\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 13, pp. 1--11.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/13\hfil Limit behavior] {Limit behavior of monotone and concave skew-product semiflows with applications} \author[B.-G. Wang \hfil EJDE-2015/13\hfilneg] {Bin-Guo Wang} \address{Bin-Guo Wang \newline School of Mathematics and Statistics, Lanzhou University, Lanzhou, Gansu 730000, China} \email{wangbinguo@lzu.edu.cn} \thanks{Submitted October 21, 2013. Published January 19, 2015.} \subjclass[2000]{34C12, 34D08, 34D45} \keywords{Monotone; skew-product semiflow; attractor; \hfill\break\indent almost periodic equation} \begin{abstract} In this article, we study the long-time behavior of monotone and concave skew-product semiflows. We show that if there are two strongly ordered omega limit sets, then one of them is a copy of the base. Thus, we obtain a global attractor result. As an application, we consider a delay differential equation. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{definition}[theorem]{Definition} \newtheorem{corollary}[theorem]{Corollary} \allowdisplaybreaks \section{Introduction} Recently, monotone skew-product semiflows generated by nonautonomous systems, in particular almost periodic systems, have extensively investigated, see \cite{hetzer, jiang, novo2, novo1, novo3,shen11, zhao}. Hetzer and Shen \cite{hetzer} considered the convergence of positive solutions of almost periodic competitive diffusion systems. Jiang and Zhao \cite{jiang} established the $1$-covering property of the omega limit set for monotone and uniformly stable skew-product semiflows with the componentwise separating property of bounded and ordered full orbits, which is an important property for considering the long-time behavior of skew-product semiflows. Novo et al \cite{novo2, novo1, novo3} considered the skew-product semiflow generated by almost periodic systems. Under the assumption that there existed two strongly ordered minimal subsets or completely strongly ordered minimal subsets, a complete description of the long-time behavior of the trajectories was given and a global picture of the dynamics was provided for a class of monotone and convex skew-product semiflows. Zhao \cite{zhao} proved a global attractivity theory for a class of skew-product semiflows. In conclusion, the properties of the omega limit set of skew-product semiflows, especially its structure, play an important role in considering the convergent behavior of the orbit. Shen and Yi \cite{shen11} told us if the omega limit set $\mathcal{O}$ is linearly stable, then there exists an integral number $N$ such that $\mathcal{O}$ is the $(N-1)$-almost periodic extension; i.e., there exists a subset $Y_0\subset Y$ (the definition of $Y$ see Section 2) such that for any $g_0\in Y_0$, card$(\mathcal{O}\cap\pi^{-1}(g_0))=N$ ($\pi$ is the natural projector). If it is uniformly stable, then it is the extension of $Y$; i.e., card$(\mathcal{O}\cap\pi^{-1}(g))=N$ for any $g\in Y$. This is not enough to understand the structure of the omega limit set thoroughly. If we can obtain the conclusion that $\mathcal{O}$ is the copy of the base $Y$; i.e., $\operatorname{card}(\mathcal{O}\cap\pi^{-1}(g))=1$ for any $g\in Y$, it would give a complete description for the long-time behavior of the orbit. For this purpose, under the assumption of the existence of two completely strongly ordered omega limit sets and motivated by \cite{novo2,novo1}, we deduce that one of them is an equilibrium point set if monotonicity and concavity are satisfied. Naturally, it is a copy of the base. Furthermore, we establish the convergent results for skew-product semiflows. This article is organized as follows. In Section 2, we present some definitions and notation of skew-product semiflows. In Section 3, we establish global attractor results and consider an almost periodic delay differential equation. \section{Preliminaries} Let $(Y,d)$ be a compact metric space. A continuous flow $(Y,\sigma,\mathbb{R})$ is defined by a continuous mapping $\sigma$ : $Y\times \mathbb{R}\to Y$, $(g,t)\mapsto\sigma(g,t)$, which satisfies (i) $\sigma_0= id$, (ii) $\sigma_{t}\cdot\sigma_{s} =\sigma_{t+s}$, for all $t,s\in \mathbb{R}$, where $\sigma_{t}(g):=\sigma (g,t)=g\cdot t$ for $g \in Y$ and $t \in\mathbb{R}$ with $g\cdot 0=g$ and $g\cdot (s+t)=(g\cdot s)\cdot t$. A continuous flow $(Y,\sigma,\mathbb{R})$ is \textbf{distal} if for any two distinct points $g_1$ and $g_2$ in $Y$, $\inf_{t\in \mathbb{R}}d(\sigma (g_1,t),\sigma (g_2,t))>0$. A semiflow ($X$, $\Phi$, $\mathbb{R^{+}}$) on Banach space $X$ is a continuous map $\Phi:X\times \mathbb{R^{+}} \to X$, $(x,t)\mapsto\Phi(x,t)$, which satisfies (i) $\Phi_0= id$, (ii) $\Phi_{t}\cdot\Phi_{s} =\Phi_{t+s}$, where $\Phi_{t}(x):=\Phi (x,t)$ for $x \in X$ and $t \geq 0$. A compact, positively invariant subset $S$ of a semiflow ($X$, $\Phi$, $\mathbb{R^{+}}$) is \textbf{minimal} if it contains no nonempty, closed and proper positively invariant subset. If $X$ itself is minimal, then ($X$, $\Phi$, $\mathbb{R^{+}}$) is called minimal semiflow. In this article, we assume that $(X,X^{+})$ is an ordered Banach space with $\operatorname{int}X^{+}\neq\emptyset$, where $\operatorname{int} X^{+}$ denotes the interior of the cone $X^{+}$. For $x$, $y\in X$, we write $x\leq y$ if $y-x\in X^{+}$; $x< y$ if $y-x\in X^{+}\backslash \{0\}$; $x\ll y$ if $y-x\in \operatorname{int}X^{+}$. In addition, the norm of Banach space $X$ is \textbf{monotone}, namely, if $0\leq x\leq y$, then $\|x\|\leq \|y\|$ (see \cite{novo1}). The ordering on $X$ induces the ordering on $Y\times X$ in the following way: \begin{gather*} (g,x)\leq (g,y)\Leftrightarrow y-x \in X^{+},\quad \forall g\in Y, \\ (g,x)<(g,y)\Leftrightarrow y-x \in X^{+}, \; x\neq y,\quad \forall g\in Y,\\ (g,x)\ll (g,y)\Leftrightarrow y-x\in \operatorname{int}X^{+},\quad \forall g\in Y. \end{gather*} Consider a skew-product semiflow: $\Pi:\mathbb{R^{+}}\times Y \times X\to Y \times X$, \begin{equation}\label{skew} (t,g,x)\mapsto (g\cdot t, u(t,g,x)). \end{equation} We assume that $(Y,\sigma,\mathbb{R})$ is a minimal flow defined by $\sigma: Y\times \mathbb{R} \to Y$, $(g,t)\mapsto g\cdot t$ and $u$ is locally $C^{1}$ in $x\in X$; that is, $u$ is $C^{1}$ in $x$, and $u_{x}$ is continuous in $g\in Y$, $t>0$ in a neighborhood of each compact subset of $Y\times X$. Moreover, for any $v\in X$, $\lim_{t\to 0^{+}}u_{x}(t,g,x)v=v$ uniformly in every compact subset of $Y\times X$. Sometimes, we also use the notation $\Pi_{t}(g,x)\equiv\Pi(t,g,x)$. We denote $\pi:Y\times X\to Y$ as the natural projection. The forward orbit of $(g_0,x_0)$ is written as \[ O(g_0,x_0)=\{\Pi(t,g_0,x_0):t\geq0\}. \] If $u(t,g_0,x_0)$ is convergent as $t\to\infty$, we can define the omega limit set of $(g_0,x_0)$ as \[ \mathcal {O}(g_0,x_0)=\{(g,x)\in Y\times X: \exists t_n\to\infty\text{ such that } g_0\cdot t_n\to g, \; u(t_n,g_0,x_0)\to x\}. \] Given a subset $K\subset Y\times X$, let us introduce the projection set of $K$ into the fiber space \[ K_{Y}:=\{g\in Y: \text{ there exists }x\in X \text{ such that } (g,x)\in K\}\subset Y. \] An \textbf{equilibrium} is a map $a: Y\to X$ such that $a(g\cdot t)=u(t,g,a(g))$, for all $g\in Y$, $t\geq0$. A set $E\subset Y\times X$ is called an \textbf{equilibrium point set} if there exists a map $a$ such that $a(g)=x$, for all $(g,x)\in E$ and $a(g\cdot t)=u(t,g,a(g))$, for all $g\in E_{Y}$, $t\geq0$. We say that the skew-product semiflow \eqref{skew} is \textbf{monotone} if \begin{equation} u(t,g,y)\geq u(t,g,x), \quad \forall y\geq x, \; t\geq0, \end{equation} and \textbf{strongly monotone} if \[ u(t,g,y)\gg u(t,g,x),\quad \forall y\gg x, \; t\geq 0. \] The skew-product semiflow \eqref{skew} is said to be \textbf{eventually strongly monotone} if there exists $t_0>0$ such that \begin{equation}\label{2.3} u(t,g,y)\gg u(t,g,x), \quad \forall y>x, \; t>t_0 \end{equation} and it preserves the ordering; i.e., \[ u(t,g,y)>_ru(t,g,x), \quad \forall y>_rx,\; t>0, \] where $>_r$ denotes the relations $\geq$, $>$ or $\gg$. The skew-product semiflow \eqref{skew} is called \textbf{concave}, if, whenever $x\leq y$, \begin{equation}\label{la2.2} u(t,g,\lambda y+(1-\lambda)x)\geq \lambda u(t,g,y)+(1-\lambda)u(t,g,x) \end{equation} for $g\in Y$, $\lambda \in[0,1]$\ and $t\in\mathbb{R}^{+}$; \textbf{strongly} \textbf{concave}, if, whenever $x\ll y$, \begin{equation}\label{la2.3} u(t,g,\lambda y+(1-\lambda)x)\gg\lambda u(t,g,y)+(1-\lambda)u(t,g,x) \end{equation} for $g\in Y$, $\lambda \in(0,1)$ and $t\in\mathbb{R}^{+}$. From the continuous hypothesis for $u$, \eqref{la2.2} is equivalent to, whenever $y\geq x$, \[ u_{x}(t,g,x)(y-x)\geq u_{x}(t,g,y)(y-x) \] for \ $g\in Y$ and $t\in\mathbb{R}^{+}$. Similarly, \eqref{la2.3} is equivalent to, whenever $y\gg x$, \[ u_{x}(t,g,x)(y-x)\gg u_{x}(t,g,y)(y-x) \] for $g\in Y$ and $t\in\mathbb{R}^{+}$. Since $x\leq\lambda y+(1-\lambda)x$ and $\lambda y+(1-\lambda)x\leq y$, we have \begin{equation}\label{kconvex} u_{x}(t,g,y)(y-x)\leq u(t,g,y)-u(t,g,x)\leq u_{x}(t,g,x)(y-x) \end{equation} for $g\in Y$ and $t\in\mathbb{R}^{+}$. Let $y\geq x$, we have \[ u(t,g,y)-u(t,g,x)=\int_0^{1} u_{x}(t,g,\lambda y+(1-\lambda)x)(y-x)d\lambda. \] A forward orbit $\{\Pi(t,g_0,x_0)|t\geq0\}$ of the skew-product semiflow \eqref{skew} is said to be \textbf{uniformly stable} if for any $\epsilon>0$ there is a $\delta=\delta(\epsilon)>0$, such that if $s>0$ and $\|u(s,g_0,x_0)-u(s,g_0,x)\|\leq \delta(\epsilon)$, we have \[ \|u(t+s,g_0,x_0)-u(t+s,g_0,x)\|\leq \epsilon, \ \forall t\geq0. \] A forward orbit $\{\Pi(t,g_0,x_0)|t\geq0\}$ of the skew-product semiflow \eqref{skew} is said to be \textbf{uniformly asymptotically stable} if it is uniformly stable and there is $\delta_0>0$ with the following property: for each $\epsilon>0$ there exists a $t_0(\epsilon)>0$ such that if $s\geq0$ and $\|u(s,g_0,x_0)-u(s,g_0,x)\|\leq \delta_0$, we get \[ \|u(t+s,g_0,x_0)-u(t+s,g_0,x)\|\leq \epsilon,\ \forall t\geq t_0(\epsilon). \] \section{Global attractor result} In this section, we assume that the skew-product semiflow \eqref{skew} satisfies eventually strong monotonicity and (strong) concavity. Based on this, we establish the global attractor results. \begin{definition} \label{def3.1} \rm Two subsets $S_1$, $S_2$ of $Y\times X$ are ordered $S_1\leq S_2$ if for each $(g,x_1)\in S_1$, there exists $(g,x_2)\in S_2$ such that $x_1\leq x_2$. We say $S_1< S_2$ if $S_1\leq S_2$ and they are different. \end{definition} \begin{definition}\label{def3.2} \rm We say the subset $S_1$, $S_2$ of $Y\times X$ to be ordered $S_1\ll S_2$ if for each $(g,x_1)\in S_1$, there exists $(g,x_2)\in S_2$ such that $x_1\ll x_2$. \end{definition} \begin{definition}\label{def3.3} \rm Two subsets $S_1$, $S_2$ are said to be completely strongly ordered $S_1\ll_{C} S_2$ if $x_1\ll x_2$ holds for all $(g,x_1)\in S_1$ and $(g,x_2)\in S_2$. \end{definition} \begin{definition}\label{de3.3} \rm Let $M\subset Y\times X$ be a compact, positively invariant subset of the skew-product semiflow $\eqref{skew}$. For $(g,x)\in M$, we define the \textbf{Lyapunov exponent} $\lambda(g,x)$ as \[ \lambda(g,x)=\limsup_{t\to \infty}\frac{\ln \|u_{x}(t,g,x)\|}{t}. \] The number $\lambda_{M}=\sup_{(g,x)\in M}\lambda(g,x)$ is called the \textbf{upper Lyapunov exponent} on $M$. If $\lambda_{M}\leq0$, then $M$ is said to be \textbf{linearly stable}. \end{definition} In addition, the following assumptions are necessary. \begin{itemize} \item[(A1)] Every bounded forward orbit $\{\Pi(t,g,x):t\geq0\}$ is precompact. \item[(A2)] $u(t,g,0)=0$, for all $g\in Y$, $t\in\mathbb{R}^{+}$. \end{itemize} \begin{theorem}\label{th3.4} Assume that {\rm (A2)} holds and $\mathcal{O}\subset Y\times \operatorname{int}X^{+}$ with $\lambda_{\mathcal{O}}<0$. Then $\mathcal{O}$ is uniformly asymptotically stable, that is, for each $g\in Y$, the forward orbit $\{\Pi(t,g,a(g)|t\geq0\}$ is uniformly asymptotically stable. Moreover, $\mathcal{O}$ is the copy of the base $Y$, i.e., $\operatorname{card}(\mathcal{O}\cap\pi^{-1}(g))=1$, for all $g\in Y$. \end{theorem} \begin{proof} The proof of the uniformly asymptotical stability is completely similar to \cite[Theorem 8.1]{novo2}, we omit the details here. In view of the theory of \cite{shen11} about the structure of omega limit sets, we deduce that $\mathcal{O}$ is an $(N-1)$-extension of $Y$ as $\lambda_{\mathcal{O}}<0$, that is, $\operatorname{card}(\mathcal{O}\cap\pi^{-1}(g))=N$ for any $g\in Y$, where $N$ is an integral number, and hence, we denote $\mathcal{O}\cap\pi^{-1}(g)=\{x_1(g),\dots ,x_{N}(g)\}$. Since $X^{+}$ is a normal cone and $\operatorname{int}X^{+}\neq\emptyset$, it is easy to deduce that, for each $g\in Y$, the finite set $\{x_1(g),\dots ,x_{N}(g)\}$ is bounded with respect to the ordering induced by $X^{+}$. Thus, there exists the supremum \[ b(g)=\sup\{x_1(g),\dots ,x_{N}(g)\}, \] which is a continuous map on $Y$. The positive invariance and monotonicity of the semiflow imply that \begin{equation}\label{la3.1} b(g\cdot t)\leq u(t,g,b(g)),\quad \forall g\in Y,\; t\geq0. \end{equation} Furthermore, we claim that $b$ is invariant under the flow $\sigma$, that is, $b(g\cdot t)= u(t,g,b(g))$ for each $g\in Y$ and $t\geq0$. On the contrary, we assume that there exist $g\in Y$ and $s>0$ such that \begin{equation}\label{la3.2} b(g\cdot s)< u(s,g,b(g)). \end{equation} Our assumption implies that $x_{i}\gg0$, $i=1,\dots ,N$, from which we deduce that $b(g)\gg0$. For $e\gg0$ we define $e$-norm by \begin{equation}\label{enorm} \|x\|_{e}=:\inf\{\gamma>0:-\gamma e\leq_{K}x\leq_{K}\gamma e\}. \end{equation} Let $e=b(g)\gg0$ and \begin{equation}\label{la3.3} \alpha=\inf\{\|b(g)-x_{i}(g)\|_{e}:i=1,\dots ,N\}. \end{equation} Obviously, $\alpha<1$ and there exists $j\in\{1,\dots ,N\}$ such that $\alpha=\|b(g)-x_{j}(g)\|_{e}$. Hence, $b(g)-x_{j}(g)\leq\alpha b(g)$, which is equivalent to \[ x_{j}(g)\geq(1-\alpha) b(g). \] The monotonicity and concavity of the skew-product semiflow and (A2) imply that \[ u(s,g,x_{j}(g))\geq(1-\alpha)u(s,g,b(g))>(1-\alpha) b(g\cdot s). \] If $\alpha=0$, then we obtain $b(g\cdot s)\geq x_{j}(g\cdot s)=u(s,g,x_{j}(g))\geq u(s,g,b(g))$, which contradicts to \eqref{la3.2}, and hence, $\alpha$ is strictly positive. Moreover, the eventually strong monotonicity and strong concavity of the semiflow show that \[ u(s+t_0,g,x_{j}(g))\gg(1-\alpha)u(t_0,g\cdot s,b(g\cdot s)). \] The property of cones implies that we can find $0<\alpha_0<\alpha$ such that \[ u(s+t_0,g,x_{j}(g))\gg(1-\alpha_0)u(t_0,g\cdot s,b(g\cdot s)). \] Using the eventually strong monotonicity and strong concavity of the semiflow again, it then follows from \eqref{la3.1} that \[ u(t,g,x_{j}(g))\gg(1-\alpha_0)b(g\cdot t), \quad \forall t\geq s+t_0. \] Since the flow is minimal, there exists a sequence $t_n\to\infty$ such that \[ \lim_{n\to \infty}(g\cdot t_n,u(t_n,g,x_{j}(g))=(g,x_{k}(g)) \] for some $k\in\{1,\dots ,N\}$. Thus, we have \[ x_{k}(g)\geq(1-\alpha_0)b(g); \] i.e., $b(g)-x_{k}(g)\leq\alpha_0b(g)=\alpha_0 e$, which contradicts to \eqref{la3.3}. Hence, $b$ is invariant under the flow $\sigma$. Define \[ \mathcal{O}_{b}=\{(g,b(g)):g\in Y\}. \] Finally, we verify that $\mathcal{O}_{b}=\mathcal{O}$. On the contrary, assume that there exist $g\in Y$ and $j\in\{1,\dots ,N\}$ such that $b(g)>x_{j}(g)$. The eventually strong monotonicity of the semiflow implies that $b(g)\gg x_{j}(g)$, for all $g\in Y$, $j\in\{1,\dots ,N\}$, which contradicts that $b$ is the supremum. Hence, we get $\mathcal{O}_{b}=\mathcal{O}$. Furthermore, the conclusion that $\mathcal{O}$ is a copy of the base $Y$ can be obtained straight. \end{proof} \begin{corollary}\label{co3.6} Let the assumptions of Theorem \ref{th3.4} hold. Then $\mathcal{O}$ is an equilibrium point set. \end{corollary} \begin{proof} By Theorem \ref{th3.4}, we have \[ \mathcal{O}=\{(g,b(g)):g\in Y\}, \] and the map $g\mapsto b(g)$ is a bijection with $b(g\cdot t)= u(t,g,b(g))$, $\forall g\in Y$, $t\geq0$. Hence, $\mathcal{O}$ is the equilibrium point set. \end{proof} \begin{lemma}\label{le3.6} Assume that two omega limit sets satisfy $\mathcal{O}_1\ll_{C} \mathcal{O}_2$. Then there exists a positive constant $c_1$ such that \[ \|u_{x}(t,g,x_2)\|\leq c_1, \quad \forall (g,x_2)\in \mathcal{O}_2, \; t\geq0. \] \end{lemma} \begin{proof} In view of the proof of \cite[Lemma 5.6]{novo2}, we know that, for $e\gg0$ there exists a constant $\bar{c}$ (depending on $e$) such that \begin{equation}\label{enorm1} \|u_{x}(t,g,x)\|\leq \bar{c}\|u_{x}(t,g,x)e\|, \quad \forall (g,x)\in Y\times X,\; t\geq0. \end{equation} The conclusion of \cite[Lemma 5.3]{novo2} implies that there exists a positive constant $\beta>0$ such that $x_2-x_1\geq\beta e$, for all $(g,x_1)\in\mathcal{O}_1$, $(g,x_2)\in\mathcal{O}_2$. The positiveness of the linear operator $u_{x}(t,g,x_2)$ shows that \[ u_{x}(t,g,x_2)(x_2-x_1)\geq\beta u_{x}(t,g,x_2)e. \] The monotonicity and concavity of the semiflow and \eqref{kconvex} show that \[ \|u_{x}(t,g,x_2)\|\leq\frac{\bar{c}}{\beta} \|u_{x}(t,g,x_2)-u_{x}(t,g,x_1)\|,\quad \forall t\geq0. \] From the above and the compact positive invariance of $\mathcal{O}_1$ and $\mathcal{O}_2$ we can conclude that there exists a positive constant $c_1$ such that \[ \|u_{x}(t,g,x_2)\|\leq c_1, \quad \forall (g,x_2)\in \mathcal{O}_2, \; t\geq0. \] The proof is complete. \end{proof} \begin{proposition}\label{pro3.8} If $\mathcal{O}_1\ll_{C} \mathcal{O}_2$ holds, then $\mathcal{O}_2$ is a linearly stable set, i.e., $\lambda_{\mathcal{O}_2}\leq0$. \end{proposition} \begin{proof} By Definition \ref{de3.3} and Lemma \ref{le3.6}, the conclusion can be obtained immediately. \end{proof} \begin{proposition}\label{pro3.9} There exists the function $g\mapsto a(g)$ such that the set \[ Y_0=\{g\in Y:(g,a(g))\in\mathcal{O}\} \] is the continuous point set of the mapping $g\mapsto a(g)$. \end{proposition} \begin{proof} It is sufficient to prove that for any $g_{k}\to g$ there exists $g\mapsto a(g)$ such that $a(g_{k})\to a(g)$. Because of the minimality of the flow, we only to prove $a(g\cdot t_{k})\to a(g\cdot t_0)$ for any $t_{k}\to t_0$. Let $(g,x)\in\mathcal{O}$, from the definition of the omega limit set, there exists a sequence $t_n\to\infty$ such that $g_0\cdot t_n\to g$, $u(t_n,g_0,x_0)\to x$. Let \[ a(g):=\lim_{n\to \infty}u(t_n,g_0,x_0)=x. \] Then \begin{align*} a(g\cdot t_0) &=\lim_{n\to \infty}u(t_n,g_0\cdot t_0,u(t_0,g_0,x_0))\\ &=\lim_{n\to \infty}u(t_n+t_0,g_0,x_0)\\ &=\lim_{n\to \infty}u(t_0,g_0\cdot t_n,u(t_n,g_0,x_0))\\ &=u(t_0,g,x), \end{align*} and for any $k\in\mathbb{N}$, \begin{align*} \lim_{k\to \infty}a(g\cdot t_{k}) &=\lim_{k\to \infty}\lim_{n\to \infty}u(t_n,g_0\cdot t_{k},u(t_{k},g_0,x_0))\\ &=\lim_{k\to \infty}\lim_{n\to \infty}u(t_{k},g_0\cdot t_n,u(t_n,g_0,x_0))\\ &=\lim_{k\to \infty}u(t_{k},g,x)\\ &=u(t_0,g,x) =a(g\cdot t_0). \end{align*} The proof is complete. \end{proof} From \cite[Proposition 6.1]{novo2}, we have the following result . \begin{proposition}\label{p6.1} Suppose that $\mathcal{O}_1\ll_{C} \mathcal{O}_2$. If $\lambda_{\mathcal{O}_2}=0$, there exist positive constant $\hat{c}$ and $c$ such that \begin{equation}\label{ne1} \hat{c}\leq\|u_{x}(t,g,x_2)\|\leq c, \quad \forall (g,x_2)\in \mathcal{O}_2, \; t\geq0. \end{equation} \end{proposition} \begin{proposition}\label{pro3.13} Assume that $\mathcal{O}_1\ll_{C} \mathcal{O}_2$ holds and $\lambda_{\mathcal{O}_2}=0$. Then there exists a minimal subset $\mathcal{O}^{*}$ of $Y\times X$ such that $\mathcal{O}_1\ll\mathcal{O}^{*}<\mathcal{O}_2$. \end{proposition} \begin{proof} As in Proposition \ref{pro3.9}, define $Y_0=\{g\in Y:(g,a(g))\in\mathcal{O}_2\}$. Let $g_0\in Y_0$, from the definition of $Y_0$, we have $(g_0,a(g_0))\in\mathcal{O}_2$. Since $\mathcal{O}_1\ll_{C} \mathcal{O}_2$, for each $(g_0,x_1)\in\mathcal{O}_1$, we have $x_1\ll a(g_0)$. Fixed $0<\alpha<1$, define \[ y_{\alpha}=\alpha x_1+(1-\alpha)a(g_0). \] Obviously, $x_1\ll y_{\alpha}< a(g_0)$. The precompactness of the forward orbit $\{\pi(t,g_0,y_{\alpha}):t\geq\delta, \ \delta>0\}$ implies that its closure contains a minimal subset, denoted by $\mathcal{O}_{\alpha}$, i.e., \[ \mathcal{O}_{\alpha}\subset\operatorname{cls}\{(g_0\cdot t,u(t,g_0,y_{\alpha})):t\geq\delta\}. \] The monotonicity of the skew-product semiflow implies $\mathcal{O}_1\leq\mathcal{O}_{\alpha}\leq\mathcal{O}_2$. In the following, we prove that $\mathcal{O}_{\alpha}$ is required. First we check $\mathcal{O}_1\ll\mathcal{O}_{\alpha}$. For $(g,z)\in\mathcal{O}_{\alpha}$, there exist a sequence $t_n\to\infty$ such that \[ \lim_{n\to \infty}\Pi(t_n,g_0,y_{\alpha})=(g,z). \] The concavity implies that \[ u(t_n,g_0,y_{\alpha})\geq \alpha u(t_n,g_0,x_1)+(1-\alpha)u(t_n,g_0,a(g_0)). \] In addition, there exists a subsequence (assume the whole sequence), $(g,z_1)\in \mathcal{O}_1$ and $(g,z_2)\in \mathcal{O}_2$ such that \[ \lim_{n\to \infty}\Pi(t_n,g_0,x_1)=(g,z_1),\quad \lim_{n\to \infty}\Pi(t_n,g_0,a(g_0))=(g,z_2). \] Hence, we have \[ z\geq\alpha z_1 +(1-\alpha)z_2. \] Since $\mathcal{O}_1\ll_{C} \mathcal{O}_2$, $z_1\ll z_2$ holds, from which we have $z\gg z_1$, Definition \ref{def3.2} tells us $\mathcal{O}_1\ll \mathcal{O}_{\alpha}$. In the following we prove $\mathcal{O}_2\neq \mathcal{O}_{\alpha}$. On the contrary, we assume that $\mathcal{O}_2=\mathcal{O}_{\alpha}$ with $(g_0,a(g_0))\in \mathcal{O}_2\cap \mathcal{O}_{\alpha}$. Thus, there exists a sequence ${t_{k}}\to\infty$ such that $\lim_{n\to \infty}\Pi(t_{k},g_0,y_{\alpha})=(g_0,a(g_0))$. Proposition \ref{p6.1} implies that there exist a positive constant $\hat{c}>0$ such that $\hat{c}\leq\|u_{x}(t,g_0,a(g_0))\|$, $\forall t\geq0$. From the inequality \eqref{kconvex} we deduce that for all $k\in \mathbb{N}$, \begin{align*} u(t_{k},g_0,a(g_0))-u(t_{k},g_0,y_{\alpha}) &\geq u_{x}(t_{k},g_0,a(g_0))(a(g_0)-y_{\alpha})\\ &=\alpha u_{x}(t_{k},g_0,a(g_0))(a(g_0)-x_1). \end{align*} It then follows from \eqref{enorm1} and the monotonicity of the skew-product semiflow that for $e=(a(g_0)-x_1)$, we can find $l$ (which only depends on $a(g_0)$ and $x_1$) such that \[ \|a(g_0\cdot t_{k})-u(t_{k},g_0,y_{\alpha})\|\geq l>0, \quad \forall k\in \mathbb{N}. \] This contradicts that $g_0$ is a point of continuity of $a(g_0)$, which implies $\lim_{n\to \infty}(g_0\cdot t_{k},u(t_{k},g_0,y_{\alpha})) =(g_0,a(g_0))$. The proof is complete. \end{proof} \begin{theorem}\label{th4.16} If $\mathcal{O}_1\ll_{C} \mathcal{O}_2$, then $\lambda_{\mathcal{O}_2}<0$. \end{theorem} \begin{proof} Proposition \ref{pro3.8} implies that $\lambda_{\mathcal{O}_2}\leq0$, hence, it is sufficient to prove $\lambda_{\mathcal{O}_2}\neq0$. On the contrary, we assume that $\lambda_{\mathcal{O}_2}=0$. It follows from Proposition \ref{pro3.13} that there exists the subset $\mathcal{O}^{*}$ of $Y\times X$ such that $\mathcal{O}_1\ll\mathcal{O}^{*}<\mathcal{O}_2$. Let $g_0\in Y_0$, then $(g_0,a(g_0))\in\mathcal{O}_2$ and there exist $(g_0,z)\in\mathcal{O}^{*}$ and $(g_0,x_1)\in\mathcal{O}_1$ such that \[ x_1\ll z< a(g_0). \] Let $e=a(g_0)-x_1\gg0$ in \eqref{enorm} and define \[ \gamma=\inf\{\|a(g_0)-x\|_{e}:(g_0,x)\in \mathcal{O}^{*}\}. \] It is easy to see that there exists $(g_0,x)\in \mathcal{O}^{*}$ such that $\gamma=\|a(g_0)-x\|_{e}$ with $0<\gamma<1$, which implies that $a(g_0)-x\leq\gamma (a(g_0)-x_1)$; i.e., \[ x\geq(1-\gamma)a(g_0)+\gamma x_1. \] Since $a(g_0)\gg x_1$, the monotonicity and strong concavity of the skew-product semiflow implies that \begin{equation}\label{4.7} u(t,g_0,x)\gg(1-\gamma)u(t,g_0,a(g_0))+\gamma u(t,g_0,x_1). \end{equation} In view of the property of the cone, there exists $\gamma_0$ with $0<\gamma_0<\gamma$ such that \[ u(t,g_0,x)\gg(1-\gamma_0)a(g_0\cdot t)+\gamma_0 u(t,g_0,x_1), \] Hence, there exists $(g_0,y)\in \mathcal{O}^{*}$ such that \[ y\geq(1-\gamma_0)a(g_0)+\gamma_0x_1; \] i.e., $a(g_0)-y\leq\gamma_0(a(g_0)-x_1)=\gamma_0 e$, which implies that $\|a(g_0)-y\|_{e}\leq\gamma_0<\gamma$. This contradicts the definition of $\gamma$. \end{proof} \begin{theorem}\label{th4.17} If $\mathcal{O}_1\ll_{C} \mathcal{O}_2$, then $\mathcal{O}_2$ is the copy of the base $Y$, i.e., for each $g\in Y$, $\operatorname{card}(\mathcal{O}_2\cap\pi^{-1}(g))=1$. \end{theorem} \begin{proof} Since $\mathcal{O}_1\ll_{C} \mathcal{O}_2$, Theorem \ref{th4.16} tells us $\lambda_{\mathcal{O}_2}<0$, the remaining is concluded by Theorem \ref{th3.4}. \end{proof} Next, we introduce the main result of this article. \begin{theorem}\label{th4.18} If {\rm (A1)} and {\rm (A2)} hold, then for any $(g,x)\in Y\times X^{+}\setminus\{0\}$ either \begin{itemize} \item[(i)] $\lim_{t\to \infty}\|u(t,g,x)\|=+\infty$, or \item[(ii)] there exists an equilibrium point set $\mathcal{O}^{*}\subset Y\times \operatorname{int}X^{+}$ such that $\mathcal{O}(g,x)=\mathcal{O}^{*}$ and $\lim_{t\to \infty}\|u(t,g,x)-u(t,g,x^{*})\|=0$, where $(g,x^{*})=\mathcal{O}^{*}\cap \pi^{-1}(g)$. \end{itemize} \end{theorem} \begin{proof} On the contrary, we assume that (i) does not hold; i.e., the forward orbit of the skew-product semiflow is bounded, From (A1) we know $\{\Pi(t,g,x)|t\geq0\}$ is precompact. The eventually strong monotonicity implies that if $(g,x)\in Y\times (X^{+}\setminus\{0\})$, then $\mathcal{O}(g,x)=:\mathcal{O}^{*}\subset Y\times \operatorname{int}X^{+}$. It then follows from (A2) that $\mathcal{O}(g,0)=:\mathcal{O}^{0}\subset Y\times \{0\}$. Hence, $\mathcal{O}^{0}\ll_{C} \mathcal{O}^{*}$. Thus, Theorem \ref{th4.16} implies that $\lambda_{\mathcal{O}^{*}}<0$. Furthermore, Theorem \ref{th4.17} and Corollary \ref{co3.6} show that $\mathcal{O}^{*}$ is a copy of the base $Y$ and an equilibrium set, i.e., $\operatorname{card}(\mathcal{O}^{*}\cap\pi^{-1}(g))=1$, for all $g\in Y$. Next we prove that $\lim_{t\to \infty}\|u(t,g,x)-u(t,g,x^{*})\|=0$. On the contrary, we assume there exists a sequence $t_n\to\infty$ and a positive constant $\epsilon>0$ such that $\|u(t_n,g,x)-u(t_n,g,x^{*})\|>\epsilon$ for all $n\geq1$. Denote $\lim_{n\to \infty}\Pi(t_n,g,x)=(\bar{g},\bar{x}_1)$ and $\lim_{n\to \infty}\Pi(t_n,g,x^{*})=(\bar{g},\bar{x}_2)$, where $(g,x^{*})=\mathcal{O}^{*}\cap\pi^{-1}(g)$. Since $\operatorname{card}(\mathcal{O}^{*}\cap \pi^{-1}(\bar{g}))=1$, we have $\bar{x}_1=\bar{x}_2$. Thus, $0=\|\bar{x}_1-\bar{x}_2\|=\lim_{n\to \infty}\|u(t_n,g,x)-u(t_n,g,x^{*})\| \geq\epsilon$, a contradiction holds. Hence, $\lim_{t\to \infty}\|u(t,g,x)-u(t,g,x^{*})\|=0$. \end{proof} Consider the almost periodic delay differential equation \begin{equation}\label{4.4equa} \begin{gathered} y'(t)=f(t,y(t),y(t-1)), \quad \forall t\in \mathbb{R}^{+}, \\ y(s)=\phi(s), \quad \forall s\in[-1,0], \end{gathered} \end{equation} where $\phi\in C^{+}:=C([-1,0],\mathbb{R}_{+}^{n})$, the function $f=(f_1,f_2,\dots ,f_n):\mathbb{R}^{+}\times\mathbb{R}^{n}\times\mathbb{R}^{n}$ is almost periodic ( Let $(X,d)$ be metric space, a function $f\in C(\mathbb{R},X)$ is said to be \textbf{almost periodic} if for any $\epsilon>0$, there exists $l=l(\epsilon)>0$ such that every interval of $\mathbb{R}$ of length $l$ contains at least one point of the set $T(\epsilon)=\{\tau\in\mathbb{R}:d(f(t+\tau),f(t))<\epsilon,\forall t\in\mathbb{R}\}$). In addition, we propose the following properties: \begin{itemize} \item[(i)] for each $y,z\in\mathbb{R}^{n}$, $t\in\mathbb{R}$ and $i\neq j$, $\frac{\partial f_{i}}{\partial y_{j}}(t,y,z)\geq0$; If $\tilde{I}$ and $\tilde{J}$\ form a partition of $N=\{1,2,\dots ,n\}$, then there exist $\delta>0$, $i\in \tilde{I}$ and $j\in \tilde{J}$, such that \[ \big|\frac{\partial f_{i}}{\partial y_{j}}(t,y,z)\big|\geq\delta, \quad \forall y,z\in\mathbb{R}^{n}, t\in\mathbb{R}; \] \item[(ii)] for $y,z\in\mathbb{R}^{n}$, $t\in\mathbb{R}$ and $i,j\in\{1,2,\dots ,n\}$, $\frac{\partial f_{i}}{\partial z_{j}}(t,y,z)\geq0$. Furthermore, There exists $\delta>0$ such that \[ \big|\frac{\partial f_{i}}{\partial z_{j}}(t,y,z)\big|\geq\delta; \] \item[(iii)] there exists $g_0\in Y$ such that $f$ \begin{itemize} \item[(a)] is concave with respect to $(y,z)$, i.e., whenever $y^{1}\leq y^{2}$,\ $z^{1}\leq z^{2}$, \[ f(t,\lambda(y^{1},z^{1})+(1-\lambda)(y^{2},z^{2}))\geq \lambda f(t,(y^{1},z^{1}))+(1-\lambda)f(t,(y^{2},z^{2})) \] for $\lambda \in[0,1]$ and $t\in\mathbb{R}^{+}$; \item[(b)] is strongly concave with respect to $(y,z)$; i.e., whenever $y^{1}\ll y^{2}$, $z^{1}\ll z^{2}$, \[ f(t,\lambda(y^{1},z^{1})+(1-\lambda)(y^{2},z^{2}))\gg \lambda f(t,(y^{1},z^{1}))+(1-\lambda)f(t,(y^{2},z^{2})); \] for $\lambda \in(0,1)$ and $t\in[0,1]$; \end{itemize} \item[(iv)] $f(\cdot,0,0)\equiv0$. \end{itemize} We embed \eqref{4.4equa} into the skew-product semiflow $\Pi:\mathbb{R^{+}}\times Y \times C^{+}\to Y \times C^{+}$ \begin{equation}\label{4.3h} \Pi(t,g,\phi)\mapsto (\sigma_{t}(g),u(t,g,\phi))\emph{жа}, \end{equation} where for $\theta\in[-1,0]$, $u(t,g,\phi)(\theta)=y(t+\theta,g,\phi)$, and $\sigma_{t}(g(s,\cdot,\cdot))=g(s,\cdot,\cdot)\cdot t=g(t+s,\cdot,\cdot)$. $y(t,g,\phi)$ is the solution of the equation \begin{equation}\label{4.mequa} y'(t)=g(t,y(t),y(t-1)), \end{equation} and for $\theta\in[-1,0]$ and $g=(g_1,g_2,\dots ,g_n)\in Y$, $y(\theta,g,\phi)=\phi(\theta)$, where \[ Y:=\operatorname{cls}\{f_{t}|t\geq0,\quad f_{t}(s,\cdot,\cdot)=f(t+s,\cdot,\cdot)\}, \] the closure is defined in the topology of uniform convergence on compact set. From the above we deduce that $Y$ is compact metric space and $(Y,\sigma,\mathbb{R^{+}})$ is minimal. By the standard theory of delay differential equations (refer to \cite{hale,hino}), we know that for all $g\in Y $ and initial value $\phi\in C$, \eqref{4.4equa} admit a unique solution $y(t,g,\phi)$, i.e., for $\theta\in[-1,0]$, $y(\theta,g,\phi)=\phi(\theta)$. If $y(t,g,\phi)$ is the unique solution of \eqref{4.4equa} in the existence interval of $t$, then $u(t,g,\phi)$ exists for all $t>0$, and the forward orbit $\{u(t,g,\phi)|t\geq1+\delta\}$ is precompact for $\delta>0$. \begin{theorem}\label{monotonecave} The skew-product semiflow \eqref{4.3h} is eventually strongly monotone and satisfies concavity and strongly concavity, respectively; i.e., there exists $g_0\in Y$ such that \[ \lambda u(t,g,v)+(1-\lambda)u(t,g,w)\leq u(t,g,\lambda v+(1-\lambda)w) \] whenever $w\geq v$, $t\geq0$, $\lambda\in[0,1]$ and $g\in Y$, and \[ \lambda u(t,g_0,v)+(1-\lambda)u(t,g_0,w)\ll u(t,g_0,\lambda v+(1-\lambda)w) \] whenever $w\gg v$, $t\geq1$ and $\lambda\in(0,1)$. \end{theorem} \begin{proof} The eventually strong monotonicity can be obtained from \cite{novo2,novo1}. Let $\lambda\in(0,1)$ and $Z_g(t)=\lambda y(t,g,v)+(1-\lambda)y(t,g,w)$, so \[ Z'_g=\lambda g(t,y(t,g,v),v(t-1))+(1-\lambda) g(t,y(t,g,w),w(t-1)),\ \forall t\in[0,1]. \] By the monotonicity of the skew-product semiflow, if $v\leq w$, then $y(t,g,v)\leq y(t,g,w)$. It then follows from (iii)(a) that \[ Z'_g(t)\leq g(t,Z_g(t),\lambda v(t-1)+(1-\lambda)w(t-1)), \quad \forall t\in[0,1]. \] From (i), (ii) and comparison theorems for this kind of ordinary differential equation (see \cite{guzman}), we have \[ \lambda y(t,g,v)+(1-\lambda)y(t,g,w)\leq y(t,g,\lambda v+(1-\lambda)w), \quad \forall t\in[0,1] \] An inductive argument shows that for each $n\in\mathbb{N}$, \[ \lambda y(t,g,v)+(1-\lambda)y(t,g,w)\leq y(t,g,\lambda v+(1-\lambda)w), \quad \forall t\in[n,n+1]. \] Hence, \[ \lambda u(t,g,v)+(1-\lambda)u(t,g,w))\leq u(t,g,\lambda v+(1-\lambda)w), \quad \forall t\geq0. \] If $v\ll w$, the strong monotonicity implies $y(t,g_0,v)\ll y(t,g_0,w)$. From (iii)(b), for each $t\in[1,2]$, \[ z'_{g_0}(t)\ll g_0(t,z_{g_0}(t),\lambda v(t-1)+(1-\lambda)w(t-1)). \] Using a same process, comparison theorems provide $Z_{g_0}(t)\ll y(t,g_0,\lambda v+(1-\lambda)w)$. Hence, \[ \lambda y(t,g_0,v)+(1-\lambda)y(t,g_0,w))\ll y(t,g_0,\lambda v+(1-\lambda)w), \quad \forall t>0. \] That is, \[ \lambda u(t,g_0,v)+(1-\lambda)u(t,g_0,w))\ll u(t,g_0,\lambda v+(1-\lambda)w), \quad \forall t>1. \] The proof is complete. \end{proof} \begin{theorem} If \eqref{4.4equa} admits a bounded solution $y(t,\phi)$, then there exists an almost periodic solution $y^{*}(t)$, $\lim_{t\to \infty}\|y(t,\phi)-y^{*}(t)\|=0$ for $\phi\in C^{+}$ with $\phi(0)>0$. \end{theorem} \begin{proof} Theorem \ref{monotonecave} tells us that the skew-product semiflow \eqref{4.3h} is eventually strongly monotone and (strongly) concave. For any $(g,\phi)\in Y\times C^{+}$ with $\phi(0)>0$, we conclude $\mathcal{O}^{*}:=\mathcal{O}(g,\phi)\subset Y\times \operatorname{int}C^{+}$. It then follows from Theorem \ref{th4.18} that $\lim_{t\to \infty}\|y(t,\phi)-y^{*}(t)\|=0$, where $(g,y^{*}(t))=\mathcal{O}^{*}\cap \pi^{-1}(g)$. \end{proof} \subsection*{Acknowledgments} This research was supported by the NSF of China under grant 10926091, and by the Fundamental Research Funds for the Central Universities lzujbky-2010-166. \begin{thebibliography}{99} \bibitem{guzman} M. de Guzm\'an; \emph{Ecuaciones Diferenciales Ordinaries}, Ed. Alhambra, Madrid, 1975. \bibitem{hale} J. K. Hale, S. M. Verduyn Lunel, ; Introduction to Functional Differential Equations, in: Applied Mathematical Sciences, V. 99, Springer, Berlin, Heidelberg, New York, 1993. \bibitem{hetzer} G. Hetzer, W. Shen; Convergence in almost periodic competition diffusion systems, \emph{J. Math. Anal. Appl.}, \textbf{262}(2001), 307--338. \bibitem{hino} Y. Hino, S. Murakami, T. Naiko; \emph{Functional Differential Equations with Infinite Delay}, in: Lecture Notes in Mathematics, V. 1473, Springer, Berlin, Heidelberg, 1991. \bibitem{jiang} J. Jiang, X-Q. Zhao; Convergence in monotone and uniformly stable skew-product semiflows with applications, \emph{J. Reine Angew. Math.}, \textbf{589} (2005), 21--55. \bibitem{novo2} S. Novo, R. Obaya; Strictly ordered minimal subsets of a class of convex monotone skew-product semiflows, \emph{J. Differential Equations}, \textbf{196} (2004), 249--288. \bibitem{novo1} S. Novo, R. Obaya, A. M. Sanz; Attractor minimal sets for cooperative and strongly convex delay differential system, \emph{J. Differential Equations}, \textbf{208} (2005), 86--123. \bibitem{novo3} S. Novo, R. Obaya, A. M. Sanz; Attractor minimal sets for non-autonomous delay functional differential equations with applications for neural networks, \emph{Proc. Roy. Soc. London}, \textbf{461A} (2005), 2767--2783. \bibitem{shen11} W. Shen, Y. Yi; Almost Automorphic and Almost Periodic Dynamics in Skew-product Semiflows, Skew-product Semiflows, in: Memoirs of the American Mathematical Society, V. 136, No. 647, Providence, RI, 1998, 1--93. \bibitem{zhao} X-Q. Zhao; Global attractivity in monotone and subhomogeneous almost periodic systems, \emph{J. Differential Equations}, \textbf{187} (2003), 494--509. \end{thebibliography} \end{document}