\documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2004(2004), No. 64, pp. 1--18.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2004 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2004/64\hfil A limit set trichotomy for order-preserving systems] {A limit set trichotomy for order-preserving systems on time scales} \author[Christian P\"otzsche \& Stefan Siegmund \hfil EJDE-2004/64\hfilneg] {Christian P\"otzsche \& Stefan Siegmund} % in alphabetical order \address{Christian P\"otzsche \hfill\break University of Augsburg\\ Department of Mathematics\\ 86135 Augsburg\\ Germany} \email{poetzsche@math.uni-augsburg.de} \urladdr{http://www.math.uni-augsburg.de/$\sim$poetzsch} \thanks{} \address{Stefan Siegmund \hfill\break J. W. Goethe University\\ Department of Mathematics\\ 60325 Frankfurt\\ Germany} \email{siegmund@math.uni-frankfurt.de} \urladdr{http://www.math.uni-frankfurt.de/$\sim$siegmund} \date{} \thanks{Submitted April 15, 2004. Published April 27, 2004.} \thanks{The first author is supported by the ``Graduiertenkolleg: Nichtlineare Probleme in Analysis, \hfill\break\indent Geometrie und Physik'' (GRK~283) financed by the DFG and the State of Bavaria. The second\hfill\break\indent author is an Emmy Noether Fellow supported by the DFG} \subjclass[2000]{37C65. 37B55, 92D25} \keywords{Limit set trichotomy, 2-parameter semiflow, dynamic equation, \hfill\break\indent time scale} \begin{abstract} In this paper we derive a limit set trichotomy for abstract order-preserving 2-parameter semiflows in normal cones of strongly ordered Banach spaces. Additionally, to provide an example, M\"uller's theorem is generalized to dynamic equations on arbitrary time scales and applied to a model from population dynamics. \end{abstract} \maketitle \numberwithin{equation}{section} \newcommand{\ang}[1]{\langle#1\rangle} \newcommand{\norm}[1]{\|#1\|} \newcommand{\set}[1]{\{#1\}} \newtheorem{thm}{Theorem}[section] \newtheorem{cor}[thm]{Corollary} \newtheorem{lem}[thm]{Lemma} \newtheorem{defn}[thm]{Definition} \newtheorem{example}[thm]{Example} \newtheorem{rem}[thm]{Remark} \renewcommand{\labelenumi}{(\alph{enumi})} \section{Introduction} In certain relevant situations it happens that a dynamical system preserves a (partial) order relation on its state space. These systems are called \emph{order-preserving} or \emph{monotone} and the ground for their qualitative theory was laid by Krasnoselskii in his two books \cite{kra64,kra68}. Meanwhile many others made further important contributions for different types of such dynamical systems like (semi-)flows of ordinary differential equation \cite{hir82,hir84,hir85,smi86}, functional differential equations \cite{smi87,ab90}, semilinear parabolic equations \cite{hir88,tak91}, ordinary difference equations \cite{hir85,tak90}, \cite{kn93,kr92}, \cite{pt91,hp93}, random dynamical systems \cite{ac98} or general skew-product flows \cite{chu01}; compare also the monographs \cite{smi96} and \cite{chu02} for numerous examples and applications. The essential property of order-preserving dynamical systems is that they possess a surprisingly simple asymptotic behavior. In fact Krause et al.~\cite{kn93,kr92} proved a so-called \emph{limit set trichotomy} (cf.~also \cite{nes99} for nonautonomous difference equations or \cite{ac01} for random dynamical systems), describing the only three asymptotic scenarios of such systems under a certain kind of concavity. In this paper we prove such a limit set trichotomy for a general model of nonexpansive dynamical processes, namely \emph{$2$-parameter semiflows} in normal cones on time scales. They include the solution operators of dynamic equations on time scales (cf.~\cite{hil90,bp01}) and in particular of nonautonomous difference and differential equations. Beyond the unification aspect, dynamic equations on time scales are predestinated to describe the interaction of biological species with hibernation periods. The crucial point is that we provide sufficient criteria for the nonexpansiveness of such solution operators in terms of concavity and cooperativity conditions on the right-hand sides of the corresponding equations. On this occasion we generalize the classical theorem of M\"uller (cf., e.g., \cite{mue26}) to dynamic equations in real Banach spaces. Thereby we closely follow the arguments of \cite{wal01}, who considers finite-dimensional ordinary differential equations and orderings with respect to arbitrary cones. However, although our state spaces are allowed to be infinite-dimensional, we have to make the assumption that cones have nonempty interior. The use of arbitrary order cones instead of $\mathbb{R}_+^d$ even in finite dimensions has the advantage that certain equations are cooperative (see Definition~\ref{defcoop}) with respect to an ordering different from the component-wise. \section{Preliminaries} \label{sec2} Let $\mathbb{T}$ be an arbitrary \emph{time scale}, i.e., a canonically ordered closed subset of the real axis $\mathbb{R}$. Since we are interested in the asymptotic behavior of systems on such sets $\mathbb{T}$, it is reasonable to assume that $\mathbb{T}$ is unbounded above in the whole paper. Moreover, $\mathbb{T}$ is called \emph{homogeneous}, if $\mathbb{T}=\mathbb{R}$ or $\mathbb{T}=h\mathbb{Z}$, $h>0$. A \emph{$\mathbb{T}$-interval} is the intersection of a real interval with the set $\mathbb{T}$, for $a,b\in\mathbb{R}$ we write $[a,b]_\mathbb{T}:=[a,b]\cap\mathbb{T}$ and (half-)open intervals are defined analogously. $(X,d)$ denotes a metric space from now on. \begin{defn}\label{sflow} \rm A mapping $\varphi:\set{(t,\tau)\in\mathbb{T}^2:\,\tau\leq t}\times X\to X$ is denoted as a \emph{$2$-parameter semiflow on $X$}, if the mappings $\varphi(t,\tau,\cdot)=\varphi(t,\tau) : X \to X$, $\tau\leq t$, satisfy the following properties: \begin{itemize} \item[(i)] $\varphi(\tau,\tau)x=x$ for all $\tau\in\mathbb{T}$, $x\in X$, \item[(ii)] $\varphi(t,s)\varphi(s,\tau)=\varphi(t,\tau)$ for all $\tau,s,t\in\mathbb{T}$, $\tau\leq s\leq t$, \item[(iii)] $\varphi(\,\cdot,\cdot)x:\set{(t,\tau)\in\mathbb{T}^2:\,\tau\leq t} \to X$ is continuous for all $x\in X$. \end{itemize} \end{defn} \begin{rem} \rm % rmk2.2 \emph{(1)}\; Sometimes $2$-parameter semiflows are also called \emph{(evolutionary) processes} (cf., e.g., \cite[p.~100, Definition~1.1]{hv93}). \emph{(2)}\; To provide some concepts from classical ($1$-parameter) semiflows, we denote a point $x_0\in X$ as an \emph{equilibrium} of $\varphi$, if $\varphi(t,\tau)x_0=x_0$ for all $\tau\le t$ holds. Moreover, for $\tau\in\mathbb{T}$ and $x\in X$, the \emph{orbit emanating from $(\tau,x)$} is $$ \gamma_\tau^+(x):=\set{\varphi(t,\tau)x\in X:\,\tau\le t} $$ and the \emph{$\omega$-limit set of $(\tau,x)$} is given by $$ \omega_\tau^+(x) := \bigcap_{\tau\le t}\mathop{\rm closure}\set{\varphi(s,\tau)x\in X:\,t\le s}. $$ Equivalently, $\omega_\tau^+(x)$ consists of all the points $x^\ast\in X$ such that there exists a sequence $t_n\to\infty$ in $\mathbb{T}$ with $x^\ast=\lim_{n\to\infty}\varphi(t_n,\tau)x$. A subset $U\subset X$ is denoted as \emph{forward invariant}, if $\varphi(t,\tau)U\subset U$ holds for $\tau\leq t$. \end{rem} \begin{example}\label{fex} \rm \emph{(1)}\; For homogeneous time scales $\mathbb{T}$, any strongly continuous discrete $(\mathbb{T}:=h\mathbb{Z},\,h>0)$ or continuous $(\mathbb{T}:=\mathbb{R})$ $1$-parameter semiflow $\set{\phi_t}_{t\geq 0}$ evidently generates a $2$-parameter semiflow $\varphi$ via $\varphi(t,\tau):=\phi_{t-\tau}$. \emph{(2)}\; Let $X$ be some Banach space $V$ and $f:\mathbb{T}\times V\to V$. Then the standard examples for $2$-parameter-semiflows are the solution operators $\varphi(t,\tau,\cdot):V\to V$, $\tau\leq t$, of nonautonomous difference equations $v(t+1)=f(t,v(t))$, $t\in\mathbb{T} := \mathbb{Z}$, or of nonautonomous ordinary differential equations $\dot{v}(t)=f(t,v(t))$, $t\in\mathbb{T} := \mathbb{R}$ in $V$, provided that in the ODE case, $f$ is e.g.~measurable in $t$, (locally) Lipschitzian in $v$, and satisfies a certain growth condition to exclude finite escape times. The general situation, where $\mathbb{T}$ is an arbitrary closed subset of the reals, occurs in the context of dynamic equations $v^\Delta=f(t,v)$ on time scales (see\ Section \ref{sec5}). \emph{(3)}\; Let $r\geq 0$ be real, $X:=C([-r,0],\mathbb{R}^d)$ the space of continuous functions endowed with the sup-norm, and $f:\mathbb{R}\times X\to\mathbb{R}^d$ be continuous and (locally) Lipschitzian in the second argument. Then, if no finite escape times appear, the solution $v(\cdot,\tau,v^0):[\tau,\infty)\to\mathbb{R}^d$ of the retarded functional differential equation \begin{gather*} \dot{v}(t)=f(t,v_t),\\ v_t(\theta):=v(t+\theta) \quad\mbox{for all }\theta\in[-r,0] \end{gather*} satisfying the initial condition $v_\tau=v^0$ for $\tau\in\mathbb{R}$, $v^0\in X$, defines a $2$-parameter semiflow on $X$ with $\mathbb{T}=\mathbb{R}$ via $\varphi(t,\tau)v^0:=v_t(\cdot,\tau,v^0)$ (cf.~\cite{hv93}). \emph{(4)}\; Criteria for more abstract nonautonomous evolutionary equations to generate a $2$-parameter semiflow can be found in \cite{am96} and the references therein. \end{example} We need some further terminology. A self-mapping $\Phi:X\to X$ will be called \emph{nonexpansive} (on $(X,d)$), if $d(\Phi x,\Phi\bar{x})\leq d(x,\bar{x})$ for all $x,\bar{x}\in X$, and $\Phi$ will be called \emph{contractive}, if $d(\Phi x,\Phi\bar{x}) 0$ such that $\alpha^{-1}u\leq v\leq \alpha u$ on the cone $V_+$ are called the \emph{parts} of $V_+$. \item[(ii)] Let $C$ be a part of $V_+$. Then $p:C\times C\to\mathbb{R}_+$, \begin{displaymath} p(u,v) := \inf\set{\log \alpha \,: \alpha^{-1} u\leq v\leq \alpha u} \quad\mbox{for all } u,v\in C, \end{displaymath} defines a metric on $C$ called the \emph{part metric} of $C$. \end{itemize} \end{defn} \begin{rem}\label{rem-partmetric} \rm \emph{(1)}\; $u$ and $v$ lie in the same part, if and only if $p(u,v)<\infty$. \emph{(2)}\; Clearly $\mathop{\rm int} V_+$ is a part and the closure of every part is also a convex cone in the Banach space $V$. For a proof of the fact that $p$ is a metric on $C$ and for other properties of the part metric we refer to \cite{bb69} or \cite[pp.~83--86]{chu02}. \emph{(3)}\; If the cone $V_+$ is normal, then $\mathop{\rm int} V_+$ is a complete metric space with respect to the part metric $p$ (cf.~\cite{tho63}). \end{rem} Norm distance and the part metric are related by the following inequality: \begin{lem}\label{partprop} \begin{enumerate} \item If $V_+$ is normal with monotone norm, then \begin{displaymath} \norm{v-\bar{v}} \leq \big(2e^{p(v,\bar{v})}-e^{-p(v,\bar{v})}-1) \min\set{\norm{v},\norm{\bar{v}}} \quad\mbox{for all } v,\bar{v}\in V_+^\ast, \end{displaymath} \item $p|_{\mathop{\rm int} V_+\times\mathop{\rm int} V_+}$ is continuous in the norm topology on $\mathop{\rm int} V_+\times\mathop{\rm int} V_+$. \end{enumerate} \end{lem} \begin{proof} See \cite[Lemma~2.3]{kn93} for (a), while assertion (b) can be found in \cite[Proof of Theorem~2]{nes99}. \end{proof} \section{A Limit Set Trichotomy}\label{sec3} The following theorem is a clear manifestation of the general experience that contractivity drastically simplifies the possible long-term behavior of a dynamical system. It is the main result in the abstract part of this paper. In the autonomous discrete time case a limit set trichotomy was discovered (and so named) by Krause and Ranft~\cite{kr92} and generalized in \cite{kn93} to infinite-dimensional autonomous difference equations; in addition, \cite{nes99} considers such nonautonomous systems. \begin{thm}[Limit Set Trichotomy]\label{lst} Let $V_+\subset V$ be a normal cone, $\mathop{\rm int} V_+\neq\emptyset$ and assume that $\varphi$ is a $2$-parameter semiflow on $V_+$ with the following properties: \begin{itemize} \item[(i)] There exists a real $T>0$ such that for all $t,\tau\in\mathbb{T}$ satisfying $T\leq t-\tau$, one has $\varphi(t,\tau)V_+^\ast\subset\mathop{\rm int} V_+$ and that the mapping $\varphi(t,\tau)|_{\mathop{\rm int} V_+}$ is nonexpansive, \item[(ii)] for all $(\tau,v)\in\mathbb{T}\times V_+$ every bounded orbit $\gamma_\tau^+(v)$ is relatively compact in the norm topology. \end{itemize} Then for every $\tau\in\mathbb{T}$ the following trichotomy holds, i.e., precisely one of the following three cases applies: \begin{enumerate} \item For all $v\in V_+^\ast$ the orbits $\gamma_\tau^+(v)$ are unbounded in norm, \item for all $v\in V_+$ the orbits $\gamma_\tau^+(v)$ are bounded in norm and for all $v\in V_+^\ast$ we have $\lim_{t\to\infty}\norm{\varphi(t,\tau)v}=0$, \item for all $v\in V_+$ the orbits $\gamma_\tau^+(v)$ are bounded in norm, the $\omega$-limit sets $\omega_\tau^+(v)$ are nonempty and for all $v\in V_+^\ast$ they have a nontrivial accumulation point. \end{enumerate} If, moreover, $\omega_\tau^+(v)\subset\mathop{\rm int} V_+\cup\set{0}$ for all $v\in V_+^\ast$ and the mappings $\varphi(t,\tau)|_{\mathop{\rm int} V_+}$ are uniformly contractive for all $t,\tau\in\mathbb{T}$ with $T\leq t-\tau$, then in case (c) we have \begin{equation} \lim_{t\to\infty}\big[\varphi(t,\tau)v_1-\varphi(t,\tau)v_2\big] = 0 \quad\mbox{for all }v_1,v_2\in V_+^\ast. \label{star2} \end{equation} \end{thm} \begin{rem} \rm %rmk3.2 \emph{(1)}\; Condition \eqref{star2} implies that all $\omega$-limit sets $\omega_\tau^+(v)$, $v\in V_+^\ast$, are identical, and it excludes the existence of two different equilibria of $\varphi$. In fact, if $\varphi$ possesses an equilibrium $v_0\in V_+^\ast$, then \eqref{star2} guarantees $\omega_\tau^+(v)=\set{v_0}$ for all $v\in V_+^\ast$. In the ``autonomous'' situation of a homogeneous time scale $\mathbb{T}$ and a $2$-parameter semiflow induced by a $1$-parameter semiflow (cf.\ Example \ref{fex}(1)), the assumption $\omega_\tau^+(v)\subset\mathop{\rm int} V_+\cup\set{0}$ becomes superfluous. This yields by the invariance of $\omega_\tau^+(v)$ and hypothesis (ii), i.e., $\omega_\tau^+(v)=\varphi(t,\tau)\omega_\tau^+(v)\subset\mathop{\rm int} V_+$ for $T\leq t-\tau$. \emph{(2)}\; One can show a stronger limit set trichotomy, if $\varphi$ is induced by a discrete $1$-parameter semiflow (cf.~\cite[Theorem~3.1]{kn93}). More results in the finite-dimensional situation $V_+=\mathbb{R}_+^d$ can be found in \cite[Theorems~1,~2]{kr92} and related topics are contained in \cite[Theorem~1.1]{tak90} or \cite[Theorem~4.1]{pt91}. Furthermore, \cite[Lemma 4]{nes99} provides sufficient conditions for the right-hand side of nonautonomous difference equations to generate a uniformly contractive $2$-parameter semigroup. On general time scales, stronger limit set trichotomies can be found in \cite{poe04} under the assumption that $\varphi$ is uniformly ascending. \emph{(3)}\; We also briefly comment the situation when $\varphi$ comes from an ordinary differential equation $(\mathbb{T}=\mathbb{R})$. For autonomous cooperative systems in $\mathbb{R}^2$, a prototype result has been given by \cite[Theorem~2.3]{hir82}. If $\varphi$ comes from a time-periodic equation, \cite[Theorem~3.1]{smi86} proved a ``limit set dichotomy'' under certain assumptions on the Floquet multipliers. Similar results are given by \cite[Theorems~3,~4]{kr92}; \cite[Theorem~6.8]{hir88} considers general continuous $1$-parameter semiflows, and \cite[Theorem~3.1]{chu01} proves a limit set trichotomy for order-preserving skew-product flows. \emph{(4)}\; Finally, the case of random dynamical systems is considered in \cite[Theorem~4.2]{ac01} and \cite[pp.~123--124, Theorem~4.4.1]{chu02}. \end{rem} \begin{proof} Let $\tau\in\mathbb{T}$ be arbitrary, but fixed. If (a) holds, then obviously (b) and (c) cannot hold. If (a) does not hold, then there exists a $v_1\in V_+^\ast$ such that the orbit $\gamma_\tau^+(v_1)$ is bounded, i.e., $\|\varphi(t,\tau)v_1\|\leq M$ for some $M\geq 0$ and all $t\geq\tau$. Now we show that in this case every orbit $\gamma_\tau^+(v)$, $v\in V_+$, is bounded in norm. Let the vector $v_2\in V_+$ be arbitrary. Then either (i) $\gamma_\tau^+(v_2)$ is bounded or (ii) there exists a $t'\in\mathbb{T}$ with $T\leq t'-\tau$ such that $\varphi(t',\tau)v_2\not=0$. In case (ii) it follows from assumption~(i) that $\varphi(t,\tau)v_1,\varphi(t,\tau)v_2\in\mathop{\rm int} V_+$ for $t\geq t'$. The Remarks~\ref{rem-partmetric} (1) and (2) imply $K:=p(\varphi(t',\tau)v_1,\varphi(t',\tau)v_2)<\infty$. Using the $2$-parameter semiflow property of $\varphi$ (cf.~Definition~\ref{sflow}(ii)) together with the fact that the mappings $\varphi(t,t')$, $t\geq t'+T$, are nonexpansive, we obtain \begin{displaymath} p(\varphi(t,\tau)v_1,\varphi(t,\tau)v_2)\leq K \quad \text{for } t\geq t'+T \,. \end{displaymath} Consequently, Lemma~\ref{partprop}(a) provides the estimate \begin{displaymath} \norm{\varphi(t,\tau)v_2} \leq \norm{\varphi(t,\tau)v_2-\varphi(t,\tau)v_1}+\norm{\varphi(t,\tau)v_1} \leq 2e^K M \end{displaymath} for all $t\geq t'+T$, proving that $\gamma_\tau^+(v_2)$ is bounded. Now we show that either (b) or (c) holds. By assumption (ii) the orbits $\gamma_\tau^+(v)$, $v\in V_+$, are relatively compact and therefore $\omega_\tau^+(v) \not= \emptyset$, moreover, the relation $\omega_\tau^+(v) = \set{0}$ is equivalent to $\lim_{t\to\infty}\varphi(t,\tau)v = 0$. We show that \begin{displaymath} \omega_\tau^+(v_1) = \set{0} \text{ for a single } v_1\in V_+^\ast \quad\implies\quad \omega_\tau^+(v) = \set{0} \text{ for any } v\in V_+^\ast \,. \end{displaymath} To this end, we assume that there exist $v_1,v_2,v_2^\ast\in V_+^\ast$ with $\omega_\tau^+(v_1)=\set{0}$ and $v_2^\ast\in\omega_\tau^+(v_2)\setminus\set{0}$. Then there exists a sequence $t_n\to\infty$ in $\mathbb{T}$ with \begin{displaymath} \lim_{n\to\infty}\varphi(t_n,\tau)v_1 = 0 \quad\text{and} \quad \lim_{n\to\infty}\varphi(t_n,\tau)v_2 = v_2^\ast \,, \end{displaymath} where we assume without lost of generaliry that $t_0 = \tau$ and $t_{n+1}-t_n\geq T$, which implies by assumption (i) that $\varphi(t_n,\tau)v_i\in\mathop{\rm int} V_+$ for $i=1,2$ and $n\in\mathbb{N}$. Using the 2-parameter semiflow property and the fact that the mappings $\varphi(t_{n+1},t_n)$, $n\in\mathbb{N}$, are nonexpansive, we get \begin{displaymath} p(\varphi(t_n,\tau)v_1,\varphi(t_n,\tau)v_2) \leq \dots \leq p(\varphi(t_1,\tau)v_1,\varphi(t_1,\tau)v_2) \leq p(v_1,v_2) \end{displaymath} for $n\in\mathbb{N}$. Choosing $N\in\mathbb{N}$ such that $\|\varphi(t_n,\tau)v_1\|\leq \|\varphi(t_n,\tau)v_2\|$ for $n\geq N$, Lemma~\ref{partprop}(a) implies the contradiction \begin{align*} \|\varphi(t_n,\tau)v_2\| & \leq \|\varphi(t_n,\tau)v_2-\varphi(t_n,\tau)v_1\| + \|\varphi(t_n,\tau)v_1\| \\ & \leq 3 e^{p(v_1,v_2)} \|\varphi(t_n,\tau)v_1\| \to 0 \quad\text{for } n\to\infty \,, \end{align*} proving that either (b) or (c) is true. It remains to show \eqref{star2} under the additional assumptions that the mappings $\varphi(t,s)$, $T\leq t-s$, are uniformly contractive, and that $\omega_\tau^+(v)\in\mathop{\rm int} V_+\cup\set{0}$ for all $v\in V_+^\ast$. Assume that \eqref{star2} does not hold. Then there exists $v_1, v_2\in V_+^\ast$, an $\varepsilon > 0$ and a sequence $t_n\to\infty$ in $\mathbb{T}$ with \begin{equation} \norm{\varphi(t_n,\tau)v_1-\varphi(t_n,\tau)v_2} \geq \varepsilon \quad\mbox{for all } n\in\mathbb{N}, \label{star3} \end{equation} where we assume without lost of generaliry $t_1\geq\tau+T$ and $t_{n+1}-t_n\geq T$, which implies that $\varphi(t_n,\tau)v_i \not= 0$ for $i=1,2$ and $n\in\mathbb{N}$. Since the orbits $\gamma_\tau^+(v_1)$ and $\gamma_\tau^+(v_2)$ are relatively compact there exists a subsequence of $(t_n)_{n\in\mathbb{N}}$, which we denote by $(t_n)_{n\in\mathbb{N}}$ again, such that the limits \begin{displaymath} v_1^\ast := \lim_{n\to\infty}\varphi(t_n,\tau)v_1 \quad\text{and} \quad v_2^\ast := \lim_{n\to\infty}\varphi(t_n,\tau)v_2 \end{displaymath} exist. By assumption $v_1^\ast,v_2^\ast\in\mathop{\rm int} V_+\cup\set{0}$ and by \eqref{star3} $v_1^\ast\neq v_2^\ast$. We can also rule out that $v_1^\ast = 0$ and $v_2^\ast\in\mathop{\rm int} V_+$, since in this case, choosing $N\in\mathbb{N}$ such that $\|\varphi(t_n,\tau)v_1\|\leq \|\varphi(t_n,\tau)v_2\|$ for $n\geq N$, Lemma~\ref{partprop}(a) would imply \begin{align*} \norm{\varphi(t_n,\tau)v_2} & \leq \norm{\varphi(t_n,\tau)v_2-\varphi(t_n,\tau)v_1}+\norm{\varphi(t_n,\tau)v_1} \\ & \leq 3 e^{p(\varphi(t_1,\tau)v_1,\varphi(t_1,\tau)v_2)} \|\varphi(t_n,\tau)v_1\| \to 0 \quad\text{for } n\to\infty \,, \end{align*} contradicting $v_1^\ast \not= v_2^\ast$. Hence we have $v_1^\ast,v_2^\ast\in\mathop{\rm int} V_+$. The $2$-parameter semiflow property and the fact that the mappings $\varphi(t_{n+1},t_n)$, $n\in\mathbb{N}$, are uniformly contractive, imply the estimates \begin{gather*} p(\varphi(t_n,\tau)v_1,\varphi(t_n,\tau)v_2) \leq c(\varphi(t_n,\tau)v_1,\varphi(t_n,\tau)v_2) < \dots \\ < p(\varphi(t_1,\tau)v_1,\varphi(t_1,\tau)v_2) \leq c(\varphi(t_1,\tau)v_1,\varphi(t_1,\tau)v_2) \,. \end{gather*} Since monotone and bounded sequences converge and the mappings $p$ and $c$ are continuous in $(v_1^\ast,v_2^\ast)$ by Lemma~\ref{partprop}(b) and assumption, respectively, we get \begin{displaymath} p(v_1^\ast,v_2^\ast) = c(v_1^\ast,v_2^\ast) \end{displaymath} in contradiction to $c(v_1^\ast,v_2^\ast) < p(v_1^\ast,v_2^\ast)$, thus proving that \eqref{star2} holds. \end{proof} \section{Subhomogeneous Order-Preserving Semiflows} The results of Section \ref{sec3} are only helpful, if one has verifiable conditions which guarantee that a $2$-parameter semiflow is nonexpansive with respect to the part-metric. Sufficient conditions will be given by the following notions. \begin{defn} \rm %def4.1 Let $U\subset V_+$. A $2$-parameter semiflow $\varphi$ on $V_+$ is said to be \begin{itemize} \item[(i)] \emph{order-preserving on $U$} if \begin{displaymath} u,v\in U,\,u\leq v \quad \implies \quad \varphi(t,\tau)u\leq \varphi(t,\tau)v \quad\mbox{for all }\tau\leq t; \end{displaymath} \item[(ii)] \emph{strictly order-preserving on $U$}, if it is order-preserving on $U$ and \begin{displaymath} u,v\in U,\,u < v \quad\implies\quad \varphi(t,\tau)u <\varphi(t,\tau)v \quad\mbox{for all }\tau\leq t; \end{displaymath} \item[(iii)] \emph{strongly order-preserving on $U$}, if $\mathop{\rm int} V_+\neq\emptyset$, if $\varphi$ is order-preserving on $U$ and \begin{displaymath} u,v\in U,\,u \ll v \quad\implies\quad \varphi(t,\tau)u \ll\varphi(t,\tau)v \quad\mbox{for all }\tau\leq t. \end{displaymath} \end{itemize} \end{defn} We now introduce a class of order-preserving $2$-parameter semiflows which possess a certain concavity property we call \emph{subhomogeneity} (sometimes also named sublinearity). Subhomogeneity means concavity for the particular case in which one of the reference points is $0$, hence asks less and is thus more general than classical concavity. The autonomous version of this property plays an important role in many studies and applications, see \cite{kra64,kra68}, \cite{kn93,kr92}, \cite{smi86,tak90} and the references therein. \begin{defn} \rm %def4.2 A $2$-parameter semiflow $\varphi$, which is order-preserving on $V_+$, is said to be \begin{itemize} \item[(i)] \emph{subhomogeneous}, if for any $v\in V_+$ and for any $\alpha\in (0,1)$ we have \begin{equation}\label{sublin} \alpha\varphi(t,\tau)v \leq \varphi(t,\tau)\alpha v \quad\mbox{for all }\tau 1$ we have \begin{equation}\label{no33} \varphi(t,\tau)\alpha v \leq \alpha \varphi(t,\tau)v \quad\mbox{for all }\tau\tau \end{displaymath} and this leads to the contractivity of $\varphi$. \end{proof} \section{Order-Preserving Dynamic Equations}\label{sec5} Supplementing our explanations from Section \ref{sec2}, we need some further terminology. $I_V$ denotes the identity map on $V$. The \emph{dual cone} $V_+'$ of $V_+$ is the set of all linear and continuous mappings $v':V\to\mathbb{R}$ such that $\ang{v,v'}\geq 0$ for all $v\in V_+$, where $\ang{v,v'}:=v'(v)$ is the duality mapping. If $V$ is a Hilbert space, then $V_+'$ can be identified with a subset of $V$ through the Riesz representation theorem (cf.~\cite[p.~104, Theorem~2.1]{lan93}). The elements of $V_+^\star:=V_+'\setminus\set{0}$ are called \emph{supporting forms} and we define $\mathcal{L}(V_+):=\set{T\in\mathcal{L}(V):\,T(V_+)\subset V_+}$. Any such operator $T\in\mathcal{L}(V_+)$ is called \emph{positive}, and \emph{strictly positive}, if $Tv=0$ implies $v=0$ for any $v\in V_+$. First of all, we can characterize the (interior) points of a cone in terms of linear functionals. \begin{lem}\label{lem51} For any $v\in V$ the following holds: \begin{enumerate} \item $v\in V_+$ $\Leftrightarrow$ $\ang{v,v'}\geq 0$ for all $v'\in V_+^\star$, \item $v\in\mathop{\rm int} V_+$ $\Leftrightarrow$ $\ang{v,v'}>0$ for all $v'\in V_+^\star$. \end{enumerate} \end{lem} \begin{proof} See~\cite[p.~221, Proposition 19.3]{dei85}. \end{proof} At this point we introduce some further notation concerning the calculus on time scales (see also~\cite{bp01}). Remember that $\mathbb{T}$ is a closed subset of $\mathbb{R}$ which is assumed to be unbounded above. $\sigma(t):=\inf\set{s\in\mathbb{T}:\,t0$. Thus, let $\nu$ denote a solution of \eqref{system} starting at $\tau\in\mathbb{T}$ in $\nu(\tau)\in\mathop{\rm int} V_+$. If the claim were false, then there exists a finite $t^\ast\in\mathbb{T}$ given by \[ t^\ast:=\sup T,\quad T:=\set{t\geq\tau:\nu(s)\in V_+\text{ for all }s\in[\tau,t]_{\mathbb{T}}}. \] Since the cone $V_+$ is closed we have $\nu(t^\ast)\in V_+$. The point $t^\ast$ is right-dense, because otherwise $\nu(\sigma(t^\ast))=\nu(t^\ast)+ \mu(t^\ast)f(t^\ast,\nu(t^\ast))\in V_+$ would yield the contradiction $t^\ast<\sigma(t^\ast)\in T$. Moreover, $\nu(t^\ast)\in\partial V_+$, because the assumption $\nu(t^\ast)\in\mathop{\rm int} V_+$ would imply the existence of a neighborhood $U\subset V_+$ of $\nu(t^\ast)$ and a $\mathbb{T}$-neighborhood $N$ of $t^\ast$ with $\nu(t)\in U$ for $t\in N$, since $\nu$ is continuous as the solution of \eqref{system}. This again contradicts the definition of $t^\ast$. Now by Lemma~\ref{lem51}(b) there exists a supporting form $v'\in V_+^\star$ such that $\ang{\nu(t^\ast),v'}=0$ and by definition $t^\ast$ is the time when the solution $\nu$ leaves $V_+$, which by Lemma~\ref{lem51}(a) gives us $\ang{\nu(t),v'}\leq 0$ for $t$ from a right-sided neighborhood of $t^\ast$. One finds $\ang{\frac{\nu(t)-\nu(t^\ast)}{t-t^\ast},v'}\leq 0$ and in the limit $t\searrow t^\ast$, we therefore obtain the contradiction \begin{displaymath} 0 \geq \ang{\nu^\Delta(t^\ast),v'} = \ang{f(t^\ast,\nu(t^\ast)),v'} > 0 \end{displaymath} with a view to the above (strengthened) hypothesis. \textit{(IV)} The verification of the assumption under the general hypothesis yields as follows: For arbitrary reals $\varepsilon>0$ and some fixed $e\in\mathop{\rm int} V_+$ one can apply the above step (III) to the solution $\nu_\varepsilon$ of the dynamic equation \begin{displaymath} v^\Delta=f(t,v)+\varepsilon e \end{displaymath} and consequently, for any $\tau\in\mathbb{T}$ and $v_0\in V_+$ one obtains $\nu_\varepsilon(t)\in V_+$ for $\tau\leq t$, provided that $\nu_\varepsilon(\tau)\in V_+$. Now let $\nu$ denote the solution of \eqref{system} satisfying $\nu(\tau)=\nu_\varepsilon(\tau)$. One should bear in mind that $\nu_\varepsilon(t)$ is continuous in $(\varepsilon,t)$ (cf.~\cite[p.~39, Satz~1.2.17]{poe02}) and consequently uniformly continuous on the set $K\times[0,\varepsilon_0]$, where $K\subset[\tau,\infty)_{\mathbb{T}}$ is a compact $\mathbb{T}$-interval and $\varepsilon_0>0$ arbitrary. By a standard argument, the solutions $\nu_\varepsilon$ converge to $\nu$ uniformly on $K$ as $\varepsilon\searrow 0$. \end{proof} \begin{cor}\label{inv} Let $V_+\subset V$ be a normal cone and assume that \begin{itemize} \item[(H3)] in any left-dense $t_0\in\mathbb{T}$ there exists a left-sided $\mathbb{T}$-neighborhood $N_0(t_0)$ of $t_0$ such that $f(s,0)\in V_+$ for all $s\in N_0(t_0)$. \end{itemize} Then $V_+^\ast$ is forward invariant for \eqref{system}, if and only if every right-dense $t_0\in\mathbb{T}$, any $v\in\partial V_+$, $v'\in V_+^\star$ such that $\ang{v,v'}=0$ satisfy $\ang{f(t_0,v),v'}\geq 0$, and, for every right-scattered $t_0\in\mathbb{T}$ any $v\in V_+^\ast$ satisfies $v+\mu(t_0)f(t_0,v)\in V_+^\ast$. \end{cor} \begin{proof} We have to show two directions: $(\Rightarrow)$ If $V_+^\ast$ is a forward invariant set, then the assertion can be shown analogously to step (II) in the proof of Lemma~\ref{lem52}. $(\Leftarrow)$ Using the induction principle (cf.~\cite[p.~4, Theorem~1.7]{bp01}) we deduce the statement \begin{displaymath} \mathcal{A}(t): v\neq 0 \quad\implies\quad \varphi(t,\tau)v\neq 0 \quad\mbox{for all }\tau\leq t. \end{displaymath} Above all, choose $v\in V_+^\ast$ arbitrarily. \begin{itemize} \item $\mathcal{A}(\tau)$ obviously holds since $\varphi(\tau,\tau)v=v$. \item Let $t\geq\tau$ be right-scattered and $\mathcal{A}(t)$ be true. Then by the $2$-parameter semiflow property and the assumption one immediately gets \begin{displaymath} \varphi(\sigma(t),\tau)v = \varphi(t,\tau)v+\mu(t)f(t,\varphi(t,\tau)v) \neq 0 \end{displaymath} i.e., $\mathcal{A}(\sigma(t))$ holds. \item Let $t\geq\tau$ be right-dense and assume that $\mathcal{A}(t)$ is valid. Then $\varphi(t,\tau)v\neq 0$ implies that $\varphi(s,\tau)v\neq 0$ in a right-sided $\mathbb{T}$-neighborhood $N$ of $t$. Hence $\mathcal{A}(t)$ yields $\varphi(s,\tau)v\neq 0$ for $s\in N$. \item Let $t\geq\tau$ be left-dense and $\mathcal{A}(s)$ be true for $st_n$, is also $\mathbb{R}_+^d$-cooperative if the matrix $I_{\mathbb{R}^d}+(t_{n+1}-t_n)D_2f(t_n,u)$ is nonnegative. Since the off-diagonal elements of $D_2f(t,u)$ are nonnegative, this is true, if the diagonal entries $a_{ii}(t_n,u)$, $i=1,\dots,d$, of the matrix $D_2f(t_n,u)$ satisfy the condition $a_{ii}(t_n,u)\geq\frac{-1}{t_{n+1}-t_n}$ with the stepsize $\mu(t_n) = t_{n+1}-t_n$ of the Euler discretization. \end{rem} \begin{thm}[M\"uller's Theorem]\label{muller} Let $U\subset V$ be $V_+$-convex. \begin{enumerate} \item If \eqref{system} is $V_+$-cooperative on $U$, then $\varphi$ is order-preserving on $U$. \item Conversely, if $\varphi$ is order-preserving on $U$, then the dynamic equation \eqref{system} is $V_+$-cooperative on $U$. \end{enumerate} \end{thm} \begin{proof} (a) Choose $u,v\in U$ with $u\leq v$ and because of the $V_+$-convexity of $U$ one has $u+h(v-u)\in U$ for $h\in[0,1]$. Then the mean value theorem (cf.~\cite[p.~341, Theorem~4.2]{lan93}) yields \begin{displaymath} \varphi(t,\tau,v)-\varphi(t,\tau,u) = \int_0^1D_3\varphi(t,\tau,u+h(v-u))(v-u)\,dh \quad \text{for all }\tau\le t \end{displaymath} and since $D_3\varphi(t,\tau,w)\in\mathcal{L}(V_+)$, $w\in U$, we obtain $D_3\varphi(t,\tau,w)(v-u)\in V_+$. Now convexity of the integral implies the claim $\varphi(t,\tau,v)\leq\varphi(t,\tau,u)$. (b) Using the fact that $D_3\varphi(t,\tau,v)$ solves the dynamic equation \eqref{var} in $\mathcal{L}(V_+)$, we get the assertion (b) with a view to Corollary~\ref{cor53}. \end{proof} The next part of this section is dedicated to sufficient conditions for strictly and strongly order-preserving mappings. Since we have not assumed regressivity of $f$ (cf.~\cite[pp.~321--322, Definition~8.14(ii)]{bp01}) the mapping $\varphi(t,\tau):V\to V$, $\tau\leq t$, needs not to be a homeomorphism. Hence the arguments of~\cite[pp.~32--33, Proof of Proposition 1.1]{smi96} do not apply directly. \begin{cor}\label{cor1} Let $U\subset V$ be $V_+$-convex. If for a $V_+$-cooperative system \eqref{system} on $U$ one of the following conditions \begin{itemize} \item[(i)] $I_V+\mu(t)f(t,\cdot):V\to V$ in one-to-one on $U$ for any $t\in\mathbb{T}$, \item[(ii)] $I_V+\mu(t)f(t,\cdot):V\to V$ is strictly order-preserving on $U$ for any $t\in\mathbb{T}$ \end{itemize} holds, then $\varphi$ is strictly order-preserving on $U$. \end{cor} \begin{proof} We proceed in two steps: \textit{(I)} To show that (i) implies (ii), fix arbitrary $t\in\mathbb{T}$, $u\in U$ and abbreviate $F(u):=u+\mu(t)f(t,u)$. Observing the fact \begin{equation} \varphi(\sigma(t),t)u=u+\mu(t)f(t,u)=F(u), \label{star} \end{equation} it is evident that $F$ is strictly order-preserving on $U$. \textit{(II)} We apply the induction principle (cf.~\cite[p.~4, Theorem~1.7]{bp01}) to the statement \begin{displaymath} \mathcal{A}(t): u0. \end{displaymath} Thus the subhomogeneity of the mapping $\varphi$ is equivalent to the property that $\ang{\alpha D_3\varphi(t,\tau,\alpha v)v-\varphi(t,\tau,\alpha v),v'}\leq 0$, i.e., by Lemma~\ref{lem51}(a) to the condition \eqref{cond}. (b) Now let $\tau0$ the external ``seed''. The equation \eqref{dis} guarantees that we have the inclusion $v(\tau_{n+1})\in\mathop{\rm int}\mathbb{R}_+^d$ after each winter --- independent of $v(t_n)\in\mathbb{R}_+^d$. For the continuous growth we lean on \cite{kr92} and consider the ODEs \begin{equation} \dot{v}_i=v_iF_i(v,t) \quad\mbox{for all }i=1,\dots,d, \label{con} \end{equation} on the intervals $[\tau_n,t_n]$, $n\in\mathbb{N}_0$, where the mappings $F_i:\mathbb{R}^d\times\mathbb{T}\to\mathbb{R}$ are continuously differentiable in each state space variable $v_1,\dots,v_d$. Obviously the boundary $\partial\mathbb{R}_+^d$ is forward invariant with respect to \eqref{con} and therefore any solution of \eqref{con} cannot leave the standard cone $\mathbb{R}_+^d$ for times $t\in[\tau_n,t_n]$, $n\in\mathbb{N}_0$. Combining both situations, we arrive at a dynamic equation \eqref{system} with right-hand side $f=(f_1,\dots,f_d)$ and \begin{displaymath} f_i(t,v) := \begin{cases} v_iF_i(v,t) &\mbox{for } t\in[\tau_n,t_n)\\ \frac{q_i(t)-1}{\mu(t)}v_i+\frac{p_i(t)}{\mu(t)} &\mbox{for } t=t_n\,. \end{cases} \end{displaymath} If we assume that the ODE \eqref{con} has no finite escape times, then the mapping $f$ satisfies the assumptions (H1)--(H2). In addition, the standard cone $\mathbb{R}_+^d$ is forward-invariant with respect to \eqref{system}. As a canonical state space for \eqref{system} we consider the cone $V_+=\mathbb{R}_+^d$, which evidently satisfies the assumption (H0), and is $\mathbb{R}_+^d$-convex, since the nonnegative orthant is convex. Under the assumption \begin{itemize} \item[$(C)$] $D_jF_i(u,t)\geq 0$ for all $u\in\mathbb{R}_+^d$, $i\neq j$, and $t\in\bigcup_{n\in\mathbb{N}_0}[\tau_n,t_n)$, \end{itemize} the system \eqref{system} is $\mathbb{R}_+^d$-cooperative on $\mathbb{R}_+^d$ and we obtain from Theorem~\ref{muller}(a) that its solution $\varphi$ is order-preserving. On the other hand, if we suppose \begin{itemize} \item[$(S)$] $\sum_{j=1}^{d}v_jD_jF_i(v,t)\leq 0$ for all $v\in\mathbb{R}_+^d$, $i=1,\dots,d$ and $t\in\bigcup_{n\in\mathbb{N}_0}[\tau_n,t_n)$, \end{itemize} then using Theorem~\ref{subt}(a) one can show that $\varphi$ is also subhomogeneous. So, due to Lemma~\ref{sub}(a), $\varphi(t,\tau)$, $\tau\leq t$, must be nonexpansive with respect to the part metric on $\mathbb{R}_+^d$. Finally, since each $\mathbb{T}$-interval of length greater or equal than $T$ contains a right-scattered point, we have $\varphi(t,\tau,v)\in\mathop{\rm int}\mathbb{R}_+^d$ for $T\leq t-\tau$ and $v\in\mathbb{R}_+^d$. Therefore the assumptions of Theorem~\ref{lst} are satisfied and our limit set trichotomy applies. In particular, if $p_i$ is bounded away from zero, we can exclude case (b) of Theorem~\ref{lst} and all solutions of the general nonautonomous dynamic equation \eqref{system} are either unbounded, or bounded with nonempty $\omega$-limit sets. \begin{example}[Kolmogorov systems] \rm A particularly relevant special case of the symbiotic interaction discussed above, are so-called \emph{Kolmogorov systems} which have the following biological interpretation (cf.~\cite{fs95}): Think of a hierarchy of species $v_1,\dots,v_d$, where $v_i(t)$ is the biomass of the $i$th species. In this hierarchy, $v_i$ interacts only with $v_{i-1}$ and $v_{i+1}$. Such a hierarchy may occur in steep mountain side or in a lake, where each population dominates a specific altitude or depth, respectively, but is obliged to cooperate with other populations in the (narrow) overlap of their zones of dominance. So we only modify the law for the continuous growth and consider the system of ODEs \begin{equation} \begin{gathered} \dot{v}_1 = v_1F_1(v_1,v_2,t)\\ \dot{v}_i = v_iF_i(v_{i-1},v_i,v_{i+1},t) \quad\mbox{for all }i=2,\dots,d-1\\ \dot{v}_d = v_dF_d(v_{d-1},v_d,t) \end{gathered} \label{last} \end{equation} to describe the behavior on the intervals $[\tau_n,t_n]$, where the mappings $F_1,\dots,F_d$ are continuously differentiable in their state space variables. Furthermore, the conditions $(C)$ and $(S)$ reduce to \begin{itemize} \item $D_2F_1(v_1,v_2,t),D_1F_i(v_1,v_2,v_3,t),D_3F_i(v_1,v_2,v_3,t), D_1F_d(v_1,v_2,t)\geq 0$ for all $v_1,v_2,v_3\in\mathbb{R}_+$, $i=2,\dots,d-1$, and $t\in\bigcup_{n\in\mathbb{N}_0}[\tau_n,t_n)$, \item $\sum_{j=1}^{2}v_jD_jF_1(v_1,v_2,t)\leq 0$ and $\sum_{j=1}^{3}v_jD_jF_i(v_1,v_2,v_3,t)\leq 0$, as well as $\sum_{j=1}^{2}v_jD_jF_d(v_1,v_2,t)\leq 0$ for all $v_1,v_2,v_3\in\mathbb{R}_+$, $i=2,\dots,d-1$, and $t\in\bigcup_{n\in\mathbb{N}_0}[\tau_n,t_n)$, \end{itemize} respectively. They guarantee that the right-hand side of \eqref{system} is $\mathbb{R}_+^d$-cooperative and generates a subhomogeneous $2$-parameter semiflow. Consequently our limit set trichotomy from Theorem~\ref{lst} applies. Explicit biological systems modelled by \eqref{last} can be found in \cite{fs95}. \end{example} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \begin{thebibliography}{00} \bibitem{ab90} O.~Arino and F.~Bourad, \emph{On the asymptotic behavior of the solutions of a class of scalar neutral equations generating a monotone semi-flow}, Journal of Differential Equations \textbf{87} (1990), 84--95. \bibitem{ac98} L.~Arnold and I.~Chueshov, \emph{Order-preserving random dynamical systems: Equilibria, attractors, applications}, Dynamics and Stability of Systems \textbf{13} (1998), 265--280. \bibitem{ac01} L.~Arnold and I.~Chueshov, \emph{A limit set trichotomy for order-preserving random systems}, Positivity \textbf{5(2)} (2001), 95--114. \bibitem{am96} B.~Aulbach and N.~Van Minh, \emph{Nonlinear semigroups and the existence and stability of semilinear nonautonomous evolution equations}, Abstract and Applied Analysis \textbf{1(4)} (1996), 361--380. \bibitem{bb69} H.~Bauer and H. S. Bear, \emph{The part metric in convex sets}, Pacific Journal of Mathematics \textbf{30(1)} (1969), 15--33. \bibitem{bp01} M.~Bohner and A.~Peterson, \emph{Dynamic Equations on Time Scales --- An Introduction with Applications}, Birk\-h\"auser, Boston, 2001. \bibitem{chu01} I.~Chueshov, \emph{Order-preserving skew-product flows and nonautonomous parabolic systems}, Acta Applicandae, Mathematicae \textbf{65} (2001), 185--205. \bibitem{chu02} I.~Chueshov, \emph{Monotone Random Systems. Theory and Applications}, Sprin\-ger-Verlag, Berlin, 2002. \bibitem{dei77} K.~Deimling, \emph{Ordinary Differential Equations in Banach Spaces}, Lecture Notes in Mathematics, 596, Springer-Verlag, Berlin, 1977. \bibitem{dei85} K.~Deimling, \emph{Nonlinear Functional Analysis}, Springer-Verlag, Berlin, 1985. \bibitem{fs95} H. I. Freedman and H. L. Smith, \emph{Tridiagonal competitive-cooperative Kolmogorov systems}, Differential Equations and Dynamical Systems, \textbf{3(4)} (1995), 367--382. \bibitem{hv93} J. K. Hale and S. M. Verduyn Lunel, \emph{Introduction to Functional Differential Equations}, Springer-Verlag, New York, 1993. \bibitem{hp93} P.~Hess and P.~Pol\'{a}\v{c}ik, \emph{Boundedness of prime periods of stable cycles and convergence to fixed points in discrete dynamical systems}, SIAM Journal of Mathematical Analysis \textbf{24(5)} (1993), 1312--1330. \bibitem{hil90} S.~Hilger, \emph{Analysis on measure chains --- a unified approach to continuous and discrete calculus}, Results in Mathematics \textbf{18} (1990), 18--56. \bibitem{hir82} M. W. Hirsch, \emph{Systems of differential equations which are competitive or cooperative. i: Limit sets}, SIAM Journal of Mathematical Analysis \textbf{13(2)} (1982), 167--179. \bibitem{hir84} M. W. Hirsch, \emph{The dynamical systems approach to differential equations}, Bulletin of the American Mathematical Society \textbf{11} (1984), 1--64. \bibitem{hir85} M. W. Hirsch, \emph{Systems of differential equations that are competitive or cooperative. ii: Convergence almost everywhere}, SIAM Journal of Mathematical Analysis \textbf{16(3)} (1985), 426--439. \bibitem{hir88} M. W. Hirsch, \emph{Stability and convergence in strongly monotone dynamical systems}, Journal f\"ur die reine und angewandte Mathematik \textbf{383} (1988), 1--53. \bibitem{kra64} M. A. Krasnoselskii, \emph{Positive Solutions of Operator Equations}, Noordhoff, Groningen, 1964. \bibitem{kra68} M. A. Krasnoselskii, \emph{The Operator of Translation Along Trajectories of Differential Equations}, Translations of Mathematical Monographs, Vol.~19, American Mathematical Society, Providence, Rhode Island, 1968. \bibitem{kn93} U.~Krause and R. D. Nussbaum, \emph{A limit set trichotomy for self-mappings of normal cones in Banach spaces}, Nonlinear Analysis, Theory, Methods \& Applications \textbf{20(7)} (1993), 855--870. \bibitem{kr92} U.~Krause and P.~Ranft, \emph{A limit set trichotomy for monotone nonlinear dynamical systems}, Nonlinear Analysis, Theory, Methods \& Applications \textbf{16(4)} (1992), 375--392. \bibitem{lan93} S.~Lang, \emph{Real and Functional Analysis}, Graduate Texts in Mathematics, 142, Springer-Verlag, Berlin, 1993. \bibitem{mue26} M.~M\"uller, \emph{\"Uber das Fundamentaltheorem in der Theorie der gew\"ohnlichen Differentialgleichungen}, Mathematische Zeitschrift \textbf{26} (1926), 619--645. \bibitem{nes99} T.~Nesemann, \emph{A limit set trichotomy for positive nonautonomous discrete dynamical systems}, Journal of Mathematical Analysis and Applications \textbf{237} (1999), 55--73. \bibitem{pt91} P.~Pol\'{a}\v{c}ik and I.~Tere\v{s}\v{c}\'{a}k, \emph{Convergence to cycles as a typical asymptotic behavior in smooth strongly monotone discrete-time dynamical systems}, Archive for Rational Mechanics and Analysis \textbf{116} (1991), 339--360. \bibitem{poe02} C.~P\"otzsche, \emph{Langsame Faserb\"undel dynamischer Gleichungen auf Ma{\ss}ketten} (in german), Ph.D.~Thesis, Universit\"at Augsburg, Berlin, 2002. \bibitem{poe04} C.~P\"otzsche, \emph{A limit set trichotomy for abstract $2$-parameter semiflows on time scales}, Functional Differential Equations \textbf{11(1--2)} (2004), 133--140. \bibitem{sch71} H. H. Schaefer, \emph{Topological Vector Spaces}, Springer-Verlag, Berlin, 1971. \bibitem{smi86} H. L. Smith, \emph{Cooperative systems of differential equations with concave nonlinearities}, Nonlinear Analysis, Theory, Methods \& Applications \textbf{10(10)} (1986), 1037--1052. \bibitem{smi87} H. L. Smith, \emph{Monotone semiflows generated by functional differential equations}, Journal of Differential Equations \textbf{66} (1987), 420--422. \bibitem{smi96} H. L. Smith, \emph{Monotone Dynamical Systems. An Introduction to the Theory of Competitive and Cooperative Systems}, American Mathematical Society, Providence, Rhode Island, 1996. \bibitem{tak90} P.~Tak\'{a}\v{c}, \emph{Asymptotic behavior of discrete-time semigroups of sublinear, strongly increasing mappings with applications to biology}, Nonlinear Analysis, Theory, Methods \& Applications \textbf{14(1)} (1990), 35--42. \bibitem{tak91} P.~Tak\'{a}\v{c}, \emph{A fast diffusion equation which generates a monotone local semiflow. i: Local existence and uniqueness} and \emph{ii: Global existence and asymptotic behavior}, Differential and Integral Equations \textbf{4(1)} (1991), 151--174, 175--187. \bibitem{tho63} A. C. Thompson, \emph{On certain contraction mappings on a partially ordered vector space}, Proceedings of the American Mathematical Society \textbf{14} (1963), 438--443. \bibitem{wal01} S.~Walcher, \emph{On cooperative systems with respect to arbitrary orderings}, Journal of Mathematical Analysis and Applications \textbf{263} (2001), 543--554. \end{thebibliography} \end{document}