\documentclass[twoside]{article} \usepackage{amssymb, amsmath} \pagestyle{myheadings} \markboth{\hfil Nonautonomous attractors of skew-product flows \hfil EJDE--2001/58} {EJDE--2001/58\hfil R. A. Johnson \& P. E. Kloeden \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol. {\bf 2001}(2001), No. 58, 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 (login: ftp)} \vspace{\bigskipamount} \\ % Nonautonomous attractors of skew-product flows with digitized driving systems % \thanks{ {\em Mathematics Subject Classifications:} 34D35, 37B55. \hfil\break\indent {\em Key words:} pullback attractor, digitization, Bebutov flow. \hfil\break\indent \copyright 2001 Southwest Texas State University. \hfil\break\indent Submitted July 05, 2001. Published August 29, 2001. \hfil\break\indent Supported by the GNAFA and the DFG Forschungsschwerpunkt ``Ergodentheorie, \hfil\break\indent Analysis und effiziente Simulation dynamischer Systeme" } } \date{} % \author{R. A. Johnson \& P. E. Kloeden} \maketitle \begin{abstract} The upper semicontinuity and continuity properties of pullback attractors for nonautonomous differential equations are investigated when the driving system of the generated skew-product flow is digitized. \end{abstract} \newtheorem{theorem}{Theorem}[section] \newtheorem{proposition}[theorem]{Proposition} \renewcommand{\theequation}{\thesection.\arabic{equation}} \catcode`@=11 \@addtoreset{equation}{section} \catcode`@=12 \section{Introduction}\label{intro} The objective of this paper is to study the semicontinuity and continuity properties of pullback attractors for nonautonomous differential equations \begin{equation}\label{eq1} x' = f(t,x), \quad x \in \mathbb{R}^d, t \in \mathbb{R}, \end{equation} under perturbation of the driving system through a digitization procedure. We will formulate this problem in concrete terms using the language of skew-product flows \cite{CKS,S,SY}, in particular, the Bebutov approach to the skew-product flow concept. Thus let $\mathcal{F}$ be some topological vector space of mappings $f:\mathbb{R} \times \mathbb{R}^d \to \mathbb{R}^d$ and, for each $t\in\mathbb{R}$, let $\theta_t:\mathcal{F}\to\mathcal{F}$ be the translation operator defined by $\theta_t(f)(s,x):= f(t+s,x)$ for all $s\in\mathbb{R}$ and $x\in\mathbb{R}^d$. Now let $P\subset \mathcal{F}$ be a metrizable compact set which is invariant in the sense that, if $p\in P$, then $\theta_t(p)\in P$ for all $t\in \mathbb{R}$. Suppose further that the mapping $\mathbb{R} \times P\to P$ defined by $(t,p)\mapsto\theta_t(p)$ is continuous. This just says that $(P,\{\theta_t : t \in \mathbb{R}\})$ is a topological flow or autonomous dynamical system. Now consider the family of differential equations \begin{equation}\label{eq2} x' = p(t,x). \end{equation} We assume, for each $x_0\in\mathbb{R}^d$ and $p \in P$, that the solution $x(t)=\phi(t,x_0,p)$ of (\ref{eq2}) satisfying $x(0)=x_0$ exists locally and is unique. We also assume that, if $t\in\mathbb{R}$, $x_0\in \mathbb{R}^d$ and $p\in P$, then the correspondence $(t,x_0,p)\mapsto (\phi(t,x_0,p),\theta_t(p))$ is continuous on its domain of definition $V\subset \mathbb{R} \times \mathbb{R}^d \times P$. This correspondence $(t,x_0,p) \mapsto (\phi(t,x_0,p),\theta_t(p)): V\to\mathbb{R}^d \times P$ is the skew product (local) flow on $\mathbb{R}^d \times P$ defined by the family of equations (\ref{eq2}). The flow $(P,\{\theta_t : t \in\mathbb{R} \})$ is referred to as the \textbf{driving system} of equations (\ref{eq2}). For each $p\in P$, the solutions $x(t)$ of (\ref{eq2}) can be viewed as projections to $\mathbb{R}^d$ of trajectories of the above skew-product flow. When the equations (\ref{eq2}) satisfy an appropriate dissipativity condition, they induce a global pullback attractor ${\bf A}=\bigcup_{p \in P} (A_p \times \{p\}) \subset\mathbb{R}^d\times P$ , where each fiber $A_p$ is defined by a procedure which amounts to a nonlinear version of the classical Weyl limit-point construction. The set ${\bf A}$ is compact and is invariant with respect to the above skew-product system on $\mathbb{R}^d \times P$ defined by the solutions of the equations (\ref{eq2}). For each $p\in P$, one can think of $A_p$ as a ``pointwise" pullback attractor; see Section \ref{prel} for details. The problem we pose is that of studying the semicontinuity, resp. continuity, properties of ${\bf A}$ and the individual fibers $A_p$, as the compact translation-invariant set $P$ is varied within $\mathcal{F}$. For technical simplicity, throughout the paper, we assume that the equations (\ref{eq2}) contract a fixed large ball in $\mathbb{R}^d$. In this way the concepts of local pullback attractor and global pullback attractor are made equivalent. This problem has been considered by Kloeden and Kozyakin \cite{KK2001,KK2002}, who, in particular, studied the upper semicontinuity in the Hausdorff sense of the $A_p$ under perturbation in $P$ when the driving system $(P,\{\theta_t:t \in\mathbb{R}\})$ has the shadowing property. Here we will not assume that shadowing holds. Rather, we will instead study perturbations in certain spaces $\mathcal{F}$ of a general compact, metrizable, Bebutov-invariant subset $P$. Among these perturbations are those determined by \textbf{digitizing} a given time-varying vector field $f(t,x)$. By this, we mean the following: the time-axis is decomposed into half-open intervals of lengths, say, between $\delta/2$ and $\delta$ for each $\delta>0$. On each such interval, the time-varying vector field $f(t,x)$ is replaced by an autonomous vector field$\bar{f}(x)$ (which usually depends on the particular interval). For example, one may choose $\bar{f}(x)$ to be the time-average of $f(t,x)$ over the given (or previous) interval, or as some particular value $f(t_{*},x)$, or in some other way as well. If now there is a compact, metrizable, Bebutov-invariant subset $P$ of $\mathcal{F}$ such that $f\in P$, then we can identify (\ref{eq1}) as one equation in the family (\ref{eq2}). In this way one is led to identify appropriate perturbations $P_{\delta}\subset\mathcal{F}$ of $P$. We will formulate and prove results to the effect that if $${\bf A}^{\delta}=\bigcup\left(A_{p_{\delta}} \times \{p_{\delta}\} : p_{\delta} \in P_{\delta}\right\}$$ is the pullback attractor for the skew-product dynamical system defined by $(P_{\delta},\{\theta_t : t \in\mathbb{R} \})$, and if $p_{\delta}\in P_{\delta}$ corresponds to a digitized version of equation (\ref{eq1}), then $A^{\delta}_{p_{\delta}}$ converges upper semicontinuously to $A_f$ as $\delta \to 0$, with respect to the Hausdorff metric on compact subsets of $\mathbb{R}^d$. We will also provide sufficient conditions that $A^{\delta}_{p_{\delta}}$ converges continuously to $A_f$ as $\delta\to0$, with respect to the Hausdorff metric. We will make use of two distinct sets of methods in our study study ${\bf A}$ and its fibers $A_p$. The first is drawn from the area of topological dynamics and dynamical systems. In applying these methods, we will make systematic use of the the skew-product local flow on $\mathbb{R}^d \times P$ induced by the solutions of equations (\ref{eq2}). The second is taken from the fixed point theory of nonlinear mappings on Banach spaces. Fixed point theory can be applied here for the following reason (among others): the set $A_p$ is equal, under our hypotheses, to the set of initial conditions $x_0\in\mathbb{R}^d$ for which the corresponding solution $x(t)=\phi(t, p,x_0)$ of (\ref{eq2}) exists and is bounded for all $t\in \mathbb{R}$. These solutions can be viewed as the fixed points of an appropriate nonlinear mapping. In order to illustrate the points made above, we consider the example \begin{equation}\label{eq3} x' = f(t,x) = - x + h(t,x), \quad x \in \mathbb{R}, t \in \mathbb{R}, \end{equation} where $h$ is uniformly continuous and uniformly Lipschitz in $x$ on each subset of the form $\mathbb{R} \times K$ where $K\subset \mathbb{R}^d$ is compact. Assume that there are positive numbers $a$ and $\sigma$, such that for each $t\in\mathbb{R}$ one has $$ a+h(t,-a) \geq \sigma, \quad -a+h(t,a) \leq - \sigma. $$ Let $f_{\delta}(t,x)$ be obtained by digitizing $f$: thus, for example, we might set $$ f_{\delta}(t,x) = -x + \frac{1}{\delta} \int_{n\delta}^{(n+1)\delta} h(s,x) ds, \quad t \in [n\delta,(n+1)\delta) $$ for each $n\in\mathbb{Z}$. Using methods of dynamical systems, we will first show that $A^{\delta}_{f_{\delta}}$ exists and tends upper semicontinuously to $A_f$ as $\delta \to 0$, and moreover that this convergence is uniform in $t$ when $f$ is replaced by $\theta_t(f)$ for any $t\in\mathbb{R}$. On the other hand, a bounded solution of (\ref{eq3}) is expressible as a fixed point of the mapping $T$ defined for each $t\in\mathbb{R}$ by $$ T[x](t) = - \int_t^{\infty} e^{(t-s)} f(s,x(s)) \, ds. $$ This mapping is defined and continuous on the Banach space $C_b$ of bounded, continuous real-valued functions on $\mathbb{R}$. If $|h(t,x)|\leq a$ whenever $|x|\leq a$ and $t\in\mathbb{R}$, and if $\left|\frac{\partial h}{\partial x}\right|\leq\alpha<1$ for all $t$ $\in$ $\mathbb{R}$ and $|x|\leq a$, then $T$ is a contraction on the ball $\{x(\cdot) \in C_b :\|x\|_{\infty} \leq a\}$. One has that $A_f=\{x_0 \in \mathbb{R} : \mbox{the solution $x(\cdot)$ of (\ref{eq3}) with $x(0)=x_0$ is bounded on $(-\infty,\infty)$} \}$, and that the set of solutions $x(\cdot)$ of (\ref{eq3}) which are bounded on $(-\infty,\infty)$ is the fixed point set of $T$. We will show that the digitization gives rise to continuous convergence in the Hausdorff sense $A^{\delta}_{f_{\delta}}$ to $A_f$ as $\delta\to0$, and moreover the convergence is uniform in $t$ when $f$ is replaced by $\theta_t(f)$ for any $t\in\mathbb{R}$. \section{Preliminaries} \label{prel} In this section, we formulate the concept of driving system in a way which is convenient for present purposes. We define the notion of pullback attractor when the driving system is compact and when a uniform dissipativity condition is valid. Finally we introduce various basic definitions which will be needed later on. Let $\mathcal{F}$ denote the vector space of all functions (time-varying vector fields) $f:\mathbb{R}\times \mathbb{R}^d\to\mathbb{R}^d$ which satisfy the following properties: \begin{enumerate} \item[i)] For each compact set $K\subset\mathbb{R}^d$, $f$ is uniformly continuous on $\mathbb{R} \times K$; \item[ii)] For each compact set $K\subset\mathbb{R}^d$, there exists a constant $L_K$ (also depending on $f$) so that $$ \|f(t,x) - f(t,y)\| \leq L_K \, \|x-y\| \quad \mbox{for all } x,y \in K, \ t \in \mathbb{R}. $$ \end{enumerate} The second condition states that $f$ is uniformly Lipschitz in $x$ on each set of the form $\mathbb{R}\times K$ where $K\subset \mathbb{R}^d$ is compact. It will be clear that all of our results can be formulated and proved when the vector fields $f$ in question satisfy less stringent conditions. However, Conditions i) and ii) are not particularly restrictive and they permit a simple exposition of the facts we wish to present. We put the topology of uniform convergence on compact sets on $\mathcal{F}$. Thus a sequence $\{f_n\}$ of elements of $\mathcal{F}$ converges to $f\in\mathcal{F}$ if and only if $f_n(t,x)\to f(t,x)$ uniformly on each set $D\times K\subset\mathbb{R}\times\mathbb{R}^d$ when $D$ and $K$ are compact. This topology is metrizable, but not complete. There is a natural flow (Bebutov flow) $\{\theta_t: t \in \mathbb{R} \}$ defined on $\mathcal{F}$ by translation of the $t$-variable, specifically $\theta_t(f)(s,x)$ $:=$ $f(t+s,x)$ for all $f$ $\in$ $\mathcal{F}$, $x\in \mathbb{R}^d$ and $s$, $t$ in $\mathbb{R}$. A simple but basic observation can now be made: if $f\in \mathcal{F}$ (i.e., if $f$ satisfies the conditions i) and ii)), then the orbit closure $P:=\rm{cls} \{\theta_t(f): t \in \mathbb{R} \}\subset \mathcal{F}$ is compact. This follows from the uniform continuity conditions on $f$. Moreover, condition ii) holds for each $p\in P$ with the same set of Lipschitz constants $\{L_K\}$. See \cite{S} for further details. Consider the family of differential equations \begin{equation}\label{eq4} x' = p(t,x), \end{equation} where $p$ ranges over $P$. If $x_0\in\mathbb{R}^d$ and $p\in P$, then equation (\ref{eq4}) admits a unique maximally defined solution $x(t)=\phi(t,x_0,p)$ satisfying $x(0)=x_0$. We define a local flow $\{\pi_t: t \in \mathbb{R} \}$ on $\mathbb{R}^d \times P$ by setting $\pi_t(x_0,p)=(\phi(t,x_0,p),\theta_t(p))$ for all triples $(t,x_0,p)$ such that the right-hand side is well-defined. This local flow is said to be of skew-product type because the component $\theta_t(p)$ does not depend on $x_0$. As noted in the Introduction, the flow $(P,\{\theta_t: t \in \mathbb{R}\})$ is referred to as the \textbf{driving system} for the family (\ref{eq4}). Of course, a skew-product local flow on $\mathbb{R}^d \times P$ can be constructed for any compact, translation-invariant subset of $\mathcal{F}$; it is not required that $P$ be generated as above by the translates of a fixed vector field $f$ (when $P$ is so generated, it is referred to as the \textbf{hull} of $f$). Next we turn to the concept of pullback attractor. Let us begin with a fixed vector field $f\in\mathcal{F}$ and the corresponding differential equation \begin{equation}\label{eq5} x' = f(t,x). \end{equation} To simplify matters, we assume that there exists $R>0$ such that \begin{equation}\label{eq6} \left < 0 \end{equation} for all $t\in\mathbb{R}$ and $x$ $\in$ $\mathbb{R}^d$ satisfying $\|x\|$ $\geq$ $R$, where $\left<\cdot,\cdot\right>$ denotes the standard inner product on $\mathbb{R}^d$. Using the uniform continuity properties of $f$, one sees that condition (\ref{eq6}) actually implies that, for some $\eta>0$, $$ \left \leq - \eta $$ for all $t\in\mathbb{R}$ and $x$ $\in$ $\mathbb{R}^d$ satisfying $\|x\|=R$. Passing to the hull $P$ of $f$, we see that, for each $p\in P$, the closed ball $B_R=\{ x\in \mathbb{R}^d:\|x\|\leq R\}$ is positively invariant with respect to the solutions of the differential equation (\ref{eq4}). In particular, if $p\in P$ and $x_0\in B_R$, then the solution $x(t)$ of (\ref{eq4}) satisfying $x(0)=x_0$ exists and lies in $B_R$ for all $t\geq 0$. Let us write $\phi(t,x_0,p)$ for the solution $x(t)$ of (\ref{eq4}) satisfying $x(0)=x_0$. If $D\subset \mathbb{R}^d$, we write $\phi(t,D,p)=\{\phi(t,x_0,p) : x_0 \in D\}$. Define \begin{equation}\label{eq7} A_p = \bigcap_{t\geq 0} \phi\left(t,B_R,\theta_{-t}(p)\right), \end{equation} then put ${\bf A}=\bigcup_{p \in P} \left(A_p \times \{p\} \right)\subset\mathbb{R}^d \times P$. We refer to $A_p$ as the pullback attractor for the single equation (\ref{eq4}), and to ${\bf A}$ as the pullback attractor for the family (\ref{eq4}), or more simply as the global pullback attractor. Clearly, the sets $A_p$ are all nonempty and compact, and $\phi$-invariant in the sense that $\phi\left(t,A_p,p\right)=A_{\theta_t(p)}$. We will consider the sets $A_p$ and ${\bf A}$ in more detail in the following sections. In the remainder of the present section we briefly recall two basic definitions, namely those of exponential dichotomy and of the Hausdorff distance. Let $P$ be a compact, translation-invariant subset of $\mathcal{F}$. Let us suppose that $P$ consists entirely of linear vector fields: $p(t,x)=P(t)x$, where we permit an abuse of notation. The function $P(\cdot)$ takes values in the set $\mathcal{M}_d$ of $d\times d$ real matrices and is uniformly continuous on $\mathbb{R}$. Let $\Psi_p(t)$ be the fundamental matrix of the linear ordinary differential equation \begin{equation}\label{eq8} x' = P(t) x ; \end{equation} thus $\Psi_p(t)$ is the $d\times d$-matrix solution of (\ref{eq8}) such that $\Psi_p(0)=I$, the $d\times d$ identity matrix. Let $\mathcal{Q}$ be the set of linear projections $Q: \mathbb{R}^d\to\mathbb{R}^d$; this set has finitely many connected components determined by the possible dimensions of the image of $Q$. \paragraph{Definition} Say that the family (\ref{eq8}) has an \textbf{exponential dichotomy} (ED) over $P$ if there are positive constants $\gamma$, $L$ and a continuous projection-valued function $Q$ defined by $p \mapsto Q_p\in\mathcal{Q}$ such that for all $p\in P$, \begin{gather*} \left\| \Psi_p(t) Q_p \Psi_p(s)^{-1} \right\| \leq L \exp^{-\gamma (t-s)} \quad (t \geq s), \\ \left\| \Psi_p(t) (I-Q_p) \Psi_p(s)^{-1} \right\| \leq L \exp^{\gamma (t-s)} \quad (t \leq s)\,. \end{gather*} Now let $(\mathcal{X},d)$ be a metric space and let $A$ and $B$ be nonempty compact subsets of $\mathcal{X}$. Define $\mathop{\rm dist}(a,B):=\min_{b\in B} d(a,b)$ and then define the Hausdorff semi-distance $$ H^{*}(A,B) := \max_{a\in A} \mathop{\rm dist}(a,B). $$ Thus, if $\epsilon>0$ and $H^{*}(A,B)<\epsilon$, then for each $a\in A$ there exists $b\in B$ such that $d(a,b)<\epsilon$. Finally, define the Hausdorff distance $H(A,B)$ between $A$ and $B$ as follows: $$ H(A,B) := \max \left\{H^{*}(A,B), H^{*}(B,A)\right\}. $$ Note that if $B$ happens to be a singleton, $B=\{b\}$, then $H^{*}(A,B) =H(A,B)$. \section{Semicontinuity results}\label{USC} We begin by fixing a compact, translation-invariant subset $P$ of the topological vector space $\mathcal{F}$ described in the preceding section. We assume that there exist numbers $R>0$, $\eta>0$ so that, if $\|x\|=R$ and $p\in P$, then \begin{equation}\label{eq9} \left \leq - \eta. \end{equation} Let $B_R=\left\{x \in \mathbb{R}^d:\|x\| \leq R \right\}$ be the closed ball in $\mathbb{R}^d$ centered at the origin with radius $R$. We make some remarks about the pullback attractor $A_p$ and the global pullback attractor ${\bf A}$, first when $P$ is held fixed and then when it is varied in some systematic way. First of all, it follows from (\ref{eq9}) that, if $t>s$, then $\phi\left(t,B_R,\theta_{-t}(p)\right) \subset \phi\left(s,B_R,\theta_{-s}(p)\right)$. Thus the intersection in (\ref{eq7}) is over a decreasing collection of sets. Using the continuity property of the reduced \v{C}ech homology functor \v{H} in the category of compact spaces together with the fact that each set $\phi\left(t,B_R,\theta_{-t}(p)\right)$ is homeomorphic to a ball, we have \v{H}$(A_p)=0$. In particular, we have: \begin{proposition} \label{prop1} For each $p\in P$, the space $A_p$ is connected; in fact $A_p$ is $\infty$-proximally connected in the sense of \cite{Gorn}. \end{proposition} We record a second fact which also follows quickly from (\ref{eq9}) and from the definition of $A_p$. \begin{proposition} \label{prop2} For each $p\in P$, one has $A_p=\{x_0 \in \mathbb{R}$: the solution $x(\cdot)$ of (\ref{eq2}) with $x(0)=x_0$ is defined on the entire real axis and satisfies $\|x(t)\|\leq R$ for all $t$ in $\mathbb{R}\}$. \end{proposition} \paragraph{Proof.} Let $x_0\in A_p$. Then for each $t<0$, there exists $\bar{x}\in B_R$ such that $\phi\left(t,\bar{x},\theta_{-t}(p)\right)=x_0$, and one has $x(t)=\bar{x}$. It follows that $x(t)$ is defined and satisfies $\|x(t)\|\leq R$ for all $t\in\mathbb{R}$. It is equally easy to see that, if the solution $x(t)$ of (\ref{eq2}) satisfying $x(0)=x_0$ exists and is bounded on $\mathbb{R}$, then $x_0\in A_p$. \hfil$\Box$ \medskip It follows from Proposition \ref{prop2} that ${\bf A}\subset \mathbb{R}^d \times P$ is compact, and from this one sees that $\check{H}({\bf A})= \check{H}(P)$ because of the continuity of the \v{C}ech homology functor on compact spaces. One also sees that, if $\mathcal{K}(\mathbb{R}^d)$ is the space of all nonempty compact subsets of $\mathbb{R}^d$, then the mapping $p\mapsto A_p:P\to\mathcal{K}(\mathbb{R}^d)$ is upper semicontinuous in the sense that (using the notation of Section \ref{prel}): $$ H^{*}\left(A_{p_n},A_p\right) \to 0 \quad \mbox{whenever} \quad p_n \to p \mbox{ in } P. $$ Next let $f\in\mathcal{F}$ be a vector field satisfying condition (\ref{eq6}). We want to consider the upper semicontinuity properties of the pullback attractor $A_f$ when $f$ is digitized. It will be convenient and informative to study the upper semicontinuity properties of the pullback attractor $A_f$ using the language of skew-product flows. First we introduce some terminology. By a \textbf{digitization} we mean a procedure which, to each $f\in\mathcal{F}$ and each real number $\delta>0$, assigns the following data with the indicated properties: \begin{enumerate} \item[I)] There is a collection $\mathcal{I}^{\delta}$ $=$ $\{I^{\delta}_j \ : \ j \in \mathbb{Z}\}$ of nonempty half-open intervals in $\mathbb{R}$ such that $\cup_{j=-\infty}^{\infty} I^{\delta}_j=\mathbb{R}$, and such that each interval $I^{\delta}_j$ has length $\leq$ $\delta$ and (say) $\geq$ $\delta/2$. \item[II)] To each $f\in\mathcal{F}$ there is associated a collection $\{f_{\delta}^j : \delta > 0, j \in \mathbb{Z}\}$ of autonomous vector fields. There is a positive function $\omega=\omega(\epsilon)$, defined for positive values of $\epsilon$ and tending to zero as $\epsilon\to0+$, such that for each interval $I^{\delta}_j \in \mathcal{I}^{\delta}$ and each $x\in\mathbb{R}^d$ the following property holds: if $\epsilon_x=\sup \left\{ \|f(r,x) - f(s,x)\| : r,s \in I^{\delta}_j \right\}$, then $$ \left\|f_{\delta}^j (x) - f(t,x)\right\| \leq \omega(\epsilon_x), \quad t \in I^{\delta}_j. $$ \item[III)] There is a positive function $\omega_1=\omega_1(M)$, which is defined for positive values of $M$ and which depends only on $M$, such that, if $x$, $y$ in $\mathbb{R}^d$ satisfy $\|f(t,x)-f(t,y)\|\leq M$ for all $t$ in some interval $I^{\delta}_j$, then $$ \left\|f_{\delta}^j (x) -f_{\delta}^j (y)\right\| \leq \omega_1(M) \|x-y\| $$ for all $\delta>0$. \item[IV)] There is a positive function $\omega_2=\omega_2(\eta)$, defined for positive values of $\eta$ and tending to zero as $\eta\to0+$, such that, if $J\subset\mathbb{R}$ is an interval and if $x\in\mathbb{R}^d$ is a point, and if $f$, $\tilde{f}\in \mathcal{F}$ satisfy $\|f(t,x)-\tilde{f}(t,x)\|\leq \eta$ for all $t\in J$, then $$ \left\|f_{\delta}^j (x) -\tilde{f}_{\delta}^j (x)\right\| \leq \omega_2(\eta) $$ for all $\delta>0$ and all $j$ such that $I^{\delta}_j\subset J$. Although these properties are cumbersome to state, they are reasonable requirements on a digitization scheme. Now let $f\in\mathcal{F}$ be a vector field satisfying (\ref{eq6}). Put $f_{\delta}(t,x)=f_{\delta}^j(x)$ for $t\in I^{\delta}_j$, $j\in\mathbb{Z}$. Abusing language slightly, we call $\{f_{\delta}:\delta >0\}$ a digitization of $f$. The vector fields $f_{\delta}$ discussed in the Introduction are obtained by procedures for which I)--IV) are satisfied, so these $f_{\delta}$ are digitizations in our sense. In fact , the subintervals $\mathcal{I}^{\delta}$ in I) for each fixed $\delta>0$ of such digitizations often also satisfies the following recurrence condition, in which case it will be called a \textbf{recurrent digitization}. \item[V)] Fix $\delta>0$. To each $\eta>0$ there corresponds a number $T$ (which may depend on $\delta$ as well as on $\eta$) so that each interval $[a,a+T] \subset\mathbb{R}$ contains a number $s$ such that $\mathop{\rm dist} (\mathcal{I}^{\delta},\mathcal{I}^{\delta}+s)<\eta$. Here $\mathcal{I}^{\delta}$ $+$ $s$ is the $s$-translate of $\mathcal{I}^{\delta}$ and $\mathop{\rm dist}$ is the Hausdorff distance on $\mathbb{R}$. \end{enumerate} % Now consider the differential equation \begin{equation}\label{eq10} x' = f_{\delta}(t,x) \end{equation} for each $\delta>0$. Though $f_{\delta}$ is only piecewise continuous in $t$, it nevertheless admits a unique local solution $x(t,x_0)$ for each initial condition $x(0,x_0)=x_0\in \mathbb{R}^d$; moreover $x(t,x_0)$ is jointly continuous on its domain of definition. Using property II) and condition (\ref{eq6}) on $f$, we see that $f_{\delta}$ also satisfies condition (\ref{eq6}) for small $\delta>0$. It follows that the pullback attractor $A_{f_{\delta}} \subset \mathbb{R}^d$ of the equation (\ref{eq10}), which is defined by the formula (\ref{eq7}), is contained in $B_R$ and is compact for small $\delta>0$. For each $\delta>0$ and $t\in\mathbb{R}$, let $\left(\theta_t(f)\right)_{\delta}$ be the digitization of the $t$-translate of $f$ (we remark parenthetically that $\theta_t(f_{\delta})\neq\left(\theta_t(f)\right)_{\delta}$ in general). We want to prove that $H^{*}\left(A_{\left(\theta_t(f)\right)_{\delta}}, A_{\theta_t(f)}\right)$ converges to zero as $\delta \to 0$, uniformly in $t\in\mathbb{R}$. That is, we want to prove that $A_{\left(\theta_t(f)\right)_{\delta}}$ tends to $A_{\theta_t(f)}$ upper semicontinuously, uniformly in $t\in\mathbb{R}$. Actually we will prove more. Let $P\subset\mathcal{F}$ be the hull of $f$ and let $p_{\delta}$ be the digitization of $p$ for each $p\in P$; then $H^{*}\left(A_{p_{\delta}},A_{p}\right)$ tends to zero as $\delta\to0$, uniformly in $p\in P$. To prove this, it will be convenient to work in an enlarged topological vector space $\mathcal{G}$ which contains $\mathcal{F}$ together with the (in general, temporally discontinuous) vector fields $p_{\delta}$. We define $\mathcal{G}$ to be the class of jointly Lebesgue measurable mappings $g\in\mathbb{R} \times \mathbb{R}^d\to \mathbb{R}^d$ which satisfy the following conditions: \begin{enumerate} \item[a)] For each compact set $K\subset\mathbb{R}^d$, one has $$ \sup_{x \in K} \sup_{t \in \mathbb{R} } \int_t^{t+1} \|g(s,x)\| \, ds < \infty ; $$ \item[b)] For each compact set $K\subset\mathbb{R}^d$ there is a constant $L_K$ (depending on $g$) so that, for almost all $t$ $\in$ $\mathbb{R}$: $$ \|g(t,x)-g(t,y)\| \leq L_K \, \|x-y\|, \quad x,y \in K. $$ \end{enumerate} Now, for each $r=1$, $2$, $3$, $\ldots$ and each $N=1$, $2$, $3$, $\ldots$ introduce a pseudo-metric $d_{r,N}$ on $\mathcal{G}$: $$ d_{r,N}(g_1,g_2) = \sup_{\|x\| \leq r} \int_{-N}^{N} \|g_1(s,x)-g_2(s,x)\| \, ds. $$ Then put $$ d_r(g_1,g_2) = \sum_{N=1}^{\infty} \frac{1}{2^N} \frac{d_{r,N}(g_1,g_2)}{1+d_{r,N}(g_1,g_2)}, $$ and finally set $$ d(g_1,g_2) = \sum_{r=1}^{\infty} \frac{1}{2^r} \frac{d_r(g_1,g_2)}{1+d_r(g_1,g_2)}. $$ We identify two elements of $\mathcal{G}$ if their $d$-distance is zero, thereby obtaining a metric space which we also call $\mathcal{G}$. Observe that, if $g\in\mathcal{G}$, then the Cauchy problem \begin{equation}\label{eq11} x' = g(t,x), \quad x(0) = x_0 \end{equation} admits a unique, maximally-defined local solution $x(t,x_0)$ for each $x_0\in\mathbb{R}^d$; moreover, $x(t,x_0)$ depends continuously on $(t,x_0)$ on its domain of definition. This can be proved using the standard Picard iteration method to solve (\ref{eq11}). We will write $$ x(t,x_0) = \phi(t,x_0,g) $$ to maintain consistency with notation used previously. Observe further that, for each $t\in\mathbb{R}$, the translation $\theta_t:\mathcal{G}\to\mathcal{G}$, i.e., $\theta_t(g)(s,x)=g(t+s,x)$ is well-defined. Let $\mathcal{G}_1\subset\mathcal{G}$ be a translation invariant subset such that the supremum in a) and the constants $L_K$ in b) of the definition of $\mathcal{G}$ are uniform in $g$ $\in$ $\mathcal{G}_1$, for each compact $K\subset\mathbb{R}^d$. Then $(t,g)\to\theta_t(g): \mathbb{R} \times \mathcal{G}_1\to\mathcal{G}_1$ is continuous. Next let $\delta_0>0$. We will show that the set $\mathcal{U}=P \cup \{ p_{\delta}:p \in P, 0 < \delta \leq\delta_0\}\subset \mathcal{G}$ is equi-uniformly continuous in the sense that, to each $\epsilon>0$, there corresponds $\eta>0$ such that, if $|t-s|<\eta$, then $d(\theta_t(p),\theta_s(p))<2\epsilon$ and $d(\theta_t(p_{\delta}),\theta_s(p_{\delta}))<2\epsilon$ for all $p$, $p_{\delta}\in\mathcal{U}$ and for all $t$, $s\in\mathbb{R}$. To do this, fix $\epsilon>0$. Recall that $P\subset\mathcal{F}\subset\mathcal{G}$ is the hull of the uniformly continuous function $f$. Hence if $B\subset\mathbb{R}^d$ is a ball centered at the origin and if $N\geq1$, then we can find $\eta_1>0$ such that, if $|t-s|<\eta_1$, then $$ \int_{-N}^{N} \|\theta_t(p)(v,x)-\theta_s(p)(v,x)\| \, dv < \epsilon/3 $$ for all $x\in B$. Then, taking account of the definition of the distance $d$, we see that it is sufficient to prove that, for some sufficiently large ball $B\subset\mathbb{R}^d$ and some sufficiently large $N$, there exists $\eta_2\in(0,\eta_1]$ such that \begin{equation}\label{eq12} \sup_{x\in B} \int_{-N}^{N} \|\theta_t(p_{\delta})(v,x)-\theta_s(p_{\delta})(v,x)\| \,dv < \epsilon \end{equation} whenever $|t-s|<\eta_2$, $0<\delta\leq \delta_0$. Let us write $d_B(\theta_t(p_{\delta}),\theta_s(p_{\delta}))$ for the quantity on the left hand side of (\ref{eq12}). To prove (\ref{eq12}), we use the properties I)--IV) of a recurrent digitization. Choose $\epsilon_1>0$ so that $\omega(\epsilon_1)<\epsilon/3$, then choose $\delta_1$ so that, if $0<\delta\leq \delta_1$ then $\epsilon_x\leq \epsilon_1/(2N)$ for all $x\in B$. Using property II), we see that, if $0<\delta\leq \delta_1$ and $p\in P$, then \begin{eqnarray*} d_B(\theta_t(p_{\delta}),\theta_s(p_{\delta})) & \leq & d_B(\theta_t(p_{\delta}),\theta_t(p)) + d_B(\theta_t(p),\theta_s(p)) + d_B(\theta_s(p),\theta_s(p_{\delta})) \\ & < & 3 \cdot \epsilon/3 = \epsilon \end{eqnarray*} whenever $|t-s|<\eta$. If $\delta_1\geq\delta_0$, we set $\eta=\eta_1$ and stop. If $\delta_1<\delta_0$ and if $\delta_1<\delta\leq \delta_0$, we first choose $\eta_2\leq \delta/100$, then note that on each interval $[u,u+N]\subset\mathbb{R}$ of length $N$, the difference $p_{\delta}(t,x)-p_{\delta}(s,x)$ is zero except on at most $\left[2N/\delta\right]+1$ subintervals of length $2\eta_2$, where $[\cdot]$ denotes the integer part of a positive number. Using the uniform boundedness of the vector fields $p_{\delta}$ $\in$ $\mathcal{U}$ on $\mathbb{R}\times B$, we can determine $\eta_3\leq \eta_2$ so that, if $|t-s|<\eta_3$, then $d_B(\theta_t(p_{\delta}),\theta_s(p_{\delta}))<\epsilon$. So if $\eta=\min \{\eta_1,\eta_3\}$ we obtain (\ref{eq12}) for all $p$, $p_{\delta}\in\mathcal{U}$. Now let $\delta\in(0,\delta_0]$. For each $p\in P$, let $P_{\delta}(p)=\rm{cls} \{\theta_t(p_{\delta}): t \in \mathbb{R} \}$. Then $P_{\delta}(p)$ is compact (this uses the recurrence condition V) of a recurrent digitization) and translation invariant in $\mathcal{G}$. Moreover, using property III) of a digitization and a Gronwall-type argument, one shows that the map $(t,x_0,g)$ $\mapsto$ $(\phi(t,x_0,g), \theta_t(g))$ defines a (continuous) skew-product flow on $\mathbb{R}^d \times P_{\delta}(p)$. Choose $\delta_0$ so that each $p_{\delta}$ satisfies (\ref{eq6}) for all $p\in P$ and $0<\delta\leq \delta_0$. Then the pullback attractor $A_{p_{\delta}}$ exists and equals $\{x_0 \in \mathbb{R}^d$ $:$ $\phi(t,x_0,p_{\delta})$ is defined on all of $\mathbb{R}$ and satisfies $\|\phi(t,x_0,p_{\delta})\|\leq R$ $\}$; see Proposition \ref{prop2}. In fact, $A_{p_{\delta}}$ is then the $p_{\delta}$-fiber of a global pullback attractor ${\bf A}^{\delta}\subset\mathbb{R}^d\times P_{\delta}(p)$. Now $p_{\delta} \to p$ in $\mathcal{G}$ as $\delta\to0$, so using property III) of a digitization and a Gronwall argument, together with the characterization of $A_{p_{\delta}}$ in terms of bounded solutions of $x'=p_{\delta}(t,x)$, one shows that $H^{*}(A_{p_{\delta}},A_p)\to0$ as $\delta\to0$. However, more is true. Using property IV) of a digitization one has the following: if $p_n\top\in P$ and if $\delta_n\to 0$, then $d(p_{(n,\delta_n)},p)\to0$. Again using II) together with a Gronwall argument and the above characterization of $A_{p_{(n,\delta_n)}}$, one sees that $H^{*}(A_{(p_{n,\delta_n)}},A_p)\to0$ as $n\to \infty$. This is a strong uniformity statement, and implies \begin{proposition}\label{prop3} $H^{*}(A_{p_{\delta}},A_p)\to0$ as $\delta\to0$, uniformly in $p\in P$. In particular, $$ H^{*}(A_{(\theta_t(f)_{\delta}},A_{\theta_t(f)}) \to 0 \quad \mbox{as } \delta \to 0, $$ uniformly in $t\in\mathbb{R}$. \end{proposition} We use the recurrence condition V) to prove that the sets $P_{\delta}(p)$ are compact. This would seem to be a basic requirement to be satisfied when one sets about computing the pullback attractor, because otherwise the convergence in (\ref{eq7}) of the intersection to $A_{p_{\delta}}$ cannot be expected to have any uniformity properties. We note, however, that Proposition \ref{prop3} could be proved without asumptions ensuring that the sets $P_{\delta}(p)$ are compact; one only needs the uniform continuity in $t$ on $\mathbb{R}$ (uniform on compact $x$ subsets of $\mathbb{R}^d$) of the vector field $f(t,x)$ and the continuity of the Bebutov flow on $(\mathcal{G},d)$. \section{Continuity results} In this section, we continue to investigate the perturbation properties of pullback attractors, this time with the goal of giving a sufficient condition for the Hausdorff continuity (and not just upper semicontinuity) of the sets $A_p$, resp. ${\bf A}$, as the base space $P$ is varied in some functional space. To be specific, let $\mathcal{F}$ be the topological space introduced in Section \ref{prel}. Let $P$ be a compact, translation-invariant subset of $\mathcal{F}$ (which need not be the hull of any one element $f\in\mathcal{F}$). Let us assume that each $p\in P$ can be written in the form $$ p(t,x) = L_p(t) x + h_p(t,x), $$ where $L_p(\cdot)$ is a uniformly continuous function with values in the set $\mathcal{M}_d$ of real $d\times d$ matrices. We assume that the mappings $(p,t,x)\mapsto L_p(t) x$ and $(p,t,x)\mapsto h_p(t,x)$ are uniformly continuous on compact subsets of $P\times \mathbb{R} \times \mathbb{R}^d$. A sufficient condition that this is the case is the following. Consider the metric space $P$; put $F(p,x)=p(0,x)$ for each $p\in P$ and $x\in\mathbb{R}^d$. Note that $F(\theta_t(p),x)=p(t,x)$ for all $t\in\mathbb{R}$, $p\in P$. Suppose that the Jacobian $\frac{\partial F}{\partial x}(p,0)$ is continuous as a function of $p$. Then, setting $$ L_p(t) x = \frac{\partial F}{\partial x}(\theta_t(p),0) x, \quad h_p(t,x) = p(t,x) - L_p(t) x, $$ we obtain the desired decomposition. We now impose the following hypothesys. \begin{enumerate} \item[(H1)] The family of linear systems $$ x' = L_p(t) x, \quad p \in P, $$ admits an exponential dichotomy over $P$ with constants $L>0$, $\gamma>0$ and continuous family of projections $\{Q_p:p \in P\}$. \end{enumerate} Let $C_b=C_b(\mathbb{R},\mathbb{R}^d)$ be the Banach space of bounded continuous functions $x:\mathbb{R}\to \mathbb{R}^d$ with the norm $\|x\|_{\infty}=\sup_{t \in \mathbb{R}}\|x(t)\|$. For each $p\in P$, define a nonlinear operator $T_p:C_b\to C_b$ as follows: \begin{eqnarray*} T_p[x](t) &=& \int_t^{\infty} \Psi_p(t) Q_p \Psi_p(s)^{-1} h_p(s,x(s)) \, ds \\ &&+ \int_{-\infty}^t \Psi_p(t) (Q_p-I) \Psi_p(s)^{-1} h_p(s,x(s)), \, ds \end{eqnarray*} where $\Psi_p(t)$ is the fundamental matrix with initial value $\Psi_p(0)=I$ (identity matrix) of the linear equation $x'=L_p(t) x$. Assume from now on that condition (\ref{eq6}) holds for all $p\in P$. We further impose a condition of uniform contractivity. \begin{enumerate} \item[(H2)] For all $p\in P$ and for all $x$, $y$ in $B_R$, one has $$ \left\|h_p(t,x) -h_p(t,x) \right\| \leq k \|x-y\|, $$ where $k<{\gamma}/{2L}$. \end{enumerate} To simplify the analysis, we now modify each $h_p(t,x)$ outside the ball $B_R$ so that $h_p$ satisfies Hypothesis (H2) for all $x \in \mathbb{R}^d$ and so that $h_p(t,x)=0$ whenever $\|x\|\geq R+1$. This means that condition (\ref{eq6}) does not hold if $\|x\|$ $\geq$ $R+1$, but it will clear that this will have no effect on our analysis of the pullback attractors of the equations (\ref{eq13}). \begin{proposition} \label{prop4} For each $p\in P$, the equation \begin{equation}\label{eq13} x' = p(t,x), \end{equation} admits a unique solution $x_p(t)$ which is bounded on all of $\mathbb{R}$. \end{proposition} \paragraph{Proof.} The argument is standard (see, e.g., Fink \cite{Fink}). The operator $T_p$ is a contraction on $C_b$ and hence admits a unique fixed point $x_p(\cdot)$, which is a bounded solution of (\ref{eq13}). Since each fixed point of $T_p$ in $C_b$ is a bounded solution of (\ref{eq13}) the proposition is proved. \hfil$\Box$ \smallskip Using Propositions \ref{prop2} and \ref{prop4}, we see that, for each $p\in P$, the pullback attractor $A_p$ $=$ $\{x_p(0)\}$, i.e., each $A_p$ contains exactly one point. Then from continuity with respect to parameters of the fixed point of a contractive mapping, we see that ${\bf A}=\{(x_p(0),p): p\in P\}\subset\mathbb{R}^d \times P$ is compact. One verifies that ${\bf A}\subset\mathbb{R}^d \times P$ is the global pullback attractor for the family (\ref{eq13}). Now, if $\{x_0\}$ is a singleton subset of a metric space $\mathcal{X}$, and if $B\subset\mathcal{X}$ is compact, then $H^{*}(B,\{x_0\})$ coincides with the Hausdorff distance $H(B,\{x_0\})$. This fact will allow us to prove that, if $p\in P$, then $A_p$ is a point of Hausdorff continuity for the pullback attractors $A_{\tilde{p}}$ of equations $x'=\tilde{p}(t,x)$ obtained by appropriate perturbations $\tilde{p}$ of $p$. We will formulate a fairly general continuity result whose hypotheses are satisfied in particular by the digitizations of Section \ref{USC}. We view the compact metric space $P$ as a subset of $\mathcal{G}$. Choose fixed values for the suprema in a) and for the constants $L_K$ in b) of the definition of $\mathcal{G}$, and let $\mathcal{G}_1\subset\mathcal{G}$ be the set of all $g\in\mathcal{G}$ for which a) and b) hold with these fixed values. As in Section \ref{USC}, write $\theta_t(g)(s,x)= g(t+s,x)$, and let $\phi(t,x_0,g)$ denote the solution of the Cauchy problem (\ref{eq11}) for each $g$ $\in$ $\mathcal{G}_1$. Using a Gronwall-type inequality, one verifies that the mapping $(t,x_0,g)$ $\mapsto$ $(\phi(t,x_0,g), \theta_t(g))$ is continuous on its domain of definition $V\subset\mathbb{R} \times \mathbb{R}^d \times \mathcal{G}_1$, and defines a continuous local skew-product flow on $\mathbb{R}^d \times\mathcal{G}_1$. Next let $\tilde{P}\in\mathcal{G}_1$ be a \textbf{compact} translation-invariant set. The Hausdorff semi-metric $H^{*}(\tilde{P},P)$ is defined relative to the metric $d$ in $\mathcal{G}_1$. Let $\eta_{*}>0$ be a constant so that \begin{equation}\label{eq14} \left \leq - \eta_{*} \end{equation} for all $p\in P$, $t\in\mathbb{R}$, and $x\in\mathbb{R}^d$ with $\|x\|=R$. One can show that there exists $\delta>0$ so that, if $H^{*}(\tilde{P},P)<\delta$, and if $\tilde{p}\in\tilde{P}$, then for each $x_0\in\mathbb{R}^d$ with $\|x_0\|=R$, the solution $x(t)$ of the Cauchy problem $$ x' = \tilde{p}(t,x), \quad x(0) = x_0, $$ satisfies $\|x(t)\|0$. The proof uses the continuity of the local skew-product flow on $\mathbb{R}^d \times\mathcal{G}_1$. This means that the family $\{\mbox{(\ref{eq13})}: \ p \in P\}$ admits a global pullback attractor $\tilde{{\bf A}}$, which lies in $B_R\times \tilde{P}$. Since $A_p$ is a singleton set, we have $H^{*}(A_p,A_{p_{\delta}})\leq H^{*}(A_{p_{\delta}},A_p)$, whether the sets $A_{p_{\delta}}$ are singleton sets or not. Thus in Proposition \ref{prop3} we actually have continuous convergence in this case. \begin{proposition}\label{prop5} For each each $\delta>0$, let $\{p_{\delta}: p \in P\}$ be a subset of $\mathcal{G}$ such that $d(p_{\delta},p)\to 0$ as $\delta\to 0$, uniformly with respect to $p\in P$. Then $H\left(A_{p_{\delta}},A_p\right)\to0$ as $\delta \to 0$, uniformly in $p\in P$. \end{proposition} Note that the equations $x'=p_{\delta}(t,x)$ need not give rise to contractions in $C_b$, so this result cannot be proved using the continuity properties of fixed points of contraction mappings. \section{Example} We give an example to illustrate the strength of Proposition \ref{prop5}. We will work with \textbf{quasi-periodic} vector fields. Let $\mathbb{T}^k$ be the $k$-torus, $k\geq 2$, and let $\gamma=(\gamma_1,\ldots,\gamma_k)$ be a rationally independent vector in $\mathbb{R}^k$. Let $(\phi_1$, $\ldots$, $\phi_k)$ be angular coordinates $\mod 2\pi$ on $\mathbb{T}^k$. Introduce the corresponding Kronecker flow $\{\theta_t: t \in\mathbb{R} \}$ on $\mathbb{T}^k$ by setting $\theta_t(\phi_1, \ldots, \phi_k) =(\phi_1+\gamma_1 t , \ldots, \phi_k +\gamma_k t)$. For brevity we set $\phi=(\phi_1, \ldots, \phi_k)$ and $\theta_t(\phi)=\phi+\gamma t$. Let $F:\mathbb{T}^k\times \mathbb{R}^d \to \mathbb{R}^d$ be a continuous function such that the Jacobian $\frac{\partial F}{\partial x}(\phi,0)$ is defined and is a continuous function of $\phi\in\mathbb{T}^k$. Let $L_{\phi}(t)=\frac{\partial F}{\partial x}(\theta_t(\phi),0)$ and $h_{\phi}(t,x)=F(\theta_t(\phi),x)-L_{\phi}(t)x$. Suppose that the family of linear equations $$ x' = L_{\phi}(t) x, \quad \phi \in \mathbb{T}^k, $$ admits an exponential dichotomy over $\mathbb{T}^k$ with constants $L>0$, $\gamma > 0$ and a continuous family of projections $\{Q_{\phi}: \phi \in \mathbb{T}^k \}$. Suppose further that there exists $R>0$ so that $\left<0$ for all $\phi\in\mathbb{T}^k$ and $x \in \mathbb{R}^d$ with $\|x\|\geq R$. Suppose finally that there is a constant $k<\gamma/(2L)$ such that $$ \|h_{\phi}(t,x) - h_{\phi}(t,y)\| \leq k \|x-y\| $$ for all $\phi\in\mathbb{T}^k$ and $x$, $y$ in $\mathbb{R}^d$ with $\|x-y\|\leq R$. Let $G_n=G_n(\phi,x)$ be any sequence of continuous functions on $\mathbb{T}^k\times \mathbb{R}^d $ with values in $\mathbb{R}^d$ such that \begin{enumerate} \item[1)] $G_n\to0$ uniformly on $\mathbb{T}^k\times B_R$ as $n\to\infty$; \item[2)] There is a real number $M$ such that $$ \|G_n(\phi,x) - G_n(\phi,y)\| \leq M \|x-y\| $$ for all $\phi\in\mathbb{T}^k$ and $x$, $y\in B_R$ and $n$ $\geq$ $1$. \end{enumerate} We do not assume that $\frac{\partial G_n}{\partial x}(\phi,0)$ exists, nor that $M$ is small. Hence, if we consider the vector functions $$ F_n(\phi,x) = F(\phi,x) + G_n(\phi,x), $$ it may not be the case that the family of equations \begin{equation}\label{eq15} x' = F_n(\theta_t(\phi),x), \quad \phi \in \mathbb{T}^k, \end{equation} generates a family of fixed-point mappings $\{T_{\phi}\}$, and even if it does, there is no guarantee that any $T_{\phi}$ is a contraction on $C_b$. The $F_n$ are of course perturbations of $F$. Let us now introduce a further perturbative element. Namely, let $\gamma^{(n)}=(\gamma^{(n)}_1,\ldots,\gamma^{(n)}_k)$ be a sequence of frequency vectors such that $\gamma^{(n)} \to \gamma$ in $\mathbb{R}^k$. Of course, it is not assumed that the $\gamma^{(n)}$ are rationally independent. Write $\theta^{(n)}_t(\phi_1, \ldots, \phi_k)= (\phi_1+\gamma^{(n)}_1 t , \ldots, \phi_k +\gamma^{(n)}_k t)$ for the corresponding Kronecker flow with $n=1,2,3,\ldots$. Now let \begin{gather*} P = \left\{ p \in \mathcal{F}: \ p(t,x) = F(\theta_t(\phi),x) \mbox{ for some } \phi \in \mathbb{T}^k \right\}, \\ P^{(n)} = \left\{p \in \mathcal{F}: \ p(t,x) = F_n(\theta^{(n)}_t(\phi),x) \mbox{ for some } \phi \in \mathbb{T}^k \right\}, \quad n \geq 1. \end{gather*} These are all compact translation-invariant subsets of $\mathcal{F}$. One can verify that $H(P^{(n)},P)\to 0$ as $n\to\infty$; this is true even though the frequency vector has been perturbed. Since condition (\ref{eq6}) is satisfied by the family (\ref{eq15}) for all sufficiently large $n$, there are pullback attractors ${\bf A}^{(n)}\subset \mathbb{R}^d\times P^{(n)}$ for each such $n$ defined by the respective families of equations (\ref{eq15}). Using the arguments preceding Proposition \ref{prop5}, one has that $H({\bf A}^{(n)},{\bf A})\to0$ as $n \to \infty$, where ${\bf A}$ is the pullback attractor defined by the equations $$ x' = F(\theta_t(\phi),x), \quad \phi \in \mathbb{T}^k. $$ Let us now use this example to illustrate how information can be obtained concerning convergence of pullback attractors under digitization. Let the letters $F$, $F_n$, $G_n$, $\gamma^{(n)}$ have the significance attributed to them above. Suppose we are given a digitization scheme satisfying the conditions I)--IV) of Section \ref{USC}. Let $P=\{ p \in \mathcal{G}:p(t,x)=F(\theta_t(\phi),x)$ for some $\phi\in\mathbb{T}^k\}$; thus $P$ is the same as before except that now it is viewed as a subset of $\mathcal{G}$. Let $\{\delta_n\}$ be a sequence of positive numbers which converges to zero. For each $\phi \in \mathbb{T}^k$, let $p_{\phi}^{(n)}$ be the $\delta_n$-digitization of the time-varying vector field $(t,x)\to F_n(\phi+\gamma^{(n)} t,x)$. Now, for large $n$, each $p_{\phi}^{(n)}$ ($\phi$ in $\mathbb{T}^k$) satisfies condition (\ref{eq6}), so the pullback attractor $A^{(n)}_{\phi}$ defined by the equation $x'= p_{\phi}^{(n)}(t,x)$ is contained in $B_R$. Let ${\bf A}\subset\mathbb{R}^d\times P$ be the global pullback attractor defined by the family of equations (\ref{eq2}). Each $p\in P$ corresponds to (at least) one point $\phi\in\mathbb{T}^k$. Let us write $p=p_{\phi}$ if $p$ corresponds to $\phi$, then write $A_{\phi}$ instead of $A_p$ for the fiber of ${\bf A}$ at $p$. Each fiber $A_{\phi}$ is a singleton subset of $\mathbb{R}^d$. Finally, arguing as in Section \ref{USC}, we can show that $H(A^{(n)}_{\phi},A_{\phi})\to0$ as $n \to \infty$. In fact, the convergence here is uniform with respect to $\phi \in\mathbb{T}^k$. \begin{thebibliography}{99} \frenchspacing \bibitem{CKS} D. N. Cheban, P. E. Kloeden and B. Schmalfu\ss, The relationship between pullback, forwards and global attractors of nonautonomous dynamical systems. \emph{Nonlinear Dynam. \& Systems Theory } (to appear). \bibitem{CF} H. Crauel and F. Flandoli, Attractors for random dynamical systems, \emph{Probab. Theory Relat. Fields.} \textbf{100} (1994), 365--393. \bibitem{Fink} A. M. Fink, \emph{Almost Periodic Differential Equations,} Springer Lecture Notes in Math. Vol. 377, Springer-Verlag, Berlin, 1974. \bibitem{Gorn} L. Gorniewicz, Topological approach to differential inclusions, in \emph{Topological Methods in Differential Equations and Inclusions,} Kluwer Academic Publishers, Series C: Math.\& Physical Sciences, Vol. 412 (1995), pp. 129--190. \bibitem{J1} R. A. Johnson, An application of topological dynamics to bifurcation theory, in \emph{Contemporary Mathematics,} Vol. 215, Amer. Math. Soc., Providence (1998), pp. 323--334. \bibitem{J2} R. A. Johnson, Some questions in random dynamical sytsems involving real noise processes, in \emph{Stochastic Dynamics,} Festschrift zur Ehre L. Arnolds, Springer--Verlag, Berlin (1999), pp. 147--180. \bibitem{JM1} R. A. Johnson and F. Mantellini, Nonautonomous differential equations, {\em CIME Lecture Notes, \/} to appear \bibitem{JM2} R. A. Johnson and F. Mantellini, A non-autonomous transcritical bifurcation problem with an application to quasi-periodic bubbles, {\em Preprint, Universit\`{a} di Firenze.\/} \bibitem{JY} R. A. Johnson and Y.F. Yi, Hopf bifurcation from non-periodic solutions of differential equations, {\em J. Differ. Eqns. \/} {\bf 107} (1994), 310--340. \bibitem{K1} P. E. Kloeden, Lyapunov functions for cocycle attractors in nonautonomous differential equations, {\em Elect. J. Diff. Eqns. \/} Conference {\bf 05} (2000), 91--102. \bibitem{KK2000} P. E. Kloeden and V.S. Kozyakin, The inflation of nonautonomous systems and their pullback attractors. \emph{Transactions of the Russian Academy of Natural Sciences, Series MMMIU.} \textbf{4}, No. 1-2 (2000), 144--169. \bibitem{KK2001} P. E. Kloeden and V. S. Kozyakin, Uniform nonautonomous attractors under discretization \emph{DANSE}-Preprint 36/00, FU Berlin, 2000. \bibitem{KK2002} P. E. Kloeden and V.S. Kozyakin, Nonautonomous attractors of skew-product flows with a shadowing driving system, \emph{Discrete \& Conts. Dynamical Systems, Series A} (2002), to appear. \bibitem{KS1} P. E. Kloeden and B. Schmalfu{\ss}, Lyapunov functions and attractors under variable time-step discretization, \emph{Discrete \& Conts. Dynamical Systems.} \textbf{2} (1996), 163--172. \bibitem{SS} R. J. Sacker and G.R Sell, Lifting properties in skew-product flows with application to differential equations, \emph{Memoirs of the American Mathematical Society.} \textbf{190} (1977). \bibitem{S} G. R. Sell, {\em Lectures on Topological Dynamics and Differential Equations.\/} Van Nostrand--Reinbold, London, 1971. \bibitem{SY} W.-X. Shen and Y.-F. Yi, {\em Almost Automorphic and Almost Periodic Dynamics in Skew-Product Flows \/} Memoirs Amer. Math. Soc., Vol. 647 (1998). \end{thebibliography} \noindent\textsc{R. A. Johnson}\\ Dipartimento di Sistemi e Informatica \\ Universit\`{a} di Firenze \\ I-50139 Firenze, Italy \\ e-mail: \texttt{johnson@ingfi1.ing.unifi.it} \medskip \noindent\textsc{P. E. Kloeden}\\ Fachbereich Mathematik\\ Johann Wolfgang Goethe Universit\"at \\ D-60054 Frankfurt am Main, Germany \\ e-mail: \texttt{kloeden@math.uni-frankfurt.de} \end{document}