\documentclass[twoside]{article} \usepackage{amssymb} % font used for R in Real numbers \pagestyle{myheadings} \markboth{\hfil Periodic and almost periodic solutions \hfil EJDE--2000/24} {EJDE--2000/24\hfil E. Hanebaly \& B. Marzouki \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol.~{\bf 2000}(2000), No.~24, pp.~1--16. \newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu \quad ftp ejde.math.unt.edu (login: ftp)} \vspace{\bigskipamount} \\ % Periodic and almost periodic solutions for multi-valued differential equations in Banach spaces \thanks{ {\em Mathematics Subject Classifications:} 34A60, 34C25, 34C27, 47H10. \hfil\break\indent {\em Key words and phrases:} Multi-valued differential equation, Hyper-accretive, \hfil\break\indent Almost periodicity, Banach space. \hfil\break\indent \copyright 2000 Southwest Texas State University and University of North Texas. \hfil\break\indent Submitted July 16, 1999. Published March 30, 2000.} } \date{} % \author{ E. Hanebaly \& B. Marzouki } \maketitle \begin{abstract} It is known that for $\omega$-periodic differential equations of monotonous type, in uniformly convex Banach spaces, the existence of a bounded solution on ${\mathbb R}^+$ is equivalent to the existence of an $\omega$-periodic solution (see Haraux \cite{B.H} and Hanebaly \cite{H1,H4}). It is also known that if the Banach space is strictly convex and the equation is almost periodic and of monotonous type, then the existence of a continuous solution with a precompact range is equivalent to the existence of an almost periodic solution (see Hanebaly \cite{H2} ). In this note we want to generalize the results above for multi-valued differential equations. \end{abstract} \newtheorem{theorem}{Theorem} \newtheorem{lemma}{Lemma} \section{Preliminaries} Let $X$ and $Y$ be Banach spaces, and $2^Y$ denote the collection of subsets of $Y$. For a multi-valued map $F: X\to 2^Y$ we define the following conditions: $F$ is {\bf upper semi-continuous} (u.s.c.) in $X$ if for every $x_0$ in $X$ and every open set $G\subset Y$ with $Fx_0\subset G$ there exists a neighborhood $U$ of $x_0$ such that $Fx_0\subset G$ for all $x\in U$. In practice $F$ is u.s.c. at $x_0$ means that $Fx\subset Fx_0 + B_\varepsilon (0)$ for all $x$ sufficiently close to $x_0$ and for $\varepsilon$ sufficiently small. $F$ is {\bf bounding} if it maps bounded subsets of $X$ into bounded subsets of $Y$. $F$ is {\bf dissipative} if $X=Y$ and $$\langle Fx-Fy,x-y \rangle_- \leq 0 \quad \forall x\in X , \forall y\in X\,.$$ This implies that for all $x_1\in Fx$ and all $y_1\in Fy$, $$\langle x_1-y_1,x-y \rangle_-\leq 0\,,$$ where the lower semi-inner product on $X$ introduced by Lumer \cite{L} is defined as $$\langle x,y \rangle_-=\|y\|\lim_{h\to 0^-}{{\|y+hx\|-\|y\|}\over h}\,. $$ $F$ is {\bf accretive} if $\langle Fx-Fy,x-y \rangle_+ \geq 0 $ where the upper semi-inner product on $X$ is defined as $$ \langle x,y \rangle_+=\|y\|\lim_{h\to 0^+}{{\|y+hx\|-\|y\|}\over h}\,.$$ We denote by $\rightharpoonup$ the convergence for the weak topology $\sigma (X,X^*)$. Recall that $x : J\subset {\mathbb R}\to X$ is said to be absolutely continuous (a.c. for short) if for each $\varepsilon > 0$ there is $\delta > 0$ such that $\sum\|x(t_i)-x(s_i)\|\leq \varepsilon$ whenever the finitely many intervals $[s_i,t_i] \subset J$ do not overlap and $\sum | t_i-s_i|\leq \delta$. In particular every Lipschitzean map is a.c. When $X$ is of finite dimension it is known that $x$ is a.c. if and only if $x$ is differentiable almost everywhere (a.e. for short) and $x'\in L^1(J,X)$, but if $X$ is of infinite dimension and $X$ is not reflexive, then an a.c. function need not be differentiable at any point (see e.g Deimling \cite{D} p.138). By a solution of the Cauchy problem \begin{equation} x'\in F(t,x)\,;\quad x(t_0)=x_0 \end{equation} in some interval $I$ (with $t_0\in I$), we mean a continuous function on $I$, a.c. in every compact subset of $I$, differentiable a.e., and that satisfies (1) a.e. on $I$. The collection non-empty compact convex subsets of $X$ will be denoted by $CV(X)$. \section{Boundedness and periodicity of solutions} We begin by giving a result concerning the existence of a global solutions. Let $(X,\|.\|)$ be a real reflexive Banach space. Consider the multi-valued Cauchy problem \begin{eqnarray} &x'(t) \in F(t,x(t))& \label{2}\\ &x(0) = x_0\,,& \label{3} \end{eqnarray} where $F:{\mathbb R}^+\times X \to CV(X)$ is u.s.c. and bounding. \begin{theorem} \label{thm1} If for all $(t,x,y)\in {\mathbb R}^+\times X\times X$, $\langle F(t,x)-F(t,y),x-y \rangle_-\leq 0$, then the Cauchy problem (2)-(3) has a unique solution defined on ${\mathbb R}^+$. \end{theorem} \paragraph{Remark.} This theorem is well known for the inclusion of type $$x'\in -Ax+f(t)$$ where A is a multi-valued maximal monotone operator on a Hilbert space and $f$ is a uni-valued map (see Brezis \cite{B}). \paragraph{Proof of Theorem 1.} Since $F$ is u.s.c. with convex values, by the approximate selection theorem (see Cellina \cite{A.C}) for each $n\geq 0$ there exists a locally lipschitzean map $f_n:{\mathbb R}^+\times X \to X$ such that $$ f_n(t,x)\in F({\mathbb R}^+\times X\cap B_{1/n}(t,x))+B_{1\over n}(0) \quad \forall (t,x)\in {\mathbb R}^+\times X\,,$$ where $B_{1/n}(t,x)$ is a ball in ${\mathbb R}^+\times X$ and $ B_{1/n}(0)$ is a ball in $X$. Since $F$ is u.s.c. at $(t,x)$, for each $\varepsilon > 0$ there exists $\delta > 0$ such that $$ F({\mathbb R}^+\times X\cap B_\delta (t,x)) \subset F(t,x)+B_\varepsilon (0)\,.$$ Then for $n$ large we can choose $\delta$ such that $ B_{1/n}(t,x)\subset B_\delta (t,x)$ and $$ F({\mathbb R}^+\times X\cap B_{1/n}(t,x))\subset F(t,x)+ B_\varepsilon (0)\,. $$ Consequently, for $\varepsilon =1/m$ with $m\geq n$ we obtain $$f_n(t,x)\in F(t,x)+B_{1/n} (0)+B_{1/n}(0)\subset F(t,x)+B_{2/n}(0)\,. $$ Now we show that for any $a > 0$ the uni-valued Cauchy problem \begin{eqnarray} &x'(t) = f_n(t,x(t))& \label{4} \\ &x(0) = x_0 &\label{5} \end{eqnarray} has a unique solution $x_n$ on $[0,a]$ and the sequence $x_n$ converges uniformly to the solution of the Cauchy problem (2)-(3). Consider $f_n$ from $[0,a]\times X$ to $X$, then $f_n$ satisfies\\ i) $f_n$ is continuous and locally lipschitzean with respect to $x$.\\ ii) $\langle f_n(t,x)-f_n(t,y),x-y \rangle_-\leq {4\over n}\|x-y\|$.\\ For proving ii), we take $f_n(t,x)\in F(t,x)+B_{2/n}(0)$ and $f_n(t,y)\in F(t,y)+B_{2/n}(0)$, so that $f_n(t,x)=a+\alpha_n$ and $f_n(t,y)=b+\beta_n$ with $a\in F(t,x), b\in F(t,y)$ and $\alpha_n,\beta_n \in B_{2/n}(0)$. Then \begin{eqnarray*} \langle f_n(t,x)-f_n(t,y),x-y \rangle_-&=&\langle a+\alpha_n-b-\beta_n,x-y \rangle_- \\ & \leq & \langle a-b,x-y \rangle_-+\langle \alpha_n-\beta_n,x-y \rangle_+ \\ & \leq & \langle \alpha_n-\beta_n,x-y \rangle_- \\ & \leq & \|\alpha_n-\beta_n\|\|x-y\|\\ & \leq & {4\over n}\|x-y\|. \end{eqnarray*} It is well known that by i) the uni-valued Cauchy problem (4)-(5) has a unique local solution $x_n$, and that by ii) this solution can be extended on $[0,a]$. This statement is proven by the standard procedure of bounding the derivative of $x_n$. Taking $y=0$ in ii) , we obtain $$\langle f_n(t,x)-f_n(t,0),x \rangle_-\leq {4\over n}\|x\|.$$ Therefore, \begin{eqnarray*} \langle x'_n(t),x_n(t)\rangle_-&=&\langle f_n(t,x_n(t))-f_n(t,0)+f_n(t,0),x_n(t)\rangle_-\\ &\leq &\langle f_n(t,x_n(t))-f_n(t,0),x_n(t)\rangle_-+\langle f_n(t,0),x_n(t)\rangle_+\\ &\leq &{4\over n}\|x_n(t)\|+\|f_n(t,0)\|\|x_n(t)\|\\ &\leq &( 1+\sup\limits_{t\in [0,a]}\|f_n(t,0)\| )\|x_n(t)\|. \end{eqnarray*} We deduce that (see appendix II) $$D^-\|x_n(t)\|\leq 1+\sup\limits_{t\in [0,a]}\|f_n(t,0)\|=k_n $$ with $k_n$ a constant which does not depend on $t$. This follows because there is $t^n_0 \in [0,a]$ such that $$\sup\limits_{t\in [0,a]}\|f_n(t,0)\|=\|f_n(t^n_0,0)\|. $$ consequently, we have a sequence $x_n \in C([0,a],X)$ that satisfies \begin{equation} x'_n(t)\in F(t,x_n(t))+B_{2/n}(0)\,. \label{6} \end{equation} Next we show that $x_n$ is a Cauchy sequence. Let $\Phi_{n,m} (t)=\|x_n(t)-x_m(t)\|$. Then $\Phi_{n,m} (0)=0$ and using the same technique as for proving ii) we deduce that \begin{eqnarray*} \Phi_{n,m} (t)D^-\Phi_{n,m} (t)& = & \langle x'_n(t)-x'_m(t),x_n(t)-x_m(t)\rangle_-\\ & \leq & ({2\over n}+{2\over m})\Phi_{n,m} (t)\,. \end{eqnarray*} Therefore, $\Phi_{n,m} (t) \leq ({2\over n}+{2\over m})a$ and then $$\sup\limits_{t\in [0,a]}\|x_n(t)-x_m(t)\| \to 0 \quad \mbox{as } n,m\to +\infty$$ Let $x$ be the limit of $x_n$. Then we have in particular $x(0)=x_0$, now we have to show that $x$ is a.e. differentiable and satisfies $$x'(t)\in F(t,x(t)) \quad \mbox{a.e. in } [0,a].$$ Since $F$ is u.s.c. and $x_n\to x$ uniformly on $[0,a]$, we deduce that for $n$ large, $$F(t,x_n(t))\subset F(t,x(t)) + B_1(0)\,.$$ Since $F$ is bounding, by (6) we have $\|x'_n(t)\|\leq c$ uniformly on $[0,a]$ for some $c > 0$. Put $J=[0,a]$, then we have $x'_n\in L^\infty (J,X)\subset L^2(J,X)$. Since $L^2(J,X)$ is reflexive (because $X$ is reflexive), there is a subsequence (which we denote by the same symbol) such that $x'_n\rightharpoonup y\in L^2(J,X)$ so \begin{eqnarray*} x_n(t)&=&x_0+\int_{0}^{t}x'_n(s)\,ds = x_0+\int_{J}\chi_{[0,t]}(s)x'_n(s)\,ds \\ &\rightharpoonup& x_0+\int_{J}\chi_{[0,t]}(s)y(s)\,ds = x_0+\int_{0}^{t}y(s)ds\,. \end{eqnarray*} Since $x_n(t)\to x(t)$, it follows that $x_n(t)\rightharpoonup x(t)$. Consequently $$x(t)=x_0+\int_{0}^{t}y(s)ds \quad \mbox {and} \quad x'(t)=y(t) \quad \mbox{ a.e. in} J.$$ We deduce that $x'_n \rightharpoonup x'$ in $L^2(J,X)$ for the weak topology $\sigma(L^2(J,X),L^2 (J,X^*)).$ Let $\varepsilon > 0$ and put $$A_\varepsilon =\big \{ z\in L^2(J,X) : z(t)\in F(t,x(t)) + \overline{B}_\varepsilon(0) \mbox{ a.e. } \big\}$$ Then $A_\varepsilon$ is nonempty (because $x_n(t)\to x(t) $ and $F$ is u.s.c., so $x'_n\in A_\varepsilon$ for n large), $A_\varepsilon$ is closed and convex, hence $A_\varepsilon$ is weakly closed. Since $x'_n\in A_\varepsilon$ and $x'_n\rightharpoonup x'$ we deduce that $$x'(t)\in \overline{F(t,x(t))}=F(t,x(t)) \mbox{ a.e.}$$ So $x$ is a solution of the Cauchy problem (2)-(3). Since $a > 0$ is arbitrary we deduce that the sequence $x_n$ converges in the Banach space $C({\mathbb R}^+,X)$ equipped with the topology of uniform convergence in compact subsets of ${\mathbb R}^+$. That $x$ is unique follows from the dissipativeness of $F$. Indeed let $x$ and $y$ be two solutions of the Cauchy problem (2)-(3), then we have $$\langle x'(t) - y'(t),x(t) - y(t) \rangle_- \leq 0 \quad \mbox{and}\quad {{1}\over 2}D^-\|x(t) - y(t)\|^2 \leq 0\,.$$ Hence the map $t \mapsto \|x(t) - y(t)\|^2$ is non increasing, and consequently \begin{equation} \label{7} \|x(t) - y(t)\| \leq \|x(0) - y(0)\|. \end{equation} Now we present a result that gives us the relationship between the existence of bounded solution and the existence of an $\omega$-periodic solution of (2) when $F$ is $\omega$-periodic. Observe that under the hypothesis of Theorem 1 the condition: There exists a positive $R$ such that $$_-\leq 0 \quad\mbox{for } \|x\| > R$$ ensures the existence of a bounded solution on $[0,+\infty[$ (see Browder \cite{Br} and Hanebaly \cite{H2}). \begin{theorem} Under the hypothesis of Theorem 1, assuming that $X$ is uniformly convex, and $F(t+\omega,x)=F(t,x)$ ($\omega > 0$), the equation (2) has an $\omega$-periodic solution if and only if it has a bounded solution on $[0,+\infty[$. \end{theorem} \paragraph{Proof.} The necessity condition is obvious because a continuous periodic map is bounded. Conversely we consider the Poincar\'e map $P : X \to X$ defined by $Px_0 = x(\omega)$ where $x_0$ is given in $X$ and $x$ is a solution of (2) which satisfies $x(0) = x_0$. The map P is well defined because of the uniqueness of solutions for the Cauchy problem (2)-(3). Now let $x$ be the solution of (2) which is bounded on $[0,+\infty [$ and put \begin{eqnarray*} x_1 &=& Px_0 = x(\omega)\\ x_2 &=& Px_1 = x(2\omega)\\ &\vdots& \\ x_n &=& Px_{n-1} = x(n\omega) \end{eqnarray*} Note that the solution $x$ is bounded, so the sequence $x_n$ is bounded, and that $P$ is non-expansive. Indeed, let $y$ and $z$ be two solutions of (2) such that $y(0) = y_0$ and $z(0) = z_0$ so by dissipativeness of $F$ and the inequality (7) we have $$\|y(t) - z(t)\| \leq \|y(0) - z(0)\| = \|y_0 - z_0\|. $$ Taking $t=\omega$ we deduce that $$\|Py_0 - Pz_0\| \leq \|y_0 - z_0\|.$$ So by the Browder-Petryshyn's fixed point theorem (see Petryshyn \cite{B.P}), $P$ has a fixed point. So there is a solution $\widetilde{x}$ of (2) which satisfies $\widetilde{x}(0) = \widetilde{x}(\omega)$ and $\widetilde{x}$ is $\omega$-periodic. Indeed, put $\widetilde{y}(t) = \widetilde{x}(t+\omega)$ then $$\widetilde{y'}(t) = \widetilde{x'}(t+\omega)\in F(t+\omega,\widetilde{x}(t+\omega)) = F(t,\widetilde{y}(t))\,.$$ Now since $\widetilde{y}(0) = \widetilde{x}(\omega) = \widetilde{x}(0)$, by (7) we deduce that $$\widetilde{x}(t) = \widetilde{y}(t) = \widetilde{x}(t+\omega)$$ hence $\widetilde{x}$ is $\omega$-periodic.\hfill$\diamondsuit$ \paragraph{Remark.} Let $x$ be an $\omega$-periodic solution of (2), if $y$ is another $\omega$-periodic solution (respectively an $T$-periodic solution with $\frac{\omega}{T}\notin {\mathbb Q}$) then $\|x(t)-y(t)\|$ is constant for all $t\in {\mathbb R}^+$. From the dissipativeness of $F$ it follows that the map $t\mapsto \|x(t)-y(t)\|$ is decreasing. Since it is continuous and periodic (respectively almost-periodic) we conclude that it is constant. \paragraph{Example.} Consider $({\mathbb R}^n,\|.\|)$ with $\|.\|$ the Euclidean norm and $\langle .,. \rangle$ the associated inner product. We consider the differential equation $$x'+x\|x\|^{\alpha}+\beta \mathop{\rm sgn}(x) = f(t)$$ where $\alpha \geq 0$, $\beta \geq 0$, $f :{\mathbb R}^+ \to {\mathbb R}^n$ is continuous and $\omega$-periodic, and $$ \mathop{\rm sgn}(x) = \left \{ \begin{array}{ll} \frac{x}{\|x\|} &\mbox {if $x\neq 0$} \\[5pt] \overline{B}(0,1) &\mbox{if $x = 0$} \end{array}\right. $$ Then the above equation becomes $x'\in F(t,x)$ where $F(t,x)=f(t)-x\|x\|^{\alpha}-\beta \mathop{\rm sgn}(x)$ is a bounding multi-valued map with compact and convex values. To conclude that the inclusion has an $\omega$-periodic solution, we have to prove the following lemma. \begin{lemma} 1) F is upper semi-continuous on ${\mathbb R}^+ \times {\mathbb R}^n$.\\ 2) There exist a positive $c_{\alpha}$ and $r_{\alpha}\geq 2$ such that for all $(t,x)\in {\mathbb R}^+ \times {\mathbb R}^n $, $$\langle F(t,x)-F(t,y),x-y \rangle\leq -c_{\alpha}\|x-y\|^{r_{\alpha}}\,.$$ In particular $F$ is dissipative with respect to $x$ \\ 3) Every solution of the inclusion $x'\in F(t,x)$ is bounded. \end{lemma} \paragraph{Proof of 1)} We have to show that for every closed $A\subset {\mathbb R}^n$, the set $$F^{-1}(A)=\{ (t,x)\in {\mathbb R}^+\times {\mathbb R}^n : F(t,x)\cap A\neq \emptyset \}$$ is closed in ${\mathbb R}^+\times{\mathbb R}^n$. Let $(t_n,x_n)\in {\mathbb R}^+\times {\mathbb R}^n$ be such that $(t_n,x_n)\to (t,x)$ and $F(t_n,x_n)\cap A\neq\emptyset$. We have to show that $F(t,x)\cap A\neq\emptyset.$ Let $y_n\in F(t_n,x_n)\cap A$, then $y_n=f(t_n)-x_n\|x_n\|^{\alpha}-\beta\gamma_n$ with $\|\gamma_n\|\leq 1$, $\gamma_n$ has a subsequence (which we denote by the same) such that $\gamma_n\to \gamma$ with $(\|\gamma \|\leq 1)$, so $$y_n=f(t_n)-x_n\|x_n\|^{\alpha}-\beta\gamma_n\to y:=f(t)-x\|x\|^{\alpha}-\beta\gamma\in F(t,x)\cap A.$$ Hence $F(t,x)\cap A\neq \emptyset$ and F is upper semi-continuous on ${\mathbb R}^+\times {\mathbb R}^n$. \paragraph{Proof of 2)} It is easy to see that for all $x,y\in {\mathbb R}^n$, $\langle \mathop{\rm sgn}(x)-\mathop{\rm sgn}(y),x-y \rangle\geq 0$. Now let $x,y\in {\mathbb R}^n$, then \begin{eqnarray*} \lefteqn{\langle x\|x\|^{\alpha}-y\|y\|^{\alpha},x-y \rangle} \\ &=&\langle x\|x\|^{\alpha}-y\|x\|^{\alpha} +x\|y\|^{\alpha}- y\|y\|^{\alpha} +y\|x\|^{\alpha}-x\|y\|^{\alpha},x-y \rangle\\ &=&\|x-y\|^2(\|x\|^{\alpha}+\|y\|^{\alpha})+\langle y\|x\|^{\alpha}-x\|y\|^{\alpha},x-y \rangle\\ &=&\frac{1}{2}\|x-y\|^2(\|x\|^{\alpha}+\|y\|^{\alpha}) +\frac{1}{2}\langle (x+y)\|x\|^{\alpha}-(x+y)\|y\|^{\alpha},x-y\rangle\\ &=& \frac{1}{2}\|x-y\|^2(\|x\|^{\alpha}+\|y\|^{\alpha})+ \frac{1}{2}(\|x\|^{\alpha}-\|y\|^{\alpha})(\|x\|^{2}-\|y\|^{2})\\ &\geq& \frac{1}{2}\|x-y\|^2(\|x\|^{\alpha}+\|y\|^{\alpha}) \end{eqnarray*} The last inequality comes from the fact that the map $\varphi(t)=t^{\alpha}$ is increasing on ${\mathbb R}^+$, so $(\|x\|^{\alpha}-\|y\|^{\alpha})(\|x\|-\|y\|)\geq 0$. Hence for $\alpha =0$, $$\langle x\|x\|^{\alpha}-y\|y\|^{\alpha},x-y \rangle\geq \|x-y\|^2\,.$$ If $0<\alpha \leq 1$ then $\|x\|^{\alpha}+\|y\|^{\alpha}\geq (\|x\|+\|y\|)^{\alpha}\geq \|x-y\|^{\alpha}$, (because the map $\varphi(t)=1+t^{\alpha}-(1+t)^{\alpha}$ is positive on ${\mathbb R}^+$), so $$\langle x\|x\|^{\alpha}-y\|y\|^{\alpha},x-y \rangle\geq \frac{1}{2}\|x-y\|^{\alpha +2} $$ If $\alpha \geq 1$ then the map $\varphi(t)=t^{\alpha}$ is convex on ${\mathbb R}^+$, so $$\|x\|^{\alpha}+\|y\|^{\alpha}\geq \frac{1}{2^{\alpha -1}}(\|x\|+\|y\|)^{\alpha}\geq \frac{1}{2^{\alpha -1}}\|x-y\|^{\alpha}\,.$$ Hence $$\langle x\|x\|^{\alpha}-y\|y\|^{\alpha},x-y \rangle\geq \frac{1}{2^\alpha}\|x-y\|^{\alpha +2}$$ \paragraph{Proof of 3)} From 2) we deduce that $$\langle F(t,x)-F(t,0),x \rangle\leq -c_{\alpha}\|x\|^{r_{\alpha}}\,,$$ where $$\begin{array}{ll} c_\alpha =1 \mbox{ and } r_\alpha =2 &\mbox { if $\alpha = 0$} \\ c_\alpha = 1/2 \mbox{ and } r_\alpha =\alpha +2 & \mbox { if $0<\alpha\leq 1$} \\ c_\alpha =1/ 2^\alpha \mbox{ and } r_\alpha =\alpha +2 & \mbox { if $\alpha\geq 1$} \end{array} $$ Let $x$ be a solution of $x'\in F(t,x)$, and let $a\in F(t,0)$. Then $a=f(t)-\beta\gamma$, $(\|\gamma\|\leq 1)$, and we have \begin{eqnarray*} \langle x'(t),x(t) \rangle&=&\langle x'(t)-a+a,x(t) \rangle\\ &=&\langle x'(t)-a,x(t) \rangle+\langle a,x(t) \rangle\\ &\leq& -c_{\alpha}\|x(t)\|^{r_{\alpha}} + ( M+\beta )\|x(t)\|\,, \end{eqnarray*} where $M=\sup\limits_{t\in {\mathbb R}}\|f(t)\|$. Therefore, $$\frac{d}{2dt}\|x(t)\|^2\leq 0 \quad\mbox{for } \|x(t)\|\geq \big( \frac{M+\beta}{c_{\alpha}} \big )^{1/(r_{\alpha}-1)}\,.$$ Consequently $$\sup\limits_{t\in {\mathbb R}}\|x(t)\|\leq \max \bigl [ \|x(0)\|, (\frac{M+\beta}{c_{\alpha}} )^{1/(r_{\alpha}-1)} \bigr ]$$ because the map $t\mapsto \|x(t)\|^2$ is decreasing outside $B\big(0,(\frac{M+\beta}{c_{\alpha}})^{1/(r_{\alpha}-1)}\big )$. \section {Almost periodic solutions} Let $(E,\|.\|)$ be a uniformly convex Banach space with $E^*$ uniformly convex. We consider the problem \begin{equation} \label{8} x'(t) \in -Ax(t)+f(t)\,, \end{equation} where $f:{\mathbb R}\to E$ is a continuous almost periodic function (see appendix I for the definition of almost periodicity) and $A:E\to 2^E\setminus \emptyset$ is a hyper-accretive multi-valued map which means that for all $\lambda > 0$, Im$(I+\lambda A)=E$ and $ \langle Ax-Ay , x-y \rangle_+ \geq 0$ for all $x,y\in E$. \begin{theorem} Problem (8) has a solution on $[t_0,+\infty[$ ($t_0\in {\mathbb R}$), which is uniformly continuous with precompact range if and only if it has a weak almost periodic solution. \end{theorem} \paragraph{Remark.} Since a continuous almost periodic map is uniformly continuous with precompact range, it is convenient to relate the existence of a solution to that of uniformly continuous with the precompact range. \paragraph{Proof of Theorem 3.} The proof will be divided into four steps. \paragraph{Step 1.} The Cauchy problem \begin{eqnarray} &x'(t) \in -Ax(t) +f(t) & \label{9}\\ & x(t_0) = x_0& \label{10} \end{eqnarray} has a unique weak solution on $ [t_0,+\infty[$. (weak solution means that there are sequences $x_n$ and $f_n$ where $x_n$ is a strong solution and $x_n\to x$ uniformly in every compact subset J of $[t_0,+\infty[$ and $f_n \to f$ in $L^1(J,E)$ ). Indeed, Since $E$ and $E^*$ are uniformly convex, the Cauchy problem \begin{eqnarray*} &x'(t) \in -Ax(t)& \\ &x(t_0) = x_0& \end{eqnarray*} has a unique strong solution on $[t_0,+\infty[$ (see Deimling \cite{D}). Since $f$ is almost periodic, $f \in L^1(J,E)$ for every compact $J\subset [t_0,+\infty[$, with $t_0\in J$, so there is a sequence $f_n$ of stairs functions which converges uniformly to $f$, hence $f_n\to f$ in $L^1(J,E)$. On the other hand for every $f_n$ there is $x_n$ such that \begin{eqnarray*} &x_n'(t) \in -Ax_n(t) +f_n(t) & \\ &x_n(t_0) = x_0\,.& \end{eqnarray*} Because if $g$ is a stair function defined on $a=b_0 a$ ) by $g(t)=y_i$ on $[b_{i-1} , b_i[$ the Cauchy problem \begin{eqnarray*} &x'(t) \in -Ax(t) + g(t)& \\ &x(t_0) = x_0 & \end{eqnarray*} has also a unique strong solution $x$ defined by $x(t)=S_i(t-b_{i-1}).x(b_{i-1})$ for $t\in [b_{i-1} , b_i]$ and $x(t_0)=x_0$ where $S_i(t)$ is the semigroup generated by the hyper-accretive operator $-( A - y_i )$. Let us show that $(x_n)$ is a Cauchy sequence in the Banach space $C([t_0,+\infty[,E)$ equipped with the topology of uniformly convergence in compact subsets. Since $-A$ is dissipative, we have $$\langle x_n'(t)-f_n(t)-x'_p(t)+f_p(t) , x_n(t)-x_p(t) \rangle_-\leq 0\,,$$ $E^*$ is uniformly convex, $\langle .,. \rangle_-= \langle .,. \rangle _+$, and $\langle .,. \rangle_-$ is linear on the first argument. Then \begin{eqnarray*} \lefteqn{ \frac{d^-}{2dt} \|x_n(t)-x_p(t)\|^2}\\ &=&\langle x'_n(t)-x'_p(t) , x_n(t)-x_p(t) \rangle_- \\ & = &\langle x'_n(t)-f_n(t)-x'_p(t)+f_p(t)+f_n(t)-f_p(t) , x_n(t) -x_p(t) \rangle_- \\ & = & \langle x'_n(t)-f_n(t)-x'_p(t)+f_p(t) , x_n(t)-x_p(t) \rangle_- \\ &&+ \langle f_n(t)-f_p(t) , x_n(t)-x_p(t) \rangle_-\\ & \leq & \langle f_n(t)-f_p(t) , x_n(t)-x_p(t) \rangle_- \\ &\leq & \|f_n(t)-f_p(t)\|\|x_n(t)-x_p(t)\| \end{eqnarray*} Hence \begin{eqnarray*} \|x_n(t)-x_p(t)\|&\leq& \|x_n(t_0)-x_p(t_0)\|+\int_{t_0} ^{t} \|f_n(s)-f_p(s)\|ds \\ &=& \int_{t_0}^{t}\|f_n(s)-f_p(s)\|ds \to 0 \quad \mbox{as } n,p\to +\infty\,. \end{eqnarray*} Without loss of generality, we can assume that the Cauchy problem (9)-(10) has a strong solution. Let $x:[t_0,+\infty[ \to E$ be the uniformly continuous solution of the Cauchy problem (9)-(10) with $x([t_0,+\infty [)$ precompact. Since $f$ is almost periodic, there is $t_n\to +\infty$ such that $f(t+t_n)\to f(t)$ uniformly on ${\mathbb R}$ (see appendix I). Consider the sequences of translated functions $$x_n(t)=x(t+t_n) \quad\mbox{and}\quad f_n(t)=f(t+t_n)$$ which are defined on the real interval $[a,+\infty[$ when $n\geq n(a)$. Since $x([t_0,+\infty[)$ is precompact, we deduce that $\{ x_n(t), t\geq a, n\geq n(a) \}$ is also precompact. On the other hand that $\{ x_n , n\geq n(a) \}$ is equi-continuous follows from the following lemma which is easy to proof. \begin{lemma} Let E be a Banach space, $J\subset {\mathbb R}$ be an interval and ${\cal M}$ a bounded subset of the Banach space $C_b(J,E)$ of continuous bounded functions. Then ${\cal M}$ is uniformly equi-continuous if and only if the mapping $(\psi,t) \mapsto \psi(t)$ of ${\cal M}\times J\subset C_b(J,E)\times {\mathbb R}$ into E is uniformly continuous on ${\cal M}\times J$. \end{lemma} Now applying Ascoli's theorem in the intervals $[-N,N]$, $N=1,2,\dots$ and using the diagonal procedure (see Zaidman \cite{Z}) it is possible to find a subsequence which converges uniformly in every compact subset $J$ of ${\mathbb R}$. But $f_n$ is almost periodic, so $f_n\to f$ in $L^1(J,E)$. Therefore, we obtain a weak solution $x^*$ of (8) defined on ${\mathbb R}$ which is uniformly continuous with range contained in the closure of $x([t_0,+\infty [)$, hence with precompact range. \paragraph{Step 2.} Put $K_0=\overline {Co} ( x^*({\mathbb R}) )$, so that $K_0$ is a compact convex subset of $E$. Let $$\Omega =\big \{ x: {\mathbb R}\to E | x({\mathbb R})\subset K_0 \big \}$$ with $x$ a uniformly continuous solution of (8) and $ J:\Omega \to {\mathbb R}^+$ defined by $Jx=\sup\limits_{t\in {\mathbb R}} \|x(t)\|$. Put $\mu=\inf\limits_{x\in \Omega }Jx$, so there is $x_n\in \Omega $ such that $J(x_n)\to \mu $. By Lemma 2 and Ascoli's theorem there is a subsequence of $x_n$ which converges uniformly in every compact subset of ${\mathbb R}$, let $\widetilde {x}$ be this limit, then $\widetilde {x}\in \Omega $ and $J\widetilde {x}=\mu$. \paragraph{Step 3.} We show that $\widetilde{x}$ is unique. Assume that there are $x_1$ and $x_2$ in $\Omega$ such that $Jx_1=Jx_2=\mu $. Since $f$ is almost periodic, there is $t_n\to -\infty$ such that $f(t+t_n)\to f(t)$ uniformly on ${\mathbb R}$. By Ascoli's theorem, we can extract from $t_n$ a subsequence (which we denote by the same symbol) such that $x_1(t+t_n)$ and $x_2(t+t_n)$ converge uniformly in every compact subset of ${\mathbb R}$. Let $$y_1=\lim x_1(t+t_n)\quad \mbox{and} \quad y_2=\lim x_2(t+t_n\,.)$$ Then $y_1$ and $y_2$ are weak solutions of (8), and $y_1,y_2\in \Omega$ with $J(y_1)=J(y_2)=\mu$. Now since $$x'_1(t+t_n)\in -Ax_1(t+t_n)+f(t+t_n)$$ and $$x'_2(t+t_n)\in -Ax_2(t+t_n)+f(t+t_n)$$ and $-A$ is dissipative, we deduce that $$\langle x'_1(t+t_n)-x'_2(t+t_n),x_1(t+t_n)-x_2(t+t_n) \rangle_-\leq 0$$ So $$\frac{d^-}{2dt}\|x_1(t+t_n)-x_2(t+t_n)\|^2\leq 0$$ Consequently the map $t\longmapsto \|x_1(t+t_n)-x_2(t+t_n)\|$ is non increasing. Since $x_i({\mathbb R})\subset K_0$ for i=1,2, we deduce that \begin{eqnarray} \|y_1(t)-y_2(t)\|&=&\lim\limits_{n\to +\infty}\|x_1(t+t_n)-x_2(t+t_n)\| \nonumber\\ &=&\lim\limits_{\tau\to -\infty} \|x_1(\tau )-x_2(\tau )\| \label{11}\\ &=&\sup\limits_{t\in {\mathbb R}}\|x_1(t)-x_2(t)\| \nonumber\\ &=&\mbox{a constant} \nonumber \end{eqnarray} To continue, we need the following lemma. \begin{lemma} Let $E$ be a strictly convex Banach space, $C$ a closed convex subset of $E$. Let $T:C\to C$ be a non expansive map and $x_0,y_0$ in $C$ such that $$\|Tx_0-Ty_0\|=\|x_0-y_0\|\,.$$ Then $$T({{x_0+y_0}\over 2})={{Tx_0+Ty_0}\over 2}.$$ \end{lemma} Let the operator $T_t:E\to E$ be defined by by $T_tx(0)=x(t)$ where $x(.)$ is a weak solution of (8). Then $T_ty_1(0)=y_1(t)$ and $T_ty_2(0)=y_2(t)$ where $y_1$ and $y_2$ are in $\Omega$. By (11), $$\|T_ty_1(0)-T_ty_2(0)\|= \|y_1(t)-y_2(t)\|=\|y_1(0)-y_2(0)\|\,.$$ So that by Lemma 3, $$T_t({{y_1(0)+y_2(0)}\over 2})= {{T_ty_1(0)+y_2(0)}\over 2}= {{y_1(t)+y_2(t)}\over 2}$$ and $y(t):=\frac{y_1(t)+y_2(t)}{2}$ is also a solution of (8) satisfying $y(0)=\frac{y_1(0)+y_2(0)}{ 2}$. Since $K_0$ is convex, $y({\mathbb R})\subset K_0$ and $y\in \Omega$. We have $$Jy_1=Jy_2=\mu$$ So, $\mu =\inf\limits_{x\in \Omega }Jx$ and $\frac{y_1+y_2}{2}\in \Omega$. We deduce that $$\mu \leq J({{y_1+y_2)}\over 2}) \leq {{Jy_1}\over 2}+{{Jy_2}\over 2} =\mu $$ and consequently $Jy=\mu$. Since $J(\frac{y_1+y_2}{2})=\frac{Jy_1}{ 2}+\frac{Jy_2}{ 2}$ we have $$\sup\limits_{t\in {\mathbb R}}\big\|{{y_1(t)+y_2(t)}\over 2}\big\|= \frac{1}{2} \sup\limits_{t\in {\mathbb R}}\|y_1(t)\|+ \frac{1}{2} \sup\limits_{t\in {\mathbb R}}\|y_2(t)\|$$ So there is $s_n \in {\mathbb R}$ such that \begin{eqnarray*} \mu -\frac{1}{n} &<& \big\|{{y_1(s_n)+y_2(s_n)}\over 2}\big\| \\ &\leq& \frac{\|y_1(s_n)\|}{2} + \frac{\|y_2(s_n)\|}{2}\\ &\leq& \mu \end{eqnarray*} and since $y_1(s_n)\in K_0; y_2(s_n)\in K_0$ there is a subsequence (which we denote by the same symbol) such that $ y_1(s_n)\to l_1$ and $y_2(s_n)\to l_2$. Then $$\big\|{{l_1+l_2}\over 2}\big\|={{\|l_1\|}\over 2}+{{\|l_2\|}\over2}= \mu\,.$$ On the other hand $\|y_i(s_n)\|\leq \mu$ implies $\|l_i\| \leq \mu$ and ${{\|l_1\|}\over 2}+{{\|l_2\|}\over2}=\mu$ implies $ \|l_i\|\geq \mu$ for $i=1,2$. Hence $\|l_1\|=\|l_2\|=\mu$. Since the norm of $E$ is strictly convex, we deduce that $l_1=l_2$ and consequently \begin{eqnarray*} \|l_1-l_2\|&=&\|y_1(t)-y_2(t)\| \\ &=& \lim\limits_{\tau \to -\infty}\|x_1(\tau) - x_2(\tau)\| \\ &=& \|x_1(-\infty) - x_2(-\infty)\| \\ &=& \sup\limits_{t\in {\mathbb R}}\|x_1(t)-x_2(t)\| \end{eqnarray*} So $x_1(t)=x_2(t)$ for every $t\in {\mathbb R}$. \paragraph{Remark.} In the case of a Hilbert space, by the parallelogram formula and by (11), we deduce directly that $x_1(t)=x_2(t)$ for all $t\in {\mathbb R}$. \paragraph{Step 4.} Finally we show that $\widetilde {x}$ the unique element of $\Omega $ which satisfies $J\widetilde {x}=\inf_{x\in \Omega}Jx$ is almost periodic. For this purpose, we use the $2^{nd}$ Bochner's characterization of almost periodicity (see appendix I). Let $t_n$ and $s_n$ be two real sequences, then by Ascoli's theorem there is a subsequence of $t_n$ (which we denote by the same symbol) such that $\widetilde {x}(t+t_n) \to y(t)$ uniformly in every compact subset of ${\mathbb R}$. Then $y(.)$ is a weak solution of \begin{equation} \label{12} x'\in -Ax+g(t)\,, \end{equation} where $g(t)=\lim f(t+t_n)$. Now consider $\widetilde {x}(t+t_n+s_n)$ and $y(t+s_n)$, then by Ascoli's theorem we can extract from $t_n$ and $s_n$ sub-sequences such that $\widetilde {x}(t+t_n+s_n)\to z_1(t)$ and $y(t+s_n)\to z_2(y)$, but $f(t+t_n+s_n)$ and $g(t+s_n)$ have the same limit which we denote by $h(t)$. Then $z_1(.)$ and $z_2(.)$ are weak solutions of \begin{equation} \label{13} x'\in -Ax+h(t) \end{equation} so $\mu=J_f(K_0)=J_h(K_0)\leq Jz_i \quad i=1,2$ where $$J_f(K_0)=\inf \big \{Jx : x \mbox {is a weak solution of (8)}, x({\mathbb R})\subset K_0 \big \}$$ and $$J_h(K_0)=\inf \big \{Jx : x \mbox {is a weak solution of (13)},x({\mathbb R})\subset K_0 \big \}\,.$$ We have $\mu =Jz_1=Jz_2$, but the equation (13) has the same property as the equation (8) because the map $h(.)$ is almost periodic. Therefore, there is a unique solution which satisfies $$J_h(K_0)=\inf \big \{Ju : u \mbox { is a weak solution of (13)}, u({\mathbb R})\subset K_0 \big \}.$$ Consequently $z_1=z_2$. Also $\widetilde {x}(t+t_n+s_n)$ and $y(t+s_n)$ have the same limit, hence $\widetilde{x}$ is almost periodic.\hfill$\diamondsuit$ \paragraph{Example.} Let $E=({\mathbb R}^n,\|.\|)$ with the Euclidean norm $\|.\|$, and let $\varphi (x)=\|x\|$. Consider $$Ax=\partial \varphi (x) + kx$$ where $k>0$ and $\partial \varphi$ is the sub-differential of $\varphi$. Since $\varphi $ is continuous and convex, $$\overbrace{\mathop{\rm Dom}(\varphi)}^{\circ} \subset \mathop{\rm Dom}(\partial \varphi) \quad\mbox{so}\quad \mathop{\rm Dom}(A)={\mathbb R}^n.$$ The problem $$x'\in -Ax+f(t)\,,$$ with $f:{\mathbb R}\to {\mathbb R}^n$ continuous and almost periodic, has a strong solution defined on $[t_0,+\infty [$ ( $t_0\in {\mathbb R}$ ) (see Brezis \cite{B}). Now since $0\in \partial \varphi (0)$ we have $$\langle f(t)-kx-x'(t) , x(t) \rangle\geq 0\,.$$ Therefore, \begin{eqnarray*} \langle x'(t),x(t) \rangle & \leq & \langle f(t),x(t) \rangle -k\|x(t)\|^2 \\ & \leq & ( M-k\|x(t)\| )\|x(t)\| \end{eqnarray*} where $M=\sup\limits_{t\in {\mathbb R}}\|f(t)\|$. We deduce that $$D^-\|x(t)\|\leq M$$ and $$\frac{d}{2dt}\|x(t)\|^2\leq 0 \quad\mbox{for } \|x(t)\|\geq \frac{M}{k}.$$ The first inequality shows that $x$ is lipschitzean, hence uniformly continuous and the second one shows that the map $t\longmapsto \|x(t)\|$ is non increasing outside of the ball $B(0,\frac{M}{k})$. Consequently $$\|x(t)\|\leq \sup(\|x(t_0)\|,\frac{M}{k} ) \quad \forall t\geq t_0.$$ So that the problem $x'\in -Ax+f(t)$ has a uniformly continuous solution which is bounded, hence with precompact range, so it has an almost periodic solution. \hfill$\diamondsuit$ \subsection*{Appendix I} Let $E$ be a real Banach space, a map $f: {\mathbb R} \to E$ is said to be almost periodic if for each $\varepsilon > 0$ there exists $l_\varepsilon$ such that for all $a\in {\mathbb R}$ there exists $\tau\in [a,a+l_\varepsilon]$ such that $$ \|f(t+\tau)-f(t)\|\leq \varepsilon \quad \forall t\in {\mathbb R}.$$ If $f$ is almost periodic then there exist $t_n\to +\infty$ and $s_n\to -\infty$ such that $f(t+t_n)\to f(t)$ and $f(t+s_n)\to f(t)$ uniformly on ${\mathbb R}$. In practice, we use the following Bochner's characterizations of almost periodicity (Yoshisawa \cite {Y}). \paragraph{First characterization.} $f\in C({\mathbb R},E)$ is almost periodic if and only if from every real sequence $t'_n$ one can extract a subsequence $t_n$ such that $\lim f(t+t_n)$ exists uniformly on the real line, furthermore the limit is also almost periodic. \paragraph{Second characterization.} $f\in C({\mathbb R},E)$ is almost periodic if and only if for every pair of real sequences $h'_n$ and $k'_n$ there are sub-sequences $h_n$ and $k_n$ such that $f(t+h_n)$ has a pointwise limit $g(t)$ on ${\mathbb R}$, and $f(t+h_n+k_n)$ and $g(t+k_n)$ have a same limit $h(t)$ on ${\mathbb R}$, and $h$ is also almost periodic. \subsection*{Appendix II} Let E be a real Banach space and $x : [a,b]\subset {\mathbb R}\to E$ differentiable, and put $\Phi(t)=\|x(t)\|$. Then $$\Phi(t)D^-\Phi(t)=_-$$ where $$D^-\Phi(t)=\limsup\limits_{h\to 0^-}\frac{\Phi(t+h)-\Phi(t)}{h}$$ is the upper Dini's derivative of $\Phi$ (see e.g Deimling \cite{D}). \begin{thebibliography}{10} {\frenchspacing \bibitem{A.C} J. Aubin, A. Cellina, {\em Differential inclusions multivalued maps and viability theory}, Springer-Verlag (1984). \bibitem{B.P} F. Browder, W. V. Petryshyn, {\em The solution by iteration of nonlinear functional equations in Banach spaces}, Bul Amer Math Soc, 72 (1966), 571--575. \bibitem{Br} F. Browder, {\em Periodic solutions of nonlinear equations of evolution in infinite dimensional spaces}, \lq\lq Lecture in differential equations" A.K Aziz. ed, V1 Van Norstrand, Newyork,(1969), 71--96. \bibitem{B} H. Brezis, {\em Op\'erateurs maximaux monotones et semigroupes de contraction dans les espaces de Hilbert}, North-Holand Amsterdam (1973). \bibitem{B.H} J. Baillon A. Haraux, {\em Comportement \`a l'infinie pour les \'equations d'\'evolution avec forcing p\'eriodique}, Arch. Rat. Mec. Anal, 67 (1977), 101--109. \bibitem{D} K. Deimling, {\em Nonlinear functional analysis}, Springer-Verlag (1985). \bibitem{H1} E. Hanebaly, {\em Solutions p\'eriodiques d'\'equations diff\'erentielles non lin\'eaires en dimension infinite}, C.R.A.S, Serie A (1979), 623--626. \bibitem{H2} E. Hanebaly, {\em Solutions presque-p\'eriodiques d'\'equations diff\'erentielles monotones}, C.R.A.S, 296 Serie I (1983), 263. \bibitem{H3} E. Hanebaly, {\em Contribution \`a l'\'etude des solutions p\'eriodiques et presque-p\'eriodiques des \'equations diff\'erentielles non lin\'eaires sur les espaces de Banach}, Th\`ese d'\'etat Universit\'e de Pau I.U.R.S (1988). \bibitem{H4} E. Hanebaly, {\em Un th\'eor\`eme du point fixe et solutions p\'eriodiques d'\'equations diff\'erentielles V-dissipatives}, $\pi$P.R.N.M.S (252) Fixed point theory and applications, (1991), 221--230. Edit=82 par J.B.Baillon M.A. thera. \bibitem{L} G. Lumer, {\em Semi-inner product spaces}, Trans. Math. Math .Soc, (1961), 29--43. \bibitem{Y} T. Yoshisawa, {\em stability theory and the existence of periodic solutions and almost periodic solutions}, Spriger-Verlag (1975). \bibitem{Z} S. Zaidman, {\em Solution of almost periodic abstract differential equations with relatively compact range}, J. Nonlinear analysis, Vol. 8, No. 9 (1984), 1091--1093. }\end{thebibliography} \bigskip \noindent{\sc E. Hanebaly } \\ Universit\'e Mohammed V Facult\'e des Sciences\\ D\'epartement de Math\'ematiques, Rabat, Maroc.\\ e-mail: hanebaly@fsr.ac.ma \smallskip \noindent{\sc Brahim Marzouki} \\ Universit\'e Mohammed I Facult\'e des Sciences\\ D\'epartement de Math\'ematiques, Oujda, Maroc.\\ e-mail: marzouki@sciences.univ-oujda.ac.ma \end{document}