\documentstyle[twoside]{article} \pagestyle{myheadings} \markboth{\hfil HOMOCLINIC ORBITS \hfil EJDE--1994/01}% {EJDE--1994/01\hfil P. Korman \& A.C. Lazer\hfil} \begin{document} \ifx\Box\undefined \newcommand{\Box}{\diamondsuit}\fi \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}\newline Vol. {\bf 1994}(1994), No. 01, pp. 1-10. Published February 15, 1994.\newline ISSN 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp (login: ftp) 147.26.103.110 or 129.120.3.113 } \vspace{\bigskipamount} \\ Homoclinic Orbits for a Class of Symmetric Hamiltonian Systems \thanks{ {\em 1991 Mathematics Subject Classifications:} 34B15, 34A34.\newline\indent {\em Key words and phrases:} Homoclinic orbits, mountain-pass lemma. \newline\indent \copyright 1994 Southwest Texas State University and University of North Texas.\newline\indent Submitted: October 22, 1993.\newline\indent Supported in part by the Taft Faculty Grant at the University of Cincinnati (P. K.)\newline\indent Supported in part by NSF under Grant DMS-91023 (A. C. L.)} } \date{} \author{Philip Korman\\ and \\ Alan C. Lazer} \newtheorem{thm}{Theorem}[section] \newtheorem{lma}{Lemma}[section] \newtheorem{cor}{Corollary} \newtheorem{ex}{Example} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \newcommand{\beq}{\begin{equation}} \newcommand{\eeq}{\end{equation}} \newcommand{\s}{\;\;\;} \newcommand{\ol}{\overline} \maketitle \begin{abstract} We study existence of homoclinic orbits for a class of Hamiltonian systems that are symmetric with respect to independent variable (time). For the scalar case we prove existence and uniqueness of a positive homoclinic solution. For the system case we prove existence of symmetric homoclinic orbits. We use variational approach. \end{abstract} \section{Introduction} \setcounter{equation}{0} Recently variational techniques have been applied to obtain existence of homoclinic orbits of the Hamiltonian systems \beq u'' - L(t)u + V_u(t,u) = 0 \, , \label{1.1} \eeq see e.g. P.H. Rabinowitz \cite{4} and W. Omana and M. Willem \cite{3}. Here $L(t)$ is a positive definite $n\times n$ matrix, $V$ is assumed to be superquadratic at infinity and subquadratic at zero in $u\,$, and the solution $u(t) \in H^1(R,R^n)$ is homoclinic at zero, i.e., $\lim_{t\rightarrow\pm\infty} u(t) = 0 \,$. The technical difficulty in applying the mountain pass lemma on the infinite interval $t\in(-\infty,\infty)\,$, is in verifying the Palais-Smale or ``compactness'' condition. In \cite{3} a new compact embedding theorem was used to verify the Palais-Smale condition. However, one had to assume that the smallest eigenvalue of $L(t)$ tends to $\infty$ as $|t|\rightarrow\infty\,$, which is a rather restrictive and not very natural condition, as it excludes e.g. the case of constant $L \,$. In this paper we show that this assumption is not necessary in case $L(t)$ and $V(t,u)$ are even in $t\,$. For the scalar case we also show existence and uniqueness of a positive homoclinic orbit. We begin by considering the equation (\ref{1.1}) on a finite interval $(-T,T)$ together with the boundary conditions \beq u(-T) = u(T) = 0 \, . \label{1.2} \eeq Using the mountain pass lemma we show that the problem (\ref{1.1}-{1.2}) has a nontrivial solution. Moreover, using variational approach, we derive uniform in $T$ estimate of $H^1$ norm of the solution. This is the crucial step, which allows us to obtain homoclinic orbits by letting $T\rightarrow\infty \,$. \section{Positive homoclinics for a scalar equation} \setcounter{equation}{0} In this section we will prove existence and uniqueness of positive homoclinics for a model problem with a cubic nonlinearity. Namely, we are looking for a positive solution of \beq u'' - a(x)u + b(x)u^3 = 0 \,, \s -\infty < x < \infty \, , \label{2.1} \eeq \beq u(-\infty) = u(\infty) = u'(-\infty) = u'(\infty) = 0 \, . \label{2.2} \eeq We assume that the functions $a(x), b(x) \in C^1(-\infty,\infty)$ are strictly positive on $(-\infty,\infty)\,$, i.e. $a(x) \geq a_0 > 0$ and $b(x) \geq b_0 > 0$ and moreover we assume that $a(x)$ and $b(x)$ are even with respect to some real number $c \,$. Without loss of generality we will assume that $c=0\,$, i.e. $a(x)$ and $b(x)$ are even functions. We assume additionally that $xa'(x) > 0$ and $xb'(x) < 0$ for all $x \neq 0 \,$. We shall obtain the solution of (\ref{2.1}-\ref{2.2}) as the limit as $T \rightarrow \infty$ of the solutions of \beq u'' - a(x)u + b(x)u^3 = 0 \s \mbox{for} \s x\in(-T,T), \s u(-T) = u(T) = 0 \, . \label{2.3} \eeq We shall need the following lemma from P. Korman and T. Ouyang \cite{2}. (Except for the last assertion, this lemma is also included in B. Gidas, W.-M. Ni and L. Nirenberg \cite{1}). \begin{lma} Consider the problem \beq u'' + f(x,u) = 0 \s \mbox{for} \s x\in(-T,T), \s u(-T) = u(T) = 0 \, . \label{2.4} \eeq Assume that the function $f\in C^1([-T,T] \times R_+)$ is such that \beq f(-x,u) = f(x,u) \s \mbox{for} \s x\in(-T,T) \s \mbox{and} \s u>0 \, , \label{2.5} \eeq \beq xf_x(x,u) < 0 \s \mbox{for} \s x\in(-T,T) \backslash \{0\} \s \mbox{and} \s u>0 \, . \label{2.6} \eeq Then any positive solution of (\ref{2.4}) is an even function with $u'(x) < 0$ on $(0,T)\,$. Moreover any two positive solutions of (\ref{2.4}) cannot intersect on $(-T,T)$ (and hence they are strictly ordered on $(-T,T)$). \end{lma} Clearly, under our conditions, the Lemma 2.1 applies to the problem (\ref{2.3}). We are looking for solution of (\ref{2.3}) which is strictly positive on $(-T,T)$, and so we can work with an equivalent problem \beq \s u'' - a(x)u + b(x) (u^+)^3 = 0 \s \mbox{for $x\in(-T,T)$ , $u(-T) = u(T) = 0 \, ,$} \label{2.7} \eeq where $u^+ = \max(u,0)$. \begin{lma} The problem (\ref{2.3}) has under our conditions a unique positive solution for any $T\geq 1 \,$. Moreover, for this solution we have an estimate \beq \int^T_{-T} \left(u'^2 + u^2 \right) \, dx \, \leq c \s \mbox{uniformly in} \s T\geq 1 \, . \label{2.8} \eeq \end{lma} \noindent {\bf Proof.} \hspace{.15in} We shall work with the problem (\ref{2.7}) using the space $H^1_0[-T,T]$ of absolutely continuous functions, which vanish at $\pm T\,$, with the norm $\|u\|^2 = \int^T_{-T}(u'^2 + u^2)\, dx \,$. We consider a functional $f: H^1_0 [-T,T] \rightarrow R\,$, defined by \[ f(u) = \int^T_{-T} \left[ \frac{u'^2}{2} + a(x) \frac{u^2}{2} - b(x) \frac{(u^+)^4}{4} \right] \, dx \, . \] Clearly $f(0) = 0\,$, and it is standard to verify that $f(u)$ satisfies the Palais-Smale condition, and that $f(u)$ has a strict local minimum at $u=0\,$. Starting with any function $u_0(x) \in H^1_0 [-1,1]\,$, such that $u^+_0 \not\equiv 0\,$, we define $u^* \in H^1_0[-T,T]$ as follows: $u^*(x) = \lambda u_0(x)$ for $x\in [-1,1]$, and $u^*(x) = 0$ for $x\in[-T,T]\backslash [-1,1]\,$. Then $f(u^*) < 0$ for $\lambda$ sufficiently large, and we conclude by the well-known mountain pass theorem that $f(u)$ has a nontrivial critical point $u(x) \in H^1_0[-T,T]\,$, which is then easily seen to be a strictly positive solution of (\ref{2.3}). The variational approach also allows us to derive the estimate (\ref{2.8}). Indeed, if we define the set of paths \[ \Gamma_T = \{ g(\tau):[0,1] \rightarrow H^1_0[-T,T] \; | \; g(0) = 0, \s g(1) = u^* \} \, , \] then the solution of $u(x)$ of (\ref{2.3}) is the point where \beq \inf_{g\in\Gamma_T} \; \max_{\tau\in[0,1]} \; f(g(\tau)) \equiv c_T \, , \label{2.9} \eeq is achieved. Let now $T_1 > T\, $. Then $\Gamma_T \subset \Gamma_{T_1}$, since any function in $H^1_0[-T,T]$ can be regarded as belonging to $H^1_0[-T_1,T_1]\,$, if one extends it by zero in $[-T_1,T_1] \backslash [-T,T]\, $. Hence for $T_1$ the set of competing paths in (\ref{2.9}) is greater than that for $T\,$, which implies that \beq c_{T_1} \leq c_T \leq c_1 \s (T_1 > T \geq 1) \, . \label{2.10} \eeq So that for the positive solution of (\ref{2.3}), \beq \int^T_{-T} \left( \frac{u'^2}{2} + a(x) \frac{u^2}{2} - b(x) \frac{u^4}{4} \right) \, dx \, \leq c_1 \s \mbox{uniformly in} \s T \geq 1 \, . \label{2.11} \eeq Multiply (\ref{2.3}) by $u$ and integrate, \beq \frac{1}{4} \int^T_{-T} \left( u'^2 + a(x)u^2 - b(x) u^4 \right) \, dx \, = 0 \, . \label{2.12} \eeq Subtracting (\ref{2.12}) from (\ref{2.11}), we establish the estimate (\ref{2.8}). Turning to the uniqueness, we notice that any two different solutions $u(x)$ and $v(x)$ of (\ref{2.3}) would have to be ordered by Lemma 2.1, i.e. $u(x) < v(x)$ for all $x\in (-T,T)\,$, which leads to a contradiction by an application of Sturm comparison theorem. \begin{thm} The problem (\ref{2.1}-\ref{2.2}) has under our conditions exactly one positive solution. Moreover this solution is an even function with $u'(x) < 0$ for $x>0\, $. \end{thm} \noindent {\bf Proof.} \hspace{.15in} Take a sequence $T_n \rightarrow \infty\,$, and consider the problem (\ref{2.3}) on the interval $(-T_n,T_n)\,$, i.e. consider \beq \s u'' - a(x)u + b(x)u^3 = 0 \s \mbox{on} \s (-T_n,T_n), \s u(-T_n) = u(T_n) = 0 \, . \label{2.3u} \eeq By Lemma 2.2 the problem (\ref{2.3u}) has a unique positive solution $u_n(x)\,$, and \beq \int^{T_n}_{-T_n} \left( u'^2_n + u^2_n \right) \, dx \, \leq c \s \mbox{uniformly in} \s n \, . \label{2.13} \eeq By Lemma 2.1, $u_n(x)$ takes its maximum at $x=0\,$, which implies that $u''_n(0) \leq 0\,$, and then from the equation (\ref{2.3u}), \beq u_n(0) \geq \sqrt{\frac{a(0)}{b(0)}} \, . \label{2.14} \eeq Writing \[ u_n(x_1) - u_n(x_2) = \int^{x_2}_{x_1} u'_n \,dx\, \leq \sqrt{x_2 - x_1} \left( \int^{x_2}_{x_1} u'^2_n \, dx \, \right)^{1/2} \, , \] we conclude that the sequence $\{u_n(x)\}$ is equicontinuous and uniformly bounded on every interval $[-T_n,T_n]\,$. Hence it has a uniformly convergent subsequence on every $[-T_n,T_n]\,$. So let $\{u^1_{n_k}\}$ be a subsequence of $\{u_n\}$ that converges on $[-T_1,T_1]\,$. Consider this subsequence on $[-T_2,T_2]$ and select a further subsequence $\{u^2_{n_k}\}$ of $\{u^1_{n_k}\}$ that converges uniformly on $[-T_2,T_2]\,$. Repeat this procedure for all $n\,$, and then take a diagonal sequence $\{u_{n_k}\}\,$, which consists of $u^1_{n_1}, u^2_{n_2}, u^3_{n_3}, \ldots\,$. Since the diagonal sequence is a subsequence of $\{u^p_{n_k}\}$ for any $p\geq 1\,$, it follows that it converges uniformly on any bounded interval to a function $u(x)\,$. Expressing $u''_{n_k}$ from the equation (\ref{2.3u}), we conclude that the sequence $\{u''_{n_k}\}\,$, and then also $\{u'_{n_k}\}\,$, converge uniformly on bounded intervals. Writing \[ u_{n_k}(x) = \int^x_a (x-\xi) u''_{n_k} (\xi)\,d\xi \s \mbox{with} \s a = -T_{n_k} - 1 \, , \] we conclude that $u(x) \in C^2(-\infty,\infty)\,$, and that $u''_{n_k} \rightarrow u''$ uniformly on bounded intervals. Hence, we can pass to the limit in the equation (\ref{2.3u}), and we conclude that $u(x)$ solves (\ref{2.1}). By (2.15) $u(x)$ is not identically zero. Writing \[ u^2(x) = \int^x_0 2uu' \, dx \, + u'^2(0) \, , \] we conclude that the limits of $u(x)$ as $x\rightarrow\pm\infty$ exist $(uu' \in L^1(-\infty,\infty))\,$. The only possibility is $u(\pm\infty) = 0\,$. Next we notice from (\ref{2.1}) that \beq |u''(x)| < c \s \mbox{for all real $x$ and some $c>0$ \, .} \label{2.15} \eeq We claim that $u'(\pm\infty) = 0\,$. If not, there is an $\varepsilon > 0$ and a sequence of $x_n \rightarrow \infty$, such that \[ |u'(x_n)| \geq \varepsilon \s \mbox{for all} \s n \, . \] By (\ref{2.15}) we can find $a$ $\delta\,$, such that \[ |u'(x)| \geq \frac{\varepsilon}{2} \s \mbox{for all $x\in(x_n-\delta, x_n+\delta)$ and all $n$} \, . \] This implies that $\int^{T_{n_k}}_{-T_{n_k}} u'^2(x)\, dx \,$ becomes large with $k\,$, say, $\int^{T_{n_k}}_{-T_{n_k}} u'^2(x)\, dx \, \geq 2c$ for some $k\,$, where $c$ is the constant from (\ref{2.13}). Then by fixing $j>k$ so large that $u_{n_j}(x)$ and $u'_{n_j}(x)$ are uniformly close to $u(x)$ and $u'(x)$ respectively on the interval $(-T_{n_k},T_{n_k})$ we get, using (\ref{2.13}), \[ \int^{T_{n_k}}_{-T_{n_k}} u'^2\, dx \, \simeq \int^{T_{n_k}}_{-T_{n_k}} u'^2_{n_j} \, dx \, \leq \int^{T_{n_j}}_{-T_{n_j}} u'^2_{n_j} \, dx \, \leq c \, , \] which is a contradiction. So that $u(x)$ is a solution of our problem (\ref{2.1}-\ref{2.2}). Since by Lemma 2.1 the functions $u_{n_k}(x)$ are even, with the only maximum at $x=0\,$, the same is true for their limit $u(x)\,$. That $u'(x) < 0$ for $x>0$ is easily seen by differentiating (\ref{2.1}) (a similar argument can be found in \cite{2}). Turning to uniqueness, let $v(x)$ be another positive solution of (2.1), (2.2) (which is also an even function with the only maximum at $x=0\,$, as follows by an easy modification of the proof of lemma 1 in [2]). Since \[ \int^\infty_{-\infty} b(x) uv(u^2 - v^2) \, dx \, = 0 \, , \] it follows that the solutions $u(x)$ and $v(x)$ cannot be ordered, and so have to intersect. Two cases are possible: either $u(x)$ and $v(x)$ have at least two positive points of intersection, or only one positive point of intersection. Assume first $\xi_1 > 0$ is the smallest positive point of intersection and $\xi_2 > \xi_1$ the next one, and $u(x) < v(x)$ on $(\xi_1,\xi_2)\,$. Multiply the equation (\ref{2.1}) by $u'$ and integrate from $\xi_1$ to $\xi_2\,$. Denoting by $x=x_1(u)$ the inverse function of $u(x)$ on $(\xi_1,\xi_2)\,$, we obtain denoting $f(x,u) = -a(x)u + b(x)u^3\,$, and $u_1 = u(\xi_1) = v(\xi_1)\,$, $u_2 = u(\xi_2) = v(\xi_2)\,$, \beq \frac{1}{2} u'^2(\xi_2) - \frac{1}{2} u'^2(\xi_1) + \int^{u_2}_{u_1} f(x_1(u),u)\,du = 0 \, . \label{2.16} \eeq Doing the same for $v(x)\,$, and denoting its inverse on $(\xi_1,\xi_2)$ by $x = x_2(v)\,$, we obtain \beq \frac{1}{2} v'^2(\xi_2) - \frac{1}{2} v'^2(\xi_1) + \int^{u_2}_{u_1} f(x_2(v),v)\, dv = 0 \, . \label{2.17} \eeq Subtracting (\ref{2.17}) from (\ref{2.16}), \begin{eqnarray} \frac{1}{2}\left(u'^2(\xi_2) - v'^2(\xi_2)\right) + \frac{1}{2} \left(v'^2(\xi_1) - u'^2(\xi_1)\right) \\ + \int^{u_1}_{u_2} \left[f(x_2(u),u) - f(x_1(u),u)\right]\, du = 0 \, . \nonumber \label{2.18} \end{eqnarray} Notice, $u_2 < u_1$ and $x_2(u) > x_1(u)$ for all $u\in(u_2,u_1)\,$. Keeping in mind that $f(x,u)$ is decreasing in $x\,$, and using uniqueness theorem for initial value problems, we conclude that all three terms on the left in (2.19) are negative. This is a contradiction, which rules out the case of two positive intersection points. If $\xi_1$ is the only intersection point, we integrate from $\xi_1$ to $\infty\,$, obtaining a similar contradiction. Uniqueness of solution follows, completing the proof of the theorem. \section{Homoclinic orbits for a class of Hamiltonian systems} \setcounter{equation}{0} We are looking for nontrivial solutions $u\in H^1(R^n,R)$ of the system \beq u'' - L(t)u + \nabla V(u) = 0 \label{3.1} \eeq \beq u(\pm\infty) = u'(\pm\infty) = 0 \, . \label{3.2} \eeq We assume that $V \in C^1(R^n,R)\,$, and the following conditions \begin{equation} \begin{array}{l} L(t) \s \mbox{is a positive definite matrix with entries of class $C^1(R)$,} \\ \mbox{and $L(-t) = L(t)$ for all $t$}\, , \end{array} \label{3.3} \end{equation} \beq (L'(t)\xi,\xi) \geq 0 \s \mbox{for all $\xi\in R^n$ and $t\geq 0$} \, , \label{3.4} \eeq \beq \; \; 0 < \gamma V(\xi) \leq (\nabla V(\xi),\xi) \s \mbox{for some constant $\gamma > 2$ and all $\xi \in R^n$} \, . \label{3.6} \eeq \begin{thm} Under assumptions (3.3-3.5) the problem (\ref{3.1}-\ref{3.2}) has a nontrivial solution $u(t)\,$, with $u(-t) = u(t)$ for all $t\,$. \end{thm} We postpone the proof of the theorem, and present two examples, which show that our result on scalar equations cannot be expected to carry over to systems. \vspace{.15in}\noindent {\bf Example 1} \hspace{.15in} On some interval $(a,b)$ consider a system \begin{equation} \begin{array}{rcl} u'' - 2u + u(u^2+v^2) = 0 &\mbox{ for }& x\in(a,b), \s u(a) = u(b) = 0\\ v'' - v + v(u^2+v^2) = 0 &\mbox{ for }& x\in(a,b), \s v(a) = v(b) = 0\,. \end{array}\label{3.7} \end{equation} This system has no positive solution (i.e. solution with $u>0$ and $v>0$ on $(a,b)$). Indeed, we can regard the first equation in (3.6) as a linear equation of the form $u'' + c(x)u = 0$, and the second one as $v'' + d(x)v = 0\,$. Since $d(x) > c(x)$, the claim follows by the Sturm's comparison theorem. This system is of type (\ref{3.1}) with $V = \frac{1}{4} \left(u^4 + v^4\right) + \frac{1}{2} u^2 v^2\,$. \vspace{.15in} \noindent {\bf Example 2} \hspace{.15in} The problem \begin{equation}\begin{array}{rcl} u'' - u + u(u^2+v^2) = 0 & \mbox{for} & x\in(a,b), \s u(a) = u(b) = 0 \\ v'' - v + v(u^2+v^2) = 0 & \mbox{for} & x\in(a,b), \s v(a) = v(b) = 0 \, , \end{array} \label{3.8}\end{equation} has infinitely many positive solutions, all of the form $u = \alpha v\,$, where $\alpha$ is an arbitrary positive constant. Indeed, regarding $u^2 + v^2$ as a known function, we see that $u$ and $v$ are positive solutions of the same linear equation, and so have to be multiples of one another. Setting $u = \alpha v\,$, we find $u$ to be the unique (in view of Lemma 2.2) positive solution of \beq u'' - u + u^3 \left( 1 + \frac{1}{\alpha^2} \right) = 0 \s \mbox{for} \s x\in (a,b), \s u(a) = u(b) = 0 \, , \label{3.9} \eeq while $v$ is the unique positive solution of \beq v'' - v + v^3(\alpha^2 + 1) = 0 \s \mbox{for} \s x\in (a,b), \s v(a) = v(b) = 0 \, . %\label{3.10} \eeq Setting $u = \alpha v$ in (\ref{3.9}), we obtain (3.9), so that the pair $(u,v)$ is indeed a solution of (3.7). \vspace{.25in} {\bf Proof of the Theorem 3.1.} We begin by showing that our condition (3.5) implies that \beq \frac{V(\xi)}{|\xi|^2} \rightarrow 0 \s \mbox{as} \s |\xi| \rightarrow 0 \, . \label{3.10} \eeq Indeed, write (3.5) at $r\xi$ with some constant $r>0$, \[ r(\nabla V(r\xi),\xi) \geq \gamma V(r\xi) \] or \[ \frac{d}{dr} V(r\xi) - \frac{\gamma}{r} V(r\xi) \geq 0 \, . \] Multiplying by $r^{-\gamma}$ and integrating over $(\epsilon, 1)\, ,$ $0 < \epsilon < 1\,$, \[ V(\xi) - \frac{V(\epsilon \xi)}{\epsilon^\gamma} \geq 0 \, . \] Let now $|\xi| = 1$ and set $\eta = \epsilon \xi \, , \, |\eta| = \epsilon \, $. Then \[ \frac{|V(\eta)|}{|\eta|^\gamma} \leq c \, , \] and (3.10) follows. As in the previous section, we approximate the solution of (\ref{3.1}-\ref{3.2}) by the problem \beq u'' - L(t)u + \nabla V(u) = 0 \; \mbox{for} \; t\in (-T,T), \; u(-T) = u(T) = 0 \, , \label{3.11} \eeq \beq u(-t) = u(t) \s \mbox{for all real $t$} \, . \label{3.12} \eeq The key step is to show the existence of $\delta > 0\,$, such that any nontrivial solution of (3.11), (3.12) satisfies \beq |u(0)| > \delta \s \mbox{independently of $T>0$} \, . \label{3.13} \eeq To prove (\ref{3.13}) we introduce the ``energy'' function for the solution $u(t)$ of (\ref{3.11}), \[ E(t) = \frac{1}{2} |u'(t)|^2 - \frac{1}{2}(L(t)u,u) + V(u(t)) \, . \] Differentiating $E(t)\,$, and using the equation (\ref{3.11}) and the condition (\ref{3.4}), we express \[ E'(t) = -\frac{1}{2}(L'(t)u,u) \leq 0 \s \mbox{for all} \s 0\leq t \leq T \, , \] and hence \[ E(0) \geq E(T) = \frac{|u'(T)|^2}{2} \geq 0 \, . \] Since $u(t)$ is even, $u'(0) = 0\,$, and then \[ E(0) = V(u(0)) - \frac{1}{2} (L(0)u(0), u(0)) \geq 0 \, , \] or \[ V(u(0)) \geq \frac{1}{2} (L(0)u(0),u(0)) \geq c|u(0)|^2 \, , \] \beq \frac{V(u(0))}{|u(0)|^2} \geq c \, . \label{3.14} \eeq Comparing (\ref{3.14}) with (\ref{3.10}), we conclude the estimate (\ref{3.13}). The rest of the proof is the same as that of the Theorem 2.1, except that we use (3.13) instead of (2.15), so we only sketch it. We take a sequence $\{T_k\} \rightarrow \infty$ as $k \rightarrow \infty\,$, $0 < T_k < T_{k+1}$ for all $k\,$. By $E_k$ we denote the subspace of $H^1_0(-T_k,T_k)\,$, consisting of even functions. By taking zero extensions, we see that $E_k \subset E_{k+1}\,$. We consider the functionals $f_k: E_k \rightarrow R\,$, defined by \[ f_k(u) = \int^{T_k}_{-T_k} \left[ \frac{1}{2} |u'|^2 + \frac{1}{2} (L(t)u,u) - V(u) \right] \, dt \,. \] The mountain pass theorem applies to $f_k(u)\,$, producing $u_k \in E_k\,$, which is an even nontrivial solution of \beq u''_k - L(t)u_k + \nabla V(u_k) = 0 \s \mbox{for} \s t\in(-T_k,T_k) \label{3.15} \eeq \beq u_k(-T) = u_k(T) = 0 \, . \label{3.16} \eeq The critical values $c_k = f_k(u_k) > 0$ are non-increasing in $k\,$. And $u_k(0) > \delta$ uniformly in $k\,$, in view of (\ref{3.13}). We show next that the $H^1$ norm of the solution of (\ref{3.15}-\ref{3.16}) is bounded uniformly in $k\,$. Multiply the equation (\ref{3.15}) by $u_k$ and integrate, \beq \int^{T_k}_{-T_k} \left[ |u'_k|^2 + (L(t)u_k,u_k) - (\nabla V(u_k),u_k) \right] \, dt \, = 0 \, . \label{3.17} \eeq On the other hand by the definition of $c_k$ and (\ref{3.6}), \begin{eqnarray} \int^{T_k}_{-T_k} \left[ \frac{1}{2} |u'_k|^2 + \frac{1}{2} (L(t)u_k,u_k) \right] \, dt \, = \int^{T_k}_{-T_k} V(u_k) \, dt \, + c_k \\ \leq \frac{1}{\gamma} \int^{T_k}_{-T_k} (\nabla V(u_k),u_k)\, dt \, + c_k \,. \nonumber \label{3.18} \end{eqnarray} Using (\ref{3.17}) in (3.18), we obtain \beq \left(\frac{1}{2} - \frac{1}{\gamma}\right) \int^{T_k}_{-T_k} \left[ |u'_k|^2 + (L(t)u_k,u_k) \right] \, dt \, \leq c_k \leq c_1 \, . \label{3.19} \eeq Since $\frac{1}{2} - \frac{1}{\gamma} > 0\,$, and \[ (L(t)u_k,u_k) \geq (L(0)u_k,u_k) \geq c|u_k|^2 \, , \] we conclude from (\ref{3.19}) that \[ \int^{T_k}_{-T_k} \left(|u'_k|^2 + |u_k|^2 \right) \, dt \, \leq c \s \mbox{uniformly in $k$} \, . \] The rest of the proof is the same as in the Theorem 2.1. \vspace{.15in} \noindent {\bf Remark.} We remark that we can prove a similar result with $V$ depending on $t\,$, provided the condition (3.5) is uniform in $t\,$, and (3.4) is replaced by \[ (L'(t)\xi, \xi) - V_t(t,\xi) \geq 0 \; \; \mbox{for all} \; \; \xi \in R^n \; \; \mbox{and} \; \; t \geq 0 \, . \] \begin{thebibliography}{99} \bibitem[1]{1} B. Gidas, W.-M. Ni, L. Nirenberg, {\em Symmetry and related properties via the maximum principle}, Commun. Math. Phys. {\bf 68}(1979), 209-243. \bibitem[2]{2} P. Korman and T. Ouyang, {\em Exact multiplicity results for two classes of boundary value problems}, Differential and Integral Equations, {\bf 6}(1993), 1507-1517. \bibitem[3]{3} W. Omana and M. Willem, {\em Homoclinic orbits for a class of Hamiltonian systems}, Differential and Integral Equations {\bf 5}(1992), 1115-1120. \bibitem[4]{4} P.H. Rabinowitz, {\em Some recent results on heteroclinic and other connecting orbits of Hamiltonian systems}, In Progress in variational methods in Hamiltonian systems and elliptic equations, M. Girardi, M. Matzeu and F. Pacella (Eds.), Longman Scientific \& Technical (1992). \end{thebibliography} {\sc Philip Korman\newline Institute for Dynamics and Department of Mathematical Sciences\newline University of Cincinnati\newline Cincinnati, OH 45221-0025} \newline E-mail address: korman@ucbeh.san.uc.edu \par {\sc Alan C. Lazer\newline Department of Mathematics \& Computer Science\newline University of Miami\newline Coral Gables, FL 33124}\newline E-mail address: lazer@mthvax.cs.miami.edu \end{document}