\input amstex \documentstyle{amsppt} \loadmsbm \hcorrection{18mm} \vcorrection{5mm} \nologo \TagsOnRight \NoBlackBoxes \headline={\ifnum\pageno=1 \hfill\else% {\tenrm\ifodd\pageno\rightheadline \else \leftheadline\fi}\fi} \def\rightheadline{EJDE--2001/29\hfil A priori estimates for global solutions \hfil\folio} \def\leftheadline{\folio\hfil Pavol Quittner \hfil EJDE--2001/29} \def\pretitle{\vbox{\eightrm\noindent\baselineskip 9pt % Electronic Journal of Differential Equations, Vol. {\eightbf 2001}(2001), No. 29, pp. 1--17.\hfil\break ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \hfill\break ftp ejde.math.swt.edu (login: ftp)\bigskip} } \topmatter \title A priori estimates for global solutions and multiple equilibria of a superlinear parabolic problem involving measures \endtitle \thanks {\it 2000 Mathematics Subject Classifications:} 35B45, 35J65, 35K60. \hfil\break\indent {\it Key words:} superlinear parabolic equation, semilinear elliptic equation, multiplicity, \hfil\break\indent singular solutions. \hfil\break\indent \copyright 2001 Southwest Texas State University. \hfil\break\indent Submitted December 14, 200. Published May 2, 2001. \endthanks \author Pavol Quittner \endauthor \address Pavol Quittner \hfill\break Institute of Applied Mathematics, Comenius University, Mlynsk\'a dolina, SK -- 84248 Bratislava, Slovakia \endaddress \email quittner\@fmph.uniba.sk \endemail \abstract We consider a noncoercive elliptic problem in a bounded domain with a power nonlinearity and measure data. It is known that the problem possesses a stable solution and we prove existence of three further solutions. The proof is based on uniform bounds of global solutions of the corresponding parabolic problem and on a topological degree argument. \endabstract \endtopmatter \document \def\eq#1{\ifmmode \eqno{(#1)}\else \hbox{\rm(#1)}\fi} \def\W#1#2{W^{#1}_{#2,\gamma}} \def\Tmax{T_{\hbox{\sevenrm max}}} \head{1. INTRODUCTION} \endhead In this paper we consider the problem $$ \gathered u_t = \Delta u+|u|^{p-1}u+\mu, \quad x\in\Omega,\ t>0, \\ u =0, \quad x\in\partial\Omega,\ t>0,\\ u(x,0) = u_0(x), \quad x\in\bar\Omega, \endgathered \eq{1.1} $$ where $\Omega\subset{\Bbb R}^n$ is a smoothly bounded domain with $n\geq2$, $\mu$ is a positive bounded Radon measure on $\Omega$ and $$ p>1,\quad p<\frac{n}{n-2}\ \hbox{ if }n>2.\eq{1.2} $$ The restriction $p(n-2)1$, $p(n-2)0\,:\, \hbox{\eq{1.1} has a positive equilibrium}\}. $$ It follows from \cite{7} (see also \cite{4} for a more general setting) that $a^*>0$. Assuming $$ 01$, $p(n-2)v_1>v_3$ and $v_4-v_1$ changes sign. Assumption \eq{1.4} is crucial also for the proof of a priori estimates of global solutions of \eq{1.1}: instead of estimating the singular solution $u(t)$ (which need not be even continuous, in general), we estimate the difference $w(t)=u(t)-v_1$ which turns out to be a H\"older continuous function. Let us mention that a priori estimates of global or periodic solutions of similar superlinear parabolic problems with regular data were already used for the proof of existence of positive stationary solutions (see \cite{17}, \cite{10}, \cite{24}), sign-changing stationary solutions (see \cite{19}, \cite{13}), infinitely many stationary solutions (see \cite{22}), periodic solutions (see \cite{11}, \cite{12}, \cite{16}), for establishing the blow-up rate of nonglobal solutions (see \cite{15}, \cite{14}), and for the study of the boundary of domains of attraction of stable equilibria (see \cite{18}). This paper is organized as follows. Section~2 deals with existence and regularity of solutions of \eq{1.1}. Main results of the paper are stated in Theorem~3.1 (a~priori estimates) and Theorem~4.5 (existence of multiple stationary solutions). %______________________________________________________________________________ \head{2. PRELIMINARIES} \endhead Let $q\in(1,\infty)$, $q'=q/(q-1)$, let $W^z_q(\Omega)$, $z\geq0$, denote the usual Sobolev-Slobodeckii space and $\gamma$ the trace operator, $\gamma:W^z_q(\Omega)\to W^{z-1/q}_q(\partial\Omega)$ for $z>1/q$. For $\theta\in I_q:=[-2,2]\setminus\{1/q+m\,:\,m\in{\Bbb Z}\}$ put $$ W^\theta:=\W\theta q:=\left.\cases \{u\in W^\theta_q(\Omega)\,:\,\gamma u=0 \} &\hbox{if }1/q<\theta,\\ W^\theta_q(\Omega) &\hbox{if }0\leq \theta<1/q,\\ \big(\W{-\theta}{q'}\big)' &\hbox{if }\theta<0, \endcases \quad \right\}\eq{2.1}$$ and let $|\cdot|_{\theta,q}$ denote the norm in $\W\theta q$. The norm in $\W 0q=L_q(\Omega)$ will be denoted simply by $|\cdot|_q$. The norm in the H\"older space $C^{0,\alpha}(\bar\Omega)$ will be denoted by $\|\cdot\|_{0,\alpha}$. Let $M=M(\Omega)$ be the space of bounded Radon measures on $\Omega$. The spaces $\W\theta q$ are ordered Banach spaces and $M(\Omega)$ is a Banach lattice (cf.~\cite{4, Section~5}). Moreover, $M(\Omega)\hookrightarrow\W\theta q$ provided $\theta<-n/q'$. For $u,v\in\W\theta q$, we write $u1,\quad z\geq0, \quad z\in I_q. \eq{2.4}$$ The existence of a unique $u$ satisfying (2.3) can be proved directly in the following way. Condition \eq{2.4} guarantees $\W zq\hookrightarrow L_p(\Omega)$ and $L_1(\Omega)\hookrightarrow M(\Omega)\hookrightarrow\W{z-2+\varepsilon}q$ for some $\varepsilon>0$, hence the mapping $F:\W zq\to\W{z-2+\varepsilon}q:u\mapsto |u|^{p-1}u+\mu$ is well defined and Lipschitz continuous. Now using \eq{2.2} we obtain $$\aligned |e^{-tA_z}u_0|_{z,q} &\leq C|u_0|_{z,q}\leq C, \\ |e^{-tA_z}F(u)|_{z,q} &\leq Ct^{-1+\varepsilon/2}|F(u)|_{z-2+\varepsilon,q} \\ &\leq Ct^{-1+\varepsilon/2}\big(1+|u|_{z,q}^p\big), \\ |e^{-tA_z}\big(F(u)-F(v)\big)|_{z,q} &\leq Ct^{-1+\varepsilon/2}\big(1+|u|_{z,q}^{p-1}+|v|_{z,q}^{p-1}\big) |u-v|_{z,q}. \endaligned \eq{2.5}$$ These inequalities easily imply that the operator $$ R(u)(t) = e^{-tA_z}u_0+ \int_0^te^{-(t-\tau)A_z} F\big(u(\tau)\big)\,d\tau$$ is a contraction in an appropriate ball of the Banach space $C\big([0,T],\W zq\big)$ if $T$ is small enough. The fixed point of $R$ is the solution of \eq{2.3}, hence of \eq{1.1}. Solutions of \eq{1.1} are not continuous, in general. Anyhow, if $u,v:[0,T]\to\W zq$ are two solutions of \eq{1.1} with initial conditions $u_0,v_0$, respectively, then the difference $w(t)=u(t)-v(t)$ is H\"older continuous for $t>0$ and its $C^{0,\alpha}(\bar\Omega)$-norm (where $\alpha>0$ is sufficiently small) can be estimated by the $\W zq$-norm of $w(0)$. More precisely, the following lemma is true. \proclaim{Lemma~2.1} Let $u,v:[0,T]\to\W zq$ be two solutions of \eq{1.1} with initial conditions $u(\cdot,0)=u_0$, $v(\cdot,0)=v_0$. Put $w=u-v$ and denote $$K_u=\sup_{\tau\in[0,T]}|u(\tau)|_{z,q}.\eq{2.6}$$ There exist $r>n$ and $\alpha>0$ such that $w(t)\in C^{0,\alpha}(\bar\Omega)\cap \W1r$ for any $t>0$ and $$ |w(t)|_{z,q}+|w(t)|_{1,r}+\|w(t)\|_{0,\alpha} \leq c(t_0,T,K_u\vee K_v)|w(0)|_{z,q} \eq{2.7}$$ for any $t\in[t_0,T]$ and $t_0>0$. Moreover, $w\in C^{0,\tilde\alpha}([t_0,T],C^{0,\alpha}(\bar\Omega)\cap \W1r)$ for some $\tilde\alpha>0$ and the norm of $w$ in this space can be bounded by a constant depending on $t_0,T,K_u\vee K_v$. \endproclaim \demo{Proof} Let $\tilde z,q$ satisfy \eq{2.4} (with $z$ replaced by $\tilde z$), $\tilde z>z$. Estimating the $\W {\tilde z}q$-norm in \eq{2.3} we obtain $$ \align |u(t)|_{\tilde z,q} &\leq Ct^{z-\tilde z}|u_0|_{z,q} +C\int_0^t e^{-c(t-\tau)}(t-\tau)^{-1+\tilde\varepsilon/2} \big|F\big(u(\tau)\big)\big|_{\tilde z-2+\tilde\varepsilon}\,d\tau \\ &\leq C(K_u)(1+t^{z-\tilde z}), \endalign $$ where $\tilde\varepsilon>0$ is small enough. Using the imbedding $\W{z_1}{q_1}\hookrightarrow\W{z_2}{q_2}$ if $z_1-\frac{n}{q_1}>z_2-\frac{n}{q_2}$ and $z_1\geq z_2$ and repeating the estimate above with different $z,\tilde z,q$, if necessary, we get $$ |u(t)|_{\tilde z,\tilde q} \leq C(\delta,K_u) \quad \hbox{for any } t\in[\delta,T],\eq{2.8}$$ whenever $\tilde z,\tilde q$ satisfy \eq{2.4}. Analogous estimates and the generalized Gronwall inequality \cite{1, Theorem~II.3.3.1} imply $$ |w(t)|_{\tilde z,\tilde q} \leq C(\delta,T,K_u,K_v)|w(0)|_{z,q} \quad \hbox{for any } t\in[\delta,T].\eq{2.9}$$ Estimates \eq{2.8} and \eq{2.9} imply that we may assume both $z=1$ and the boundedness of $u(\tau),v(\tau)$, $\tau\in[0,T]$, in $\W{\tilde z}{\tilde q}$ for any $\tilde z,\tilde q$ satisfying \eq{2.4}. In particular, $u(\tau),v(\tau)$ are bounded in $\W1q$ for any $q\frac n2$. Put $Q=q$, $R=1$ and choose $\beta>0$ and $\varepsilon>0$ small. Then $$ L_R(\Omega)\hookrightarrow \W{\beta-1+\varepsilon}Q \quad \hbox{and}\quad \W{1+\beta}Q\hookrightarrow L_{sR/(s-R)} $$ due to $$ \frac1R<\frac1Q+\frac{1-\beta}n \quad \hbox{and}\quad \frac1R\geq\frac1Q+\frac1s-\frac{1+\beta}n, \quad s>R, \eq{2.10}$$ hence $$\align |\Phi\big(u(\tau),v(\tau)\big)|_{\beta-1+\varepsilon,Q} &\leq C|\Phi\big(u(\tau),v(\tau)\big)|_R = C|w(\tau)\varphi(\tau)|_R \\ &\leq C|w(\tau)|_{sR/(s-R)}|\varphi(\tau)|_s \leq C|w(\tau)|_{1+\beta,Q} . \endalign $$ This and the variation-of-constants formula imply $$\align |w(t)|_{1+\beta,Q} &\leq Ct^{-\beta}|w_0|_{1,Q}+C\int_0^t(t-\tau)^{-1+\varepsilon/2} |\Phi\big(u(\tau),v(\tau)\big)|_{\beta-1+\varepsilon,Q}\,d\tau \\ &\leq Ct^{-\beta}|w_0|_{1,Q} + C\int_0^t(t-\tau)^{-1+\varepsilon/2} |w(\tau)|_{1+\beta,Q}, \endalign$$ so that the Gronwall inequality implies $$|w(t)|_{1+\beta,Q}\leq C_1(t)|w_0|_{1,Q},$$ where $C_1(t)$ is bounded for $t$ lying in compact subsets of $(0,T]$. Now $\W{1+\beta}Q \hookrightarrow\W1{\tilde Q}$ for some $\tilde Q>Q$. If $\tilde Q\leq n$ then repeating the estimates above with $Q$ replaced by $\tilde Q$, $R$ by $\tilde R$ and $\beta$ by $\tilde\beta$ such that \eq{2.10} remains true, we obtain $$|w(2t)|_{1+\tilde\beta,\tilde Q}\leq C_2(t)|w(t)|_{1,\tilde Q} \leq CC_1(t)C_2(t)|w_0|_{1,Q}.$$ A standard bootstrap argument yields the estimate of $w$ in $\W1{Q}$ for some $Q>n$, hence in $C^{0,\alpha}(\bar\Omega)$ for some $\alpha>0$. This shows \eq{2.7}. Notice that an upper bound for the bootstrap procedure is given by $1/Q>1/s-1/n$. The H\"older continuity of $w:[t_0,T]\to \W1Q$ for some $Q>n$ follows from the variation-of-constants formula, the estimates above and the estimates $$\align \Big|\int_{t_1}^{t_2}e^{-(t_2-\tau)A_z}\Phi(u(\tau),v(\tau))\,d\tau\Big|_{1,Q} &\leq C\int_{t_1}^{t_2}(t_2-\tau)^{-1+\varepsilon/2}\,d\tau \leq C(t_2-t_1)^{\varepsilon/2}, \\ |e^{-(t_2-t_1)A_z}f-f|_{1,Q} &\leq C(t_2-t_1)^{\beta/2}|f|_{1+\beta,Q}, \endalign $$ where $T\geq t_2>t_1\geq t_0$ and $f:=\int_0^{t_1}e^{-(t_1-\tau)A_z}\Phi(u(\tau),v(\tau))\,d\tau$ (cf.~also \cite{1, Theorem~II.5.3.1}). \qed\enddemo \proclaim{Remark~2.2}\rm Let $u$ and $v$ be solutions of \eq{1.1} on $[0,T]$ with initial conditions $u(\cdot,0)=u_0$ and $v(\cdot,0)=v_0$, where $u_0,v_0\in\W zq$ and $z,q$ satisfy \eq{2.4}. Let $\mu_k\to\mu$ in $M(\Omega)$ and $u_{0,k}\to u_0$, $v_{0,k}\to v_0$ in $\W zq$. Let $u_k,v_k$ be solutions of \eq{1.1} with $\mu$ replaced by $\mu_k$ and initial conditions $u_k(\cdot,0)=u_{0,k}$, $v_k(\cdot,0)=v_{0,k}$. Put $w=u-v$, $w_k=u_k-v_k$. Then the variation-of-constants formula, estimates \eq{2.5}, Gronwall's inequality and obvious modifications of estimates in the proof of Lemma~2.1 imply that $u_k,v_k$ are well defined on $[0,T]$ for $k$ large enough, $\sup_{t\in[0,T]}|u(t)-u_k(t)|_{z,q}\to0$ as $k\to\infty$ and $\sup_{t\in[t_0,T]}|w(t)-w_k(t)|_{1,r}\to0$ as $k\to\infty$ for some $r>n$ and any $t_0>0$. \endproclaim The following theorem follows from \cite{21, Theorem~3.1}. \proclaim{Theorem~2.3} Let $z,q$ satisfy \eq{2.4} and $u\in C\big([0,T),\W zq\big)$ be the maximal solution of \eq{1.1}. Let $u(t)$ be bounded in $L_r(\Omega)$ for $t\in[0,T)$, where $r>\frac{n}{2}(p-1)$, $r>1$. Then $T=+\infty$ and $u(t)$ is bounded in $\W zq$ for $t\in[0,\infty)$. \endproclaim %______________________________________________________________________________ \head{3. A PRIORI ESTIMATES} \endhead The main result of this section is the following \proclaim{Theorem~3.1} Assume \eq{1.2},\eq{1.3},\eq{1.4}. Let $u$ be a global solution of \eq{1.1} and let $z,q$ satisfy \eq{2.4}. Then $|u(t)|_{z,q}\leq c$, where $c$ depends only on the norm of $u_0$ in $\W zq$. \endproclaim \demo{Proof} Assumption \eq{1.4} guarantees existence of the minimal positive stationary solution $v_1$. Let $u$ be a global solution of \eq{1.1} and put $w(t):=u(t)-v_1$. The functions $v_1$ and $w$ are solutions of the following problems $$ \gathered 0 = \Delta v_1 + v_1^p + \mu, \quad x\in\Omega, \\ v_1 = 0, \quad x\in\partial\Omega, \endgathered \eq{3.1} $$ and $$ \gathered w_t = \Delta w+h(w), \quad x\in\Omega,\ t>0, \\ w = 0, \quad x\in\partial\Omega,\ t>0, \\ w(x,0) = u_0(x)-v_1(x), \quad x\in\bar\Omega \endgathered \eq{3.2} $$ where $$ h(w):=|w+v_1|^{p-1}(w+v_1)-v_1^p.$$ Due to Lemma~2.1, we have $$w(t)\in \W1r\hookrightarrow C^{0,\alpha}(\bar\Omega)$$ for some $r>n$, $\alpha>0$ and any $t>0$. Moreover, putting $$ f(w):= \frac1{p+1}\big(|w+v_1|^{p+1}-v_1^{p+1}\big)-wv_1^p, $$ the regularity of $w$ and the mean value theorem imply $$ |f(w)|\leq p\big(|w|+|v_1|\big)^{p-1}|w|^2 \leq C(w)\big(1+|v_1|^{p-1}\big) \in L_s(\Omega)$$ for some $s>\frac{n}{2}$, since $v_1\in L_r(\Omega)$ for any $r<\frac n{n-2}$ (see \cite{4}). Consequently, the energy functional $$ E(w):=\frac12\int_\Omega|\nabla w|^2\,dx-\int_\Omega f(w)\,dx\eq{3.3}$$ is well defined along the solution $w=w(t)$. Multiplying the equation in \eq{3.2} by $w$ and integrating by parts yields $$ \frac12\frac{d}{dt}\int_\Omega w(t)^2\,dx =-2E\big(w(t)\big)+\int_\Omega g\big(w(t)\big) \,dx,\eq{3.4}$$ where $$ g(w) = \frac{p-1}{p+1}|w+v_1|^{p+1}-|w+v_1|^{p-1}(w+v_1)v_1+wv_1^p+\frac2{p+1}v_1^{p+1}.$$ We shall show in Lemma~3.2 that there exist positive constants $c_0,c_1,\dots,c_4$ such that $$ \gathered g(w) \geq c_0|w|^{p+1}-c_1w^2v_1^{p-1}, \quad \\ c_2|w|^{p+1}+c_3w^2v_1^{p-1}\geq f(w) \geq c_4|w|^{p+1}. \endgathered \eq{3.5} $$ Assume that $\varepsilon>0$. Integrating inequalities in \eq{3.5} and using the estimate $$ \int_\Omega w^2v_1^{p-1}\,dx \leq |w|_{r_1}^2|v_1|_{r_2}^{p-1} = C|w|_{r_1}^2 \leq \varepsilon|\nabla w|_2^2+C_\varepsilon|w|_2^2 \eq{3.6}$$ (where $r_1<\frac{2n}{n-2}$ and $r_2<\frac{n}{n-2}$ are suitable exponents required by the corresponding H\"older inequality) one obtains $$ \gathered \int_\Omega g(w)\,dx \geq \int_\Omega\big(C_0 |w|^{p+1} -C_1 w^2-\varepsilon|\nabla w|^2\big)\,dx , \\ \int_\Omega\big(C_2 |w|^{p+1}+C_3 w^2+\varepsilon|\nabla w|^2\big)\,dx \geq \int_\Omega f(w)\,dx\geq \int_\Omega C_4 |w|^{p+1}\,dx , \endgathered \eq{3.7} $$ where $C_0,C_1,\dots,C_4$ are positive constants (and $C_1,C_3$ depend on $\varepsilon$). The choice of $r_1,r_2$ in \eq{3.6} is possible due to $$ 2\frac{n-2}{2n}+(p-1)\frac{n-2}n<1. $$ Now \eq{3.4}, \eq{3.7} and the choice $\varepsilon\leq1/4\wedge C_0/(8C_2)$ imply $$ \aligned \frac12\frac{d}{dt}\int_\Omega w(t)^2\,dx &\geq -2(1+2\varepsilon)E\big(w(t)\big)+\tilde C_0\int_\Omega|w|^{p+1}\,dx - \tilde C_1\int_\Omega w^2\,dx \\ &\geq -2(1+2\varepsilon)E\big(w(t)\big)+\hat C_0\Bigl(\int_\Omega w^2\,dx\Bigr)^{(p+1)/2} - \hat C_1. \endaligned \eq{3.8}$$ Let $t_12$. If $n=2$ then one can make a bootstrap argument based on maximal regularity as in \cite{20} to get a priori bound in $L_r(\Omega)$ for any $r>1$. Since it is not completely clear which estimate corresponds to \cite{20, (16)} in our case (and also for the reader's convenience) we repeat the whole argument from \cite{20}. We already know by \eq{3.11} that $$ -C\leq \frac12\int_\Omega|\nabla w(t)|^2\,dx-\int_\Omega f\big(w(t)\big)\,dx \leq C. \eq{3.16} $$ Moreover, \eq{3.15} implies $$ \sup_{t\geq t_0} \int_t^{t+1}|w(s)|_{p+1}^{(p+1)r}\,ds < C \eq{3.17} $$ for any $t_0>0$ and $r=2$. The interpolation theorem in \cite{10, Appendice}, \eq{3.13} and \eq{3.17} imply $$ \sup_{t\geq t_0}|w(t)|_\lambda < C \eq{3.18} $$ for any $$ \lambda<\lambda(r):=p+1-\frac{p-1}{r+1}. $$ Due to Theorem~2.3 and the definition of $w$, estimate \eq{3.18} guarantees the required bound in $\W zq$ if $$ \lambda(r)>\frac n2(p-1)=p-1, $$ or, equivalently, $$ pr$ (with the difference $\tilde r-r$ bounded away from zero). Thus, after finitely many steps we prove \eq{3.17} with some $r$ satisfying $2r+3>p$ which will conclude the proof. Hence, let \eq{3.17} be true for some $r\geq2$. Then \eq{3.18} is true for $\lambda<\lambda(r)$. Choose $\lambda\in\big(2,\lambda(r)\big)$ and denote $$ \theta=\frac{p+1}{p-1}\frac{\lambda-2}\lambda, \quad \lambda'=\frac\lambda{\lambda-1} \quad \hbox{ and }\quad p_1=\frac{p+1}p.$$ Then $\theta\in(0,1)$ and $\lambda'\in(p_1,2)$ due to $\lambda0$ small ($\delta_1\leq\delta/4$) such that $u$ and $w$ stay bounded on $(\tau_1,\tau_1+2\delta_1)$ in $\W 1{\tilde q}$ by a constant $C_6=C_6(C_5,{\tilde C}_5,\delta_1)$. By Lemma~2.1, $w(t)$ stays bounded in $C(\bar\Omega)$ on $(\tau_1+\delta_1,\tau_1+2\delta_1)$ by a constant $C_7=C_7(C_6,\delta_1)$. Since $v_1\in L_s(\Omega)$ for any $s$ and $|h(w)|\leq\tilde C|v_1|^{p-1}$ if $|w|\leq C$, the function $h$ stays bounded in $L_\rho(\Omega)$ (for some $\rho>p_1$) on $(\tau_1+\delta_1,\tau_1+2\delta_1)$ by a constant $C_8=C_8(C_7,v_1,\rho)$. Now standard estimates in the variation-of-constants formula for $w$ on $(\tau_1+\delta_1,\tau_1+2\delta_1)$ imply $$ |w(\tau_1+2\delta_1)|_{2-\varepsilon,\rho} \leq C_9, $$ where $C_9=C_9(C_8,\delta_1,\varepsilon)$ and $\varepsilon>0$ is small. Choose $\varepsilon<2/p_1-2/\rho$. Then $W^{2-\varepsilon}_\rho(\Omega)\hookrightarrow X_P$, where $X_P:=(E_0,E_1)_{1-1/P,P}$ is the real interpolation space between $E_0=L_{p_1}(\Omega)$ and $E_1=W^2_{p_1}(\Omega)$, and $P>1$ is arbitrary. Consequently, $$ \|w(\tau)\|_{X_P} \leq C_{10}, \eq{3.20} $$ where $\tau:=\tau_1+2\delta_1\in(t-\delta,t)$. Notice that given $t\geq t_0$ and $\delta\in(0,t_0/2)$ we have found $\tau\in(t-\delta,t)$ and $C_{10}=C_{10}(\delta,v_1,|u_0|_{z,q},P)$ such that \eq{3.20} is true and $C_{10}$ is independent of $w$ and $t$. We have $1-\theta=\frac2{p-1}\big(\frac{p+1}\lambda-1\big)<\frac2r$ for $\lambda$ sufficiently close to $\lambda(r)$ since the last inequality is satisfied for $\lambda=\lambda(r)$. Now choose $\tilde r>r$ such that $$ \beta:=\frac2{(1-\theta)\tilde r}>1$$ and notice that $\theta\tilde r\beta'>1$ where $\beta'=\beta/(\beta-1)$. Next we use \eq{3.7} and \eq{3.16}, then \eq{3.19}, H\"older's inequality, \eq{3.13}, maximal Sobolev regularity (see \cite{1, Theorem~III.4.10.7}), \eq{3.20} and inequality $|h(w)|\leq C(|w|^p+|v_1|^p)$ to get $$ \eqalign{ \int_\tau^{t+1}|w(s)|_{p+1}^{\tilde r(p+1)}\,ds &\leq C\Bigl(1+\int_\tau^{t+1}|w(s)|_{1,2}^{2\tilde r}\,ds\Bigr) \cr &\leq C\Bigl(1+\int_\tau^{t+1}|w_t(s)|_{p_1}^{\theta\tilde r} |w_t(s)|_2^{(1-\theta)\tilde r}\,ds\Bigr) \cr &\leq C\Bigl(1+ \Bigl(\int_\tau^{t+1}|w_t(s)|_{p_1}^{\theta\tilde r\beta'}\,ds \Bigr)^{1/\beta'} \underbrace{\Bigl(\int_\tau^{t+1}|w_t(s)|_2^2\,ds \Bigr)^{1/\beta}}_{\leq C}\Bigr) \cr &\leq C\Bigl(1+\Bigl( \int_\tau^{t+1}|h(w(s))|_{p_1}^{\theta\tilde r\beta'}\,ds \Bigr)^{1/\beta'} + \|w(\tau)\|_{X_P}^{\theta\tilde r}\Bigr) \cr &\leq C\Bigl(1+\Bigl( \int_\tau^{t+1}|w(s)|_{p+1}^{p\theta\tilde r\beta'}\,ds \Bigr)^{1/\beta'} \Bigr), }$$ where $P=\theta\tilde r\beta'$. Now we see that the last estimate implies \eq{3.17} with $\tilde r$ instead of $r$ provided $p\theta\tilde r\beta'\leq \tilde r(p+1)$, that is if $\theta\beta'\leq p_1$. This condition is equivalent to $$ p\leq \frac{\lambda(\tilde r-1)-\tilde r}{\tilde r-2}. \eq{3.21}$$ Considering $\tilde r\to r+$ and $\lambda\to\lambda(r)-$ we see that it is sufficient to verify $$ p(r-2)< \lambda(r)(r-1)-r,$$ which is equivalent to $(p-1)2r>0$. Consequently, the sufficient condition for bootstrap is satisfied and we are done. Note that the possibility of choosing $\tilde r-r$ bounded away from zero follows by an easy contradiction argument. \qed\enddemo \proclaim{Lemma~3.2} The functions $f,g$ from the proof of Theorem~3.1 satisfy \eq{3.5} for any $w\in{\Bbb R}$ and $v_1>0$. \endproclaim \demo{Proof} Since $f$ and $g$ can be viewed as positively homogeneous functions of two variables $w,v_1$ and $v_1>0$ one can put $v_1=1$. Consequently, we have to show $$ \align g_1(w) &\geq c_0|w|^{p+1}-c_1w^2, \\ c_2|w|^{p+1}+c_3w^2\geq f_1(w) &\geq c_4|w|^{p+1}, \endalign $$ where $$ \align f_1(w) &= \frac1{p+1}\big(|w+1|^{p+1}-1\big)-w, \\ g_1(w) &= \frac{p-1}{p+1}|w+1|^{p+1}-|w+1|^{p-1}(w+1)+w+\frac2{p+1}. \endalign $$ First let us show $f_1(w)\geq c_4|w|^{p+1}$. If $w>-1$ then $f_1'(w)=(w+1)^p-1$ has the same sign as $w$ and $f_1(w)=0$, hence $f_1(w)>0$ if $w>-1$, $w\ne0$. Obviously, $f_1(w)\geq-\frac1{p+1}-w>0$ if $w\leq-1$. Since $f_1(w)\approx\frac p2w^2$ as $w\to0$, there exists $\delta_1>0$ such that $f_1(w)\geq |w|^{p+1}$ for $|w|\leq\delta_1$. Since $f_1(w)/|w|^{p+1}\to\frac1{p+1}$ as $|w|\to\infty$, there exists $K_1>\delta_1$ such that $f_1(w)\geq\frac1{2(p+1)}|w|^{p+1}$ for $|w|\geq K_1$. The function $f_1$ is positive and continuous on the compact set $M_1:=[-K_1,-\delta_1]\cup[\delta_1,K_1]$, hence there exists $\varepsilon>0$ such that $f_1(w)\geq\varepsilon K_1^{p+1}\geq\varepsilon|w|^{p+1}$ for $w\in M_1$. Consequently, it is sufficient to choose $c_4=1\wedge\varepsilon\wedge\frac1{2(p+1)}$. The same arguments as above show $f_1(w)\leq c_2|w|^{p+1}+c_3w^2$ if $c_2,c_3$ are sufficiently large. The inequality for $g_1$ is equivalent to $G_1(w)+c_1w^2\geq G_2(w)$, where $$ \align G_1(w) &= \frac{p-1}{p+1}|w+1|^{p+1}+w+\frac2{p+1}, \\ G_2(w) &= |w+1|^{p-1}(w+1)+c_0|w|^{p+1}. \endalign $$ Fix $c_0<\frac{p-1}{p+1}$ and assume $c_1\geq1$. Since $G_1(w)-G_2(w)=o(w^2)$ as $w\to0$, there exists $\delta_2>0$ such that $G_1(w)+c_1w^2\geq G_2(w)$ for $|w|\leq\delta_2$ (and $\delta_2$ does not depend on $c_1\geq1$). Since $G_1(w)/G_2(w)\to\frac{p-1}{(p+1)c_0}>1$ as $|w|\to\infty$, there exists $K_2>\delta_2$ such that $G_1(w)\geq G_2(w)$ for $|w|\geq K_2$. Since the function $G_2$ is bounded on the compact set $M_2:=[-K_2,-\delta_2]\cup[\delta_2,K_2]$ by some constant $D_2$, the choice $c_1>D_2/\delta_2^2$ guarantees $c_1w^2\geq D_2\geq G_2(w)$ on $M_2$. \qed\enddemo %______________________________________________________________________________ \head{4. STATIONARY SOLUTIONS} \endhead In this section we consider the problem $$ \gathered 0 = \Delta u + |u|^{p-1}u + \mu, \quad x\in\Omega, \\ u = 0, \quad x\in\partial\Omega, \endgathered \eq{4.1} $$ where $\Omega\subset{\Bbb R}^n$ is a smoothly bounded domain, $n\geq2$, $p$ satisfies \eq{1.2}, and $\mu$ satisfies \eq{1.3} and \eq{1.4}. Recall that assumption \eq{1.4} guarantees the existence of the minimal positive solution $v_1$ of \eq{4.1}. We fix $z=1$ and $q$ satisfying \eq{2.4} and denote $$X=\W zq,\quad Y=\W{z-2}{q}\quad\hbox{and}\quad Z=L_p(\Omega).$$ Notice that $X\hookrightarrow Z\hookrightarrow L_1(\Omega)\hookrightarrow M(\Omega)\hookrightarrow Y$. Recall also from Section~2 that $A:=A_z:X\to Y$ is a linear isomorphism and denote $$ F(u)=|u|^{p-1}u+\mu \quad \hbox{and}\quad S=A^{-1}F.$$ The results of \cite{4} imply that $A^{-1}\geq0$, $F:Z\to Y$ and $S:Z\to X$ are nondecreasing, $S$ is compact. The solutions of \eq{4.1} correspond to the fixed points of the operator $S|_X:X\to X$. We denote by ${\Cal E}$ the set of all solutions of \eq{4.1}. In our study we shall use also the semiflow generated by problem \eq{1.1}. The considerations in Section~2 imply that this semiflow can be considered both in $X$ and in $Z$. Due to \cite{4, Theorem~5.1} and \cite{1, Theorem~II.6.4.1}, this semiflow is order preserving. We call $u\in Z$ a {\bf supersolution} of \eq{4.1} if $u\geq S(u)$ and $(1-e^{-tA})(u-S(u))\geq 0$ for all $t>0$. If $u\in X$ then these conditions may be replaced by a single condition $Au\geq Fu$: this follows from the following facts: $A^{-1}\geq0$, $e^{-tA}\geq0$, $\frac{1-e^{-tA}}t w\to Aw$ if $t\to0$, $w\in X$, and $(1-e^{-tA})w=\int_0^te^{-sA}Aw\,ds\geq0$ if $w\in X$, $Aw\geq0$. The subsolution is defined in an analogous way. One of the basic properties of sub- and supersolutions is formulated in the following \proclaim{Proposition 4.1} If $u^+\in Z$ is a supersolution of \eq{4.1} and $u_0\in X$, $u_0\leq u^+$, then the solution $u:[0,\Tmax)\to X$ of \eq{1.1} satisfies $u(t)\leq u^+$ for any $t\in[0,\Tmax)$, where $\Tmax$ is the maximal existence time of this solution. Analogous assertion is true for subsolutions. \endproclaim \demo{Proof} The solution $u$ can be (locally) obtained as the limit of the sequence $\{u_k\}$, where $u_1(t)\equiv u_0$ and $$ u_{k+1}(t) = e^{-tA}u_0+\int_0^te^{-(t-s)A}F\big(u_k(s)\big)\,ds.$$ (cf.~the existence proof in Section~2). We shall show by induction that $u_k(t)\leq u^+$. Obviously, $u_1(t)\leq u^+$. Hence assume $u_k(t)\leq u^+$. Then $F(u_k(s))\leq F(u^+) = AS(u^+)$, so that $$ \eqalign{ u_{k+1}(t) &\leq e^{-tA}u^+ + \int_0^te^{-(t-s)A}AS(u^+)\,ds \cr &= e^{-tA}u^+ + S(u^+) - e^{-tA}S(u^+) \leq u^+.\ \qed }$$ \enddemo In what follows, we shall construct a subsolution $v_\varepsilon$ and a supersolution $v^\varepsilon$ such that $$ v^\varepsilon\geq v_1+\varepsilon,\quad v_\varepsilon\leq-\varepsilon\quad \hbox{and}\quad [v_\varepsilon,v^\varepsilon]\cap{\Cal E}=\{v_1\}.$$ Due to \cite{4, Section~12}, the operator $F:Z\to Y$ is of the class $C^1$ and the operator $u\mapsto u-S'(v_1)u$ is an isomorphism considered both as an operator $X\to X$ and $Z\to Z$. Consequently, $v_1$ is an isolated stationary solution of \eq{1.1} both in $X$ and in $Z$. Similarly, the operator $\tilde F:Z\to Y:u\mapsto |u|^p+\mu$ is $C^1$ and $\tilde F'(v_1)=F'(v_1)$, hence the implicit function theorem guarantees the unique solvability of the equation $ u=A^{-1}\tilde F(u+\varepsilon) $ in the neighbourhood of $v_1$ if $\varepsilon>0$ is small enough. Denoting this solution by $u^\varepsilon$, the function $v^\varepsilon:=u^\varepsilon+\varepsilon$ satisfies $0=\Delta v^\varepsilon+|v^\varepsilon|^p+\mu$ in $\Omega$ and $v^\varepsilon=\varepsilon$ on $\partial\Omega$. Since $\Delta v^\varepsilon\leq0$, we have $u^\varepsilon=v^\varepsilon-\varepsilon\geq0$, hence $u^\varepsilon=S(u^\varepsilon+\varepsilon)$ and $v^\varepsilon-S(v^\varepsilon)=\varepsilon$. Consequently, $v^\varepsilon$ is a supersolution of problem \eq{4.1}. Next we show that $$ \hbox{any positive supersolution $v^+$ of \eq{4.1} fulfils $v^+\geq v_1$.} \eq{4.2}$$ Indeed, assuming the contrary and denoting $y=v_1\wedge v^+$ we have $S(y)\leq S(v_1)=v_1$ and $S(y)\leq S(v^+)\leq v^+$, hence $S(y)\leq y$. Since $0$ is a subsolution of \eq{4.1}, and the operator $S:[0,y]\to[0,y]$ is nondecreasing and compact, there exists a solution of the problem $u=S(u)$ in the order interval $[0,y]$ (see \cite{3, Corollary~6.2}). Since $y0$ is small enough. Since $\Delta v_\varepsilon=-v_\varepsilon|v_\varepsilon|^{p-1}>0$ in $\Omega$ and $v_\varepsilon=-\varepsilon$ on $\partial\Omega$, we have $v_\varepsilon\leq-\varepsilon$ in $\bar\Omega$. As before, the order interval $[v_\varepsilon,0]$ does not contain any solution of \eq{4.3} different from $0$ if $\varepsilon$ is small enough. Next we show that $$ \hbox{any supersolution $v^+$ of \eq{4.1} satisfying $v^+\geq v_\varepsilon$ fulfils $v^+\geq v_1$.} \eq{4.4}$$ Assume the contrary and let $v^+\geq v_\varepsilon$ be a supersolution of \eq{4.1}, $v^+\not\geq v_1$. Then $v^+$ cannot be positive due to \eq{4.2}. Since $v^+$ is also a supersolution of \eq{4.3}, the function $y:=v^+\wedge0<0$ is a supersolution of \eq{4.3} as well. Consequently, we can find a solution of \eq{4.3} between the subsolution $v_\varepsilon$ and the supersolution $y$, a contradiction. Notice that \eq{4.4} and $[v_1,v^\varepsilon]\cap{\Cal E}=\{v_1\}$ imply $[v_\varepsilon,v^\varepsilon]\cap{\Cal E}=\{v_1\}$. Now denote $D_A$ the domain of attraction of the equilibrium $v_1$ (that is $D_A$ is the set of all initial conditions $u_0\in X$ for which the solution $u(t)$ of \eq{1.1} exists globally and tends to $v_1$ in $X$ as $t\to+\infty$). Summarizing the considerations above we obtain the following \proclaim{Lemma~4.2} The sets $\{v\in X\,:\,v\geq v_\varepsilon\}$ and $\{v\in X\,:\,v\leq v^\varepsilon\}$ are invariant under the semiflow defined by \eq{1.1}. The set $[v_\varepsilon,v^\varepsilon]\cap X$ is a subset of $D_A$. The set $D_A$ is open in $X$. \endproclaim \demo{Proof} The invariance follows from Proposition~4.1. If $u_0\in[v_\varepsilon,v^\varepsilon]$ then the corresponding solution $u(t)$ of \eq{1.1} stays in the same order interval, hence it is global due to Theorem~2.3. The existence of the Lyapunov functional (see \eq{3.10}), the boundedness of $u(t)$ and the compactness of the semiflow imply that the $\omega$-limit set $\omega(u_0)$ of this solution is a nonempty compact set consisting of equilibria. Since $\omega(u_0)\subset[v_\varepsilon,v^\varepsilon]$ and $[v_\varepsilon,v^\varepsilon]\cap{\Cal E}=\{v_1\}$, we have $\omega(u_0)=\{v_1\}$. Now let $u_0\in D_A$, $\delta>0$, $K:=\|v_1\|_X+1$ and $\eta=\frac\varepsilon{2c}$, where $c=c(\delta,\delta,K)$ is the constant from \eq{2.7}. Since $u(t)\to v_1$ in $X$ as $t\to\infty$, there exists $t_1>0$ such that $\|u(t)-v_1\|_X\leq\eta$ if $t\geq t_1$. Put $T=t_1+\delta$. Then \eq{2.7} implies $$\|u(T)-v_1\|_{0,\alpha}\leq c(\delta,\delta,K)\|u(t_1)-v_1\|_X\leq\frac\varepsilon2.$$ If $y_0\in X$, $\|y_0-u_0\|_X\leq\frac12$, and $y$ is the solution of \eq{1.1} with the initial condition $y_0$, then there exists $\beta>0$ (independent of $y_0$) such that $y(t)$ exists and $\|y(t)-u(t)\|_X\leq1$ for $t\leq\beta$. If we require $\|y_0-u_0\|_X\leq\frac\varepsilon{2c}$, where $c=c(\beta,T,K_u+2)$ is the constant from \eq{2.7}, then \eq{2.7} implies existence of $y(t)$ for $t\leq T$ and the estimate $$\|y(T)-u(T)\|_{0,\alpha}\leq c(\beta,T,K_u+2)\|u_0-y_0\|_X\leq\frac\varepsilon2,$$ hence $$\|y(T)-v_1\|_{0,\alpha}\leq \|y(T)-u(T)\|_{0,\alpha}+\|u(T)-v_1\|_{0,\alpha}\leq\varepsilon,$$ so that $y(T)\in[v_\varepsilon,v^\varepsilon]\cap X\subset D_A$. This implies that the set $D_A$ is open in $X$. \qed\enddemo \proclaim{Lemma~4.3} Let $V$ be a finite dimensional subspace of $X$. Then the set $D_A\cap V$ is bounded. \endproclaim \demo{Proof} We shall proceed by contradiction. Assume $u_k\in D_A\cap V$, $\|u_k\|_X\to\infty$ as $k\to\infty$. Then $\|u_k-v_1\|_X\to\infty$ as well. Since the solution $U_k(t)$ of \eq{1.1} with the initial condition $u_k$ fulfils $U_k(t)-v_1\in \W12$ for $t>0$ and $\W12\hookrightarrow X$, we may assume $u_k-v_1\in\W12$ and $\alpha_k:=|u_k-v_1|_{1,2}\to\infty$. Denote $w_k=(u_k-v_1)/\alpha_k$, $A_k=\int_\Omega|\nabla w_k|^2\,dx$, $B_k=\int_\Omega|w|_k^{p+1}\,dx$. The sequence $w_k$ is bounded in $\W12$ and belongs to a finite dimensional subspace, hence it is relatively compact and we may assume $w_k\to w$ in $\W12$, $|w|_{1,2}=1$. Moreover, we have $A_k\leq1$, $B_k\to \int_\Omega|w|^{p+1}\,dx>0$. Using \eq{3.3},\eq{3.7} we obtain $$ E(u_k-v_1) = E(\alpha_k w_k) \leq \alpha_k^2A_k-\alpha_k^{p+1}C_4B_k \leq -\hat C_1 $$ for $k$ sufficiently large, where $\hat C_1$ is the constant from \eq{3.8}. Since $t\mapsto E\big((U_k-v_1)(t)\big)$ is nonincreasing, \eq{3.8} implies $$ \frac12\frac{d}{dt}\int_\Omega(U_k-v_1)^2\,dx \geq \hat C_0\Bigl(\int_\Omega(U_k-v_1)^2\,dx\Bigr)^{(p+1)/2}+\hat C_1,$$ so that $U_k$ cannot exist globally, a contradiction. \qed\enddemo In what follows put $$ V=\{v_1+\alpha_1y_1+\alpha_2y_2\,:\,\alpha_1,\alpha_2\in{\Bbb R}\}, $$ where $y_1,y_2\in X$ are continuous functions such that $y_1>0$ and $y_2$ changes sign. Denote ${\Cal E}^+=\{v\in{\Cal E}\,:\,v>v_1\}$ and ${\Cal E}^-=\{v\in{\Cal E}\,:\,v0$, inequality \eq{2.7} implies $\|u(t_k+\delta)-v^+\|_{0,\alpha}\to0$ as $k\to\infty$, hence $$ u(t_k+\delta)\geq v^+-\frac\varepsilon2\geq v_\varepsilon+\frac\varepsilon2\geq v_\varepsilon \eq{4.5}$$ for $k$ large enough. Fix $k$ and put $T=t_k+\delta$. We have $u(t)\geq v_\varepsilon$ for $t\geq T$, hence $\omega(u_0)\subset{\Cal E}^+$. Moreover, if $y$ denotes the solution of \eq{1.1} with the initial condition $y_0\in\partial_VD_A$ then \eq{2.7} implies $$ \|(u-y)(T)\|_{0,\alpha}\leq \frac\varepsilon2 \quad\hbox{provided}\quad \|u_0-y_0\|_X<\eta,\eq{4.6}$$ where $\eta>0$ is small enough (cf.~the proof of Lemma~4.2). Estimates \eq{4.5} and \eq{4.6} imply $y\geq v_\varepsilon$, hence $\omega(y_0)\subset{\Cal E}^+$. Consequently, the set $\partial_VD_A^+$ is open in $\partial_VD_A$. The proofs in the case $\partial_VD_A^-$ are analogous. \qed\enddemo \proclaim{Theorem~4.5} Assume \eq{1.2}, \eq{1.3}, \eq{1.4} and let $v_1$ be the minimal positive solution of \eq{4.1}. Then there exist solutions $v_2,v_3,v_4$ of \eq{4.1} such that $v_2>v_1>v_3$ and the function $v_4-v_1$ changes sign. \endproclaim \demo{Proof} The proof is similar to the corresponding proof in \cite{19}. Due to Lemmata~4.2-4.3, there exists $u_0\in\partial D_A$, $u_0>v_1$. The a priori estimates from Section~3 guarantee that the $\omega$-limit set $\omega(u_0)$ is a nonempty compact subset consisting of equlibria (cf.~the proof of Lemma~4.2). Proposition~4.1 and \eq{4.4} imply $\omega(u_0)\subset{\Cal E}^+$, hence there exists $v_2\in{\Cal E}^+$. Similarly one obtains the existence of $v_3\in{\Cal E}^-$. The existence of $v_4$ will be shown by contradiction. Hence, assume the contrary; then the compact set $\partial_VD_A$ can be written as a union of two open disjoint subsets $\partial_VD_A^\pm$. Consequently, both $\partial_VD_A^+=\partial_VD_A\setminus\partial_VD_A^-$ and $\partial_VD_A^-$ are compact and their distance is positive. Moreover, $\partial_VD_A^\pm\cap\{v_1\mp\lambda y_1\,:\,\lambda>0\}=\emptyset$ so that the following homotopy $$ H(t,u) =\cases v_1+(1-2t)(u-v_1)+2ty_1,& u\in \partial_VD_A^+,\ t\in[0,1/2]\\ v_1+(1-2t)(u-v_1)-2ty_1,& u\in \partial_VD_A^-,\ t\in[0,1/2]\\ v_1+(2-2t)y_1 +(2t-1)y_2,& u\in \partial_VD_A^+,\ t\in[1/2,1]\\ v_1+(2-2t)(-y_1)+(2t-1)y_2,& u\in \partial_VD_A^-,\ t\in[1/2,1], \endcases $$ fulfils $H(t,u)\ne v_1$ for $u\in\partial_VD_A$. The homotopy invariance property of the topological degree in $V$ yields $$ 1 = \hbox{deg}\,(H(0,\cdot),v_1,D_A\cap V) = \hbox{deg}\,(H(1,\cdot),v_1,D_A\cap V) = 0,$$ which is a contradiction. \qed\enddemo %______________________________________________________________________________ \proclaim{\bf Acknowledgement} \rm The author was partially supported by the Swiss National Science Foundation and by VEGA Grant 1/7677/20. \endproclaim %______________________________________________________________________________ \Refs \widestnumber\key{23} % \ref\key{1} \by H. Amann \book Linear and Quasilinear Parabolic Problems \bookinfo Volume I: Abstract Linear Theory \publ Birk\-h\"auser %Verlag \publaddr Basel %-Boston-Berlin \yr 1995 \endref % \ref\key{2} \by H. Amann \paper Parabolic evolution equations and nonlinear boundary conditions \jour J. Differ. Equations \vol 72 \yr 1988 \pages 201-269 \endref % \ref\key{3} \by H.~Amann \paper Fixed point equations and nonlinear eigenvalue problems in ordered Banach spaces \jour SIAM Review \vol 18 \yr 1976 \pages 620-709 \endref % \ref\key{4} \by H. Amann and P. Quittner \paper Elliptic boundary value problems involving measures: existence, regularity, and multiplicity \jour Advances in Diff. Equations \yr 1998 \vol 3 \pages 753-813 \endref % \ref\key{5} \by A.~Bahri and H.~Berestycki \paper A perturbation method in critical point theory and applications \jour Trans. Amer. Math. Soc. \vol 267 \yr 1981 \pages 1-32 \endref % \ref\key{6} \by A.~Bahri and P.L.~Lions \paper Morse index of some min-max critical points I. Application to multiplicity results \jour Commun. Pure \& Appl. Math. \vol 41 \yr 1988 \pages 1027-1037 \endref % \ref\key{7} \by P.~Baras and M.~Pierre \paper Crit\`ere d'existence de solutions positives pour des \'equations semi-lin\'eaires non monotones \jour Analyse Non Lin\'eaire, Ann.\ Inst.\ H.~Poincar\'e \vol 2 \yr 1985 \pages 185-212 \endref % \ref\key{8} \by T. Bartsch, K.-C. Chang and Z.-Q. Wang \paper On the Morse indices of sign changing solutions of nonlinear elliptic problems \jour Math. Z. \vol 233 \yr 2000 \pages 655-677 \endref % \ref\key{9} \by Ph. Bolle, N. Ghoussoub and H. Tehrani \paper The multiplicity of solutions in non-homogeneous boundary value problems \jour Manuscripta Math. \vol 101 \yr 2000 \pages 325-350 \endref % \ref\key{10} \by T. Cazenave and P.-L. Lions \paper Solutions globales d'\'equations de la chaleur semi li\-n\'eaires \jour Comm. Partial Diff. Equ. \vol 9 \yr 1984 \pages 955-978 \endref % \ref\key{11} \by M.J. Esteban \paper On periodic solutions of superlinear parabolic problems \jour Trans. Amer. Math. Soc. \vol 293 \yr 1986 \pages 171-189 \endref % \ref\key{12} \by M.J. Esteban \paper A remark on the existence of positive periodic solutions of superlinear parabolic problems \jour Proc. Amer. Math. Soc. \vol 102 \yr 1988 \pages 131-136 \endref % \ref\key{13} \by M. Fila and P. Quittner \paper Global solutions of the Laplace equation with a nonlinear dynamical boundary condition \jour Math. Meth. Appl. Sci. \vol 20 \yr 1997 \pages 1325-1333 \endref % \ref\key{14} % Theorem B \by S. Filippas and F. Merle \paper Modulation theory for the blowup of vector-valued nonlinear heat equations \jour J. Differ. Equations \vol 116 \yr 1995 \pages 119-148 \endref % \ref\key{15} % Theorem 3.1 \by Y. Giga and R.V. Kohn \paper Characterizing blowup using similarity variables \jour Indiana Univ. Math. J. \vol 36 \yr 1987 \pages 1-40 \endref % \ref\key{16} \by N. Hirano and N. Mizoguchi \paper Positive unstable periodic solutions for superlinear parabolic equations \jour Proc. Amer. Math. Soc. \vol 123 \yr 1995 \pages 1487-1495 \endref % \ref\key{17} \by W.-M. Ni, P.E. Sacks and J. Tavantzis \paper On the asymptotic behavior of solutions of certain quasilinear parabolic equations \jour J. Differ. Equations \vol 54 \yr 1984 \pages 97-120 \endref % \ref\key{18} \by P. Pol\'a\v cik \paper Domains of attraction of equilibria and monotonicity properties of convergent trajectories in parabolic systems admitting strong comparison principle \jour J. reine angew. Math. \vol 400 \yr 1989 \pages 32-56 \endref % \ref\key{19} \by P. Quittner \paper Signed solutions for a semilinear elliptic problem \jour Differ. Integral Equations \vol 11 \yr 1998 \pages 551-559 \endref % \ref\key{20} \by P. Quittner \paper A priori bounds for global solutions of a semilinear parabolic problem \jour Acta Math. Univ. Comenianae \vol 68 \yr 1999 \pages 195-203 \endref % \ref\key{21} \by P. Quittner \paper Global existence for semilinear parabolic problems \jour Adv. Math. Sci. Appl. \vol 10 \yr 2000 \pages 643-660 \endref % \ref\key{22} \by P. Quittner \paper Boundedness of trajectories of parabolic equations and stationary solutions via dynamical methods \jour Diff. Integral Equations \vol 7 \yr 1994 \pages 1547-1556 \endref % \ref\key{23} \by P.H. Rabinowitz \paper Multiple critical points of perturbed symmetric functionals \jour Trans. Amer. Math. Soc. \vol 272 \yr 1982 \pages 753-770 \endref % \ref\key{24} \by Ph. Souplet and Q.S. Zhang \paper Existence of ground states for semilinear elliptic equations with decaying mass: a parabolic approach \finalinfo Preprint \endref % \ref\key{25} \by M. Struwe \paper Infinitely many critical points for functionals which are not even and applications to superlinear boundary value problems \jour Manuscripta Math. \vol 32 \yr 1980 \pages 335-364 \endref % \ref\key{26} \by L. Veron \book Singularities of Solutions of Second Order Quasilinear Equations \bookinfo Pitman Research Notes in Math. \#353 \publ Longman Sci. \& Tech. \yr 1996 \endref % \ref\key{27} \by Z.-Q. Wang \paper On a superlinear elliptic equation \jour Ann. Inst. Henri Poincar\'e - Anal. non lin\'eaire \vol 8 \yr 1991 \pages 43-57 \endref % \endRefs \enddocument %______________________________________________________________________________