0,1[ \,. \end{align*} Since the last inequality holds for any $R>0$ and $ \delta \in ]0,1[$, the claim follows. \end{proof} Theorem \ref{limits} provides an alternative version of the construction of the ``Struwe-solution'' (see \cite{Harpes1}). \begin{coro} \label{existenceLL} Let $ u_0 \in H^{1,2}(\Omega;S^2)$. Then there is a global distributional solution $ u \in H^{1,2}_{\rm loc}(\overline{\Omega}\times ]0,\infty[;S^2) \cap L^\infty (]0,\infty[; H^{1,2}(\Omega;S^2)) \mbox{ with } \partial_t u \in L^2 (\Omega\times ] 0,\infty[;\mathbb{R}^3) $ of \eqref{LL} with initial and boundary data $u_0$, which is smooth on $\Omega\times ]0,\infty[$ except at finitely many points and has decreasing and right continuous energy. If in addition $u_0 \in H^{3/2,2}(\partial\Omega;S^2)$, then $u$ is unique among the solutions of \eqref{LL} with initial and boundary data $u_0$ which are smooth except for isolated singular points and with $\lim_{t\searrow s} E(u(t)) < E(u(s)) +\epsilon_0$ for all $ s \geq 0$. (It is also unique among the $ H^{1,2}_{\rm loc}$-solutions with decreasing energy by Freire's result.) \end{coro} \begin{proof} By Theorem \ref{limits} the $\epsilon$-approximation scheme provides a smooth short time solution $$ u \in C^\infty \bigl( \Omega \times ]0,T_0[; S^2 \bigr) $$ to \eqref{LL} with boundary data $u_0$ and $ \lim_{t\searrow 0} u(\cdot,t) = u_0$ in $H^{1,2}(\Omega;\mathbb{R}^3)$. Also there are $ \{ x_1, \ldots, x_K\} \subset \Omega $, such that $$ \lim_{t \nearrow T_0} u(\cdot,t) = u(\cdot,T_0) \quad \mbox{in } C^\infty \bigl(\Omega \smallsetminus \{ x_1, \ldots, x_K\}, \mathbb{R}^3\bigr) $$ and $$ \norm{\nabla u(\cdot,T_0)}^2_{L^2(\Omega)} \leq \liminf_{t \nearrow T_0} \norm{\nabla u(\cdot,t)}^2_{L^2(\Omega)}\leq 2 E_0 \,. $$ In particular $u(\cdot,T_0) \in H^{1,2}(\Omega)$. If we now set $ \tilde{u}_0 := u(\cdot,T_0) $ and repeat the same procedure with $ \tilde{u}_0$ instead of $u_0$, we obtain step by step a global solution with point singularities. To see that $ \partial_t u \in L^2 (\Omega \times \mathbb{R}_+;\mathbb{R}^3)$, we sum up the energy inequalities of each time intervall $]t_k, t_{k+1}[$ on which $u$ is regular and use that the energy is right continuous, whereas $$ \limsup_{t\nearrow t_{k+1}} E(u(t)) \geq E(u(t_{k+1})) + \epsilon_0 $$ by Lemma \ref{en-jump}. This yields $$ \int_0^\infty \int_\Omega |\partial_t u|^2 dx dt \leq E_0 - \sum_k \epsilon_0 $$ and also shows that there can only be finitely many ``singular times'' $t_k$. Now assume we have two solutions $u_1$ and $u_2$ of \eqref{LL} with initial and boundary data $u_0$ and both with finitely many point singularities and $\lim_{t\searrow s} E(u(t)) < E(u(s)) +\epsilon_0$ for all $ s \geq 0$. By Remark \ref{unique-Struwe}, we have $u_1 = u_2$ on $\Omega \times [0,T_1[$, where $T_1$ is the maximal commun smooth existence time, i.e. either $u_1$ or $u_2$ has point singularities at $T_1$. However by Corollary 4 on the existence of smooth extensions in \cite{Harpes1}, if $u_1$ admits a smooth extension up to $T_1$, then so does $u_2$ and conversely. Moreover, since the criterion for the existence of a smooth extension is local, both solutions have the same singularities $x_1,\ldots, x_K$ at time $T_1$ and $u_1(\cdot,T_1) = u_2(\cdot,T_1)$ on $\Omega \setminus \{x_1,\ldots, x_K\}$. By Theorem 6 in \cite{Harpes1}, and the assumption on the energy, the extension of $u_1$ and $u_2$ after $T_1$ is again unique ``for a short time'' and an iteration of the previous argument leads to the claimed uniqueness. \end{proof} \subsection*{Acknowledgement} The author gratefully acknowledges encouragement and support from Michael Struwe. \begin{thebibliography}{99} \bibitem{BBH1} B\'ethuel,~F., Br\'ezis,~H.,~H\'elein,F.: Asymptotics for the minimization of a Ginzburg-Landau functional. Calc.Var. {\bf 1}, 123-148 (1993) \bibitem{Chang} Chang,~K,C.: Heat flow and boundary value problem for harmonic maps. Ann. Inst. Henri Poincar\'e, Vol. 6, No. 5, 363-395 (1989) \bibitem{Chen1} Chen,~Y.: Weak solutions to the evolution problem for harmonic maps into spheres. Math.Z. {\bf 201}, 69-74 (1989) \bibitem{Chen2} Chen,~Y.: Dirichlet boundary value problems of Landau-Lifshitz equation. Comm. Partial Diff. Eqs, {\bf 25} (1\&2), 101-124 (2000) \bibitem{Chen-Ding-Guo} Chen,~Y. Ding,~S. Guo,~B.: Partial Regularity for two dimensional Landau-Lifshitz equations. Acta Mathematica Sinica, New Series, 1998 July, Vol 14, No 13, 423-432 (1998) \bibitem{Chen-Guo} Chen,~Y. Guo,~B.: Two dimensional Landau-Lifshitz equation. J.Partial Diff. Eqs., Vol.{\bf 9}, No.4, 313-322 (1996) \bibitem{Chen-Lin1} Chen,~Y. Lin,~F.H.: Evolution of harmonic maps with Dirichlet boundary conditions, Comm.Anal.Geom.{\bf 1 }, 327-346 (1993) \bibitem{Cheng} Cheng.~X.: Estimate of the singular set of the evolution problem for harmonic maps. J.Diff.Geom.{\bf 34}, 169-174 (1991) \bibitem{Ding-Guo1} Ding,~S. Guo,~B.: Initial boundary value problem for the Landau-Lifshitz system with applied field. J.Partial Diff. Eqs., Vol.{\bf 13}, No.1, 35-50 (2000) \bibitem{Ding-Guo2} Ding,~S. Guo,~B.: Initial boundary value problem for the Landau-Lifshitz system (I) - Existence and partial regularity. Progress in natural science, Vol {\bf 8}, No.1, 11-23, February (1998) \bibitem{Ding-Guo3} Ding,~S. Guo,~B.: Initial boundary value problem for the Landau-Lifshitz system (II) - Uniqueness. Progress in natural science, Vol {\bf 8}, No.2, 11-23, April (1998) \bibitem{Evans-Gariepy} Evans,~L.C., Gariepy,~R.F: Measure theory and fine properties of functions. Studies in advanced math. CRC press (1992) \bibitem{Freire1} Freire,~A.: Uniqueness for the harmonic flow from surfaces to general targets. Comment.Math.Helvetici {\bf 70}, 310-338 (1995) \bibitem{Freire2} Freire,~A.: Corrections to `` Uniqueness \dots ''. Comment.Math.Helvetici {\bf 71}, 330-337 (1995) \bibitem{Guo-Hong1} Guo,~B. Hong,~M.C.: The Landau-Lifshitz equation of the ferromagnetic spin chain and harmonic maps. Calc. Var.{\bf 1}, 311-334 (1993) \bibitem{Harpes0} Harpes,~P.: Ginzburg-Landau type approximations for the Landau-Lifshitz flow and the harmonic map flow in two space dimensions. Phd-thesis, Eth-Z\"urich, Diss.ETH No. 14430 (2001) \bibitem{Harpes1} Harpes,~P.: Uniqueness and bubbling of the 2-dimensional Landau-Lifshitz flow. Calc.Var. {\bf 20} (2), 213-229 (2004) \bibitem{Harpes3} Harpes,~P.: Bubbling of approximations for the 2-d Landau-Lifshitz flow. Preprint (2002) \bibitem{Helein} H\'elein,~F.: R\'egularit\'e des applications faiblement harmoniques entre une surface et une vari\'et\'e riemanienne. C.R.Acad.Sci.Paris, t.312, S\'erie I, p.591-596 (1991) \bibitem{Hong} Hong,~M.C.: The Landau-Lifshitz equation with the external field- a new extension for harmonic maps with values in $S^2$. Math.Z. {\bf 220} 171-188 (1995) \bibitem{LadySolUral} Ladyzenskaya,~O.A., Solonnikov,~V.A., Ural'ceva,~N.N.: Linear- and Quasilinear Equations of Parabolic type. Transl. Math. Monogr. \bibitem{LL} Landau,~L.D. Lifshitz,~E.N.M.: On the theory of the dispersion of magnetic permeability in ferromagnetic bodies. Phys. Z. Sowj. {\bf 8}, 153 (1935); ter Haar, D. (eds.) Reproduced in Collected Papers of L.D.Landau, 101-114. New York: Pergamon Press (1965) \bibitem{Lieberman} Lieberman,~G.M: Second order parabolic differential equations. World scientific (1996) \bibitem{Lin-Wang1} Lin,~F.H, Wang,~C.: Harmonic and quasi harmonic spheres. Comm. Analysis Geom., Vol 7, No 2, 397-429 (1999) \bibitem{Lin-Wang2} Lin,~F.H, Wang,~C.: Harmonic and quasi harmonic spheres, Part II. Comm. Analysis Geom., Vol10, No 2, 341-375 (2002) \bibitem{Lin-Wang3} Lin,~F.H, Wang,~C.: Harmonic and quasi harmonic spheres, Part III. Ann. I. H. Poincare-AN 19, 2, 209-259 (2002) \bibitem{Tristan} Rivi$\grave{e}$re,~T.: Flot des applications harmoniques en dimension deux. (1993) \bibitem{Schoen} Schoen,~R.: Analytic aspects of the harmonic map problem. Seminar on Nonlinear Partial Differential Equations (Cern, ed.), Springer (1984) \bibitem{Noebi} Simon,~L.: Theorems on Regularity and Singularity of Energy Minimizing Maps. Lectures in Mathematics ETH Z\"urich, Birkh\"auser \bibitem{Struwe1} Struwe,~M.: On the evolution of harmonic maps of Riemannian surfaces. Comment. Math. Helvetici. {\bf 60}, 558-581 (1985) \bibitem{Struwe2} Struwe,~M.: Variational Methods, Series of Modern Surveys in Math., Vol.34, 2nd edition, Springer (1991) \bibitem{Struwe3} Struwe,~M.: On the evolution of harmonic maps in higher dimension. J.Diff.Geom. {\bf 28}, 485-502 (1988) \bibitem{Struwe4} Struwe,~M.: Heat flow methods for harmonic maps of surfaces and applications to free boundary problems. Lect. Notes Math. {\bf 1324}, 293-319. Springer (1988) \bibitem{Struwe5} Struwe,~M.: Geometric Evolution Problems. IAS/Park City Mathematics series, Vol 2 (1996) \bibitem{Struwe6} Struwe,~M.: Singular pertubations of geometric variational problems. Elliptic and Parabolic Methods in Geometry, A.K.Peters, Ltd. (1996) \bibitem{Struwe-Chen} Struwe,~M., Chen,~Y.: Existence and Partial Regularity for the Heat flow for Harmonic maps. Math.~Z. {\bf 201}, 83-103 (1989) \bibitem{Li-Tian2} Tian,~G. Li,~J.: The blow up locus of heat flows for harmonic maps. Acta Mathematica Sinica, English Series, Vol.{\bf 16}, No.1,p. 29-62 (2000) \end{thebibliography} \end{document}