1: \documentclass[]{preprint}
2: \usepackage{times}
3: \usepackage{mhequ}
4: \usepackage{mhfig}
5: \usepackage{mhenvs}
6: \usepackage{Symbols}
7: \usepackage{ScriptFonts}
8:
9: \def\prob{{\bf P}}
10: \def\expect{{\bf E}}
11: \def\from{:}
12:
13: {
14: \theorembodyfont{\rmfamily}
15: \newtheorem{notation}[lemma]{Notation}
16: }
17:
18: \ifx\pdfoutput\undefined
19: \else
20: \pdfinfo{
21: /Title (fBm.pdf)
22: /Author (M. Hairer)
23: /Subject (Fractional Brownian motion)
24: /Keywords (Ergodicity, fractional Brownian motion, memory)
25: /Creator (TeXShop)
26: }
27: \input protcode.tex
28: \fi
29:
30: \makeatletter
31: \def\Kern{\@ifnextchar({\Kern@par}{\Kern@nopar}}
32: \def\Kern@par(#1){(#1)^{H-{1\over 2}}}
33: \def\Kern@nopar#1{{#1}^{H-{1\over 2}}}
34: \def\maketag#1#2{(#1)\def\@currentlabel{#1}\label{#2}}
35: \makeatother
36:
37: \def\Noise{\CW}
38: \def\State{\CX}
39: \def\pNoise{{\prob_{\! w}}}
40: \def\Evol{{\cscr Q}}
41: \def\Coupl{{\cscr C}}
42: \def\BigSpace{{\cscr X}}
43: \def\Dens{{\cscr D}}
44: \def\C#1{{C^{\mathbf{A1}}_{#1}}}
45: \def\nb{\penalty10000}
46:
47: \def\iCond{{\cscr M}_1^{w}}
48: \def\pMeas{{\cscr M}_1}
49: \def\Meas{{\cscr M}_\phi}
50: \def\cadlag{c\`adl\`ag}
51: \def\Wien{{\mathsf W}}
52: \def\TV{{\mathrm TV}}
53: \def\step#1{{\textbf #1}}
54:
55: \begin{document}
56: \ifx\pdfoutput\undefined
57: \else
58: \setprotcode\font
59: {\it \setprotcode\font}
60: {\bf \setprotcode\font}
61: {\bf \it \setprotcode\font}
62: \pdfprotrudechars=1
63: \fi
64:
65: \title{Ergodicity of Stochastic Differential Equations Driven by
66: Fractional Brownian Motion}
67: % Short title: "Ergodicity of SDEs Driven by fBm"
68:
69: \author{Martin Hairer}
70: \institute{Mathematics Research Centre, University of Warwick\\
71: \email{hairer@maths.warwick.ac.uk}\\
72: Homepage: \texttt{http://www.maths.warwick.ac.uk/\char`~hairer/}}
73:
74:
75: \maketitle
76:
77: \begin{abstract}
78: We study the ergodic properties of finite-dimensional systems of SDEs driven by
79: non-degenerate additive fractional Brownian motion with
80: arbitrary Hurst parameter $H\in(0,1)$. A general framework is constructed to make
81: precise the notions of ``invariant measure'' and ``stationary state'' for such a system.
82: We then prove under rather weak dissipativity conditions that such
83: an SDE possesses a unique stationary solution and that the convergence
84: rate of an arbitrary solution towards the stationary one is (at least) algebraic. A lower bound on the
85: exponent is also given.\\
86: \\
87: Keywords: Ergodicity, fractional Brownian motion, memory\\
88: Subject Classification: 60H10, 26A33
89: \end{abstract}
90:
91: \tableofcontents
92:
93: \section{Introduction and main result}
94:
95: In this paper, we investigate the long-time behaviour of stochastic differential equations driven by fractional
96: Brownian motion. Fractional Brownian motion (or fBm for short) is a centred Gaussian process satisfying $B_H(0) = 0$ and
97: \begin{equ}[e:deffBm]
98: \expect|B_H(t) - B_H(s)|^2 = |t-s|^{2H}\;,\qquad t,s>0\;,
99: \end{equ}
100: where $H$, the Hurst parameter, is a real number in the range $H\in(0,1)$. When $H={1\over 2}$, one recovers
101: of course the usual Brownian motion, so this is a natural one-parameter family of generalisations
102: of the ``standard'' Brownian motion. It follows from \eref{e:deffBm} that fBm is also self-similar, but with the scaling law
103: \begin{equ}
104: t \mapsto B_H(at) \qquad\approx\qquad t \mapsto a^H B_H(t)\;,
105: \end{equ}
106: where $\approx$ denotes equivalence in law. Also, the sample paths of $B_H$ are $\alpha$-H\"older continuous for every
107: $\alpha < H$. The main difference between fBm and the usual Brownian motion is that it is neither
108: Markovian, nor a semi-martingale, so most standard tools from stochastic calculus
109: cannot be applied to its analysis.
110:
111: Our main motivation is to tackle the problem of ergodicity in non-Markovian systems.
112: Such systems arise naturally in several situations. In physics, stochastic forces are used to describe the
113: interaction between a (small) system and its (large) environment. There is no a-priori reason to assume that
114: the forces applied by the environment to the system are independent over disjoint time intervals.
115: In statistical mechanics, for example, a non-Markovian
116: noise term appears when one attempts to derive
117: the Langevin equation from first principles \cite{MR98h:82049,MR90a:82002}.
118: Self-similar stochastic processes like fractional Brownian motion
119: appear naturally in hydrodynamics \cite{MR39:3572}.
120: It appears that fractional Brownian motion is also useful to model long-time correlations in stock markets \cite{MR2001g:60129,fbmfinance}.
121:
122: Little seems to be known about the long-time behaviour of non-Markovian systems.
123: In the case of the non-Markovian Langevin equation (which is \textit{not} covered
124: by the results in this paper
125: due to the presence of a delay term), the stationary solution is explicitly known to be distributed according to
126: the usual equilibrium Gibbs measure. The relaxation towards equilibrium is a very hard problem that was
127: solved in \cite{MR98h:82049,MR2001c:82006}. It is however still open in the non-equilibrium case, where the
128: invariant state can not be guessed a-priori.
129: One well-studied general framework for the study of systems driven by noise with extrinsic
130: memory like the ones considered in this paper is given by the theory of Random Dynamical Systems (see the
131: monograph \cite{RDS} and the reference list therein). In that framework, the existence of random attractors,
132: and therefore the \textit{existence} of invariant measures
133: seems to be well-understood. On the other hand, the problem of \textit{uniqueness} (in an appropriate sense, see
134: the comment following \theo{theo:largeH} below) of the invariant measure on the
135: random attractor seems to be much harder, unless one can show that the system possesses a unique
136: stochastic fixed point. The latter situation was studied in
137: \cite{SchmalMasl} for infinite-dimensional evolution equations driven by fractional Brownian motion.
138:
139: The reasons for choosing fBm as driving process for \eref{e:mainequ} below are twofold. First, in particular
140: when $H > {1\over 2}$, fractional Brownian motion presents genuine long-time correlations that
141: persist even under rescaling. The second reason is that there exist simple, explicit formulae that relate
142: fractional Brownian motion to ``standard'' Brownian motion, which simplifies our analysis.
143: We will limit ourselves to the case where the memory of the system
144: comes entirely from the driving noise process, so we do not consider stochastic delay equations.
145:
146: We will only consider equations driven by non-degenerate additive noise, \ie\ we consider equations of the form
147: \begin{equ}[e:mainequ]\tag{SDE}
148: dx_t = f(x_t)\,dt + \sigma\,dB_H(t)\;,\qquad x_0 \in \R^n\;,
149: \end{equ}
150: where $x_t \in \R^n$, $f\from \R^n \to \R^n$, $B_H$ is an $n$-dimensional fractional Brownian motion with Hurst parameter $H$,
151: and $\sigma$ is a constant and invertible $n\times n$ matrix. Of course, \eref{e:mainequ} should be interpreted as an integral
152: equation.
153:
154: In order to ensure the existence of globally bounded solutions and in order to have some control on the speed at which
155: trajectories separate, we make throughout the paper the following assumptions on the components of \eref{e:mainequ}:
156: \begin{claim}[A2']
157: \item[\textbf{A1}] \textit{Stability.} There exist constants $\C{i} > 0$ such that
158: \begin{equ}
159: \scal{f(x)-f(y),x-y} \le \min\{\C1 - \C2\|x-y\|^2, \C3\|x-y\|^2\}\;,
160: \end{equ}
161: for every $x,y \in \R^n$.
162: \item[\textbf{A2}] \textit{Growth and regularity.} There exist constants $C,N>0$ such that $f$ and its derivative satisfy
163: \begin{equ}
164: \|f(x)\| \le C\sml(1+\|x\|\smr)^N\;, \qquad \|Df(x)\| \le C\sml(1+\|x\|\smr)^N\;,
165: \end{equ}
166: for every $x \in \R^n$.
167: \item[\textbf{A3}] \textit{Non-degeneracy.} The $n\times n$ matrix $\sigma$ is invertible.
168: \end{claim}
169:
170: \begin{remark}\label{rem:sigma}
171: We can assume that $\|\sigma\| \le 1$ without any loss of generality. This assumption
172: will be made throughout the paper in order to simplify some expressions.
173: \end{remark}
174:
175: One typical example that we have in mind is given by
176: \begin{equ}
177: f(x) = x - x^3\:,\qquad x\in \R\;,
178: \end{equ}
179: or any polynomial of odd degree with negative leading coefficient.
180: Notice that $f$ satisfies \textbf{A1}--\textbf{A2}, but that it is not globally Lipschitz continuous.
181:
182: When the Hurst parameter $H$ of the fBm driving \eref{e:mainequ} is bigger than $1/2$,
183: more regularity for $f$ is required, and we will then sometimes assume that the following
184: stronger condition holds instead of \textbf{A2}:
185:
186: \begin{claim}[A2']
187: \item[\textbf{A2'}] \textit{Strong regularity.} The derivative of $f$ is globally bounded.
188: \end{claim}
189:
190: Our main result is that \eref{e:mainequ} possesses a \textit{unique} stationary solution.
191: Furthermore, we obtain an explicit bound showing that every (adapted) solution to
192: \eref{e:mainequ} converges towards this stationary solution, and that this convergence
193: is at least algebraic. We make no claim concerning the optimality of this bound for the class
194: of systems under consideration.
195: Our results are slightly different for small and for large values of $H$, so we state them separately.
196:
197: \begin{theorem}[Small Hurst parameter]\label{theo:smallH}
198: Let $H\in \sml(0,{\textstyle{1\over 2}}\smr)$ and let $f$ and $\sigma$ satisfy \textbf{A1}--\textbf{A3}.
199: Then, for every initial condition, the solution to \eref{e:mainequ} converges towards a unique stationary solution
200: in the total variation norm. Furthermore, for every $\gamma < \max_{\alpha < H} \alpha(1-2\alpha)$, the difference between the solution and the stationary solution
201: is bounded by $C_\gamma t^{-\gamma}$ for large $t$.
202: \end{theorem}
203:
204: \begin{theorem}[Large Hurst parameter]\label{theo:largeH}
205: Let $H\in \sml({\textstyle{1\over 2}},1\smr)$ and let $f$ and $\sigma$ satisfy \textbf{A1}--\textbf{A3} and \textbf{A2'}.
206: Then, for every initial condition, the solution to \eref{e:mainequ} converges towards a unique stationary solution
207: in the total variation norm. Furthermore, for every $\gamma < {1\over 8}$, the difference between the solution and the stationary solution
208: is bounded by $C_\gamma t^{-\gamma}$ for large $t$.
209: \end{theorem}
210:
211: \begin{remark}
212: The ``uniqueness'' part of these statements should be understood as uni\-que\-ness in law in
213: the class of stationary solutions
214: adapted to the natural filtration induced by the two-sided fBm that drives the equation. There could in theory be other stationary solutions,
215: but they would require knowledge of the future to determine the present, so they are usually discarded as unphysical.
216:
217: Even in the context of Markov processes, similar situations do occur. One can well have uniqueness of the invariant measure,
218: but non-uniqueness of the stationary state, although other stationary states would have to foresee the future. In this sense,
219: the notion of uniqueness appearing in the above statements is similar to the notion of uniqueness of the invariant measure for Markov processes. (See
220: \eg \cite{RDS}, \cite{MR93c:60096} and \cite{MR1888416}
221: for discussions on invariant measures that are not necessarily measurable with respect to the past.)
222: \end{remark}
223:
224: \begin{remark}
225: The case $H={\textstyle{1\over 2}}$ is not covered by these two theorems, but it is well-known that the convergence
226: toward the stationary state is exponential in this case (see for example \cite{MT}).
227: In both cases, the word ``total variation'' refers to the total variation distance between measures on the space
228: of paths, see also \theo{theo:mainTheoFormal} below for a rigorous formulation of the results above.
229: \end{remark}
230:
231: \subsection{Idea of proof and structure of the paper}
232:
233: Our first task is to make precise the notions of ``initial condition'', ``invariant measure'', ``uniqueness'', and ``convergence''
234: appearing in the formulation of Theorems~\ref{theo:smallH} and \ref{theo:largeH}. This will be
235: achieved in Section~\ref{sec:general} below, where we construct a general framework for the study
236: of systems driven by non-Markovian noise. Section~\ref{sec:setting} shows how \eref{e:mainequ}
237: fits into that framework.
238:
239: The main tool used in the proof of Theorems~\ref{theo:smallH} and \ref{theo:largeH} is a coupling
240: construction similar in spirit to the ones presented in \cite{MatNS,HExp02}. More precisely, we first
241: show by some compactness argument that there exists at least one invariant measure $\mu_*$ for
242: \eref{e:mainequ}. Then, given an initial condition distributed according to some arbitrary measure $\mu$,
243: we construct a ``coupling process''
244: $(x_t,y_t)$ on $\R^n\times \R^n$ with the following properties:
245: \begin{claim}[3.]
246: \item[1.] The process $x_t$ is a solution to \eref{e:mainequ} with initial condition $\mu_*$.
247: \item[2.] The process $y_t$ is a solution to \eref{e:mainequ} with initial condition $\mu$.
248: \item[3.] The random time $\tau_\infty = \min\{t\,|\, x_s = y_s\quad \forall s \ge t\}$ is almost surely finite.
249: \end{claim}
250: The challenge is to introduce correlations between $x_s$ and $y_s$ in precisely such a way that
251: $\tau_\infty$ is finite. If this is possible, the uniqueness of the invariant measure follows immediately.
252: Bounds on the moments of $\tau_\infty$ furthermore translate into bounds on the rate of convergence towards this
253: invariant measure.
254: In Section~\ref{sec:coupl}, we expose the general mechanism by which
255: we construct this coupling. Section~\ref{sec:defCoupl} is then devoted to the precise formulation of the
256: coupling process and to the study of its properties, which will be used in Section~\ref{sec:mainproof}
257: to prove Theorems~\ref{theo:smallH} and \ref{theo:largeH}. We conclude this paper with a few
258: remarks on possible extensions of our results to situations that are not covered here.
259:
260: \begin{acknowledge}
261: The author wishes to thank Dirk Bl\"omker, David Elworthy, Xue-Mei Li, Neil O'Connell, and
262: Roger Tribe for their interest in this work and for many helpful suggestions, stimulating questions, and
263: pertinent remarks. He would also like to thank the referee for his careful reading of the manuscript.
264:
265: The author also wishes to thank the Mathematics Research Centre of the University of Warwick
266: for its warm hospitality. This work was supported by the Fonds National Suisse.
267: \end{acknowledge}
268:
269: \section{General theory of stochastic dynamical systems}
270: \label{sec:general}
271:
272: In this section, we first construct an abstract framework that can be used to model a large
273: class of physically relevant models where the driving noise is stationary. Our framework
274: is very closely related to the framework of random dynamical systems with however one
275: fundamental difference. In the theory of random dynamical systems (RDS), the abstract space
276: $\Omega$ used to model the noise part typically encodes the {\em future}
277: of the noise process. In our framework of ``stochastic dynamical systems'' (SDS)
278: the noise space $\Noise$ typically encodes the {\em past} of the noise process. As a consequence,
279: the evolution on $\Noise$ will be stochastic, as opposed to the deterministic evolution on $\Omega$
280: one encounters in the theory of RDS. This distinction may seem futile at first sight, and one could
281: argue that the difference between RDS and SDS is non-existent by adding the past of the noise process to
282: $\Omega$ and its future to $\CW$.
283:
284: The additional structure we require is that the evolution on $\Noise$
285: possesses a \textit{unique} invariant measure. Although this requirement may sound very strong, it is actually
286: not, and most natural examples satisfy it, as long as $\Noise$ is chosen in such a way that it
287: does not contain information about the future of the noise.
288: In very loose terms, this requirement of having a unique invariant
289: measure states that the noise process driving our system is stationary and that the Markov process
290: modelling its evolution captures all its essential features in such a way that it could not be
291: used to describe a noise process different from the one at hand. In particular, this means that
292: there is a continuous inflow of ``new randomness'' into the system, which is a crucial feature
293: when trying to apply probabilistic methods to the study of ergodic properties of the system.
294: This is in opposition to the RDS formalism, where the noise is ``frozen'', as soon as an
295: element of $\Omega$ is chosen.
296:
297: From the mathematical point of view, we will consider that the physical process we are interested in
298: lives on a ``state space'' $\State$ and that its driving noise belongs to a
299: ``noise space'' $\Noise$. In both cases, we only consider Polish (\ie complete, separable, and
300: metrisable) spaces. One should think of the state space as a relatively small space which contains
301: all the information accessible to a physical observer of the process. The noise space should be
302: thought of as a much bigger abstract space containing all the information needed to construct
303: a mathematical model of the driving noise up to a certain time. The information contained in the
304: noise space is not accessible to the physical observer.
305:
306: Before we state our definition of a SDS, we will recall several notations and definitions, mainly
307: for the sake of mathematical rigour. The reader can safely skip the next subsection and come back
308: to it for reference concerning the notations and the mathematically precise definitions of the concepts
309: that are used.
310:
311: \subsection{Preliminary definitions and notations}
312:
313: First of all, recall he definition of a transition semigroup:
314: \begin{definition}\label{def:kernel}
315: Let $(\CE,{\cscr E})$ be a Polish space endowed with its
316: Borel $\sigma$-field. A {\em transition semigroup} $\CP_t$ on $\CE$ is a family of maps
317: $\CP_t \from \CE \times {\cscr E} \to [0,1]$ indexed by $t \in [0,\infty)$ such that
318: \begin{claim}[iii)]
319: \item[i)] for every $x\in\CE$, the map $A \mapsto \CP_t(x,A)$ is a probability measure on $\CE$
320: and, for every $A \in {\cscr E}$, the map $x \mapsto \CP_t(x,A)$ is ${\cscr E}$-measurable,
321: \item[ii)] one has the identity
322: \begin{equ}
323: \CP_{s+t}(x,A) = \int_{\CE} \CP_s(y,A)\,\CP_t(x,dy)\;,
324: \end{equ}
325: for every $s,t > 0$, every $x\in\CE$, and every $A\in {\cscr E}$.
326: \item[iii)] $\CP_0(x,\cdot) = \delta_x$ for every $x\in\CE$.
327: \end{claim}
328: \end{definition}
329: We will freely use the notations
330: \begin{equ}
331: \sml(\CP_t \psi\smr)(x) = \int_\CE \psi(y)\,\CP_t(x,dy)\;,\qquad \sml(\CP_t \mu\smr)(A) = \int_\CE \CP_t(x,A)\,\mu(dx)\;,
332: \end{equ}
333: where $\psi$ is a measurable function on $\CE$ and $\mu$ is a measure on $\CE$.
334:
335: Since we will always work with topological spaces, we will require our transition semigroups
336: to have good topological properties. Recall that a sequence $\{\mu_n\}$ of measures on a topological
337: space $\CE$ is said to converge toward a limiting measure $\mu$ in the weak topology if
338: \begin{equ}
339: \int_\CE \psi(x)\,\mu_n(dx) \to \int_\CE \psi(x)\,\mu(dx)\;,\qquad \forall \psi \in \CC_b(\CE)\;,
340: \end{equ}
341: where $\CC_b(\CE)$ denotes the space of bounded continuous functions from $\CE$ into $\R$.
342: In the sequel, we will use the notation $\pMeas(\CE)$ to denote the space of probability
343: measures on a Polish space $\CE$, endowed with the topology of weak convergence.
344:
345: \begin{definition}\label{def:Feller}
346: A {\em transition semigroup} $\CP_t$ on a Polish space $\CE$ is {\em Feller} if it maps $\CC_b(\CE)$
347: into $\CC_b(\CE)$.
348: \end{definition}
349:
350: \begin{remark}\label{rem:cont}
351: This definition is equivalent to the requirement that $x \mapsto \CP_t(x,\cdot\,)$ is continuous from
352: $\CE$ to $\pMeas(\CE)$. As a consequence, Feller semigroups preserve the weak topology in the
353: sense that if $\mu_n \to \mu$ in $\pMeas(\CE)$, then $\CP_t \mu_n \to \CP_t \mu$ in $\pMeas(\CE)$
354: for every given $t$.
355: \end{remark}
356:
357: Now that we have defined the ``good'' objects for the ``noisy'' part of our construction, we turn
358: to the trajectories on the state space. We are looking for a space which has good topological
359: properties but which is large enough to contain most interesting examples. One such space
360: is the space of {\em \cadlag} paths (continu \`a droite, limite \`a gauche --- continuous on the right,
361: limits on the left), which can be turned into a Polish space when equipped with a suitable topology.
362:
363: \begin{definition}\label{def:cadlag}
364: Given a Polish space $\CE$ and a positive number $T$, the space $\CD([0,T],\CE)$ is the
365: set of functions $f\from[0,T]\to \CE$ that are right-continuous and whose left-limits exist at every point.
366: A sequence $\{f_n\}_{n\in\N}$ converges to a limit $f$ if and only if there exists a sequence $\{\lambda_n\}$
367: of continuous and increasing functions $\lambda_n\from[0,T] \to [0,T]$ satisfying $\lambda_n(0) = 0$,
368: $\lambda_n(T)=T$, and such that
369: \begin{equ}[e:cond1]
370: \lim_{n\to \infty}\sup_{0\le s < t \le T} \Bigl|{\log{\lambda_n(t) - \lambda_n(s) \over t-s}}\Bigr| = 0\;,
371: \end{equ}
372: and
373: \begin{equ}[e:cond2]
374: \lim_{n\to\infty} \sup_{0\le t\le T} d\sml(f_n(t), f(\lambda_n(t))\smr) = 0\;,
375: \end{equ}
376: where $d$ is any totally bounded metric on $\CE$ which generates its topology.
377:
378: The space $\CD(\R_+, \CE)$ is the space of all functions from $\R_+$ to $\CE$ such that their restrictions to $[0,T]$
379: are in $\CD([0,T],\CE)$ for all $T>0$. A sequence converges in $\CD(\R_+, \CE)$ if there exists a sequence
380: $\{\lambda_n\}$ of continuous and increasing functions $\lambda_n\from\R_+\to \R_+$ satisfying $\lambda_n(0) = 0$
381: and such that \eref{e:cond1} and \eref{e:cond2} hold.
382: \end{definition}
383:
384: It can be shown (see \eg \cite{MR88a:60130} for a proof) that the spaces $\CD([0,T],\CE)$ and $\CD(\R_+, \CE)$
385: are Polish when equipped with the above topology (usually called the Skorohod topology). Notice that the
386: space $\CD([0,T],\CE)$ has a natural embedding into $\CD(\R_+, \CE)$ by setting $f(t) = f(T)$ for $t>T$ and
387: that this embedding is continuous. However, the restriction operator from $\CD(\R_+, \CE)$ to $\CD([0,T],\CE)$
388: is not continuous, since the topology on $\CD([0,T],\CE)$ imposes that $f_n(T) \to f(T)$, which is not imposed
389: by the topology on $\CD(\R_+, \CE)$.
390:
391: In many interesting situations, it is enough to work with continuous sample paths, which live in much simpler spaces:
392:
393: \begin{definition}\label{def:continuous}
394: Given a Polish space $\CE$ and a positive number $T$, the space $\CC([0,T],\CE)$ is the
395: set of continuous functions $f\from[0,T]\to \CE$ equipped with the supremum norm.
396:
397: The space $\CC(\R_+, \CE)$ is the space of all functions from $\R_+$ to $\CE$ such that their restrictions to $[0,T]$
398: are in $\CC([0,T],\CE)$ for all $T>0$. A sequence converges in $\CC(\R_+, \CE)$ if all its restrictions converge.
399: \end{definition}
400:
401: It is a standard result that the spaces $\CC([0,T],\CE)$ and $\CC(\R_+, \CE)$ are Polish if $\CE$ is Polish. We
402: can now turn to the definition of the systems we are interested in.
403:
404: \subsection{Definition of a SDS}
405:
406: Let us recall the following standard notations.
407: Given a product space $\State \times \Noise$, we denote by $\Pi_\State$ and $\Pi_\Noise$ the maps that select the
408: first (resp.\ second) component of an element. Also, given two measurable spaces $\CE$ and $\CF$, a measurable
409: map $f\from\CE\to\CF$, and a measure $\mu$ on $\CE$, we define the measure $f^*\mu$ on $\CF$ in the natural
410: way by $f^*\mu = \mu \circ f^{-1}$.
411:
412: We first define the class of noise processes we will be interested in:
413: \begin{definition}\label{def:noise}
414: A quadruple $\sml(\Noise, \{\CP_t\}_{t\ge0}, \pNoise, \{\theta_t\}_{t\ge 0}\smr)$ is called a
415: {\em stationary noise process} if it satisfies the following:
416: \begin{claim}[iii)]
417: \item[i)] $\Noise$ is a Polish space,
418: \item[ii)] $\CP_t$ is a Feller transition semigroup on $\Noise$, which accepts $\pNoise$ as its unique
419: invariant measure,
420: \item[iii)] The family $\{\theta_t\}_{t>0}$ is a semiflow of measurable maps on $\Noise$ satisfying the property
421: $\theta_t^*\CP_t^{}(x,\cdot) = \delta_x$ for every $x\in\Noise$.
422: \end{claim}
423: \end{definition}
424:
425: This leads to the following definition of SDS, which is intentionally kept as close as possible to
426: the definition of RDS in \cite[Def.~1.1.1]{RDS}:
427: \begin{definition}\label{def:SDS}
428: A {\em stochastic dynamical system} on the Polish space $\State$ over the stationary noise
429: process $\sml(\Noise, \{\CP_t\}_{t\ge0}, \pNoise, \{\theta_t\}_{t\ge 0}\smr)$ is a mapping
430: \begin{equ}
431: \phi\from \R_+ \times \State \times \Noise \to \State \;,\qquad (t,x,w)\mapsto \phi_t(x,w)\;,
432: \end{equ}
433: with the following properties:
434: \begin{claim}[(SDS2)]
435: \item[\maketag{SDS1}{tag:SDS1}] {\em Regularity of paths:} For every $T>0$, $x\in\State$, and $w\in\Noise$, the map $\Phi_T(x,w)\from[0,T]\to \State$ defined by
436: \begin{equ}{}
437: \Phi_T(x,w)(t) = \phi_t(x,\theta_{T-t}w)\;,
438: \end{equ}
439: belongs to $\CD([0,T],\State)$.
440: \item[\maketag{SDS2}{tag:SDS2}] {\em Continuous dependence:} The maps $(x,w)\mapsto \Phi_T(x,w)$ are continuous from
441: $\State\times\Noise$ to $\CD([0,T],\State)$ for every $T>0$.
442: \item[\maketag{SDS3}{tag:SDS3}] {\em Cocycle property:} The family of mappings $\phi_t$ satisfies
443: \begin{equs}
444: \phi_0(x,w) &= x\;,\\
445: \phi_{s+t}(x,w) &= \phi_s\sml(\phi_t(x,\theta_s w),w\smr)\;, \label{e:cocycle}
446: \end{equs}
447: for all $s,t > 0$, all $x\in\State$, and all $w\in\Noise$.
448: \end{claim}
449: \end{definition}
450:
451: \begin{remark}\label{rem:Hans}
452: The above definition is very close to the definition of {\em Markovian random dynamical system}
453: introduced in \cite{MR93c:60096}. Beyond the technical differences, the main difference is a shift in
454: the viewpoint: a Markovian RDS is built on top of a RDS, so one can analyse it from both a semigroup
455: point of view and a RDS point of view. In the case of a SDS as defined above, there is no underlying RDS
456: (although one can always construct one), so the semigroup point of view is the only one we consider.
457: \end{remark}
458:
459: \begin{remark}
460: The cocycle property \eref{e:cocycle} looks different from the cocycle property for random dynamical systems.
461: Actually, in our case $\phi$ is a {\em backward cocycle} for $\theta_t$, which is reasonable since,
462: as a ``left inverse'' for $\CP_t$, $\theta_t$ actually pushes time backward.
463: Notice also that, unlike in the definition of RDS, we require some continuity property with respect to
464: the noise to hold. This continuity property sounds quite restrictive, but it is actually mainly a matter of
465: choosing a topology on $\Noise$, which is in a sense ``compatible'' with the topology on $\State$.
466: \end{remark}
467:
468:
469: Similarly, we define a continuous (where ``continuous'' should be thought of as continuous with respect
470: to time) SDS by
471:
472: \begin{definition}\label{def:contSDS}
473: A SDS is said to be {\em continuous} if $\CD([0,T],\State)$ can be replaced by $\CC([0,T],\State)$ in
474: the above definition.
475: \end{definition}
476:
477: \begin{remark}\label{rem:compat}
478: One can check that the embeddings $\CC([0,T],\State) \hookrightarrow \CD([0,T],\State)$
479: and $\CC(\R_+,\State) \hookrightarrow \CD(\R_+,\State)$ are continuous, so a continuous SDS also satisfies
480: Definition~\ref{def:SDS} of a SDS.
481: \end{remark}
482:
483:
484: Given a SDS as in Definition~\ref{def:SDS} and an initial condition $x_0 \in \State$, we now turn to the
485: construction of a stochastic process with initial condition $x_0$ constructed in a natural way from $\phi$.
486: First, given $t\ge 0$ and $(x,w) \in \State\times\Noise$,
487: we construct a probability measure $\CQ_t(x,w;\cdot\,)$ on $\State\times\Noise$ by
488: \begin{equ}[e:defQt]
489: \CQ_t(x,w; A\times B) = \int_B \delta_{\phi_t(x,w')}(A)\, \CP_t(w, dw')\;,
490: \end{equ}
491: where $\delta_x$ denotes the delta measure located at $x$. The following result is elementary:
492: %
493: %
494: \begin{lemma}\label{lem:semigroup}
495: Let $\phi$ be a SDS on $\State$ over $\sml(\Noise, \{\CP_t\}_{t\ge0}, \pNoise, \{\theta_t\}_{t\ge 0}\smr)$ and define the family
496: of measures $\CQ_t(x,w;\cdot\,)$ by \eref{e:defQt}. Then $\CQ_t$ is a Feller transition semigroup on $\State \times \Noise$.
497: Furthermore, it has the property that if $\Pi_\Noise^* \mu = \pNoise$
498: for a measure $\mu$ on $\State \times\Noise$, then $\Pi_\Noise^* \CQ_t \mu = \pNoise$.
499: \end{lemma}
500: %
501: %
502: \begin{proof}
503: The fact that $\Pi_\Noise^* \CQ_t \mu = \pNoise$
504: follows from the invariance of $\pNoise $ under $\CP_t$.
505: We now check that $\CQ_t$ is a Feller transition semigroup. Conditions {\it i)} and {\it iii)} follow immediately from
506: the properties of $\phi$. The continuity of $\CQ_t(x,w;\cdot\,)$ with respect to $(x,w)$ is a straightforward consequence
507: of the facts that $\CP_t$ is Feller and that $(x,w) \mapsto \phi_t(x,w)$ is continuous (the latter statement follows from
508: \eref{tag:SDS2} and the definition of the topology on $\CD([0,t],\State)$).
509:
510: It thus remains only to check that the Chapman-Kolmogorov equation holds. We have from the cocycle property:
511: \begin{equs}
512: \CQ_{s+t}(x,w;&A\times B) = \int_B \delta_{\phi_{s+t}(x,w')}(A)\, \CP_{s+t}(w, dw') \\
513: &= \int_B \int_\State \delta_{\phi_{s}(y,w')}(A) \delta_{\phi_{t}(x,\theta_s w')}(dy)\, \CP_{s+t}(w, dw')\\
514: &= \int_\Noise \int_B \int_\State \delta_{\phi_{s}(y,w')}(A) \delta_{\phi_{t}(x,\theta_s w')}(dy)\, \CP_{s}(w'', dw')\, \CP_{t}(w, dw'')\;.
515: \end{equs}
516: The claim then follows from the property $\theta_s^*\CP_s(w'',dw') = \delta_{w''}(dw')$ by exchanging the order of integration.
517: \end{proof}
518:
519: \begin{remark}\label{rem:manypoint}
520: Actually, \eref{e:defQt} defines the evolution of the one-point process generated by $\phi$. The $n$-points process
521: would evolve according to
522: \begin{equ}
523: \CQ_t^{(n)}(x_1,\ldots,x_n,w; A_1\times\ldots\times A_n \times B) = \int_B \prod_{i=1}^n \delta_{\phi_t(x_i,w')}(A_i )\, \CP_t(w, dw')\;.
524: \end{equ}
525: One can check as above that this defines a Feller transition semigroup on $\State^n \times \Noise$.
526: \end{remark}
527:
528:
529: This lemma suggests the following definition:
530: \begin{definition}\label{def:IC}
531: Let $\phi$ be a SDS as above. Then a probability measure $\mu$ on $\State\times\Noise$ is called a {\em generalised
532: initial condition} for $\phi$ if $\Pi_\Noise^* \mu = \pNoise$. We denote by $\Meas$ the space of generalised initial conditions
533: endowed with the topology of weak convergence. Elements of $\Meas$ that are of the form
534: $\mu = \delta_x\times \pNoise$ for some $x\in\State$ will be called {\em initial conditions}.
535: \end{definition}
536:
537: Given a generalised initial condition $\mu$, it is natural construct a stochastic process $(x_t,w_t)$ on $\State\times\Noise$ by
538: drawing its initial condition according to $\mu$ and then evolving it according to the transition semigroup $\CQ_t$.
539: The marginal $x_t$ of this process on $\State$ will be called the {\em process generated by $\phi$ for $\mu$}. We will denote
540: by $\Evol \mu$ the law of this process (\ie $\Evol\mu$ is a measure on $\CD(\R_+,\State)$ in the general case and
541: a measure on $\CC(\R_+,\State)$ in the continuous case). More rigorously, we define for every $T>0$ the measure
542: $\Evol_T\mu$ on $\CD([0,T],\State)$ by
543: \begin{equ}
544: \Evol_T\mu = \Phi_T^*\CP_t \mu\;,
545: \end{equ}
546: where $\Phi_T$ is defined as in \eref{tag:SDS1}. By the embedding $\CD([0,T],\State)\hookrightarrow\CD(\R_+,\State)$,
547: this actually gives a family of measures on $\CD(\R_+,\State)$.
548: It follows from the cocycle property that the restriction to $\CD([0,T],\State)$
549: of $\Evol_{T'}\mu$ with $T'>T$ is equal to $\Evol_T\mu$. The definition of the topology on $\CD(\R_+,\State)$
550: does therefore imply that the sequence $\Evol_T\mu$ converges weakly to a unique measure on $\CD(\R_+,\State)$
551: that we denote by $\Evol\mu$. A similar argument, combined with \eref{tag:SDS2} yields
552: %
553: %
554: \begin{lemma}\label{lem:cont}
555: Let $\phi$ be a SDS. Then, the operator $\Evol$ as defined above is continuous from $\Meas$ to
556: $\pMeas(\CD(\R_+,\State))$. \qed
557: \end{lemma}
558:
559:
560: This in turn motivates the following equivalence relation:
561: \begin{definition}\label{def:equiv}
562: Two generalised initial conditions $\mu$ and $\nu$ of a SDS $\phi$ are {\em equivalent} if the processes generated
563: by $\mu$ and $\nu$ are equal in law. In short, $\mu \sim \nu \Leftrightarrow \Evol \mu = \Evol \nu$.
564: \end{definition}
565: The physical interpretation of this notion of equivalence is that the noise space contains some redundant information
566: that is not required to construct the future of the system. Note that
567: this does not necessarily mean that the noise space could be reduced in order to have a more ``optimal'' description
568: of the system. For example, if the process $x_t$ generated by any generalised initial condition is Markov, then all the information
569: contained in $\Noise$ is redundant in the above sense (\ie $\mu$ and $\nu$ are equivalent if $\Pi_\State^*\mu = \Pi_\State^*\nu$).
570: This does of course not mean that $\Noise$ can be entirely thrown away in the above description (otherwise,
571: since the map $\phi$ is deterministic, the evolution
572: would become deterministic).
573:
574: The main reason for introducing the notion of SDS is to have a framework in which one can study
575: ergodic properties of physical systems with memory. It should be noted that it is designed to describe
576: systems where the memory is \textit{extrinsic}, as opposed to systems with \textit{intrinsic} memory like
577: stochastic delay equations.
578: We present in the next subsection a few elementary ergodic results in
579: the framework of SDS.
580:
581: \subsection{Ergodic properties}
582:
583: In the theory of Markov processes, the main tool for investigating ergodic properties is the {\em invariant measure}.
584: In the setup of SDS, we say that a measure $\mu$ on $\State\times\Noise$ is invariant for the SDS $\phi$ if it
585: is invariant for the Markov transition semigroup $\CQ_t$ generated by $\phi$. We say that a
586: measure $\mu$ on $\State\times\Noise$ is {\em stationary} for $\phi$ if one has
587: \begin{equ}
588: \CQ_t \mu \sim \mu\;,\quad \forall t>0\;,
589: \end{equ}
590: \ie if the process on $\State$ generated by $\mu$ is stationary. Following our philosophy of considering only what happens
591: on the state space $\State$, we should be interested in stationary measures, disregarding completely whether they
592: are actually invariant or not. In doing so, we could be afraid of loosing many convenient results from the
593: well-developed
594: theory of Markov processes. Fortunately, the following lemma shows that the set of invariant
595: measures and the set of stationary measures are actually the same, when quotiented
596: by the equivalence relation of Definition~\ref{def:equiv}.
597: %
598: %
599: \begin{proposition}\label{prop:invariant}
600: Let $\phi$ be a SDS and let $\mu$ be a stationary measure for $\phi$. Then, there exists
601: a measure $\mu_\star \sim \mu$ which is invariant for $\phi$.
602: \end{proposition}
603:
604: %
605: %
606: \begin{proof}
607: Define the ergodic averages
608: \begin{equ}[e:ergodic]
609: \CR_T \mu = {1 \over T}\int_0^T \CQ_t \mu\, dt\;.
610: \end{equ}
611: Since $\mu$ is stationary, we have $\Pi_\State^*\CR_T \mu = \Pi_\State^* \mu$ for every $T$.
612: Furthermore, $\Pi_\Noise^*\CR_T \mu = \pNoise$ for every $T$, therefore the sequence of measures
613: $\CR_T \mu$ is tight on $\State \times \Noise$. Let $\mu_\star$ be any of its accumulation points
614: in $\pMeas(\State \times \Noise)$. Since $ \CQ_t $ is Feller, $\mu_\star$ is invariant for $ \CQ_t $
615: and, by \lem{lem:cont}, one has $\mu_\star \sim \mu$.
616: \end{proof}
617:
618: From a mathematical point of view, it may in some cases be interesting to know whether the invariant
619: measure $\mu_\star$ constructed in \prop{prop:invariant} is uniquely determined by $\mu$. From
620: an intuitive point of view, this uniqueness property should hold
621: if the information contained in the trajectories on the state space $\State $ is sufficient to
622: reconstruct the evolution of the noise. This intuition is made rigorous by the following proposition.
623:
624: %
625: %
626: \begin{proposition}\label{prop:faithful}
627: Let $\phi$ be a SDS, define ${\cscr W}_T^x$ as the $\sigma$-field on $\Noise$ generated
628: by the map $\Phi_T(x,\cdot\,)\from\Noise\to\CD([0,T],\State)$, and set ${\cscr W}_T = \bigwedge_{x\in\State} {\cscr W}_T^x$.
629: Assume that ${\cscr W}_T \subset {\cscr W}_{T'}$ for $T < T'$ and that
630: ${\cscr W} = \bigvee_{T\ge 0} {\cscr W}_T$ is equal to the Borel
631: $\sigma$-field on $\Noise$. Then, for $\mu_1$ and $\mu_2$ two invariant measures, one has
632: the implication $\mu_1 \sim \mu_2 \Rightarrow \mu_1 = \mu_2$.
633: \end{proposition}
634: %
635: %
636:
637: \begin{proof}
638: Assume $\mu_1 \sim \mu_2$ are two invariant measures for $\phi$.
639: Since ${\cscr W}_T \subset {\cscr W}_{T'}$ if $T < T'$, their equality follows if one can show that, for every $T > 0$,
640: \begin{equ}[e:equal]
641: \expect \bigl(\mu_1\,|\, {\cscr X} \otimes {\cscr W}_T \bigr) = \expect \bigl(\mu_2\,|\, {\cscr X} \otimes {\cscr W}_T\bigr)\;,
642: \end{equ}
643: where ${\cscr X}$ denotes the Borel $\sigma$-field on $\State$.
644:
645: Since $\mu_1 \sim \mu_2$, one has in particular $\Pi_\State^*\mu_1 = \Pi_\State^*\mu_2$, so let us call
646: this measure $\nu$. Since $\Noise$ is Polish, we then have the disintegration $x \mapsto \mu_i^x$, yielding formally
647: $\mu_i(dx,dw) = \mu_i^x(dw)\,\nu(dx)$, where $\mu_i^x$ are probability measures on $\Noise$.
648: (See \cite[p.~196]{MR58:18632} for a proof.) Fix $T>0$ and define the family $\mu_i^{x,T}$ of probability measures
649: on $\Noise$ by
650: \begin{equ}
651: \mu_i^{x,T} = \int_\Noise \CP_t(w,\cdot\,)\,\mu_i^x(dw)\;.
652: \end{equ}
653: With this definition, one has
654: \begin{equ}
655: \Evol_T\mu_i = \int_\State \bigl(\Phi_T(x,\cdot\,)^*\mu_i^{x,T}\bigr)\,\nu(dx)\;.
656: \end{equ}
657: Let $e_0\from \CD([0,T],\State)\to\State$ be the evaluation map at $0$, then
658: \begin{equ}
659: \expect \bigl(\Evol_T\mu_i\,|\, e_0 = x\bigr) = \bigl(\Phi_T(x,\cdot\,)^*\mu_i^{x,T}\bigr)\;,
660: \end{equ}
661: for $\nu$-almost every $x\in\State$.
662: Since $\Evol_T\mu_1 = \Evol_T\mu_2$, one therefore has
663: \begin{equ}[e:equal_2]
664: \expect \bigl(\mu_1^{x,T} \,|\, {\cscr W}_T^x \bigr)
665: = \expect \bigl(\mu_2^{x,T} \,|\, {\cscr W}_T^x \bigr)\;,
666: \end{equ}
667: for $\nu$-almost every $x\in\State$. On the other hand, the invariance of $\mu_i$ implies that,
668: for every $A\in {\cscr X}$ and every $B \in {\cscr W}_T$, one has the equality
669: \begin{equ}
670: \mu_i(A \times B) = \int_\State \int_B \chi_A \bigl(\phi_T(x,w)\bigr)\, \mu_i^{x,T}(dw)\,\nu(dx)\;.
671: \end{equ}
672: Since $\phi_T(x,\cdot\,)$ is ${\cscr W}_T^x$-measurable and $B \in {\cscr W}_T^x$, this is equal to
673: \begin{equ}
674: \int_\State \int_B \chi_A \bigl(\phi_T(x,w)\bigr)\, \expect \bigl(\mu_i^{x,T}\,|\, {\cscr W}_T^x\bigr)(dw)\,\nu(dx)\;.
675: \end{equ}
676: Thus \eref{e:equal_2} implies \eref{e:equal} and the proof of \prop{prop:faithful} is complete.
677: \end{proof}
678:
679: The existence of an invariant measure is usually established by finding a Lyapunov function.
680: In this setting, Lyapunov functions are given by the following definition.
681:
682: \begin{definition}\label{def:Lyap}
683: Let $\phi$ be a SDS and let $F\from\State \to [0,\infty)$ be a continuous function. Then $F$
684: is a {\em Lyapunov function for $\phi$} if it satisfies the following conditions:
685: \begin{claim}[(L2)]
686: \item[(L1)] The set $F^{-1}([0,C])$ is compact for every $C \in [0,\infty)$.
687: \item[(L2)] There exist constants $C$ and $\gamma > 0$ such that
688: \begin{equ}[e:condLyap]
689: \int_{\State\times\Noise} F(x)\,\sml(\CQ_t \mu\smr)(dx,dw) \le C + e^{-\gamma t} \int_{\State} F(x)\,\sml(\Pi_\State^*\mu\smr)(dx)\;,
690: \end{equ}
691: for every $t>0$ and every generalised initial condition $\mu$ such that the right-hand side is finite.
692: \end{claim}
693: \end{definition}
694:
695: It is important to notice that one does {\em not} require $F$ to be a Lyapunov function for the
696: transition semigroup $\CQ_t$, since \eref{e:condLyap} is only required to hold for measures
697: $\mu$ satisfying $\Pi_\Noise^*\mu = \pNoise$. One nevertheless has the following result:
698:
699: \begin{lemma}\label{lem:Lyap}
700: Let $\phi$ be a SDS. If there exists a Lyapunov function $F$ for $\phi$, then there exists also
701: an invariant measure $\mu_\star$ for $\phi$, which satisfies
702: \begin{equ}[e:estInv]
703: \int_{\State\times\Noise} F(x)\,\mu_\star(dx,dw) \le C\;.
704: \end{equ}
705: \end{lemma}
706:
707: \begin{proof}
708: Let $x\in\State$ be an arbitrary initial condition, set $\mu = \delta_x \times \pNoise$, and
709: define the ergodic averages $\CR_T \mu$ as in \eref{e:ergodic}. Combining (L1) and (L2) with the fact that
710: $\Pi_\Noise^*\CR_T \mu = \pNoise$, one immediately gets the tightness of the sequence $\{\CR_T \mu\}$. By the
711: standard Krylov-Bogoloubov argument, any limiting point of $\{\CR_T \mu\}$ is an invariant measure for $\phi$.
712: The estimate \eref{e:estInv} follows from \eref{e:condLyap}, combined with the fact that $F$ is continuous.
713: \end{proof}
714:
715:
716:
717: This concludes our presentation of the abstract framework in which we analyse the
718: ergodic properties of \eref{e:mainequ}.
719:
720: %
721: %
722: %
723: \section{Construction of the SDS}
724: \label{sec:setting}
725: %
726: %
727: %
728:
729: In this section, we construct a continuous stochastic dynamical system which
730: yields the solutions to \eref{e:mainequ} in an appropriate sense.
731:
732: First of all, let us discuss what we mean by ``solution'' to \eref{e:mainequ}.
733: \begin{definition}\label{def:solution}
734: Let $\{x_t\}_{t\ge 0}$ be a stochastic process with continuous sample paths. We say
735: that $x_t$ is a {\em solution} to \eref{e:mainequ} if the stochastic process $N(t)$ defined by
736: \begin{equ}[e:fBmsol]
737: N(t) = x_t - x_0 - \int_0^t f(x_s)\,ds\;,
738: \end{equ}
739: is equal in law to $\sigma B_H(t)$, where $\sigma$ is as in \eref{e:mainequ} and $B_H(t)$ is
740: a $n$-dimensional fBm with Hurst parameter $H$.
741: \end{definition}
742:
743: We will set up our SDS in such a way that, for every generalised initial condition $\mu$, the canonical
744: process associated to the measure $\Evol\mu$ is a solution to \eref{e:mainequ}. This will be the content
745: of \prop{prop:defSDS} below. In order to achieve this,
746: our main task is to set up a noise process in a way which complies to Definition~\ref{def:noise}.
747:
748: \subsection{Representation of the fBm}
749:
750: In this section, we give a representation of the fBm $B_H(t)$ with
751: Hurst parameter $H \in (0,1)$ which is suitable for our analysis.
752: Recall that, by definition, $B_H(t)$ is a centred Gaussian process satisfying $B_H(0) = 0$ and
753: \begin{equ}[e:propfBm]
754: \expect|B_H(t) - B_H(s)|^2 = |t-s|^{2H}\;.
755: \end{equ}
756: %The usual representation for the fractional Brownian
757: %motion with Hurst parameter $H>{1\over 2}$ is given by
758: %\begin{equ}[e:usual]
759: % B_H(t) = c_H \int_0^t K_H(t,s)\,dw(s)\;,
760: %\end{equ}
761: %for some positive constant $c_H$ and
762: %\begin{equ}
763: % K_H(t,s) = s^{{1\over2}-H} \int_s^t (y-s)^{H-{3\over 2}}y^{H-{1\over 2}}\,dy\;.
764: %\end{equ}
765: %In the case $H <{1\over 2}$, a similar expression exists, but the corresponding kernel $K$ is
766: %slightly more complicated.
767: %This representation is not very natural for our purpose which is to construct and study stationary solutions
768: %to \eref{e:mainequ},
769: %since it is precisely not stationary (it is only defined for $t>0$).
770: Naturally, a {\it two-sided fractional
771: Brownian motion} by requiring that \eref{e:propfBm} holds for all $s,t\in\R$.
772: Notice that, unlike for the normal Brownian motion, the two-sided fBm is \textit{not} obtained by gluing
773: two independent
774: copies of the one-sided fBm together at $t=0$.
775: We have the following useful representation of the two-sided fBm, which is also (up to the normalisation constant)
776: the representation used in the original paper \cite{MR39:3572}.
777:
778: \begin{lemma}\label{lem:fbm}
779: Let $w(t)$, $t\in\R$ be a two-sided Wiener process and let $H \in (0,1)$. Define for some constant $\alpha_H$ the process
780: \begin{equ}[e:repres]
781: B_H(t) = \alpha_H\int_{-\infty}^0 \Kern(-r)\,\sml(dw(r+t) - dw(r)\smr)\;.
782: \end{equ}
783: Then there exists a choice of $\alpha_H$ such that $B_H(t)$ is a two-sided fractional Brownian motion with
784: Hurst parameter $H$. \qed
785: \end{lemma}
786:
787: \begin{notation}
788: Given the representation \eref{e:repres} of the fBm with Hurst parameter $H$, we call $w$ the ``Wiener process associated to $B_H$''. We
789: also refer to $\{w(t)\,:\, t\le 0\}$ as the ``past'' of $w$ and to $\{w(t)\,:\, t > 0\}$ as the ``future'' of $w$. We similarly refer to the ``past'' and the ``future'' of $B_H$. Notice the notion of future for $B_H$ is different from the notion of future for $w$ in terms of $\sigma$-algebras, since the future of $B_H$
790: depends on the past of $w$.
791: \end{notation}
792:
793: \begin{remark}
794: The expression \eref{e:repres} looks strange at first sight, but one should actually think of $B_H(t)$ as being given by
795: $B_H(t) = \tilde B_H(t) - \tilde B_H(0)$, where
796: \begin{equ}[e:Btilde]
797: \raise0.4em\hbox{``}\tilde B_H(t) = \alpha_H \int_{-\infty}^t \Kern(t-s)\,dw(s)\;.\;\raise0.4em\hbox{''}
798: \end{equ}
799: This expression is strongly reminiscent of the usual representation of the stationary Ornstein-Uhlenbeck process, but with an algebraic kernel instead of an exponential one.
800: Of course, \eref{e:Btilde} does not make any sense since $(t-s)^{H-{1\over 2}}$ is not square integrable. Nevertheless, \eref{e:Btilde}
801: has the advantage of explicitly showing the stationarity of the increments for the two-sided fBm.
802: %In can also be
803: %interpreted as the large-time limit of the usual
804: %representation \eref{e:usual}, using the fact that
805: %\begin{equ}
806: % K_H(t,s) \approx (t-s)^{H-{1\over 2}}
807: %\end{equ}
808: %for $t\gg s$.
809: \end{remark}
810:
811: %\begin{proof}[of \lem{lem:fbm}]
812: %It follows from the definition that $B_H(t)$ is a centred Gaussian process. It thus suffices to check that
813: %\begin{equ}
814: % \expect |B_H(t)-B_H(s)|^2 \propto |t-s|^{2H}\;.
815: %\end{equ}
816: %Assume $t>s$. We then have
817: %\begin{equ}
818: % \expect |B_H(t)-B_H(s)|^2 = \int_{-\infty}^s \bigl(\Kern(t-r)-\Kern(s-r)\bigr)^2\,dr + \int_s^t (t-r)^{2H-1}\,dr\;.
819: %\end{equ}
820: %The contribution of the second term is obviously proportional to $|t-s|^{2H}$. After a change of variable, the first term can be written as
821: %\begin{equ}
822: % \int_{0}^\infty \bigl(\Kern(r+t-s)-\Kern r\bigr)^2\,dr = (t-s)^{2H}\int_0^\infty \bigl(\Kern(r+1)-\Kern r\bigr)^2\,dr\;,
823: %\end{equ}
824: %thus also producing a contribution proportional to $|t-s|^{2H}$. It turns out that the constant $\alpha_H$ appearing in the statement is actually given by
825: %\begin{equ}
826: % \alpha_H = \sqrt{(2-2H)\Gamma\sml({\textstyle{3\over 2}}-H\smr) \over \Gamma(2-2H)\Gamma\sml(H+{\textstyle{1\over 2}}\smr)}\;.
827: %\end{equ}
828: %This concludes the proof of \lem{lem:fbm}.
829: %\end{proof}
830:
831: \subsection{Noise spaces}
832:
833: In this section, we introduce the family of spaces that will be used to model our noise.
834: Denote by $\CC_0^\infty(\R_-)$ the set
835: of $\CC^\infty$ function $w\from(-\infty,0] \to \R$ satisfying $w(0) = 0$ and having compact support.
836: Given a parameter $H\in(0,1)$,
837: we define for every $w\in\CC_0^\infty(\R_-)$ the norm
838: \begin{equ}[e:defN]
839: \|w\|_H = \sup_{t,s\in\R_-} {|w(t)-w(s)| \over |t-s|^{1-H\over 2} \sml(1+|t| + |s|\smr)^{1\over2}}\;.
840: \end{equ}
841: We then define the Banach space $\CH_H$ to be the closure of $\CC_0^\infty(\R_-)$ under the norm
842: $\|\nb\cdot\nb\|_H$. The following lemma is important in view of the framework exposed in Section~\ref{sec:general}:
843:
844: \begin{lemma}\label{lem:separ}
845: The spaces $\CH_H$ are separable.
846: \end{lemma}
847:
848: \begin{proof}
849: It suffices to find a norm $\|\cdot\|_\star$ which is stronger than $\|\cdot\|_H$ and such that
850: the closure of $\CC_0^\infty(\R_-)$ under $\|\cdot\|_\star$ is separable. One example of such a norm
851: is given by $\|w\|_\star= \sup_{t < 0} |t \dot w(t)|$.
852: \end{proof}
853:
854: Notice that it is crucial to define $\CH_H$ as the closure of $\CC_0^\infty$ under $\|\cdot\|_H$. If we defined
855: it simply as the space of all functions with finite $\|\cdot\|_H$-norm, it would not be separable. (Think of the
856: space of bounded continuous functions, versus the space of continuous functions vanishing at infinity.)
857:
858: In view of the representation \eref{e:repres}, we define the linear operator $\CD_H$ on functions $w\in\CC_0^\infty$ by
859: \begin{equ}[e:defD]
860: \sml(\CD_H w\smr)(t) = \alpha_H\int_{-\infty}^0 (-s)^{H-{1\over 2}}\sml(\dot w(s+t)- \dot w(s)\smr)\,ds\;,
861: \end{equ}
862: where $\alpha_H$ is as in \lem{lem:fbm}.
863: We have the following result:
864:
865: \begin{lemma}\label{lem:spaces}
866: Let $H \in (0,1)$ and let $\CH_H$ be as above. Then the operator $\CD_H$, formally defined by \eref{e:defD},
867: is continuous from
868: $\CH_H$ into $\CH_{1-H}$. Furthermore, the operator $\CD_H$ has a bounded inverse, given by the formula
869: \begin{equ}
870: \CD_H^{-1} = \gamma_H \CD_{1-H}\;,
871: \end{equ}
872: for some constant $\gamma_H$ satisfying $\gamma_H = \gamma_{1-H}$.
873: \end{lemma}
874:
875: \begin{remark}
876: The operator $\CD_H$ is actually (up to a multiplicative constant) a fractional integral of order $H-{1\over 2}$ which
877: is renormalised in such a way that one gets rid of the divergence at $-\infty$. It is therefore not surprising that
878: the inverse of $\CD_H$ is $\CD_{1-H}$.
879: \end{remark}
880:
881: \begin{proof}
882: For $H = {1\over 2}$, $\CD_H$ is the identity and there is nothing to prove. We therefore assume in the sequel that
883: $H \neq {1\over 2}$.
884:
885: We first show that $\CD_H$ is continuous from $\CH_H$ into $\CH_{1-H}$.
886: One can easily check that $\CD_H$ maps $\CC_0^\infty$ into the set of $\CC^\infty$ functions which
887: converge to a constant at $-\infty$. This set can be seen to belong to $\CH_{1-H}$ by a
888: simple cutoff argument, so it suffices
889: to show that $\|\CD_H w\|_{1-H} \le C\|w\|_H$ for $w\in\CC_0^\infty$.
890: Assume without loss of generality that $t>s$ and define $h = t-s$. We then have
891: \begin{equs}
892: \sml(\CD_H w\smr)(t)- \sml(\CD_H w\smr)(s) &= \alpha_H \int_{-\infty}^{s} \bigl(\Kern(t-r) - \Kern(s-r)\bigr)\,dw(r) \\
893: &\quad + \alpha_H\int_s^t\Kern(t-r)\,dw(r) \;.
894: \end{equs}
895: Splitting the integral and integrating by parts yields
896: \begin{equs}
897: \sml(\CD_H w\smr)(t) &- \sml(\CD_H w\smr)(s) = -\alpha_H\sml(H-{\textstyle{1\over 2}}\smr) \int_{s-h}^s (s-r)^{H-{3\over 2}}\sml(w(r)-w(s)\smr)\,dr \\
898: & +\alpha_H\sml(H-{\textstyle{1\over 2}}\smr) \int_{t-2h}^t (t-r)^{H-{3\over 2}}\sml(w(r)-w(t)\smr)\,dr \\
899: & +\alpha_H\sml(H-{\textstyle{1\over 2}}\smr)\int_{-\infty}^{s-h} \bigl((t-r)^{H-{3\over 2}} - (s-r)^{H-{3\over 2}}\bigr)\sml(w(r)-w(s)\smr)\,dr \\
900: & +\alpha_H (2h)^{H-{1\over 2}}\sml(w(t) - w(s)\smr)\\[0.5em]
901: & \equiv T_1 + T_2 + T_3 + T_4\;.
902: \end{equs}
903: We estimate each of these terms separately. For $T_1$, we have
904: \begin{equ}
905: |T_1| \le C \sml(1 + |s| + |t|\smr)^{1/2} \int_0^h r^{H-{3\over 2}+ {1-H\over 2}}\,dr \le C h^{H\over2} \sml(1 + |s| + |t|\smr)^{1/2}\;.
906: \end{equ}
907: The term $T_2$ is bounded by $C h^{H\over2} \sml(1 + |s| + |t|\smr)^{1/2}$ in a similar way. Concerning $T_3$, we bound it by
908: \begin{equs}
909: |T_3| &\le C \int_h^\infty \bigl(r^{H-{3\over 2}} - (h+r)^{H-{3\over 2}}\bigr)\sml(w(s-r) - w(s)\smr)\,dr \\
910: &\le C h \int_h^\infty r^{H-{5\over 2}}r^{1-H\over 2} \sml(1 + |s| + |r|\smr)^{1/2}\,dr \\
911: & \le C h^{H \over 2} \sml(1 + |s|\smr)^{1/2} + C h \int_h^\infty r^{{H\over 2}-2} \sml(h+r\smr)^{1/2}\,dr\\
912: & \le C h^{H \over 2} \sml(1 + |s| + h \smr)^{1/2} \le C h^{H \over 2} \sml(1 + |s| + |t|\smr)^{1/2}\;.
913: \end{equs}
914: The term $T_4$ is easily bounded by $C h^{H \over 2} \sml(1 + |s| + |t|\smr)^{1/2}$, using the fact that $w \in \CH_H$.
915: This shows that $\CD_H$ is bounded from $\CH_H$ to $\CH_{1-H}$.
916:
917: It remains to show that $\CD_H \circ \CD_{1-H}$
918: is a multiple of the identity. For this, notice that if $w \in \CC_0^\infty$, then one has in the
919: notations of \cite[pp.~94--95]{Frac} the following identities
920: \begin{equs}[2]
921: \bigl(\CD_H w\bigr)(t) &= -\alpha_H \Gamma(H + {\textstyle{1\over 2}}) \bigl(\bigl(I_+^{H-{1\over 2}} w\bigr)(t) - \bigl(I_+^{H-{1\over 2}} w\bigr)(0)\bigr)\;,&\quad H &> {\textstyle{1\over 2}}\;,\\
922: \bigl(\CD_H w\bigr)(t) &= -\alpha_H \Gamma(H + {\textstyle{1\over 2}}) \bigl(\bigl(D_+^{{1\over 2}-H} w\bigr)(t) - \bigl(D_+^{{1\over 2}-H} w\bigr)(0)\bigr)\;,&\quad H &< {\textstyle{1\over 2}}\;.
923: \end{equs}
924: Furthermore, \eref{e:defD} shows that $\CD_H w = 0$ if $w$ is a constant. The claim then follows immediately
925: from the fact that if $w \in \CC_0^\infty$ and $\alpha \in (0,1)$, one has $D_+^\alpha I_+^\alpha w = w$ and $I_+^\alpha D_+^\alpha w= w$ (see \cite[Thm.~2.4]{Frac}).
926: \end{proof}
927:
928:
929: Since we want to use the operators $\CD_H$ and $\CD_{1-H}$ to switch between Wiener processes and fractional
930: Brownian motions, it is crucial to show that the sample paths of the two-sided Wiener process belong to every
931: $\CH_H$ with probability $1$. Actually, what we show is that the Wiener measure can be constructed as
932: a Borel measure on $\CH_H$.
933:
934: \begin{lemma}\label{lem:wienerspace}
935: There exists a unique Gaussian measure $\Wien$ on $\CH_H$ which is such that the canonical process associated
936: to it is a time-reversed Brownian motion.
937: \end{lemma}
938:
939: \begin{proof}
940: We start by showing that the $\CH_H$-norm of the Wiener paths has bounded moments of all orders.
941: It follows from a generalisation of the Kolmogorov criterion \cite[Theorem 2.1]{Yor} that
942: \begin{equ}[e:locW]
943: \expect \biggl(\sup_{s,t \in [0,2]} {|w(s) - w(t)| \over |s-t|^{{1-H}\over 2}}\biggr)^p < \infty
944: \end{equ}
945: for all $p>0$.
946: Since the increments of $w$ are independent, this implies that, for every $\eps > 0$, there exists
947: a random variable $C_1$ such that
948: \begin{equ}[e:proplocw]
949: \sup_{|s-t| \le 1} {|w(s) - w(t)| \over |s-t|^{{1-H}\over 2} \sml(1 + |t| + |s|\smr)^{\eps}} < C_1\;,
950: \end{equ}
951: with probability $1$, and that all the moments of $C_1$ are bounded. We can therefore
952: safely assume in the sequel that $|t-s| > 1$. It follows
953: immediately from \eref{e:proplocw} and the triangle inequality that there exists a constant $C$ such that
954: \begin{equ}[e:proplocw2]
955: |w(s) - w(t)| \le C C_1 |t-s| \sml(1 + |t| + |s|\smr)^{\eps}\;,
956: \end{equ}
957: whenever $|t-s| > 1$. Furthermore, it follows from the time-inversion property of the Brownian motion,
958: combined with \eref{e:locW},
959: that $|w|$ does not grow much faster than $|t|^{1/2}$
960: for large values of $t$. In particular, for every $\eps' >0$, there exists a random variable $C_2$ such that
961: \begin{equ}[e:propglobw]
962: |w(t)| \le C_2\sml(1+|t|\smr)^{{1 \over 2} + \eps'}\;,\qquad \forall t\in\R\;,
963: \end{equ}
964: and that all the moments of $C_2$ are bounded.
965: Combining \eref{e:proplocw2} and \eref{e:propglobw}, we get (for some other constant $C$)
966: \begin{equ}
967: |w(s) - w(t)| \le C C_1^{1-H\over 2} C_2^{1+H \over 2} |t-s|^{1-H\over 2} \sml(1+|s| + |t|\smr)^{{H+1 \over 4} + \eps{1-H\over 2} + \eps'{1+H\over 2}}\;.
968: \end{equ}
969: The claim follows by choosing for example $\eps = \eps' = (1-H)/4$.
970:
971: This is not quite enough, since we want the sample paths to belong to the closure of $\CC_0^\infty$
972: under the norm
973: $\|\cdot\|_H$. Define the function
974: \begin{equ}
975: (s,t) \mapsto \Gamma(s,t) = {\sml(1+|t| + |s| \smr)^{2} \over |t-s|} \;.
976: \end{equ}
977: By looking at the above proof, we see that we actually proved the stronger statement that
978: for every $H\in(0,1)$, one can find a $\gamma > 0$ such that
979: \begin{equ}
980: \|w\|_{H,\gamma} = \sup_{s,t} {\Gamma(s,t)^\gamma |w(s) - w(t)| \over |s-t|^{{1-H}\over 2} \sml(1 + |t| + |s|\smr)^{1\over 2}} < \infty
981: \end{equ}
982: with probability $1$.
983: Let us call $\CH_{H,\gamma}$ the
984: Banach space of functions with finite $\|\cdot\|_{H,\gamma}$-norm. We will show that one has the continuous inclusions:
985: \begin{equ}[e:inclSpaces]
986: \CH_{H,\gamma} \hookrightarrow \CH_H \hookrightarrow \CC(\R_-,\R)\;.
987: \end{equ}
988: Let us call $\tilde\Wien$ the usual time-reversed Wiener measure on $\CC(\R_-,\R)$ equipped
989: with the $\sigma$-field ${\cscr R}$ generated by the evaluation functions.
990: Since $\CH_{H,\gamma}$ is a measurable subset of $\CC(\R_-,\R)$ and $\tilde\Wien(\CH_{H,\gamma})=1$,
991: we can restrict $\tilde\Wien$ to a measure on $\CH_H$, equipped with the restriction $\tilde {\cscr R}$
992: of ${\cscr R}$. It remains to show that $\tilde {\cscr R}$ is equal to the Borel $\sigma$-field ${\cscr B}$ on $\CH_H$.
993: This follows from the fact that the evaluation functions are ${\cscr B}$-measurable (since they are actually continuous)
994: and that a countable number of function evaluations suffices to determine the $\|\cdot\|_{H}$-norm of a function.
995: The proof of \lem{lem:wienerspace} is thus complete if we show \eref{e:inclSpaces}.
996:
997: Notice first that the function $\Gamma(s,t)$ becomes large
998: when $|t-s|$ is small or when either $|t|$ or $|s|$ are large, more precisely
999: we have
1000: \begin{equ}[e:boundGamma]
1001: \Gamma(s,t) > \max\{|s|,\,|t|,\,|t-s|^{-1}\}\;.
1002: \end{equ}
1003: Therefore, functions $w \in \CH_{H,\gamma}$
1004: are actually more regular and have better growth properties than what is needed to
1005: have finite $\|\cdot\|_H$-norm. Given $w$ with $\|w\|_{H,\gamma} < \infty$ and
1006: any $\eps > 0$, we will construct a function
1007: $\tilde w \in \CC_0^\infty$ such that $\|w - \tilde w\|_H < \eps$. Take two $\CC^\infty$ functions
1008: $\phi_1$ and $\phi_2$ with the following shape:
1009: \begin{equ}
1010: \mhpastefig[3/4]{Cutoff}
1011: \end{equ}
1012: Furthermore, we choose them such that:
1013: \begin{equ}
1014: \int_{\R_-} \phi_1(s)\,ds = 1\;,\qquad \Bigl|{d\phi_2(t) \over dt}\Bigr| \le 2\;.
1015: \end{equ}
1016: For two positive constants
1017: $r<1$ and $R>1$ to be chosen later, we define
1018: \begin{equ}
1019: \tilde w(t) = \phi_2(t/R)\int_{\R_-} w(t+s){\phi_1(s/r)\over r}\,ds\;.
1020: \end{equ}
1021: \ie we smoothen out $w$ at length scales smaller than $r$ and we cut it off at
1022: distances bigger than $R$. A straightforward estimate shows that there exists a constant $C$ such
1023: that
1024: \begin{equ}
1025: \|\tilde w\|_{H,\gamma} \le C\|w\|_{H,\gamma}\;,
1026: \end{equ}
1027: independently of $r<1/4$ and $R>1$.
1028: For $\delta > 0$ to be chosen later, we then divide the
1029: quadrant $K = \{(t,s)\,|\,t,s < 0\}$ into three regions:
1030:
1031: \noindent\hspace{1.5cm}
1032: \mhpastefig[4/5]{Regions}
1033: \begin{minipage}[b]{8cm}
1034: \begin{equs}
1035: K_1 &= \{(t,s)\,|\, |t| + |s| \ge R\} \cap K\;,\\
1036: K_2 &= \{(t,s)\,|\, |t-s| \le \delta\} \cap K \setminus K_1\;,\\
1037: K_3 &= K \setminus (K_1 \cup K_2)\;.
1038: \end{equs}
1039: \end{minipage}
1040: We then bound $\|w-\tilde w\|_H$ by
1041: \begin{equs}
1042: \|w - \tilde w\|_H &\le \sup_{(s,t) \in K_1\cup K_2}{C\|w\|_{H,\gamma} \over \Gamma(t,s)^\gamma}
1043: + \sup_{(s,t)\in K_3} {|w(s)-\tilde w(s)| + |w(t)-\tilde w(t)| \over |t-s|^{1-H\over 2}(1+|t|+|s|)^{1\over 2}}\\
1044: &\le C\sml(\delta^\gamma + R^{-\gamma}\smr) \|w\|_{H,\gamma} + 2\delta^{H-1\over 2} \sup_{0<t<R} |w(t)-\tilde w(t)|\;.
1045: \end{equs}
1046: By choosing $\delta$ small enough and $R$ large enough, the first term can be made arbitrarily small. One can then
1047: choose $r$ small enough to make the second term arbitrarily small as well.
1048: This shows that \eref{e:inclSpaces} holds and therefore the proof of \lem{lem:wienerspace} is complete.
1049: \end{proof}
1050:
1051: \subsection{Definition of the SDS}
1052:
1053: The results shown so far in this section are sufficient to construct the required SDS. We start by considering the pathwise
1054: solutions to \eref{e:mainequ}. Given a time $T>0$, an initial condition $x\in \R^n$, and a noise $b\in \CC_0([0,T],\R^n)$, we
1055: look for a function $\Phi_T(x,b) \in \CC([0,T],\R^n)$ satisfying
1056: \begin{equ}[e:defSol]
1057: \Phi_T(x,b)(t) = \sigma b(t) + x + \int_0^t f \sml(\Phi_T(x,b)(s)\smr)\,ds\;.
1058: \end{equ}
1059: We have the following standard result:
1060: \begin{lemma}\label{lem:existSol}
1061: Let $f\from\R^n\to\R^n$ satisfy assumptions \textbf{A1} and \textbf{A2}. Then, there exists a unique map
1062: $\Phi_T\from \R^n\times \CC([0,T],\R^n) \to \CC([0,T],\R^n)$ satisfying \eref{e:defSol}. Furthermore, $\Phi_T$
1063: is locally Lipschitz continuous.
1064: \end{lemma}
1065: \begin{proof}
1066: The local (\ie small $T$) existence and uniqueness of continuous solutions to \eref{e:defSol} follows from a standard contraction
1067: argument. In order to show the global existence and the local Lipschitz property, fix $x$, $b$, an $T$, and define
1068: $y(t) = x + \sigma b(t)$. Define $z(t)$ as the solution to the differential equation
1069: \begin{equ}[e:defz]
1070: \dot z(t) = f(z(t) + y(t))\;,\quad z(0) = 0\;.
1071: \end{equ}
1072: Writing down
1073: the differential equation satisfied by $\|z(t)\|^2$ and using \textbf{A1} and \textbf{A2}, one sees that \eref{e:defz} possesses
1074: a (unique) solution up to time $T$. One can then set $\Phi_T(x,b)(t) = z(t) + y(t)$ and check that it satisfies \eref{e:defSol}.
1075: The local Lipschitz property of $\Phi_T$ then immediately follows from the local Lipschitz property of $f$.
1076: \end{proof}
1077:
1078:
1079: We now define the stationary
1080: noise process. For this, we define
1081: $\theta_t\from\CH_H\to\CH_H$ by
1082: \begin{equ}
1083: \sml(\theta_t w\smr)(s) = w(s-t)-w(-t)\;.
1084: \end{equ}
1085: In order to construct the transition semigroup $\CP_t$, we define first $\tilde \CH_H$ like $\CH_H$, but with
1086: arguments in $\R_+$ instead of $\R_-$, and we write $\tilde \Wien$ for the Wiener measure on $\tilde \CH_H$,
1087: as constructed in \lem{lem:wienerspace} above. Define the function $P_t\from \CH_H \times \tilde \CH_H \to \CH_H$
1088: by
1089: \begin{equ}[e:defTrans]
1090: \sml(P_t(w,\tilde w)\smr)(s) = \cases{\tilde w(t+s) - \tilde w(t) & for $s > -t$,\cr
1091: w(t+s) - \tilde w(t) & for $s \le -t$,}
1092: \end{equ}
1093: and set $\CP_t(w,\cdot\,) = P_t(w,\cdot\,)^*\tilde \Wien$. This construction can be visualised by the following picture:
1094: \begin{equ}
1095: \mhpastefig[2/3]{Evol}
1096: \end{equ}
1097: One then has the following.
1098:
1099: \begin{lemma}\label{lem:Noise}
1100: The quadruple $\sml(\CH_H, \{\CP_t\}_{t\ge 0}, \Wien, \{\theta_t\}_{t\ge 0}\smr)$ is a stationary noise process.
1101: \end{lemma}
1102: \begin{proof}
1103: We already know from \lem{lem:separ} that $\CH_H$ is Polish. Furthermore, one has $\theta_t \circ P_t(w,\cdot\,) = w$, so
1104: it remains to show that $\CP_t$ is a Feller transition semigroup with $\Wien$ as its unique invariant measure.
1105: It is straightforward to check that it is a transition semigroup and the Feller property follows from the continuity of $P_t(w,\tilde w)$
1106: with respect to $w$. By the definition \eref{e:defTrans} and the time-reversal invariance of the Wiener process, every
1107: invariant measure for $\{\CP_t\}_{t\ge 0}$ must have its finite-dimensional distributions coincide with those of $\Wien$. Since the Borel
1108: $\sigma$-field on $\CH_H$ is generated by the evaluation functions, this shows that $\Wien$ is the only invariant
1109: measure.
1110: \end{proof}
1111:
1112: We now construct a SDS over $n$ copies of the above noise process. With a slight abuse of notation, we denote
1113: that noise process by $\sml(\Noise, \{\CP_t\}_{t\ge 0}, \Wien, \{\theta_t\}_{t\ge 0}\smr)$. We define the (continuous)
1114: shift operator $R_T\from\CC \sml((-\infty,0],\R^n\smr)\to \CC_0\sml([0,T],\R^n\smr)$ by
1115: $\sml(R_T b\smr)(t) = b(t-T)-b(-T)$ and set
1116: \begin{equa}[e:defPhi]
1117: \phi \from \R_+\times \R^n \times \Noise &\to \R^n\\
1118: (t,x,w) &\mapsto \Phi_t\bigl(x, R_t \CD_H w\bigr)(t)\;.
1119: \end{equa}
1120: From the above results, the following is straightforward:
1121: \begin{proposition}\label{prop:defSDS}
1122: The function $\phi$ of \eref{e:defPhi} defines a continuous SDS over the noise process
1123: $\sml(\Noise, \{\CP_t\}_{t\ge 0}, \Wien, \{\theta_t\}_{t\ge 0}\smr)$. Furthermore, for every generalised
1124: initial condition $\mu$, the process generated by $\phi$ from $\mu$ is a solution to \eref{e:mainequ}
1125: in the sense of Definition~\ref{def:solution}.
1126: \end{proposition}
1127: \begin{proof}
1128: The regularity properties of $\phi$ have already been shown in \lem{lem:existSol}. The cocycle property
1129: is an immediate consequence of the composition property for solutions of ODEs. The fact that the
1130: processes generated by $\phi$ are solutions to \eref{e:mainequ} is a direct consequence of \eref{e:defSol}, combined
1131: with \lem{lem:fbm}, the definition of $\CD_H$, and the fact that $\Wien$ is the Wiener measure.
1132: \end{proof}
1133:
1134: To conclude this section, we show that, thanks to the dissipativity condition imposed on the drift term $f$,
1135: the SDS defined above admits any power of the Euclidean norm on $\R^n$ as a Lyapunov function:
1136: %
1137: %
1138: \begin{proposition}\label{prop:Lyap}
1139: Let $\phi$ be the continuous SDS defined above and assume that \textbf{A1} and \textbf{A2} hold. Then,
1140: for every $p \ge 2$, the map $x \mapsto \|x\|^p$ is a Lyapunov function for $\phi$.
1141: \end{proposition}
1142: %
1143: %
1144: \begin{proof}
1145: Fix $p\ge 2$ and let $\mu$ be an arbitrary generalised initial condition satisfying
1146: \begin{equ}
1147: \int_{\R^n} \|x\|^p\,\sml(\Pi_{\R^n}^*\mu\smr)(dx) < \infty\;.
1148: \end{equ}
1149: Let $\tilde \phi$ be the continuous SDS
1150: associated by \prop{prop:defSDS} to the equation
1151: \begin{equ}[e:defy]
1152: dy(t) = - y\,dt + \sigma\,dB_H(t)\;.
1153: \end{equ}
1154: Notice that both $\phi$ and $\tilde\phi$ are defined over the same stationary noise process.
1155:
1156: We define $x_t$ as the process generated by $\phi$ from $\mu$ and $y_t$ as the process generated by $\tilde\phi$
1157: from $\delta_0 \times \Wien$ (in other words $y_0 = 0$). Since both SDS are defined over the same
1158: stationary noise process, $x_t$ and $y_t$ are defined
1159: over the same probability space.
1160: The process $y_t$ is obviously Gaussian, and a direct
1161: (but lengthy) calculation shows that its variance is given by:
1162: \begin{equ}
1163: \expect \|y_t\|^2 = 2H \tr \sml(\sigma \sigma^*\sml)\, e^{- t} \int_0^t s^{2H-1}\cosh(t-s)\,ds\;,
1164: \end{equ}
1165: In particular, one has for all $t$:
1166: \begin{equ}[e:boundy]
1167: \expect \|y_t\|^2 \le 2H \tr \sml(\sigma \sigma^*\sml)\, \int_0^\infty s^{2H-1}e^{-s}\,ds = \Gamma(2H+1)\tr \sml(\sigma \sigma^*\sml) \equiv C_\infty\;.
1168: \end{equ}
1169: Now define $z_t = x_t - y_t$. The process $z_t$ is seen to satisfy the random differential equation given by
1170: \begin{equ}
1171: {dz_t \over dt} = f\sml(z_t + y_t\smr) + y_t\;,\qquad z_0 = x_0\;.
1172: \end{equ}
1173: Furthermore, one has the following equation for $\|z_t\|^2$:
1174: \begin{equ}
1175: {d\|z_t\|^2 \over dt} = 2\scal{z_t, f\sml(z_t + y_t\smr)} + 2\scal{z_t,y_t}\;.
1176: \end{equ}
1177: Using \textbf{A2--A3} and the Cauchy-Schwarz inequality, we can estimate the right-hand side of this expression by:
1178: \begin{equ}[e:estz]
1179: {d\|z_t\|^2 \over dt} \le 2\C1 - 2\C2\|z_t\|^2 + 2 \scal{z_t,y_t + f(y_t)} \le -2\C2\|z_t\|^2 + \tilde C\sml(1+\|y_t\|^2\smr)^N\;,
1180: \end{equ}
1181: for some constant $\tilde C$. Therefore
1182: \begin{equ}
1183: \|z_t\|^2 \le e^{-2\C2 t}\|x_0\|^2 + \tilde C\int_0^t e^{-2\C2(t-s)} \sml(1+\|y_s\|^2\smr)^N\, ds\;.
1184: \end{equ}
1185: It follows immediately from \eref{e:boundy} and the fact that $y_s$ is Gaussian with bounded covariance \eref{e:boundy}
1186: that there exists a constant $C_p$ such that
1187: \begin{equ}
1188: \expect \|z_t\|^p \le C_p e^{-p\C2 t} \expect \|x_0\|^p + C_p\;,
1189: \end{equ}
1190: for all times $t>0$. Therefore \eref{e:condLyap} holds and the proof of \prop{prop:Lyap} is complete.
1191: \end{proof}
1192:
1193: \section{Coupling construction}
1194: \label{sec:coupl}
1195:
1196: We do now have the necessary formalism to study the long-time behaviour of the SDS $\phi$ we constructed
1197: from \eref{e:mainequ}. The main tool that will allow us to do that is the notion of self-coupling for stochastic dynamical
1198: systems.
1199:
1200: \subsection{Self-coupling of SDS}
1201: \label{sec:selfcoupl}
1202:
1203: The main goal of this paper is to show that the asymptotic behaviour of the solutions of \eref{e:mainequ} does not
1204: depend on its initial condition. This will then imply that the dynamics converges to a stationary state (in a suitable sense).
1205: We therefore look for a suitable way of comparing solutions to \eref{e:mainequ}. In general, two solutions starting from
1206: different initial points in $\R^n$ and driven with the same realisation of the noise $B_H$ have no reason of getting close
1207: to each other as time goes by. Condition \textbf{A1} indeed only ensures that they will tend to approach each other
1208: as long as they are sufficiently far apart. This is reasonable, since by comparing only solutions driven by the same
1209: realisation of the noise process, one completely forgets about the randomness of the system and the ``blurring'' this
1210: randomness induces.
1211:
1212: It is therefore important to compare probability measures (for example on path\-space) induced by the solutions rather
1213: than the solution themselves. More precisely, given a SDS $\phi$ and two generalised initial conditions $\mu$ and $\nu$,
1214: we want to compare the measures $\Evol \CQ_t\mu$ and $\Evol \CQ_t\nu$ as $t$ goes to infinity. The distance we will
1215: work with is the total variation distance, henceforth denoted by $\|\cdot\|_\TV$. We will actually use the following useful
1216: representation of the total variation distance. Let $\Omega$ be a measurable space and let $\prob_1$ and $\prob_2$ be two
1217: probability measures on $\Omega$. We denote by $C(\prob_1,\prob_2)$ the set of all probability measures on $\Omega\times\Omega$
1218: which are such that their marginals on the two components are equal to $\prob_1$ and $\prob_2$ respectively. Let furthermore
1219: $\Delta\subset \Omega\times\Omega$ denote the diagonal, \ie the set of elements of the form $(\omega,\omega)$. We then
1220: have
1221: \begin{equ}[e:charTV]
1222: \|\prob_1-\prob_2\|_\TV = 2 - \sup_{\prob \in C(\prob_1,\prob_2)} 2\prob(\Delta)\;.
1223: \end{equ}
1224: Elements of $C(\prob_1,\prob_2)$ will be referred to as {\em couplings} between $\prob_1$ and $\prob_2$. This leads naturally
1225: to the following definition:
1226:
1227: \begin{definition}\label{def:selfcoupl}
1228: Let $\phi$ be a SDS with state space $\State$ and let $\Meas$ be the associated space of generalised initial conditions. A {\em self-coupling}
1229: for $\phi$ is a
1230: measurable map $(\mu,\nu) \mapsto \Evol(\mu,\nu)$ from $\Meas\times\Meas$ into $\CD(\R_+,\State) \times \CD(\R_+,\State)$,
1231: with the property that for every pair $(\mu,\nu)$, $\Evol(\mu,\nu)$ is a coupling for $\Evol\mu$ and $\Evol\nu$.
1232: \end{definition}
1233:
1234: Define the shift map $\Sigma_t \from \CD(\R_+,\State)\to\CD(\R_+,\State)$ by
1235: \begin{equ}
1236: \sml(\Sigma_t x\smr)(s) = x(t+s)\;.
1237: \end{equ}
1238: It follows immediately from the cocycle property and the stationarity of the noise process
1239: that $\Evol \CQ_t\mu = \Sigma_t^*\Evol\mu$. Therefore, the
1240: measure $\Sigma_t^*\Evol(\mu,\nu)$ is a coupling for $\Evol \CQ_t\mu$ and $\Evol \CQ_t\nu$ (which is in general
1241: different from the coupling $\Evol(\CQ_t\mu,\CQ_t\nu)$). Our aim in the remainder of this paper is to construct a self-coupling
1242: $\Evol(\mu,\nu)$ for the SDS associated to \eref{e:mainequ} which has the property that
1243: \begin{equ}
1244: \lim_{t\to\infty}\sml(\Sigma_t^*\Evol(\mu,\nu)\smr)(\Delta) = 1\;,
1245: \end{equ}
1246: where $\Delta$ denotes as before the diagonal of the space $\CD(\R_+,\State) \times \CD(\R_+,\State)$.
1247: We will then use the inequality
1248: \begin{equ}[e:TVprop]
1249: \|\Evol \CQ_t \mu-\Evol\CQ_t \nu \|_\TV \le 2 - 2\sml(\Sigma_t^*\Evol(\mu,\nu)\smr)(\Delta)\;,
1250: \end{equ}
1251: to deduce the uniqueness of the stationary state for \eref{e:mainequ}.
1252:
1253: In the remainder of the paper, the general way of constructing such a self-coupling will be the following. First, we fix a Polish space
1254: $\CA$ that contains some auxiliary information on the dynamics of the coupled process we want to keep track of. We also define a
1255: ``future'' noise space $\Noise_+$ to be equal to $\tilde \CH_H^n$, where $\tilde \CH_H$ is as in \eref{e:defTrans}. There is a natural
1256: continuous time-shift operator on $\R \times \Noise \times \Noise_+$ defined for $t>0$ by
1257: \begin{equ}[e:timeShift]
1258: (s,w,\tilde w) \mapsto (s-t,P_t(w,\tilde w), S_t \tilde w)\;,\qquad \sml(S_t \tilde w\smr)(r) = \tilde w(r+t) - \tilde w(t)\;,
1259: \end{equ}
1260: where $P_t$ was defined in \eref{e:defTrans}. We then construct a (measurable) map
1261: \begin{equa}[e:notCoupl]
1262: \Coupl \from\State^2\times\Noise^2\times\CA &\to \R \times \pMeas(\CA \times \Noise_+^2)\;, \\
1263: (x,y,w_x,w_y,a) &\mapsto \sml(T(x,y,w_x,w_y,a), \Wien_2(x,y,w_x,w_y,a)\smr)\;,
1264: \end{equa}
1265: with the properties that, for all $(x,y,w_x,w_y,a)$,
1266: \begin{claim}
1267: \item[\maketag{C1}{e:C1}] The time $T(x,y,w_x,w_y,a)$ is positive and greater than $1$.
1268: \item[\maketag{C2}{e:C2}] The marginals of $\Wien_2(x,y,w_x,w_y,a)$ onto the two copies of $\Noise_+$
1269: are both equal to the Wiener measure $\Wien$.
1270: \end{claim}
1271: We call the map $\Coupl$ the ``coupling map'', since it yields a natural way of constructing a self-coupling for the SDS $\phi$.
1272: The remainder of this subsection explains how to achieve this.
1273:
1274: Given the map $\Coupl$, we can construct a Markov process on the augmented space $\BigSpace = \State^2\times \Noise^2 \times \R_+
1275: \times \CA\times \Noise_+^2 $ in the following way. As long as the component $\tau\in\R_+$ is positive, we just time-shift the
1276: elements in $\Noise^2 \times \Noise_+^2 \times \R_+$ according to \eref{e:timeShift} and we evolve in $\CX^2$ by solving
1277: \eref{e:mainequ}. As soon as $\tau$ becomes $0$, we redraw the future of the noise up to time $T(x,y,a)$ according to the
1278: distribution $\Wien_2$, which may at the same time modify the information stored in $\CA$.
1279:
1280: To shorten notations, we denote elements of $\BigSpace$ by
1281: \begin{equ}
1282: X = (x,y,w_x,w_y,\tau,a,\tilde w_x, \tilde w_y)\;.
1283: \end{equ}
1284: With this notation, the transition function $\tilde \CQ_t$ for the process we just described is defined by:
1285: \begin{claim}
1286: \item[$\bullet$] For $t < \tau$, we define $\tilde \CQ_t(X;\cdot\,)$ by
1287: \begin{equa}
1288: \tilde \CQ_t(X;\cdot\,) = &\delta_{\phi_t(x,P_t(w_x,\tilde w_x))} \times \delta_{\phi_t(y,P_t(w_y,\tilde w_y))} \times \delta_{P_t(w_x,\tilde w_x)} \\
1289: &\quad \times \delta_{P_t(w_y,\tilde w_y)} \times \delta_{\tau-t} \times \delta_{a} \times \delta_{S_t \tilde w_x}\times \delta_{S_t \tilde w_y}\;.
1290: \end{equa}
1291: \item[$\bullet$] For $t = \tau$, we define $\tilde \CQ_t(X;\cdot\,)$ by
1292: \begin{equa}[e:defQtjump]
1293: \tilde \CQ_t(X;\cdot\,) = &\delta_{\phi_t(x,P_t(w_x,\tilde w_x))} \times \delta_{\phi_t(y,P_t(w_y,\tilde w_y))} \times \delta_{P_t(w_x,\tilde w_x)} \\
1294: &\quad\times \delta_{P_t(w_y,\tilde w_y)} \times \delta_{T(x,y,P_t(w_x,\tilde w_x),P_t(w_y,\tilde w_y),a)} \\
1295: &\quad\times \Wien_2(x,y,P_t(w_x,\tilde w_x),P_t(w_y,\tilde w_y),a) \;.
1296: \end{equa}
1297: \item[$\bullet$] For $t > \tau$, we define $\tilde Q_t$ by imposing that the Chapman-Kolmogorov equations hold.
1298: Since we assumed that $T(x,y,w_x,w_y,a)$ is always greater than $1$, this procedure is well-defined.
1299: \end{claim}
1300: We now construct an initial condition for this process, given two generalised initial conditions $\mu_1$ and $\mu_2$ for $\phi$.
1301: We do this in such a way that, in the beginning, the noise component of our process lives on the diagonal of the space $\Noise^2$. In other
1302: words, the two copies of the two-sided fBm driving our coupled system have the same past. This
1303: is possible since the marginals of $\mu_1$ and $\mu_2$ on $\Noise$ coincide. Concerning the components of the initial condition in
1304: $\R_+\times\CA\times\Noise_+^2$, we just draw them according to the map $\Coupl$, with some distinguished element $a_0\in\CA$.
1305:
1306: We call $\Evol_0(\mu_1,\mu_2)$ the measure on $\BigSpace$ constructed by this procedure. Consider a cylindrical subset of $\BigSpace$
1307: of the form
1308: \begin{equ}
1309: X = X_1\times X_2 \times W_1 \times W_2 \times F\;,
1310: \end{equ}
1311: where $F$ is a measurable subset of $\R_+ \times \CA \times \Noise_+^2$.
1312: We make use of the disintegration $w \mapsto \mu_i^w$, yielding formally $\mu_i(dx,dw) = \mu_i^w(dx)\,\Wien(dw)$, and we
1313: define $\Evol_0(\mu_1,\mu_2)$ by
1314: \begin{equs}
1315: \Evol_0(\mu_1,\mu_2)(X) = \int_{W_1\cap W_2} \int_{X_1}\int_{X_2}
1316: & \sml(\delta_{T(x_1,x_2,w,w,a_0)} \times \Wien_2(x_1,x_2,w,w,a_0)\smr)(F) \\
1317: &\qquad \mu_2^w(dx_2)\, \mu_1^w(dx_1)\,\Wien(dw)\;.\label{e:definit}
1318: \end{equs}
1319: With this definition, we finally construct the self-coupling $\Evol(\mu_1,\mu_2)$ of $\phi$ corresponding to the function $\Coupl$
1320: as the marginal on $\CC(\R_+,\State) \times \CC(\R_+,\State)$ of the process generated by the initial condition $\Evol_0(\mu_1,\mu_2)$
1321: evolving under the semigroup given by $\tilde \CQ_t$. Condition \eref{e:C2} ensures that this is indeed a coupling for $\Evol\mu_1$ and
1322: $\Evol\mu_2$.
1323:
1324: The following subsection gives an overview of the way the coupling function $\Coupl$ is constructed.
1325:
1326: \subsection{Construction of the coupling function}
1327: \label{sec:constrcoupl}
1328:
1329: Let us consider that the initial conditions $\mu_1$ and $\mu_2$ are fixed once and for all and denote by $x_t$ and $y_t$
1330: the two $\State$-valued processes obtained by considering the marginals of $\Evol(\mu_1,\mu_2)$
1331: on its two $\State$ components. Define the random (but not stopping) time $\tau_\infty$ by
1332: \begin{equ}
1333: \tau_\infty = \inf \bigl\{t>0\,|\, \text{$x_s = y_s$ for all $s>t$}\bigr\}\;.
1334: \end{equ}
1335: Our aim is to find a space $\CA$ and a function $\Coupl$ satisfying \eref{e:C1} and \eref{e:C2} such that
1336: the processes $x_t$ and $y_t$ eventually meet and stay together for all times, \ie such that
1337: $\lim_{T\to \infty} \prob(\tau_\infty < T) = 1$. If the noise process driving the system
1338: was Markov, the ``stay together'' part of this statement would not be a problem, since it would suffice to start driving $x_t$
1339: and $y_t$
1340: with identical realisations of the noise as soon as they meet. Since the fBm is not Markov, it is possible to make the future
1341: realisations of two copies coincide with probability $1$ only if the past realisations also coincide.
1342: If the past realisations do not coincide for
1343: some time, we interpret this as introducing a ``cost'' into the system, which we need to master
1344: (this notion of cost will be made precise in Definition~\ref{def:cost} below). Fortunately, the memory
1345: of past events becomes smaller and smaller as time goes by, which can be interpreted as a natural tendency of the
1346: cost to decrease. This way of interpreting our system leads to the following algorithm that should be implemented by the
1347: coupling function $\Coupl$.
1348: \begin{equ}[e:algo]
1349: \mhpastefig[7/8]{Diagram}
1350: \end{equ}
1351: The precise meaning of the statements appearing in this diagram will be made clear in the sequel, but the general
1352: idea of the construction should be clear by now. One step in \eref{e:algo} corresponds to the time between two
1353: jumps of the $\tau$-component of the coupled process. Our aim is to construct the coupling function $\Coupl$ in such
1354: a way that, with probability $1$, there is a time after which step \step{2} always succeeds. This time is then precisely the
1355: random time $\tau_\infty$ we want to estimate.
1356:
1357: It is clear from what has just been exposed that
1358: we will actually never need to consider the continuous-time process on the space $\BigSpace$ given by the
1359: self-coupling described in the previous section, but it is sufficient to describe what happens at the beginning
1360: of each step in \eref{e:algo}. We will therefore only consider the discrete-time dynamic obtained by sampling
1361: the continuous-time system just before each step.
1362: The discrete-time dynamic will
1363: take place on the space $\CZ = \bigl(\State^2 \times \Noise^2 \times \CA\bigr) \times \R_+$ and we will
1364: denote its elements by
1365: \begin{equ}
1366: (Z,\tau)\;,\quad Z = (x,y,w_x,w_y,a)\;,\quad \tau \in \R_+\;.
1367: \end{equ}
1368: Since the time steps of the discrete dynamic are not equally spaced, the time $\tau$ is required to keep
1369: track of how much time really elapsed.
1370: The dynamic of the discrete process $(Z_n,\tau_n)$ on $\CZ$ is determined by the
1371: function $\Phi\from \R_+\times \CZ \times (\CA\times\Noise_+^2) \to \CZ$ given by
1372: \begin{equs}
1373: \Phi\bigl(t,(Z,\tau), (\tilde w_x, \tilde w_y, \tilde a)\bigr) &= \bigl( \phi_t(x,P_t(w_x,\tilde w_x)), \phi_t(y,P_t(w_y,\tilde w_y)), \\
1374: &\quad P_t(w_x,\tilde w_x),P_t(w_y,\tilde w_y), \tilde a, \tau + t\bigr)\;.
1375: \end{equs}
1376: (The notations are the same as in the definition of $\tilde \CQ_t$ above.)
1377: With this definition at hand, the transition function for the process $(Z_n,\tau_n)$ is given
1378: by
1379: \begin{equ}[e:transZn]
1380: {\cscr P} (Z,\tau) = \Phi\sml(T(Z),(Z,\tau), \cdot \smr)^*\Wien_2(Z)\;,
1381: \end{equ}
1382: where $T$ and $\Wien_2$ are defined in \eref{e:notCoupl}. Given two generalised
1383: initial conditions $\mu_1$ and $\mu_2$ for the original SDS, the initial condition $(Z_0,\tau_0)$ is
1384: constructed by choosing $\tau_0 = 0$ and by drawing $Z_0$ according to the measure
1385: \begin{equ}
1386: \mu_0(X) = \delta_{a_0}(A)\int_{W_1\cap W_2} \int_{X_1}\int_{X_2} \mu_2^w(dx_2)\, \mu_1^w(dx_1)\,\Wien(dw)\;,
1387: \end{equ}
1388: where $X$ is a cylindrical set of the form $X = X_1\times X_2 \times W_1\times W_2\times A$.
1389: It follows from the definitions \eref{e:defQtjump} and \eref{e:definit} that if we define $\tau_n$ as the $n$th jump of the process
1390: on $\BigSpace$ constructed above and $Z_n$ as (the component in $\State^2 \times \Noise^2 \times \CA$ of) its left-hand limit at $\tau_n$, the process we obtain
1391: is equal in law to the Markov chain that we just constructed.
1392:
1393: Before carrying further on with the construction of $\Coupl$, we make a few preliminary computations to see how
1394: changes in the past of the fBm affect its future. The formulae and estimates obtained in the next subsection are
1395: crucial for the construction of $\Coupl$ and for the obtention of the bounds that lead to Theorems~\ref{theo:smallH} and \ref{theo:largeH}. In particular, \prop{prop:pastfuture} is the main estimate that leads to the coherence of the
1396: coupling construction and to the bounds on the convergence rate towards the stationary state.
1397:
1398: \subsection{Influence of the past on the future}
1399:
1400: Let $w_x \in \CH_H$ and set $B_x = \CD_H w_x$. Consider furthermore two functions $g_w$ and
1401: $g_B$ satisfying
1402: \begin{equ}[e:condg]
1403: t \mapsto \int_0^t g_w(s)\,ds \in \CH_H\;,\qquad
1404: t \mapsto \int_0^t g_B(s)\,ds \in \CH_{1-H}\;,
1405: \end{equ}
1406: and define $B_y$ and $w_y$ by $B_y(0) = w_y(0) = 0$ and
1407: \begin{equ}[e:relBw]
1408: dB_y = dB_x + g_B\,dt\;,\qquad dw_y = dw_x + g_w\,dt\;.
1409: \end{equ}
1410: As an immediate consequence of the definition of $\CD_H$,
1411: the following relations between $g_w$ and $g_B$ will ensure that $B_y = \CD_H w_y$.
1412:
1413: \begin{lemma}\label{lem:formulas}
1414: Let $B_x$, $B_y$, $w_x$, $w_y$, $g_B$, and $g_w$ be as in \eref{e:condg}, \eref{e:relBw} and assume that
1415: $B_x = \CD_H w_x$ and $B_y = \CD_H w_y$. Then, $g_w$ and $g_B$ satisfy the following relation:
1416: \minilab{e:relG}
1417: \begin{equs}
1418: g_w(t) &= \alpha_H {{d \over dt} \int_{-\infty}^t (t-s)^{{1\over 2}-H}} g_B(s)\,ds\;, \label{e:relG1}\\
1419: g_B(t) &= \gamma_H \alpha_{1-H} {{d \over dt} \int_{-\infty}^t \Kern(t-s) g_w(s) \,ds} \;.\label{e:relG2}
1420: \end{equs}
1421: If $g_w(t) = 0$ for $t > t_0$, one has
1422: \minilab{e:relG}
1423: \begin{equ}[e:relG3]
1424: g_B(t) = \sml(H-{\textstyle{1\over 2}}\smr)\gamma_H \alpha_{1-H}\int_{-\infty}^{t_0} (t-s)^{H - {3\over 2}} g_w(s)\,ds \;,
1425: \end{equ}
1426: for $t \ge t_0$. Similarly, if $g_B(t) = 0$ for $t > t_0$, one has
1427: \minilab{e:relG}
1428: \begin{equ}[e:relG4]
1429: g_w(t) =\sml({\textstyle{1\over 2}}-H\smr)\alpha_{H} \int_{-\infty}^{t_0} (t-s)^{-H - {1\over 2}} g_B(s) \,ds \;,
1430: \end{equ}
1431: for $t \ge t_0$. If $g_w$ is differentiable for $t>t_0$ and $g_w(t) = 0$ for $t < t_0$, one has
1432: \minilab{e:relG}
1433: \begin{equ}[e:relG5]
1434: g_B(t) = {\gamma_H \alpha_{1-H} g_w(t_0) \over (t-t_0)^{{1\over 2}-H}} + \gamma_H \alpha_{1-H} \int_{t_0}^t {g_w'(s) \over (t-s)^{{1\over 2}-H}}\,ds \;,
1435: \end{equ}
1436: for $t \ge t_0$. Similarly, if $g_B$ is differentiable for $t>t_0$ and $g_B(t) = 0$ for $t < t_0$, one has
1437: \minilab{e:relG}
1438: \begin{equ}[e:relG6]
1439: g_w(t) = {\alpha_{H} g_B(t_0) \over (t-t_0)^{H-{1\over 2}}} + \alpha_{H} \int_{t_0}^t {g_B'(s) \over (t-s)^{H-{1\over 2}}}\,ds \;,
1440: \end{equ}
1441: for $t \ge t_0$.
1442: \end{lemma}
1443:
1444: \begin{proof}
1445: The claims \eref{e:relG1} and \eref{e:relG2} follow immediately from \eref{e:relBw}, using the linearity of $\CD_H$ and the inversion formula.
1446: The other claims are simply obtained by differentiating under the integral, see \cite{Frac} for a justification.
1447: \end{proof}
1448:
1449: We will be led in the sequel to consider the following situation, where $t_1$, $t_2$ and $g_1$ are assumed to be given:
1450: \begin{equ}[e:defg]
1451: \mhpastefig[6/7]{Timeline}
1452: \end{equ}
1453: In this picture, $g_w$ and $g_B$ are related by (\ref{e:relG1}--\ref{e:relG2}) as before. The boldfaced regions indicate
1454: that we consider the corresponding parts of $g_w$ or $g_B$ to be given. The dashed regions indicate that those
1455: parts of $g_w$ and $g_B$ are computed from the boldfaced regions by using the relations (\ref{e:relG1}--\ref{e:relG2}).
1456: The picture is coherent since the formulae (\ref{e:relG1}--\ref{e:relG2}) in both cases only use information about the past to compute the present.
1457: One should think of the interval $[0,t_1]$ as representing the time spent on steps \step{1} and \step{2} of the algorithm
1458: \eref{e:algo}. The interval $[t_1, t_2]$ corresponds to the waiting time, \ie step \step{3}.
1459: Let us first give an explicit formula for $g_2$ in terms of $g_1$:
1460:
1461: \begin{lemma}
1462: Consider the situation of \prop{prop:pastfuture}. Then, $g_2$ is given by
1463: \begin{equ}[e:formg2]
1464: g_2(t) = C\int_0^{t_1} { t^{{1\over 2}-H}(t_2-s)^{H-{1\over 2}} \over t + t_2 -s} g_1(s)\,ds\;,
1465: \end{equ}
1466: with a constant $C$ depending only on $H$.
1467: \end{lemma}
1468:
1469: \begin{proof}
1470: We extend $g_1(t)$ to the whole real line by setting it equal to $0$ outside of $[0,t_1]$.
1471: Using \lem{lem:formulas}, we see that, for some constant $C$ and for $t > t_2$,
1472: \begin{equs}
1473: g_2(t - t_2) &= C\int_0^{t_2} (t-s)^{-H-{1\over 2}}\,g_B(s)\,ds \\
1474: &= C\int_0^{t_2} (t-s)^{-H-{1\over 2}}{d \over ds} \int_0^{s}(s-r)^{H-{1\over 2}}g_1(r)\,dr\, ds \\
1475: &= C(t-t_2)^{-H-{1\over 2}} \int_0^{t_2} (t_2-r)^{H-{1\over 2}}g_1(r)\,dr \\
1476: &\quad- C\sml(H+{\textstyle{1\over 2}}\smr) \int_0^{t_2} (t-s)^{-H-{3\over 2}}\int_0^{s}(s-r)^{H-{1\over 2}} g_1(r)\,dr\, ds\\
1477: &\equiv C\int_0^{t_1} K(t,r) \, g_1(r)\,dr\;,
1478: \end{equs}
1479: where the integration stops at $t_1$ because $g_1$ is equal to $0$ for larger values of $t$.
1480: The kernel $K$ is given by
1481: \begin{equs}
1482: K(t,r) &= (t-t_2)^{-H-{1\over 2}} (t_2-r)^{H-{1\over 2}} - \sml(H+{\textstyle{1\over 2}}\smr) \int_r^{t_2}
1483: (t-s)^{-H-{3\over 2}}(s-r)^{H-{1\over 2}}\, ds \\
1484: &= {(t-t_2)^{-H-{1\over 2}}(t_2-r)^{H+{1\over 2}} \over t-r}\Bigl({t-r \over t_2 -r} -1\Bigr) \\
1485: &= {(t-t_2)^{{1\over 2}-H}(t_2-r)^{H-{1\over 2}} \over t-r}\;,
1486: \end{equs}
1487: and the claim follows.
1488: \end{proof}
1489:
1490: We give now estimates on $g_2$ in terms of $g_1$. To this end, given $\alpha > 0$,
1491: we introduce the following norm on functions $g\from\R_+\to \R^n$:
1492: \begin{equ}
1493: \|g\|_\alpha^2 = \int_0^\infty (1+ t)^{2\alpha} \|g(t)\|^2\,dt\;.
1494: \end{equ}
1495: The following proposition is essential to the coherence of our coupling construction:
1496:
1497: \begin{proposition}\label{prop:pastfuture}
1498: Let $t_2 > 2 t_1 > 0$, let $g_1\from[0,t_1] \to \R^n$ be a square integrable function, and define $g_2\from\R_+\to\R_+$ by
1499: \begin{equ}
1500: g_2(t) = \int_0^{t_1} { t^{{1\over 2}-H}(t_2-s)^{H-{1\over 2}} \over t + t_2 -s} \|g_1(s)\|\,ds\;.
1501: \end{equ}
1502: Then, for every $\alpha$ satisfying
1503: \begin{equ}
1504: 0 < \alpha < \min\{{\textstyle {1\over 2}}\,;\, H\}\;,
1505: \end{equ}
1506: there exists a constant $\kappa>0$ depending only on $\alpha$ and $H$ such that the estimate
1507: \begin{equ}[e:boundfuture]
1508: \|g_2\|_\alpha \le \kappa \Bigl|{t_2 \over t_1}\Bigr|^{\alpha -{1\over 2}} \|g_1\|_\alpha
1509: \end{equ}
1510: holds.
1511: \end{proposition}
1512:
1513: \begin{remark}
1514: The important features of this proposition are that the constant $\kappa$ \textit{does not} depend on $t_1$ or $t_2$ and
1515: that the exponent in \eref{e:boundfuture} is negative.
1516: \end{remark}
1517:
1518: \begin{proof}
1519: We define $r = t_2/t_1$ to shorten notations.
1520: Using \eref{e:formg2} and Cauchy-Schwarz, we then have
1521: \begin{equs}
1522: \|g_2(t)\| &\le C\|g_1\|_\alpha \sqrt{\int_0^{t_1} (1+ s)^{-2\alpha} {(rt_1-s)^{2H-1} t^{1-2H} \over \sml(t + rt_1 -s\smr)^2}\,ds } \\
1523: &= C\|g_1\|_\alpha {t_1}^{H-\alpha} t^{1-H-{1\over 2}} \sqrt{\int_0^{1} s^{-2\alpha} {(r-s)^{2H-1}\over \sml(t + rt_1 - t_1s\smr)^2}\,ds }\\
1524: &\le C \|g_1\|_\alpha {{t_1}^{H-\alpha} t^{{1\over 2}-H} r^{H-{1\over 2}} \over t + (r-1) t_1 }\;,
1525: \end{equs}
1526: where we made use of the assumptions that $2\alpha < 1$ and $r \ge 2$. Therefore, $\|g_2\|_\alpha$ is bounded by
1527: \begin{equs}
1528: \|g_2\|_\alpha &\le \kappa \|g_1\|_\alpha{t_1}^{H-\alpha} r^{H-{1\over 2}} \sqrt{\int_0^\infty {(1+ t)^{2\alpha} t^{1-2H} \over \sml(t + (r-1) t_1 \smr)^2}\,dt} \\
1529: &\le \kappa \|g_1\|_\alpha r^{\alpha - {1\over 2}} \sqrt{\int_0^\infty { t^{2\alpha} t^{1-2H} \over (t + 1)^2}\,dt} \;,
1530: \end{equs}
1531: for some constant $\kappa$, where the last inequality was obtained through the change of variables
1532: $t \mapsto (r-1)t_1 t$ and used the fact that $r \ge 2$. The convergence of the integral is obtained under the condition $\alpha < H$ which
1533: is verified by assumption,
1534: so the proof of \prop{prop:pastfuture} is complete.
1535: \end{proof}
1536:
1537: We will construct our coupling function $\Coupl$ in such a way that there always exist functions $g_w$ and $g_B$ satisfying
1538: \eref{e:condg} and \eref{e:relBw}, where $w_x$ and $w_y$ denote the noise components of our coupling process, and $B_x$ and $B_y$
1539: are obtained by applying the operator $\CD_H$ to them.
1540: We have now all the necessary ingredients for the construction of $\Coupl$.
1541:
1542: \section{Definition of the coupling function}
1543: \label{sec:defCoupl}
1544:
1545: Our coupling construction depends on a parameter $\alpha < \min\{{1\over 2},H\}$ which we fix once and
1546: for all. This parameter will then be tuned in Section~\ref{sec:mainproof}.
1547:
1548: First of all, we define the auxiliary space $\CA$:
1549: \begin{equ}[e:defA]
1550: \CA = \{0,1,2,3\} \times \N \times \N \times \R_+\;.
1551: \end{equ}
1552: Elements of $\CA$ will be denoted by
1553: \begin{equ}[e:notA]
1554: a = (S ,N, \tilde N, T_3)\;.
1555: \end{equ}
1556: The component $S$ denotes which step of \eref{e:algo} is going to be performed next (the value $0$ will be used only for the initial
1557: value $a_0$). The counter $N$ is incremented every time step \step{2} is performed and is reset to $0$
1558: every time another step is performed. The counter $\tilde N$ is incremented every time step \step{1} or step
1559: \step{2} fails. If steps \step{1} or \step{2} fail, the time $T_3$ contains the duration of the upcoming
1560: step \step{3}. We take
1561: \begin{equ}
1562: a_0 = (0,1,1,0)
1563: \end{equ}
1564: as initial condition for our coupling construction.
1565:
1566: Remember that the coupling function $\Coupl$ is a function from $\State^2\times \Noise^2 \times \CA$, representing the state of the system
1567: at the end of a step, into $\R\times \pMeas(\CA\times\Noise_+^2)$, representing the duration and the realisation of the noise
1568: for the next step. We now define $\Coupl$ for the four possible values of $S$.
1569:
1570: \subsection{Initial stage ({\bfseries\itshape S = 0})}
1571:
1572: Notice first that \textbf{A1} implies that
1573: \begin{equ}[e:consA1]
1574: {\scal{f(y)-f(x), y-x} \over \|y-x\|} \le \C4 - \C2 \|y-x\|\;,
1575: \end{equ}
1576: where we set $\C4 = \sqrt{\C1(\C2+\C3)}$.
1577:
1578: In the beginning, we just wait until the two copies of our process are within distance $1 + \sml({\C4 / \C2}\smr)$ of each other.
1579: If $x_t$ and $y_t$ satisfy \eref{e:mainequ} with the same realisation of the noise
1580: process $B_H$, and $\rho_t = y_t - x_t$, we have by for $\|\rho_t\|$ the differential inequality
1581: \begin{equ}
1582: {d \|\rho_t\| \over dt} = {\scal{f(y_t)-f(x_t), \rho_t}\over \|\rho_t\|} \le \C4 - \C2\|\rho_t\|\;,
1583: \end{equ}
1584: and therefore by Gronwall's lemma
1585: \begin{equ}
1586: \|\rho_t\| \le \|y_0 - x_0\| e^{-\C2 t} + {\C4\over\C2}\bigl(1-e^{-\C2 t}\bigr)\;.
1587: \end{equ}
1588: It is enough to wait for a time $t = \sml(\log \|y_0 - x_0\| \smr)/ \C2$ to ensure that
1589: $\|\rho_t\| \le 1 + \sml({\C4 / \C2}\smr)$, so we define the coupling function
1590: $\Coupl$ in this case by
1591: \begin{equ}[e:defCoupl0]
1592: T(Z,a_0) = \max\Bigl\{{\log \|y_0 - x_0\| \over \C2}\,,\, 1\Bigr\}\;,\qquad \Wien_2(Z,a_0) = \Delta^*\Wien\times \delta_{a'}\;,
1593: \end{equ}
1594: where the map $\Delta\from\Noise_+\to \Noise_+^2$ is defined by $\Delta(w) = (w,w)$
1595: and the element $a'$ is given by
1596: \begin{equ}
1597: a' = (1,0,0,0)\;.
1598: \end{equ}
1599: In other terms, we wait until the two copies of the process are close to each other, and then we
1600: proceed to step \step{1}.
1601:
1602: \subsection{Waiting stage ({\bfseries\itshape S = 3})}
1603:
1604: In this stage, both copies evolve with the same realisation of the underlying Wiener process.
1605: Using notations \eref{e:notA} and \eref{e:notCoupl}, we therefore define the coupling function
1606: $\Coupl$ in this case by
1607: \begin{equ}[e:defCoupl1]
1608: T(Z,a) = T_3\;,\qquad \Wien_2(Z,a) = \Delta^*\Wien\times \delta_{a'}\;,
1609: \end{equ}
1610: where the map $\Delta$ is defined as above
1611: and the element $a'$ is given by
1612: \begin{equ}
1613: a' = (1,N,\tilde N,0)\;.
1614: \end{equ}
1615: Notice that this definition is in accordance with \eref{e:algo}, \ie the counters $N$ and $\tilde N$
1616: remain unchanged, the dynamic evolves for a time $T_3$ with two identical realisations of the
1617: Wiener process (note that the realisations of the fBm driving the two copies of the system
1618: are in general different, since the {\em pasts} of the Wiener processes may differ), and then
1619: proceeds to step \step{1}.
1620:
1621: \subsection{Hitting stage ({\bfseries\itshape S = 1})}
1622:
1623: In this section, we construct and then analyse the map $\Coupl$ corresponding to the step \step{1},
1624: which is the most important
1625: step for our construction. We start with a few preliminary computations. Define $W_{1,1}$ as the space of
1626: almost everywhere differentiable functions $g$, such that the quantity
1627: \begin{equ}
1628: \|g\|_{1,1} = \int_0^1 \Bigl\|{dg_B(t) \over dt}\Bigr\|\,dt + \|g(0)\|\;,
1629: \end{equ}
1630: is finite.
1631:
1632: \begin{lemma}\label{lem:bigH}
1633: Let $g_B\from [0,1] \to \R^n$ be in $W_{1,1}$ and define $g_w$ by \eref{e:relG1} with
1634: $H\in \sml({\textstyle{1\over2}},1\smr)$. (The function $g_B$ is extended to $\R$ by setting
1635: it equal to $0$ outside of $[0,1]$ and $g_w$ is considered as a function
1636: from $\R_+$ to $\R^n$.) Then, for every $\alpha \in (0,H)$, there exists a constant $C$ such that
1637: \begin{equ}
1638: \|g_w\|_\alpha \le C\|g_B\|_{1,1}\;.
1639: \end{equ}
1640: \end{lemma}
1641:
1642: \begin{proof}
1643: We first bound the $\L^2$ norm of $g_w$ on the interval $[0,2]$.
1644: Using \eref{e:relG6}, we can bound $\|g_w(t)\|$ by
1645: \begin{equ}
1646: \|g_w(t)\| \le C \|g_B(0)\| t^{{1\over 2}-H} + C \int_0^t \|\dot g_B(s)\| (t-s)^{{1\over 2}-H}\,ds\;.
1647: \end{equ}
1648: Since $t^{{1\over 2}-H}$ is square integrable at the origin, it remains to bound the terms $I_1$ and $I_2$ given by
1649: \begin{equs}
1650: I_1 &= \int_0^2 \Bigl(\int_0^t (t-s)^{{1\over 2} -H} \|\dot g_B(s)\|\,ds \int_0^t (t-r)^{{1\over 2} -H} \|\dot g_B(r)\|\,dr \Bigr)\, dt\;,\\
1651: I_2 &= \|g_B(0)\| \int_0^2 t^{{1\over 2} -H} \int_0^t (t-s)^{{1\over 2} -H} \|\dot g_B(s)\|\,ds\, dt\;,
1652: \end{equs}
1653: We only show how to bound $I_1$, as $I_2$ can be bounded in a similar fashion. Writing $r\vee s = \max\{r,s\}$ one has
1654: \begin{equ}
1655: I_1 = \int_0^1 \int_0^1 \int_{r \vee s}^2 (t-s)^{{1\over 2} -H} (t-r)^{{1\over 2} -H}\, dt \|\dot g_B(s)\| \|\dot g_B(r)\|\,dr\,ds\;.
1656: \end{equ}
1657: Since
1658: \begin{equ}
1659: \int_{r \vee s}^2 (t-s)^{{1\over 2} -H} (t-r)^{{1\over 2} -H}\,dt \le \int_{r \vee s}^2 \sml(t-(r \vee s)\smr)^{1 - 2H}\,dt \le {2^{2-2H} \over 2-2H}\;,
1660: \end{equ}
1661: $I_1$ is bounded by $C \|g_B\|_{1,1}^2$.
1662:
1663: It remains to bound the large-time tail of $g_w$. For $t \ge 2$, one has, again by \lem{lem:formulas},
1664: \begin{equ}[e:largeT]
1665: \|g_w(t)\| \le (t-1)^{-H-{1\over 2}} \sup_{s \in [0,1]} \|g_B(s)\| \le C (t-1)^{-H-{1\over 2}} \|g_B\|_{1,1}\;.
1666: \end{equ}
1667: It follows from the definition that the $\|\cdot\|_\alpha$-norm of this function is bounded if $\alpha < H$.
1668: The proof of \lem{lem:bigH} is complete.
1669: \end{proof}
1670:
1671: In the case $H < {1\over 2}$, one has a similar result, but the regularity of $g_B$ can be weakened.
1672:
1673: \begin{lemma}\label{lem:smallH}
1674: Let $g_B\from [0,1] \to \R^n$ be a continuous function and define $g_w$ as in \lem{lem:bigH}, but with $H\in \sml(0,{\textstyle{1\over2}}\smr)$. Then, for every $\alpha \in (0,H)$, there exists a constant $C$ such that
1675: \begin{equ}
1676: \|g_w\|_\alpha \le C\sup_{t \in [0,1]} \|g_B(t)\|\;.
1677: \end{equ}
1678: \end{lemma}
1679:
1680: \begin{proof}
1681: Since $H < {1\over 2}$, one can move the derivative under the integral of the first equation in \lem{lem:formulas} to get
1682: \begin{equ}
1683: \|g_w(t)\| \le C \int_0^t (t-s)^{-H -{1\over 2}}\|g_B(s)\|\,ds \le C\sup_{t \in [0,1]} \|g_B(t)\|\;.
1684: \end{equ}
1685: This shows that the restriction of $g_w$ to $[0,2]$ is square integrable. The large-time tail can be bounded by \eref{e:largeT}
1686: as before.
1687: \end{proof}
1688:
1689: We already hinted several times towards the notion of a ``cost function'' that measures the difficulty of coupling
1690: the two copies of the process. This notion is now made precise. Denote by $Z = (x_0,y_0,w_x,w_y)$
1691: an element of $\State^2\times\Noise^2$ and assume that there exists a square integrable function
1692: $g_w\from\R_-\to\R^n$ such that
1693: \begin{equ}[e:defg_w]
1694: w_y(t) = w_x(t) + \int_t^0 g_w(s)\,ds\;,\qquad \forall\,t<0\;.
1695: \end{equ}
1696: In regard of \eref{e:formg2}, we introduce for $T>0$ the operator $\CR_T$ given by
1697: \begin{equ}
1698: \sml(\CR_T g\smr)(t) = C\int_{-\infty}^0 {t^{{1\over 2}-H} (T-s)^{H-{1\over 2}} \over t+T-s} \|g(s)\|\,ds\;,
1699: \end{equ}
1700: where $C$ is the constant appearing in \eref{e:formg2}.
1701: The cost is then defined as follows.
1702: \begin{definition}\label{def:cost}
1703: The cost function $\CK_\alpha \from \L^2(\R_-) \to [0,\infty]$ is defined by
1704: \begin{equ}[e:defcost]
1705: \CK_\alpha (g) = \sup_{T > 0} \|\CR_T g\|_\alpha + C_K\int_{-\infty}^0 (-s)^{H-{3\over 2}} \|g(s)\|\,ds\;,
1706: \end{equ}
1707: where, for convenience, we define $C_K = \sml| (2H-1)\gamma_H\alpha_{1-H}\smr|$.
1708: Given $Z$ as above, $\CK_\alpha(Z)$ is defined as $\CK_\alpha(g_w)$ if there
1709: exists a square integrable function $g_w$ satisfying \eref{e:defg_w} and as $\infty$ otherwise.
1710: \end{definition}
1711:
1712: \begin{remark}\label{rem:cost}
1713: The cost function $\CK_\alpha$ defined above has the important property that
1714: \begin{equ}[e:shiftcost]
1715: \CK_\alpha(\theta_t g) \le \CK_\alpha(g)\;,\quad \text{for all $t\ge 0$,}
1716: \end{equ}
1717: where the shifted function $\theta_t g$ is given by
1718: \begin{equ}
1719: \sml(\theta_t g\smr)(s) = \cases{g(s+t) & if $s < -t$,\cr 0&otherwise. }
1720: \end{equ}
1721: Furthermore, it is a norm, and thus satisfies the triangle inequality.
1722: \end{remark}
1723:
1724: \begin{remark}\label{rem:meancost}
1725: By \eref{e:formg2}, the first term in \eref{e:defcost} measures by how much the
1726: two realisations of the Wiener process have to differ in order to obtain identical increments
1727: for the associated fractional Brownian motions. By \eref{e:relG3},
1728: the second term in \eref{e:defcost} measures by how much the two realisations
1729: of the fBm differ if one lets the system evolve with two identical realisations of the Wiener process.
1730: \end{remark}
1731:
1732: We now turn to the construction of the process $(x_t,y_t)$ during step \step{1}. We will set up our coupling construction in such a way that, whenever step
1733: \step{1} is to be performed, the initial condition $Z$ is admissible in the following sense:
1734:
1735: \begin{definition}\label{def:defAdm}
1736: Let $\alpha$ satisfy $0<\alpha<\min\{{1\over 2};H\}$. We say that $Z = (x_0,y_0,w_x,w_y)$ is {\em admissible} if one has
1737: \begin{equ}[e:defAdm]
1738: \|x_0 - y_0\| \le 1+{1 + \C4 \over \C2}\;,
1739: \end{equ}
1740: (the constants $\C{i}$ are as in \textbf{A1} and in \eref{e:consA1}), and its cost satisfies $\CK_\alpha(Z) \le 1$.
1741: \end{definition}
1742:
1743:
1744: Denote now by $\Omega$ the space
1745: of continuous functions $\omega\from [0,1] \to \R^n$ which are the restriction to $[0,1]$ of an element of $\tilde \CH_H$.
1746: Our aim is construct two measures $\prob_Z^1$ and $\prob_Z^2$ on $\Omega \times \Omega$
1747: satisfying the following conditions:
1748: \begin{claim}
1749: \item[\textbf{B1}] The marginals of $\prob_Z^1+\prob_Z^2$ onto the two components $\Omega$ of the product space are both equal to the Wiener
1750: measure $\Wien$.
1751: \item[\textbf{B2}] Let ${\cscr B}_\kappa \subset \Omega\times\Omega$ denote the set of pairs $(\tilde w_x,\tilde w_y)$ such that there exists
1752: a function $g_w\from [0,1]\to\R^n$ satisfying
1753: \begin{equ}
1754: \tilde w_y(t) = \tilde w_x(t) + \int_0^t g_w(s)\,ds\;,\qquad \int_0^1 \|g_w(s)\|^2\,ds \le \kappa\;.
1755: \end{equ}
1756: Then, there exists a value of $\kappa$ such that, for every admissible initial condition $Z_0$, we have $\prob_Z^1({\cscr B}_\kappa) + \prob_Z^2({\cscr B}_\kappa) = 1$.
1757: \item[\textbf{B3}] Let $(x_t,y_t)$ be the process constructed by solving \eref{e:mainequ} with respective initial conditions $x_{0}$ and
1758: $y_{0}$, and with respective noise processes $P_t(w_x,\tilde w_x)$ and $P_t(w_y,\tilde w_y)$. Then, one has $x_1 = y_1$ for $\prob_Z^1$-almost every noise $(\tilde w_x, \tilde w_y)$. Furthermore, there exists a constant $\delta>0$ such that
1759: $\prob_Z^1(\Omega\times\Omega) \ge \delta$ for every admissible initial condition $Z$.
1760: \end{claim}
1761:
1762: \begin{remark}\label{rem:}
1763: Both measures $\prob_Z^1$ and $\prob_Z^2$ can easily be extended to measures on $\Noise_+^2$ in such a way
1764: that \textbf{B1} holds. Since the dynamic constructed from the coupling function $\Coupl$ will not depend on this
1765: extension, we just choose one arbitrarily and denote again by $\prob_Z^1$ and $\prob_Z^2$ the
1766: corresponding measures on $\Noise_+^2$.
1767: \end{remark}
1768:
1769: Given $\prob_Z^1$ and $\prob_Z^2$, we construct the coupling function $\Coupl$ in the following way, using notations
1770: \eref{e:notA} and \eref{e:notCoupl}:
1771: \begin{equ}[e:defCoupl3]
1772: T(Z,a) = 1\;,\qquad \Wien_2(Z,a) = \prob_Z^1 \times \delta_{a_1} + \prob_Z^2 \times \delta_{a_2}\;,
1773: \end{equ}
1774: where the two elements $a_1$ and $a_2$ are defined as
1775: \minilab{e:defa}
1776: \begin{equs}
1777: a_1 &= (2,0,\tilde N,0)\;,\label{e:defa1}\\
1778: a_2 &= (3,0,\tilde N+1,t_* \tilde N^{4/(1-2\alpha)})\;,\label{e:defa2}
1779: \end{equs}
1780: for some constant $t_*$ to be determined later in this section.
1781: Notice that this definition reflects the algorithm \eref{e:algo} and the explanation following \eref{e:notA}.
1782: The reason behind the particular choice of the waiting time in \eref{e:defa2} will become clear in \rem{rem:step1}.
1783:
1784: The way the construction of $ \prob_Z^1$ and $ \prob_Z^2$ works is very close to the binding construction in \cite{HExp02}. The main difference is
1785: that the construction presented in \cite{HExp02} doesn't allow to satisfy \textbf{B2} above. We will therefore introduce
1786: a symmetrised version of the binding construction that allows to gain a better control over $g_w$.
1787: If $\mu_1$ and $\mu_2$ are two positive measures with densities $D_1$ and $D_2$ with respect to
1788: some common measure $\mu$, we define the measure $\mu_1\wedge \mu_2$ by
1789: \begin{equ}
1790: \sml(\mu_1\wedge \mu_2\smr)(dw) = \min\{D_1(w), D_2(w)\}\,\mu(dw)\;.
1791: \end{equ}
1792: The key ingredient for the construction of $ \prob_Z^1$ and $ \prob_Z^2$ is the following lemma, the proof of which will be given later in this section.
1793: \begin{lemma}\label{lem:Psi}
1794: Let $Z = (x_0,y_0,w_x,w_y)$ be an admissible initial condition
1795: and let $H$, $\sigma$, and $f$ satisfy the hypotheses of either
1796: \theo{theo:smallH} or \theo{theo:largeH}. Then, there exists a measurable map $\Psi_Z\from\Omega\to\Omega$ with
1797: measurable inverse,
1798: having the following properties.
1799: \begin{claim}
1800: \item[\textbf{B1'}] There exists a constant $\delta > 0$ such that $\Wien \wedge \Psi_Z^*\Wien$ has mass bigger than $2\delta$
1801: for every admissible initial condition $Z$.
1802: \item[\textbf{B2'}] There exists a constant $\kappa$ such that $\{(\tilde w_x,\tilde w_y)\,|\, \tilde w_y = \Psi_Z(\tilde w_x)\} \subset {\cscr B}_\kappa$
1803: for every admissible initial condition $Z$.
1804: \item[\textbf{B3'}] Let $(x_t,y_t)$ be the process constructed by solving \eref{e:mainequ} with respective initial conditions $x_{0}$ and $y_{0}$, and with noise processes $P_t(w_x,\tilde w_x)$ and $P_t(w_y,\Psi_Z(\tilde w_x))$. Then, one
1805: has $x_1 = y_1$ for every $\tilde w_x \in \Omega$ and every admissible initial condition $Z$.
1806: \end{claim}
1807: Furthermore, the maps $\Psi_Z$ and $\Psi_Z^{-1}$ are measurable with respect to $Z$.
1808: \end{lemma}
1809:
1810: Given such a $\Psi_Z$, we first define the maps $\Psi_\uparrow$ and $\Psi_\rightarrow$
1811: from $\Omega$ to $\Omega\times\Omega$ by
1812: \begin{equ}
1813: \Psi_\uparrow(\tilde w_x) = \sml(\tilde w_x,\Psi_Z(\tilde w_x)\smr)\;,\qquad\Psi_\rightarrow(\tilde w_y) = \sml(\Psi_Z^{-1}(\tilde w_y),\tilde w_y\smr)\;.
1814: \end{equ}
1815: (See also Figure~\ref{fig:constr} below.)
1816: We also define the ``switch map'' $S\from\Omega\times\Omega \to \Omega\times\Omega$ by $S(\tilde w_x,\tilde w_y) = (\tilde w_y,\tilde w_x)$.
1817: \begin{figure}[h]
1818: \begin{center}
1819: \mhpastefig{Coupling}
1820: \end{center}
1821: \caption{Construction of $\prob_Z$.}\label{fig:constr}
1822: \end{figure}
1823:
1824: With these definitions at hand, we construct two measures $\prob_Z^1$ and $\tilde \prob_Z^1$ on $\Omega\times\Omega$ by
1825: \begin{equ}[e:constP]
1826: \prob_Z^1 = {1\over 2} \bigl(\Psi_\uparrow^*\Wien \wedge \Psi_\rightarrow^*\Wien\bigr)\;,\qquad \tilde\prob_Z^1 = \prob_Z^1 + S^*\prob_Z^1 \;.
1827: \end{equ}
1828: On Figure~\ref{fig:constr}, $\prob_Z^1$ lives on the boldfaced curve and $\tilde \prob_Z^1$ is its symmetrised version
1829: which lives on both the boldfaced and the dashed curve.
1830: Denote by $\Pi_{i}\from\Omega\times\Omega \to \Omega$ the projectors onto the $i$th component and by
1831: $\Delta\from\Omega\to\Omega \times \Omega$ the lift onto the diagonal $\Delta(w) = (w,w)$. Then, we define
1832: the measure $\prob_Z^2$ by
1833: \begin{equ}[e:defP]
1834: \prob_Z^2 = S^*\prob_Z^1 + \Delta^*\sml(\Wien - \Pi_1^* \tilde \prob_Z^1\smr)\;.
1835: \end{equ}
1836: By \eref{e:constP}, $\Wien > \Pi_1 \tilde \prob_Z^1$, so $\prob_Z^1$ and $\prob_Z^2$ are both positive
1837: and their sum is a probability measure.
1838: Furthermore, one has by definition
1839: \begin{equ}
1840: \prob_Z^1 + \prob_Z^2 = \tilde \prob_Z^1+ \Delta^*\sml(\Wien - \Pi_1^* \tilde \prob_Z^1\smr)\;.
1841: \end{equ}
1842: Since $\Pi_1^*\Delta^*$ is the identity, this immediately implies
1843: \begin{equ}
1844: \Pi_1^*\prob_Z^1 + \Pi_1^*\prob_Z^2= \Wien\;.
1845: \end{equ}
1846: The symmetry $S^*\tilde \prob_Z^1 = \tilde \prob_Z^1$ then implies that the second marginal is also equal
1847: to $\Wien$,
1848: \ie \textbf{B1} is satisfied. Furthermore, the set $\{(\tilde w_x,\tilde w_y)\,|\, \tilde w_y = \Psi_Z(\tilde w_x)\}$ has $\prob_Z$-measure
1849: bigger than $\delta$ by \textbf{B1'}, so \textbf{B3} is satisfied as well. Finally, \textbf{B2} is an immediate consequence of
1850: \textbf{B2'}. It remains to construct the function $\Psi_Z$.
1851:
1852: \begin{proof}[of \lem{lem:Psi}]
1853: As previously, we write $Z$ as
1854: \begin{equ}[e:notZ0]
1855: Z = (x_0,y_0,w_x,w_y)\;.
1856: \end{equ}
1857: In order to construct $\Psi_Z$, we proceed as in \cite[Sect.\ 5]{HExp02}, except that we want the solutions $x_t$ and $y_t$
1858: to become equal after time $1$. Let $\tilde w_x\in\Omega$ be given and define
1859: \begin{equ}[e:defBH]
1860: B_H(t) = \sml(\CD_H P_1(w_x,\tilde w_x)\smr)(t-1)\;,
1861: \end{equ}
1862: where $W$ denotes the corresponding part of the initial condition $Z_0$ in \eref{e:notZ0}. We write the solutions
1863: to \eref{e:mainequ} as
1864: \minilab{e:defxy}
1865: \begin{equs}
1866: dx_t &= f(x_t)\,dt + \sigma dB_H(t) \label{e:defxt}\;,\\
1867: dy_t &= f(y_t)\,dt + \sigma dB_H(t) + \sigma \tilde g_B(t)\,dt\;, \label{e:defyt}
1868: \end{equs}
1869: where $\tilde g_B(t)$ is a function to be determined. Notice that $x_t$ is completely determined by $\tilde w_x$ and by the
1870: initial condition $Z$. We introduce the process $\rho_t = y_t - x_t$, so we get
1871: \begin{equ}[e:equrho]
1872: {d \rho_t \over dt} = f(x_t+ \rho_t) - f(x_t) + \sigma \tilde g_B(t)\;.
1873: \end{equ}
1874: We now define $\tilde g_B(t)$ by
1875: \begin{equ}[e:defgb]
1876: \tilde g_B(t) = -\sigma^{-1}\Bigl(\kappa_1\rho_t + \kappa_2 {\rho_t \over \sqrt{\|\rho_t\|}}\Bigr)\;,
1877: \end{equ}
1878: for two constants $\kappa_1$ and $\kappa_2$ to be specified. This yields for the
1879: norm of $\rho_t$ the estimate
1880: \begin{equ}
1881: {d \|\rho_t\|^2 \over dt} \le 2(\C3 - \kappa_1)\|\rho_t\|^2 - 2\kappa_2 \|\rho_t\|^{3/2}\;.
1882: \end{equ}
1883: We choose $\kappa_1 = \C3$ and so
1884: \begin{equ}[e:estrho]
1885: \|\rho_t\| \le \cases{\bigl(6\kappa_2 t - \sqrt{\|\rho_0\|}\bigr)^2 & for
1886: $t < \sqrt{\|\rho_0\|} / (6\kappa_2)$,\cr 0& for $t \ge \sqrt{\|\rho_0\|} / (6\kappa_2)$.}
1887: \end{equ}
1888: We can then choose
1889: $\kappa_2$ sufficiently large, so that $\|\rho_t\| = 0$ for $t > 1/2$. Since the initial condition
1890: was admissible by assumption, the constant $\kappa_2$ can be chosen as a function of the constants $\C{i}$ only.
1891: Notice also that the preceding construction yields $\tilde g_B$ as a function of
1892: $Z$ and $\tilde w_x$ only.
1893:
1894: We then construct $\tilde w_y = \Psi_Z(\tilde w_x)$ in such a way that \eref{e:defxy} is satisfied with the function
1895: $\tilde g_B$ we just constructed. Define $g_w$ by \eref{e:defg_w} and construct $g_B$ by applying \eref{e:relG2}. Then,
1896: we extend $\tilde g_B$ to $(-\infty,1]$ by simply putting it equal to $g_B$ on $(-\infty,0]$. Applying the inverse formula
1897: \eref{e:relG1}, we obtain a function $\tilde g_w$ on $(-\infty,1]$, which is equal to $g_w$ on $(-\infty,0]$ and which is
1898: such that
1899: \begin{equ}
1900: \sml(\Psi_Z(\tilde w_x)\smr)(t) \equiv \tilde w_x(t) + \int_0^t \tilde g_w(s)\,ds \;,
1901: \end{equ}
1902: has precisely the required property.
1903:
1904: It remains to check that the family of maps $\Psi_Z$ constructed this way has the properties stated in \lem{lem:Psi}.
1905: The inverse of $\Psi_Z$ is constructed in the following way. Choose $\tilde w_y \in \Omega$ and consider the solution
1906: to the equation
1907: \begin{equ}
1908: dy_t = f(y_t)\,dt + \sigma dB_H'(t)\;,
1909: \end{equ}
1910: where $B_H$ is defined as in \eref{e:defBH} with $x$ replaced by $y$. Once $y_t$ is obtained, one can construct the
1911: process $\rho_t$ as before, but this time by solving
1912: \begin{equ}
1913: {d \rho_t \over dt} = f(y_t) - f(y_t - \rho_t) - \Bigl(\kappa_1\rho_t + \kappa_2 {\rho_t \over \sqrt{\|\rho_t\|}}\Bigr)\;.
1914: \end{equ}
1915: This allows to define $\tilde g_B$ as in \eref{e:defgb}. The element $\tilde w_x \equiv \Psi_Z^{-1}(\tilde w_y)$ is then obtained by the same
1916: procedure as before.
1917:
1918: Before turning to the proof of properties \textbf{B1'}--\textbf{B3'}, we give some estimate on the function $\tilde g_w$ that we just constructed.
1919:
1920:
1921: \begin{lemma}\label{lem:estg_w}
1922: Assume that the conditions of \lem{lem:Psi} hold. Then, there exists a constant $K$ such that the function $\tilde g_w(Z,\tilde w_x)$
1923: constructed above satisfies
1924: \begin{equ}
1925: \int_0^1 \sml\|\tilde g_w(Z,\tilde w_x)(s)\smr\|^2\,ds < K\;,
1926: \end{equ}
1927: for every admissible initial condition $Z$ and for every $\tilde w_x \in \Noise_+$.
1928: \end{lemma}
1929:
1930:
1931: \begin{proof}
1932: We write $\tilde g_w(t)$ for $t>0$ as
1933: \begin{equs}
1934: \tilde g_w(t) &= C\int_{-\infty}^0 {t^{{1\over 2}-H}(-s)^{H-{1\over 2}} \over t-s} g_w(s)\,ds + \alpha_H {d\over dt}\int_0^t
1935: (t-s)^{{1\over 2}-H} \tilde g_B(s)\,ds\;, \\
1936: &\equiv \tilde g_w^{(1)}(t) + \tilde g_w^{(2)}(t)\;.
1937: \end{equs}
1938: where $g_w$ is defined by \eref{e:defg_w}, $g_B$ is given by \eref{e:defgb}, and the constant $C$ is the constant appearing
1939: in \eref{e:formg2}.
1940: The $\L^2$-norm of $\tilde g_w^{(1)}$ is bounded by $1$ by
1941: the assumption that $Z$ is admissible.
1942: To bound the norm of $\tilde g_w^{(2)}$, we treat the cases $H<{1\over 2}$ and $H>{1\over 2}$ separately.
1943:
1944: \noindent \textbf{The case $\mathbf{H < {1\over 2}}$.} For this case, we simply combine \lem{lem:smallH} with the definition
1945: \eref{e:defgb} and the estimate \eref{e:estrho}.
1946:
1947: \noindent \textbf{The case $\mathbf{H>{1\over 2}}$.} For this case, we apply \lem{lem:bigH}, so we bound the $\|\cdot\|_{1,1}$-norm of $\tilde g_B$. By \eref{e:defgb}, one has
1948: \begin{equ}[e:estdgb]
1949: \Bigl\|{d \over dt} \tilde g_B(t)\Bigr\| \le C\Bigl\|{d\rho_t \over dt}\Bigr\| \bigl(1 + \|\rho_t\|^{-{1\over 2}}\bigr)\;,
1950: \end{equ}
1951: for some positive constant $C$. Using \eref{e:equrho}, the assumption about the boundedness
1952: of the derivative of $f$, and the definition \eref{e:defgb} we get
1953: \begin{equ}
1954: \Bigl\|{d\rho_t \over dt}\Bigr\| \le C \bigl(\|\rho_t\| + \sqrt{\|\rho_t\|}\bigr) \;.
1955: \end{equ}
1956: Combining this with \eref{e:estdgb} and \eref{e:estrho}, the required bound on $\|\tilde g_B\|_{1,1}$ follows.
1957: \end{proof}
1958:
1959:
1960: Property \textbf{B1'} now follows from \lem{lem:estg_w} and Girsanov's theorem in the following way. Denote by $\Dens_Z$ the density of $\Psi_Z^*\Wien$ with
1961: respect to $\Wien$, \ie $\sml(\Psi_Z^*\Wien\smr)(d\tilde w_x) = \Dens_Z(\tilde w_x)\,\Wien(d\tilde w_x)$. It is given by Girsanov's formula
1962: \begin{equ}
1963: \Dens_Z(\tilde w_x) = \exp \Bigl(\int_0^1 \bigl\langle\sml(\tilde g_w(Z,\tilde w_x)\smr)(t)\,,\,d\tilde w_x(t)\bigr\rangle -{1\over 2}\int_0^1 \sml\|\tilde g_w(Z,\tilde w_x)\smr\|^2(t)\,dt\Bigr)\;.
1964: \end{equ}
1965: One can check (see \eg \cite{MatNS}) that $\|\Wien \wedge \Psi_Z^*\Wien\|_\TV$ is bounded from below by
1966: \begin{equ}
1967: \|\Wien \wedge \Psi_Z^*\Wien\|_\TV \ge \Bigl(4\int_\Omega \Dens_Z(w)^{-2}\,\Wien(dw) \Bigr)\;.
1968: \end{equ}
1969: Property \textbf{B1'} thus follows immediately from \lem{lem:estg_w}, using the fact that
1970: \begin{equ}
1971: \int_\Omega \exp \Bigl(-2\int_0^1 \bigl\langle\sml(\tilde g_w(Z,\tilde w_x)\smr)(t)\,,\,d\tilde w_x(t)\bigr\rangle - 2\int_0^1 \sml\|\tilde g_w(Z,\tilde w_x)\smr\|^2(t)\,dt\Bigr)\,\Wien(dw) = 1\;.
1972: \end{equ}
1973: Property \textbf{B2'} is also an immediate consequence of \lem{lem:estg_w}, and property \textbf{B3'} follows by construction from \eref{e:estrho}. The proof of \lem{lem:Psi} is complete.
1974: \end{proof}
1975:
1976: Before concluding this subsection we show that, if step \step{1} fails, $t_*$ can be chosen in such a way that
1977: the waiting time $t_*\tilde N^{4/(1-2\alpha)}$ in
1978: \eref{e:defa2} is long enough so that \eref{e:defAdm} holds again after step \step{3} and so that the cost
1979: function does not increase by more than $1/(2\tilde N^2)$. By the triangle inequality, the second claim follows
1980: if we show that
1981: \begin{equ}[e:boundCost]
1982: \CK_\alpha\sml(\theta_t \tilde g_w(Z,\tilde w_x)\smr) \le {1 \over 2 \tilde N^2}\;,
1983: \end{equ}
1984: whenever $t$ is large enough (the shift $\theta_t$ is as in \eref{e:shiftcost}). Combining \eref{e:boundfuture}, \lem{lem:estg_w}, and the
1985: definition of $\CK_\alpha$, we get, for some constant $C$,
1986: \begin{equ}
1987: \CK_\alpha\sml(\theta_t \tilde g_w(Z,\tilde w_x)\smr) \le C t^{\alpha - {1\over 2}} + C t^{H-{3\over 2}}\;,\quad\text{for $t\ge 2$.}
1988: \end{equ}
1989: There thus exists a constant $t_*$ such that the bound \eref{e:boundCost} is satisfied if the waiting time
1990: is longer than $t_* \tilde N^{4/(1-2\alpha)}$. It remains to show that \eref{e:defAdm} holds after the waiting
1991: time is over. If step \step{1} failed, the realisations $\tilde w_x$ and $\tilde w_y$ are drawn either in the set
1992: \begin{equ}
1993: \tilde \Delta_1 = \{ (\tilde w_x, \tilde w_y) \in \Omega^2 \,|\, \tilde w_x = \tilde w_y \}\;,
1994: \end{equ}
1995: or in the set
1996: \begin{equ}
1997: \tilde \Delta_2 = \{(\tilde w_x, \tilde w_y) \in \Omega^2 \,|\, \tilde w_x = \Psi_Z(\tilde w_y) \}
1998: \end{equ}
1999: (see Figure~\ref{fig:constr}). In order to describe the dynamics also during the waiting time
2000: (\ie step \step{3}), we extend those sets to $\Noise_+^2$ by
2001: \begin{equs}
2002: \Delta_i &= \big\{(\tilde w_x, \tilde w_y) \in \Noise_+^2\,|\, (\tilde w_x |_{[0,1]}, \tilde w_y |_{[0,1]}) \in \tilde \Delta_i\;,
2003: \\
2004: &\qquad \text{and}
2005: \quad \tilde w_x (t) - \tilde w_y (t) = \text{const}\quad \text{for $t>1$}\big\}\;.
2006: \end{equs}
2007: Given an admissible initial condition $Z = (x_0,y_0,w_x,w_y)$ and a pair $(\tilde w_x, \tilde w_y) \in \Noise_+^2$, we consider
2008: the solutions $x_t$ and $y_t$ to \eref{e:mainequ} given by
2009: \begin{equa}[e:evolCoupl]
2010: dx_t &= f(x_t)\,dt + \sigma dB_H^x(t) \;,\\
2011: dy_t &= f(y_t)\,dt + \sigma dB_H^y(t) \;,
2012: \end{equa}
2013: where $B_H^x$ (and similarly for $B_H^y$) is constructed as usual by concatenating $w_x$ and $\tilde w_x$ and applying the operator
2014: $\CD_H$.
2015: The key observation is the following lemma.
2016:
2017: \begin{lemma}\label{lem:controldist}
2018: Let $Z$ be an admissible initial condition as above, let $(\tilde w_x, \tilde w_y) \in \Delta_1 \cup \Delta_2$, and
2019: let $x_t$ and $y_t$ be given by \eref{e:evolCoupl} for $t>0$. Then, there exists a constant $t_* > 0$ such that
2020: \begin{equ}
2021: \|x_t - y_t\| \le 1+ {1 + \C4 \over \C2}
2022: \end{equ}
2023: holds again for $t > t_*$.
2024: \end{lemma}
2025:
2026: \begin{proof}
2027: Fix an admissible initial condition $Z$ and consider the case when $(\tilde w_x, \tilde w_y) \in \Delta_2$ first.
2028: Let $g_w \from \R_- \to \R^n$ be as in \eref{e:defg_w} and define $\tilde g_w \from \R_+ \to \R^n$
2029: by
2030: \begin{equ}
2031: \tilde w_y(t) = \tilde w_x(t) + \int_0^t \tilde g_w(s)\,ds\;.
2032: \end{equ}
2033: Introducing $\rho_t = y_t - x_t$, we see that it
2034: satisfies the equation
2035: \begin{equ}[e:equrho2]
2036: {d \rho_t \over dt} = f(y_t) - f(x_t) + \sigma \CG_t\;,
2037: \end{equ}
2038: where the function $\CG_t$ is given by
2039: \begin{equ}[e:defG]
2040: \CG_t = c_1 \int_{-\infty}^0 (t-s)^{H-{3\over 2}} g_w(s)\,ds +
2041: c_2 {d \over dt} \int_{0}^t (t-s)^{H-{1\over 2}} \tilde g_w(s)\,ds\;,
2042: \end{equ}
2043: with some constants $c_1$ and $c_2$ depending only on $H$.
2044: It follows from \eref{e:equrho2}, \eref{e:consA1}, and Gronwall's lemma, that the Euclidean norm $\|\rho_t\|$ satisfies the inequality
2045: \begin{equ}[e:estrhoGron]
2046: \|\rho_t\| \le e^{-\C2 t}\|\rho_0\| + \int_0^t e^{-\C2 (t-s)}\sml(\C4 + \|\CG_s\|\smr)\,ds\;.
2047: \end{equ}
2048: Consider first the time interval $[0,1]$ and define
2049: \begin{equ}
2050: \tilde \CG_t = c_1\int_{-\infty}^0 (t-s)^{H-{3\over 2}} g_w(s)\,ds -
2051: c_2 {d \over dt} \int_{0}^t (t-s)^{H-{1\over 2}} \tilde g_w(s)\,ds\;,
2052: \end{equ}
2053: \ie, we simply reversed the sign of $\tilde g_w$.
2054: This corresponds to the case where $(\tilde w_x, \tilde w_y)$ are interchanged, and thus satisfy
2055: $\tilde w_y = \Psi_Z(\tilde w_x)$ instead of $\tilde w_x = \Psi_Z(\tilde w_y)$.
2056: We thus deduce from \eref{e:defgb} and \eref{e:estrho} that
2057: \begin{equ}[e:wantboundG]
2058: \|\tilde \CG_s\| \le \|\sigma^{-1}\|\bigl(\kappa_1 \|\rho_0\|Ê+ \kappa_2 \sqrt{\|\rho_0\|}\bigr)\;,
2059: \end{equ}
2060: for $s \in [0,1]$. This yields for $\|\CG_s\|$ the estimate
2061: \begin{equs}
2062: \|\CG_s\| &\le \|\sigma^{-1}\|\bigl(\kappa_1 \|\rho_0\|Ê+ \kappa_2 \sqrt{\|\rho_0\|}\bigr) + 2 c_1 \int_{-\infty}^0 (t-s)^{H-{3\over 2}} \|g_w(s)\|\,ds \\
2063: &\le \|\sigma^{-1}\|\bigl(\kappa_1 \|\rho_0\|Ê+ \kappa_2 \sqrt{\|\rho_0\|}\bigr) + 1 \;, \label{e:estG1}
2064: \end{equs}
2065: where we used the fact that $Z$ is admissible for the second step. Notice that \eref{e:estG1} only holds for
2066: $s \in [0,1]$, so we consider now the case $s > 1$. In this case, we can write $\CG_t$ as
2067: \begin{equ}
2068: \CG_t = c_1 \int_{-\infty}^0 (t-s)^{H-{3\over 2}} g_w(s)\,ds +
2069: c_1 \int_{0}^1 (t-s)^{H-{3\over 2}} \tilde g_w(s)\,ds\;.
2070: \end{equ}
2071: The first term is bounded by $1$ as before.
2072: In order to bound the second term, we use \lem{lem:estg_w}, so we get
2073: \begin{equ}[e:boundGt]
2074: \|\CG_t\| \le 1 + \sqrt{{K\over 2H-2} \bigl((t-1)^{2H-2} - t^{2H-2}\bigr)}\;.
2075: \end{equ}
2076: This function has a singularity at $t=1$, but this singularity is always integrable. For $t>2$ say, it behaves like
2077: $t^{H-{3\over 2}}$. Putting the estimates \eref{e:estG1} and \eref{e:boundGt} into \eref{e:estrhoGron}, we
2078: see that there exists a constant $C$ depending only on $H$ and on the parameters in assumption
2079: \textbf{A1} such that, for $t>2$, one has the estimate
2080: \begin{equ}
2081: \|\rho_t\| \le e^{-\C2 t}\|\rho_0\| + {1+\C4 \over \C2} + C t^{H-{3\over 2}}\;.
2082: \end{equ}
2083: The claim follows at once.
2084: \end{proof}
2085:
2086: \begin{remark}\label{rem:step1}
2087: To summarise, we have shown the following in this section:
2088: \begin{claim}[AA]
2089: \item[1.] There exists a positive constant $\delta$ such that if the state $Z$ of the coupled system is admissible,
2090: step \step{1} has a probability larger than $\delta$ to succeed.
2091: \item[2.] If step \step{1} fails and the waiting time for step \step{3} is chosen larger than
2092: $t_*\tilde N^{4/(1-2\alpha)}$, then the state of the coupled
2093: system is again admissible after the end of step \step{3}, provided the cost $\CK_\alpha(Z)$ at the beginning of step \step{1}
2094: was smaller than $1-{1 \over 2\tilde N^2}$.
2095: \item[3.] The increase in the cost given between the beginning of step \step{1} and the end
2096: of step \step{3} is smaller than ${1 \over 2\tilde N^2}$.
2097: \end{claim}
2098: \end{remark}
2099:
2100: In the following subsection, we will define step \step{2} and so conclude the construction and the analysis of the
2101: coupling function $\Coupl$.
2102:
2103: \subsection{Coupling stage ({\bfseries\itshape S = 2})}
2104:
2105: In this subsection, we construct and analyse the coupling map $\Coupl$ corresponding to step \step{2}.
2106: Following \eref{e:algo}, we construct it in such a way that, with positive probability, the two copies of the process
2107: \eref{e:mainequ} are driven with the same noise. In other terms, if $Z=(x_0,y_0,w_x,w_y)$ denotes the
2108: state of our coupled system at the beginning of step \step{2}, we construct a measure $\prob_Z$
2109: on $\Noise_+^2$ such that if $(\tilde w_x, \tilde w_y)$ is drawn according to $\prob_Z$, then one has
2110: \begin{equ}[e:defcoupl]
2111: \sml(\CD_H(w_x \sqcup \tilde w_x)\smr)(t) = \sml(\CD_H(w_y \sqcup \tilde w_y)\smr)(t)\;,\quad t>0\;,
2112: \end{equ}
2113: with positive probability. Here, $\sqcup$ denotes the concatenation operator given by
2114: \begin{equ}
2115: \sml(w \sqcup \tilde w\smr)(t) = \cases{w(t) & for $t < 0$,\cr\tilde w(t) & for $t \ge 0$.}
2116: \end{equ}
2117: In the notation \eref{e:notA}, step \step{2} will have a duration $2^N$ and $N$ will be incremented
2118: by $1$ every time step \step{2} succeeds.
2119:
2120: The construction of $\prob_Z$ will be similar in spirit to the construction of the previous section.
2121: We therefore introduce as before the function $\tilde g_w$ given by
2122: \begin{equ}[e:defgwtilde]
2123: \tilde w_y(t) = \tilde w_x(t) + \int_0^t \tilde g_w(s)\,ds\;.
2124: \end{equ}
2125: Our main concern is of course to get good bounds on this function $\tilde g_w$. This is
2126: achieved by the following lemma, which is crucial in the process of showing that step \step{2} will
2127: eventually succeed infinitely often.
2128:
2129: \begin{lemma}\label{lem:step1}
2130: Let $Z_0$ be an admissible initial condition and denote by $\CT$ the measure on $\State^2\times\Noise^2$
2131: obtained by evolving $Z_0$ according to the successful realisation of step \step{1}. Then, there exists a constant
2132: $\tilde K > 0$ depending only on $H$, $\alpha$, and the parameters appearing in \textbf{A1}, such that for $\CT$-almost
2133: every $Z = (x,y,w_x,w_y)$, and for every pair $(\tilde w_x,\tilde w_y)$ satisfying \eref{e:defcoupl}, we have the
2134: bounds
2135: \begin{equ}[e:boundstep2]
2136: \|\tilde g_w\|_\alpha \le \tilde K \;,\qquad \Bigl\|{d \tilde g_w \over dt}\Bigr\|_{\alpha+1} \le \tilde K \;.
2137: \end{equ}
2138: Furthermore, one has $x=y$, $\CT$-almost surely.
2139: \end{lemma}
2140:
2141: \begin{proof}
2142: It is clear from \lem{lem:Psi} that $x=y$.
2143: Let now $Z$ be an element drawn according to $\CT$
2144: and denote by $g_w\from\R_-\to \R^n$ the function formally defined by
2145: \begin{equ}[e:defgw2]
2146: dw_y(t) = dw_x(t) + g_w(t)\,dt\;.
2147: \end{equ}
2148: We also denote by $g_b\from \R_- \to \R^n$ the function such that
2149: \begin{equ}[e:defgb2]
2150: dB_y(t) = dB_x(t) + g_b(t)\,dt\;,
2151: \end{equ}
2152: where $B_x = \CD_H w_x$ and $B_y = \CD_H w_y$.
2153: (Note that $g_w$ and $g_b$ are almost surely well-defined, so we discard elements $Z$
2154: for which they can not be defined.)
2155: Since $Z$ corresponds almost surely to a successful realization of step \textbf{1}, $g_b$ is equal
2156: on the interval $[-1,0]$ (up to translation in time) to the function $\tilde g_B$ constructed in \eref{e:defgb}. By \eref{e:estrho}, there exists therefore a constant $C_g$ such that
2157: \begin{equ}[e:boundgb]
2158: \|g_b(s)\| \le \cases{C_g & for $s\in [-1,-{\textstyle{1\over 2}})$,\cr 0& for $s\in[-{\textstyle{1\over 2}},0]$.}
2159: \end{equ}
2160: Combining the linearity of $\CD_H$ with
2161: \eref{e:formg2}, one can see that if $(\tilde w_x,\tilde w_y)$ satisfy \eref{e:defcoupl}, then the
2162: function $\tilde g_w$ is given by the formula
2163: \begin{equ}[e:exprgwtilde1]
2164: \tilde g_w(t) = C_1\int_{-\infty}^{-1} {|t+1|^{{1\over 2}-H}|s + 1|^{H-{1\over 2}} \over t-s}g_w(s)\,ds
2165: + C_2\int_{-1}^{-1/2} (t-s)^{-H-{1\over 2}} g_b(s)\,ds\;,
2166: \end{equ}
2167: for some constants $C_1$ and $C_2$ depending only on $H$. Notice that the second integral
2168: only goes up to $1/2$ because of \eref{e:boundgb}.
2169:
2170: Since the initial condition $Z_0$ is admissible by assumption, the $\|\cdot\|_\alpha$-norm of the first
2171: term is bounded by $1$. The $\|\cdot\|_\alpha$-norm of the second term is also bounded by a constant,
2172: using \eref{e:boundgb} and the assumption
2173: $\alpha < H$.
2174:
2175: Deriving \eref{e:exprgwtilde1} with respect to $t$, we see that there exists a constant $K$ such that
2176: \begin{equa}[e:boundderg]
2177: \Bigl\|{d\tilde g_w(t) \over dt}\Bigr\| &\le {K \over t+1} \Big(\int_{-\infty}^{-1} {|t+1|^{{1\over 2}-H}|s + 1|^{H-{1\over 2}} \over t-s} \|g_w(s)\|\,ds
2178: \\
2179: &\qquad\qquad+ \int_{-1}^{-1/2} (t-s)^{-H-{1\over 2}} \|g_b(s)\|\,ds\Big)\;,
2180: \end{equa}
2181: and the bound on the derivative follows as previously.
2182: \end{proof}
2183:
2184: The definition of our coupling function will be based on the following lemma:
2185: \begin{lemma}\label{lem:gauss}
2186: Let $\CN$ be the normal distribution on $\R$, choose $a \in \R$, $b\ge |a|$, and define
2187: $M = \max\{4b,2\log(8/b)\}$. Then, there exists a measure
2188: $\CN_{a,b}^2$ on $\R^2$ satisfying the following properties:
2189: \begin{claim}[33]
2190: \item[1.] Both marginals of $\CN_{a,b}^2$ are equal to $\CN$.
2191: \item[2.] If $|b|\le 1$, one has
2192: \begin{equ}
2193: \CN_{a,b}^2 \bigl(\bigl\{(x,y)\,|\, y = x+a\bigr\}\bigr) > 1 - b\;.
2194: \end{equ}
2195: Furthermore, the above quantity is always positive.
2196: \item[3.] One has
2197: \begin{equ}
2198: \CN_{a,b}^2 \bigl(\bigl\{(x,y)\,|\, |y - x| \le M\bigr\}\bigr) = 1\;.
2199: \end{equ}
2200: \end{claim}
2201: \end{lemma}
2202:
2203: \begin{proof}
2204: Consider the following picture:
2205: \begin{center}
2206: \begin{minipage}{5cm}
2207: \mhpastefig{Gaussian}
2208: \end{minipage}
2209: \begin{minipage}{4cm}
2210: \begin{equs}
2211: L_1\,:\quad y &= x \;,\\
2212: L_2\,:\quad y &= -x \;,\\
2213: L_3\,:\quad y &= x+a \;.\\
2214: \quad&\quad
2215: \end{equs}
2216: \end{minipage}
2217: \end{center}
2218: Denote by $\CN_x$ the normal distribution on the set $L_x = \{(x,y)\,|\, y = 0\}$ and by $\CN_y$ the normal
2219: distribution on the set $L_y = \{(x,y)\,|\, x = 0\}$. We also define the maps $\pi_{i,x}$ (respect. $\pi_{i,y}$) from
2220: $L_x$ (respect. $L_y$) to $L_i$, obtained by only modifying the $y$ (respect. $x$) coordinate. Notice that
2221: these maps are invertible and denote their inverses by $\tilde \pi_{i,x}$ (respect. $\tilde \pi_{i,y}$).
2222: We also denote by $\CN_x|_M$ (respect. $\CN_y|_M$) the restriction of $\CN_x$ (respect. $\CN_y$)
2223: to the square $[-{M\over 2},{M\over 2}]^2$.
2224:
2225: With these notations, we define the measure $\CN_3$ on $L_3$ as
2226: \begin{equ}
2227: \CN_3 = \pi_{3,x}^* \sml(\CN_x |_M\smr)\wedge \pi_{3,y}^* \sml(\CN_y |_M\smr)\;.
2228: \end{equ}
2229: The measure $\CN_{a,b}^2$ is then defined as
2230: \begin{equ}
2231: \CN_{a,b}^2 = \CN_3 + \pi_{2,x}^*\bigl(\sml(\CN_x |_M\smr) - \tilde \pi_{3,x}^*\CN_3\bigr)
2232: + \pi_{1,x}^*\bigl(\CN_x -\sml(\CN_x |_M\smr) \bigr)\;.
2233: \end{equ}
2234: A straightforward calculation, using the symmetries of the problem, shows that property 1 is indeed satisfied.
2235: Property 3 follows immediately from the construction, so it remains to check that property 2 holds, \ie that
2236: \begin{equ}
2237: \CN_3(L_3) \ge 1- b\;,
2238: \end{equ}
2239: for $|b| < 1$, and $\CN_3(L_3)>0$ otherwise. It follows from the definition of the total variation distance $\|\cdot\|_\TV$ that
2240: \begin{equ}
2241: \CN_3(L_3) = 1-{1 \over 2}\|\sml(\CN_x |_M\smr) - \tau_a^* \sml(\CN_x |_M\smr) \|_\TV\;,
2242: \end{equ}
2243: where $\tau_a(x) = x-a$.
2244: Since $M\ge 4b \ge 4a$, is clear from the picture and from the fact that the density of the normal distribution is
2245: everywhere positive, that $\CN_3(L_3)>0$ for every $a\in\R$. It therefore suffices to consider the case
2246: $|b| \le 1$. Since
2247: $\int_M^\infty e^{-x^2/2}\,dx < b/8$, one
2248: has $\|\CN_x|_M -\CN_x \|_\TV \le b/4$, which implies
2249: \begin{equ}
2250: \CN_3(L_3) \ge 1- {b\over 4} - {1 \over 2}\|\CN_x - \tau_a^* \CN_x\|_\TV\;.
2251: \end{equ}
2252: A straightforward computation shows that, for $|a|\le 1$,
2253: \begin{equ}
2254: \|\CN_x - \tau_a^* \CN_x\|_\TV \le \sqrt{e^{a^2}-1} \le \sqrt{2} a \;,
2255: \end{equ}
2256: and the claim follows.
2257: \end{proof}
2258:
2259: We will use the following corollary:
2260:
2261: \begin{corollary}\label{cor:Girs}
2262: Let $\Wien$ be the Wiener measure on $\Noise_+$, let $g\in \L^2(\R_+)$ with $\|g\| \le b$, let
2263: $M = \max\{4b,2\log(8/b)\}$, and define
2264: the map $\Psi_g\from\Noise_+\to\Noise_+$ by
2265: \begin{equ}
2266: \sml(\Psi_g w\smr)(t) = w(t) + \int_0^t g(s)\,ds\;.
2267: \end{equ}
2268: Then, there exists a measure $\Wien_{g,b}^2$ on $\Noise_+^2$ such that
2269: the following properties hold:
2270: \begin{claim}[33]
2271: \item[1.] Both marginals of $\Wien_{g,b}^2$ are equal to the Wiener measure $\Wien$.
2272: \item[2.] If $b \le 1$, one has the bound
2273: \begin{equ}[e:goodset]
2274: \Wien_{g,b}^2 \bigl(\bigl\{(\tilde w_x, \tilde w_y)\,|\, \tilde w_y = \Psi_g(\tilde w_x)\bigr\}\bigr) \ge 1-b\;.
2275: \end{equ}
2276: Furthermore, at fixed $b >0$, the above quantity is always positive and a decreasing function of $\|g\|$.
2277: \item[3.] The set
2278: \begin{equ}
2279: \Bigl\{(\tilde w_x, \tilde w_y)\,\Big|\, \exists \kappa \,:\, \tilde w_y(t) = \tilde w_x(t) + \kappa \int_0^t g(s)\,ds\,,\,
2280: |\kappa| \| g\| \le M\Bigr\}
2281: \end{equ}
2282: has full $\Wien_{g,b}^2$-measure.
2283: \end{claim}
2284: \end{corollary}
2285:
2286: \begin{proof}
2287: This is an immediate consequence of the $\L^2$ expansion of white noise, using $g$ as one of the
2288: basis functions and applying \lem{lem:gauss} on that component.
2289: \end{proof}
2290:
2291: Given this result (and using the same notations as above), we turn to the construction of the coupling function $\Coupl$
2292: for step \step{2}. Given an initial condition $Z = (x_0,y_0,w_x,w_y)$, remember that $g_w$ is defined by
2293: \eref{e:defg_w}. We furthermore define the function $\tilde g_w\from \R_+ \to \R^n$ by
2294: \begin{equ}[e:exprgwtilde2]
2295: \tilde g_w(t) = C \int_{-\infty}^0 {t^{{1\over 2}-H}(-s)^{H-{1\over 2}} \over t-s} g_w(s)\,ds\;,
2296: \end{equ}
2297: with $C$ the constant appearing in \eref{e:formg2}. By \eref{e:formg2}, $\tilde g_w$ is the only
2298: function that ensures that \eref{e:defcoupl} holds if $\tilde w_x$ and $\tilde w_y$ are related
2299: by \eref{e:defgwtilde}. (Notice that, although \eref{e:exprgwtilde1} seems to differ substantially
2300: from \eref{e:exprgwtilde2}, they do actually define the same function.) Given $Z$ as above and
2301: $a\in\CA$, denote by $g_{a,Z}$ the restriction of $\tilde g_w$ to the interval $[0,2^N]$ (prolonged
2302: by $0$ outside). It follows from \lem{lem:step1} that there exists a constant $K$ such that if
2303: the coupled process was in an admissible state at the beginning of step \step{1}, then
2304: the a-priori estimate
2305: \begin{equ}[e:aprioristep2]
2306: \|g_{a,Z}\|^2 \equiv \int_0^{2^N}\|g_{a,Z}(s)\|^2\,ds \le C 2^{-2\alpha N} \equiv b_N^2
2307: \end{equ}
2308: holds for some constant $C$.
2309: We thus define $b = \max\{b_N,\|g_{a,Z}\|\}$ and denote by $\Wien_{Z,a}^2$ the restriction of
2310: $\Wien_{g_{a,Z},b}^2$ to the ``good'' set \eref{e:goodset} and by $\tilde\Wien_{Z,a}^2$ its restriction
2311: to the complementary set.
2312:
2313: We choose furthermore an arbitrary exponent $\beta$ satisfying the condition
2314: \begin{equ}[e:condbeta]
2315: \beta > {1\over 1-2\alpha}\;.
2316: \end{equ}
2317: With these notations at hand, we define the coupling function for step \step{2}:
2318: \begin{equ}
2319: T(Z,a) = 2^N\;, \quad \Wien_2(Z,a) = \Wien_{Z,a}^2 \times \delta_{a'} + \tilde \Wien_{Z,a}^2 \times \delta_{a''}\;,
2320: \end{equ}
2321: where
2322: \begin{equ}[e:defstep2]
2323: a' = (2,N+1,\tilde N, 0)\;,\quad a'' = (3,0,\tilde N+1, \tilde t_* 2^{\beta N}\tilde N^{4/(1-2\alpha)})\;,
2324: \end{equ}
2325: for some constant $\tilde t_*$ to be determined in the remainder of this section.
2326: The waiting time in \eref{e:defstep2} has been chosen in such a way that the following holds.
2327:
2328: \begin{lemma}\label{lem:step2}
2329: Let $(Z_0,a_0) \in \State^2\times\Noise^2\times\CA$ with $Z_0$ admissible and
2330: denote by $\CT$ the measure on $\State^2\times\Noise^2$
2331: obtained by evolving it according to the successful realisation of step \step{1}, followed by
2332: $N$ successful realisations of step \step{2}, one failed realisation of step \step{2}, and one waiting
2333: period \step{3}. There exists a constant $\tilde t_*$
2334: such that $\CT$-almost every $Z = (x,y,w_x,w_y)$ satisfies
2335: \begin{equ}
2336: \|x - y\| \le 1+ {1+\C4\over \C2}\;,\quad \CK_\alpha(Z) \le \CK_\alpha(Z_0) + {1 \over 2\tilde N^2}\;,
2337: \end{equ}
2338: where $\tilde N$ denotes the value of the corresponding component of $a_0$.
2339: \end{lemma}
2340:
2341: \begin{proof}
2342: We first show the bound on the cost function. Given $Z$ distributed according to $\CT$ as in
2343: the statement, we define $g_w$ by \eref{e:defgw2} as usual. The bounds we get on the function
2344: $g_w$ are schematically depicted in the following figure, where the time interval $[\tilde t_2,t_3]$
2345: corresponds to the failed realisation of step \step{2}.
2346: \begin{equ}[e:figBounds]
2347: \mhpastefig{Bounds}
2348: \end{equ}
2349: Notice that, except for the contribution coming from times smaller than $t_1$, we are exactly in the
2350: situation of \eref{e:defg}.
2351: Since the cost of a function is decreasing under time shifts, the contribution to $\CK_\alpha(Z)$
2352: coming from $(-\infty,t_1]$ is bounded by $\CK_\alpha(Z_0)$. Denote by $g$ the function defined
2353: by
2354: \begin{equ}
2355: g(t) = \cases{g_w(t+t_1) & for $t\in [0, t_3-t_1]$,\cr 0& otherwise.}
2356: \end{equ}
2357: %We also denote by $\tilde g$ the function defined by
2358: %\begin{equ}
2359: %\tilde g(t) = \cases{g_w(t+\tilde t_2) & for $t\in [0, t_3-\tilde t_2]$,\cr 0& otherwise.}
2360: %\end{equ}
2361: Using the definition of the cost function together with \prop{prop:pastfuture} and the Cauchy-Schwarz
2362: inequality, we obtain for some
2363: constants $C_1$ and $C_2$ the bound
2364: \begin{equ}
2365: \CK_\alpha(Z) \le \CK_\alpha(Z_0) + C_1 \sqrt{|t_3|^{2H-2}-|t_1|^{2H-2}}\|g\| + C_2 \Bigl|{t_1 \over t_3 - t_1}\Bigr|^{\alpha-{1\over 2}}\|g\|_\alpha\;,
2366: \end{equ}
2367: where $\|\cdot\|$ denotes the $\L^2$-norm.
2368: Since step \step{1} has length $1$ and the $N$th occurrence of step \step{2} has length $2^{N-1}$,
2369: we have
2370: \begin{equ}
2371: |t_3 - t_1| = 2^{N+1}\;,\quad |t_3| = \tilde t_* 2^{\beta N}\tilde N^{4/(1-2\alpha)}\;.
2372: \end{equ}
2373: In particular, one has $|t_3| > |t_3-t_1|$ if $\tilde t_*$ is larger than $1$.
2374: Since
2375: \begin{equ}
2376: \sqrt{|t_3|^{2H-2}-|t_1|^{2H-2}} \le |t_3|^{H-{3\over 2}} |t_3 - t_1|^{1\over 2} \le \Bigl|{t_3 \over t_3 - t_1}\Bigr|^{-{1\over 2}}\;,
2377: \end{equ}
2378: this yields (for a different constant $C_1$) the bound
2379: \begin{equ}
2380: \CK_\alpha(Z) \le \CK_\alpha(Z_0) + C_1 \Bigl|{t_3 \over t_3 - t_1}\Bigr|^{\alpha-{1\over 2}} \|g\|_\alpha
2381: \le\CK_\alpha(Z_0) + C_1 {\tilde t_*^{\alpha -{1\over 2}}2^{-\gamma N}\over \tilde N^2} \|g\|_\alpha\;,
2382: \end{equ}
2383: where we defined $\gamma = (\beta-1)({1\over 2}-\alpha)$. Notice that \eref{e:condbeta} guarantees that
2384: $\gamma > \alpha$.
2385:
2386: We now bound the $\|\cdot\|_\alpha$-norm of $g$. We know from \lem{lem:step1} that
2387: the contribution coming from the time interval $[t_1,\tilde t_2]$ is bounded by some constant $K$.
2388: Furthermore, by \eref{e:aprioristep2}, we have for the contribution coming from the interval
2389: $[\tilde t_2, t_3]$ a bound of the type
2390: \begin{equ}
2391: \int_{\tilde t_2}^{t_3}\|\tilde g(s)\|^2\,ds \le C (N+1)^2\;,
2392: \end{equ}
2393: for some positive constant $C$. This yields for $g$ the bound
2394: \begin{equ}
2395: \|g\|_\alpha \le C (N+1)2^{\alpha N}\;,
2396: \end{equ}
2397: for some other constant $C$. Since $\gamma > \alpha$, there exists a constant $C$ such that
2398: \begin{equ}
2399: \CK_\alpha(Z) \le \CK_\alpha(Z_0) + C {\tilde t_*^{\alpha -{1\over 2}}\over \tilde N^2}\;.
2400: \end{equ}
2401: By choosing $\tilde t_*$ sufficiently large, this proves the claim concerning the increase of the total cost.
2402:
2403: It remains to show that, at the end of step \step{3}, the two realisations of \eref{e:mainequ}
2404: didn't drift to far apart.
2405: Define $g_b$ by \eref{e:defgb2} as usual and notice that, by construction, $x_t = y_t$ for $t = \tilde t_2$.
2406: Writing as before $\rho_t = y_t - x_t$, one has for $t > \tilde t_2$ the estimate
2407: \begin{equ}[e:defrhostep2]
2408: \|\rho_t\| \le {\C4 \over \C2} + \int_{\tilde t_2}^t e^{-\C2(t-s)} \|g_b(s)\|\,ds\;.
2409: \end{equ}
2410: We first estimate the contribution coming from the time interval $[\tilde t_2, t_3]$. Denote by
2411: $\tilde g\from [\tilde t_2, t_3] \to \R^n$ the value $g_w$ would have taken, had the last occurence of step
2412: \step{2} succeeded and not failed
2413: (this corresponds to the dashed curve in \eref{e:figBounds}).
2414: Defining $\hat g = g_w - \tilde g$, we have by \eref{e:relG5} that, on the interval $t\in [\tilde t_2,t_3]$,
2415: \begin{equ}[e:defgbstep2]
2416: g_b(t) = C_1 {\hat g(\tilde t_2) \over (t-\tilde t_2)^{{1\over 2}-H}} + C_2 \int_{\tilde t_2}^t {{d\hat g\over ds}(s) \over (t-s)^{{1\over 2}-H}}\,ds\;.
2417: \end{equ}
2418: By Corollary~\ref{cor:Girs} and the construction of the coupling function,
2419: $\hat g$ is proportional to $g_w$ and, by \eref{e:aprioristep2},
2420: we also have for $\hat g$ a bound of the type $\|\hat g\| \le C (N+1)$ (the norm is the $\L^2$-norm over the
2421: time interval $[\tilde t_2, t_3]$). Furthermore, \eref{e:boundderg} yields $\|{d\hat g\over ds}\| \le C (N+1) 2^{-N}$.
2422: Recall that every differentiable function defined on an interval of length $L$ satisfies
2423: \begin{equ}
2424: |f(t)| \le {\|f\| \over \sqrt L} + \Bigl\|{df \over dt}\Bigr\|\sqrt L \;.
2425: \end{equ}
2426: (The norms are $\L^2$-norms.) Using this to bound
2427: the first term in \eref{e:defgbstep2} and the Cauchy-Schwarz inequality for the second term,
2428: we get a constant $C$ such that $g_b$ is bounded by
2429: \begin{equ}
2430: \|g_b(t)\| \le C (N+1) \bigl(1 + 2^{-{N\over 2}} (t-\tilde t_2)^{H-{1\over 2}}\bigr)\;.
2431: \end{equ}
2432: From this and \eref{e:defrhostep2}, we get an other constant $C$ such that $\|\rho_t\| \le C (N+1)$ at the
2433: time $t = t_3$. We finally turn to the interval $[t_3,0]$. It follows from \eref{e:relG3} that, for some constant
2434: $C$, we have
2435: \begin{equ}
2436: \|g_b(t)\| \le {1\over 2} + C |t-t_3|^{H-1} \|g\|\;,
2437: \end{equ}
2438: where the term ${1\over 2}$ is the contribution from the times smaller than $t_1$. Since we know by
2439: \eref{e:aprioristep2} and Corollary~\ref{cor:Girs} that the $\L^2$-norm of $g$ is bounded by $C(N+1)$ for some constant
2440: $C$, we obtain the required estimate by choosing $\tilde t_*$ sufficiently large.
2441: \end{proof}
2442:
2443: \begin{remark}\label{rem:step2}
2444: To summarise this subsection, we have shown the following, assuming that the coupled system
2445: was in an admissible state before performing step \step{1} and that step \step{1} succeeded:
2446: \begin{claim}[AA]
2447: \item[1.] There exists constants $\delta' \in (0,1)$ and $K > 0$ such that the $N$th consecutive
2448: occurrence of step \step{2} succeeds with probability larger than $\max\{\delta', 1-K2^{-\alpha N}\}$.
2449: This occurrence has length $2^{N-1}$.
2450: \item[2.] If the $N$th occurrence of step \step{2} fails and the waiting time for step \step{3} is chosen longer
2451: than $\tilde t_* 2^{\beta N}\tilde N^{4/(1-2\alpha)}$, then the state of the coupled system is again admissible after the end of step \step{3}, provided that
2452: the cost $\CK_\alpha(Z)$ at the beginning of step \step{1} was smaller than $1 - {1\over 2 \tilde N^2}$.
2453: \item[3.] The increase in the cost given between the beginning of step \step{1} and the end
2454: of step \step{3} is smaller than ${1 \over 2\tilde N^2}$.
2455: \end{claim}
2456: \end{remark}
2457:
2458: Now that the construction of the coupling function $\Coupl$ is completed,
2459: we can finally turn to the proof of the results announced in the introduction.
2460:
2461: \section{Proof of the main result}
2462: \label{sec:mainproof}
2463:
2464: Let us first reformulate Theorems~\ref{theo:smallH} and \ref{theo:largeH} in a more precise way,
2465: using the notations developed in this paper.
2466:
2467: \begin{theorem}\label{theo:mainTheoFormal}
2468: Let $H \in (0,1) \setminus \{{1\over 2}\}$, let $f$ and $\sigma$ satisfy \textbf{A1}--\textbf{A3} if $H<{1\over 2}$ and
2469: \textbf{A1}, \textbf{A2'}, \textbf{A3} if $H>{1\over 2}$, and let $\gamma < \max_{\alpha < H} \alpha(1-2\alpha)$.
2470: Then, the SDS defined in Proposition~\ref{prop:defSDS} has a unique invariant measure $\mu_*$.
2471: Furthermore, there exist positive constants $C$ and $\delta$ such that, for every generalised initial
2472: condition $\mu$, one has
2473: \begin{equ}[e:convergence]
2474: \|\Evol \CQ_t \mu - \Evol \mu_*\|_\TV \le 2\mu\bigl(\{\|x_0\| > e^{\delta t}\}\bigr) + C t^{-\gamma}\;.
2475: \end{equ}
2476: \end{theorem}
2477:
2478: \begin{proof}
2479: The existence of $\mu_*$ follows from \prop{prop:Lyap} and \lem{lem:Lyap}. Furthermore, the assumptions
2480: of \prop{prop:faithful} hold by the invertibility of $\sigma$, so the uniqueness of $\mu_*$ will follow
2481: from \eref{e:convergence}.
2482:
2483: Denote by $\phi$ the SDS constructed in Proposition~\ref{prop:defSDS}, and consider the self-coup\-ling
2484: $\Evol(\mu,\mu_*)$ for $\phi$ constructed in Section~\ref{sec:defCoupl}. We denote by $(x_t,y_t)$
2485: the canonical process associated to $\Evol(\mu,\mu_*)$ and we define a random time $\tilde \tau_\infty$ by
2486: \begin{equ}
2487: \tilde \tau_\infty = \inf \bigl\{t > 0 \,|\, x_s = y_s\, \forall \, s \ge t\bigr\}\;.
2488: \end{equ}
2489: It then follows immediately from \eref{e:TVprop} that
2490: \begin{equ}
2491: \|\Evol \CQ_t \mu - \Evol \mu_*\|_\TV \le 2\prob \sml(\tilde \tau_\infty > t\smr)\;.
2492: \end{equ}
2493: Remember that $\Evol(\mu,\mu_*)$ was constructed as the marginal of the law of a Markov
2494: process with continuous time, living on an augmented phase space $\BigSpace$.
2495: Since we are only interested
2496: in bounds on the random time $\tilde \tau_\infty$ and since we know that $x_s = y_s$ as long
2497: as the coupled system is in the state \step{2}, it suffices to consider the Markov chain $(Z_n,\tau_n)$
2498: constructed in \eref{e:transZn}. It is clear that $\tilde \tau_\infty$ is then dominated by the random time
2499: $\tau_\infty$ defined as
2500: \begin{equ}
2501: \tau_\infty = \inf \bigl\{\tau_n\,|\, S_m = 2\,\forall\, m\ge n\}\;,
2502: \end{equ}
2503: where $S_n$ is the component of $Z_n$ indicating the type of the corresponding step.
2504: Our interest therefore only goes to the dynamic of $\tau_n$ and $S_n$. We define the
2505: sequence of times $t(n)$ by
2506: \begin{equ}[e:deftn]
2507: t(0) = 1\;,\quad t(n+1) = \inf\{m > t(n)\,|\, S_m = 1\}\;,
2508: \end{equ}
2509: and the sequence of durations $\Delta\tau_n$ by
2510: \begin{equ}
2511: \Delta\tau_n = \tau_{t(n+1)} - \tau_{t(n)}\;,
2512: \end{equ}
2513: with the convention $\Delta\tau_n = +\infty$ if $t(n)$ is infinite (\ie if the set in \eref{e:deftn} is empty).
2514: Notice that we set $t(0) = 1$ and not $0$ because we will treat step \step{0} of the coupled process
2515: separately.
2516: The duration $\Delta\tau_n$ therefore measures the time needed by the coupled system starting
2517: in step \step{1} to come back again to step \step{1}. We define the sequence $\xi_n$ by
2518: \begin{equ}
2519: \xi_0 = 0\;,\quad \xi_{n+1} = \cases{-\infty & if $\Delta\tau_n = +\infty$,\cr \xi_n + \Delta\tau_n& otherwise.}
2520: \end{equ}
2521: By construction, one has
2522: \begin{equ}[e:proptauC]
2523: \tau_\infty = \tau_1 + \sup_{n\ge 0}\xi_n\;,
2524: \end{equ}
2525: so we study the tail distribution of the $\Delta\tau_n$.
2526:
2527: For the moment, we leave the value $\alpha$ appearing throughout the paper free, we will tune it at the
2528: end of the proof. Notice also that, by Remarks~\ref{rem:step1} and \ref{rem:step2}, the cost increases by
2529: less than ${1 \over 2\tilde N^2}$ every time the counter $\tilde N$ is increased by $1$. Since the initial
2530: condition has no cost (by the choice \eref{e:definit} of its distribution), this implies that, with probability
2531: $1$, the system is in an admissible state every time step \step{1} is performed.
2532:
2533: Let us first consider the probability of $\Delta\tau_n$ being infinite. By \rem{rem:step1}, the probability for
2534: step \step{1} to succeed is always greater than $\delta$. After step \step{1}, the $N$th occurrence of step \step{2}
2535: has length $2^{N-1}$, and a probability greater than $\max\{\delta', 1-K2^{-\alpha N}\}$ of succeeding.
2536: Therefore, one has
2537: \begin{equ}
2538: \prob\sml(\Delta\tau_n \ge 2^N\smr) \ge \delta \prod_{k=0}^N \max\{\delta', 1-K2^{-\alpha k}\}\;.
2539: \end{equ}
2540: This product always converges, so there exists a constant $p_* > 0$ such that
2541: \begin{equ}
2542: \prob\sml(\Delta\tau_n = \infty\smr) \ge p_*\;,
2543: \end{equ}
2544: for every $n > 0$. Since our estimates are uniform over all admissible initial conditions and
2545: the coupling is chosen in such a way that the system is always in an admissible state at the beginning
2546: of step \step{1}, we actually just proved that the conditional probability of $\prob\sml(\Delta\tau_n = \infty\smr)$
2547: on any event involving $S_m$ and $\Delta\tau_m$ for $m < n$ is bounded from below by $p_*$.
2548:
2549: For $\Delta\tau_n$ to be finite, there has to be a failure of step \step{2} at some point (see \eref{e:algo}).
2550: Recall that if step \step{2} succeeds exactly $N$ times, the corresponding value for $\Delta\tau_n$ will
2551: be equal to $2^N + \tilde t_* 2^{\beta N} (1+n)^{4/(1-2\alpha)}$ for $N>0$ and to $t_* (1+n)^{4/(1-2\alpha)}$ for
2552: $N=0$. This follows from \eref{e:defa2} and \eref{e:defstep2}, noticing that $\tilde N$ in those formulae
2553: counts the number of times step \step{1} occurred and is therefore equal to $n$. We also know that
2554: the probability of the $N$th occurrence of step \step{2} to fail is bounded from above by $K2^{-\alpha N}$.
2555: Therefore, a very crude estimate yields a constant $C$ such that
2556: \begin{equ}
2557: \prob\bigl((1+n)^{-4/(1-2\alpha)} \Delta\tau_n \ge C2^{\beta N} \,\,\text{and}\,\,\Delta\tau_n \neq \infty\bigr) \le K\sum_{k>N}2^{-\alpha k}\;.
2558: \end{equ}
2559: This immediately yields for some other constant $C$
2560: \begin{equ}[e:boundDeltaT]
2561: \prob\bigl((1+n)^{-4/(1-2\alpha)} \Delta\tau_n \ge T \,\,\text{and}\,\,\Delta\tau_n \neq \infty\bigr) \le C T^{-\alpha/\beta}\;.
2562: \end{equ}
2563: As a consequence, the process $\xi_n$ is stochastically dominated by the Markov chain
2564: $\zeta_n$ defined by
2565: \begin{equ}
2566: \zeta_0 = 0\;,\quad \zeta_{n+1} = \cases{-\infty & with probability $p_*$,\cr\zeta_n + (n+1)^{4/(1-2\alpha)}p_n &
2567: with probability $1-p_*$,}
2568: \end{equ}
2569: where the $p_n$ are positive i.i.d.\ random variables with tail distribution $C T^{-\alpha/\beta}$, \ie
2570: \begin{equ}
2571: \prob(p_n \ge T) = \cases{C T^{-\alpha/\beta} & if $C T^{-\alpha/\beta} < 1$, \cr 1 & otherwise.}
2572: \end{equ}
2573: With these notations and using the representation \eref{e:proptauC}, $\tau_\infty$ is bounded by
2574: \begin{equ}[e:boundtauC]
2575: \prob \sml(\tau_\infty > t\smr) \le \prob(\tau_1 > t/2) + \prob \Bigl(\sum_{n=0}^{n_*} (n+1)^{4/(1-2\alpha)} p_n > t/2\Bigr)\;,
2576: \end{equ}
2577: where $n_*$ is a random variable independent of the $p_n$ and such that
2578: \begin{equ}[e:boundnstar]
2579: \prob(n_* = k) = p_*(1-p_*)^k\;.
2580: \end{equ}
2581: In order to bound the second term in \eref{e:boundtauC}, it thus suffices to estimate terms of the form
2582: $\sum_{n=0}^{k} (n+1)^{4/(1-2\alpha)} p_n$ for fixed values of $k$.
2583: Using the Cauchy-Schwartz inequality, one obtains the existence of positive constants $C$
2584: and $N$ such that
2585: \begin{equ}
2586: \prob \Bigl(\sum_{n=0}^{k} (n+1)^{4/(1-2\alpha)} p_n > t/2\Bigr) \le C (k+1)^{N} t^{-\alpha/\beta}\;.
2587: \end{equ}
2588: Combining this with \eref{e:boundnstar} and \eref{e:boundtauC} yields, for some other constant $C$,
2589: \begin{equ}
2590: \prob \sml(\tau_\infty > t\smr) \le \prob(\tau_1 > t/2) + C t^{-\alpha/\beta}\;.
2591: \end{equ}
2592: By the definition of step \step{0} \eref{e:defCoupl0}, we get for $\tau_1$:
2593: \begin{equ}
2594: \prob(\tau_1 > t/2) \le \mu\bigl(\{\|x_0\| > e^{\C2 t/2}/2\}\bigr) + \mu_*\bigl(\{\|y_0\| > e^{\C2 t/2}/2\}\bigr)\;.
2595: \end{equ}
2596: Since, by \prop{prop:Lyap}, the invariant measure $\mu_*$ has bounded moments, the second
2597: term decays exponentially fast. Since $\alpha < \min\{{1\over 2}, H\}$ and $\beta > (1-2\alpha)^{-1}$
2598: are arbitrary, one can realise $\gamma = \alpha/\beta$ for $\gamma$ as in the statement.
2599:
2600: This concludes the proof of \theo{theo:mainTheoFormal}.
2601: \end{proof}
2602:
2603: %As a consequence, we get
2604:
2605: %\begin{proof}[of \theo{theo:existunique}]
2606: %We first show that, even if assumption \textbf{A2'} is dropped, \eref{e:mainequ} possesses a unique
2607: %invariant measure $\mu_*$. The existence of an invariant measure $\mu_*$ follows as above
2608: %from \prop{prop:Lyap} and \lem{lem:Lyap}.
2609: %Given $N>0$, define a cutoff function $\chi_N\from \R^n \to \R$ by
2610: %\begin{equ}
2611: %\chi_N(x) = \chi(\|x\|^2/N^2)\;,
2612: %\end{equ}
2613: %where $\chi\from \R_+ \to \R$ is a smooth, positive, and decreasing function satisfying $\chi(x) = 1$ for $x \le 1$
2614: %and $\chi(x) = 0$ for $x \ge 2$. We define then $f_N\from\R^n\to\R^n$ by
2615: %\begin{equ}
2616: %f_N(x) = \chi_N(x) f(x) - \C2 \sml(1-\chi_N(x)\smr)x\;.
2617: %\end{equ}
2618: %Since $f_N$ satisfies the conditions of \theo{theo:mainTheoFormal}, there exists a unique invariant
2619: %measure $\mu_*^N$ for the SDS associated to the equation
2620: %\begin{equ}[e:mainequN]
2621: %dx = f_N(x)\,dt + \sigma\,dB_H(t)\;.
2622: %\end{equ}
2623:
2624: %\end{proof}
2625:
2626:
2627: We conclude this paper by discussing several possible extensions of our result. The first two extensions
2628: are straightforward and can be obtained by simply rereading the paper carefully and (in the
2629: second case) combining its results with the ones obtained in the references.
2630: The two other extensions are less obvious and merit further investigation.
2631:
2632: \subsection{Noise with multiple scalings}
2633:
2634: One can consider the case where the equation is driven by several independent fBm's with different values of
2635: the Hurst parameter:
2636: \begin{equ}
2637: dx_t = f(x_t)\,dt + \sum_{i=1}^m \sigma_i \, dB^i_{H_i}(t)\;.
2638: \end{equ}
2639: It can be seen that in this case, the invertibility of $\sigma$ should be replaced by the condition that
2640: the linear operator
2641: \begin{equ}
2642: \sigma = \sigma_1 \oplus \sigma_2 \oplus \ldots \oplus \sigma_m \from \R^{mn}\to \R^n\;,
2643: \end{equ}
2644: has rank $n$. The condition on the convergence exponent $\gamma$ then becomes
2645: \begin{equ}
2646: \gamma < \min\sml\{{\gamma_1,\ldots,\gamma_m}\smr\}\;,
2647: \end{equ}
2648: where $\gamma_i = \max_{\alpha < H_i} \alpha(1-2\alpha)$.
2649:
2650: \subsection{Infinite-dimensional case}
2651:
2652: In the case where the phase space for \eref{e:mainequ} is infinite-dimensional, the question of
2653: global existence of solutions is technically more involved and was tackled in \cite{MaslFBM}.
2654: Another technical difficulty arises from the fact that one might want to take for $\sigma$ an
2655: operator which is not boundedly invertible, so \textbf{A3} would fail on a formal level.
2656: One expects to be able to overcome this difficulty at least in the case where the equation is semilinear
2657: and parabolic, \ie of the type
2658: \begin{equ}
2659: dx = Ax\,dt + F(x)\,dt + Q\,dB_H(t)\;,
2660: \end{equ}
2661: with the domain of $F$ ``larger'' (in a sense to be quantified) than the domain of $A$
2662: and $B_H$ a cylindrical fBm on some Hilbert space $\CH$ on which the solution
2663: is defined, provided the eigenvalues of $A$ and of $Q$ satisfy some compatibility
2664: condition as in \cite{ZDP1,Cerr,EH3}.
2665:
2666: On the other hand, it is possible in many cases to split the phase space into
2667: a finite number of ``unstable modes'' and an infinite number of ``stable modes'' that
2668: are slaved to the unstable ones. In this situation, it is sufficient to construct step \step{1}
2669: in such a way that the unstable modes meet, since the stable ones will then automatically
2670: converge towards each other. A slight drawback of this method is that the convergence
2671: towards the stationary state no longer takes place in the total variation distance.
2672: We refer to \cite{MatNS,ArmenKuk,HExp02} for implementations
2673: of this idea in the Markovian case.
2674:
2675: \subsection{Multiplicative noise}
2676:
2677: In this case, the problem of existence of global solutions can already be hard. In the case $H>1/2$, the
2678: fBm is sufficiently regular, so one obtains pathwise existence of solutions by rewriting \eref{e:mainequ} in
2679: integral form and interpreting the stochastic integral pathwise as a Riemann-Stieltjes integral.
2680: In the case $H\in ({1\over 4}, {1\over 2})$,
2681: it has been shown recently \cite{MR96b:60150,MR2000c:60089,MR2003c:60066} that pathwise
2682: solutions can also be obtained by realising the fBm as
2683: a geometric rough path. More refined probabilistic estimates are required in the analysis
2684: of step \step{1} of our coupling construction. The equivalent of equation \eref{e:equrho}
2685: then indeed contains a multiplicative noise term, so the deterministic estimate \eref{e:estrho} fails.
2686:
2687: \subsection{Arbitrary Gaussian noise}
2688:
2689: Formally, white noise is a centred Gaussian process $\xi$ with correlation function
2690: \begin{equ}
2691: \expect \xi(s)\xi(t) = C_w(t-s) = \delta(t-s)\;.
2692: \end{equ}
2693: The derivative of the fractional Brownian motion with Hurst parameter $H$ is formally also a centred Gaussian
2694: process, but its correlation function is proportional to
2695: \begin{equ}
2696: C_H(t-s) = |t-s|^{2H-2}\;,
2697: \end{equ}
2698: which should actually be interpreted as the second derivative of $|t-s|^{2H}$ in the sense of distributions.
2699:
2700: A natural question is whether the results of the present paper also apply to differential equations driven by
2701: Gaussian noise with an arbitrary correlation function $C(t-s)$. There is no conceptual obstruction to
2702: the use of the method of proof presented in this paper in that situation, but new estimates are required.
2703: It relies on the fact that the driving process is a fractional Brownian motion only to be able to explicitly
2704: perform the computations of Section~\ref{sec:defCoupl}.
2705:
2706: \bibliographystyle{Martin}
2707: \markboth{\sc \refname}{\sc \refname}
2708: \bibliography{refs}
2709: \end{document}