0805.2002/m.tex
1: 
2: \documentclass[12pt]{article}
3: 
4: \newcommand{\ignore}[1]{}
5: 
6: \usepackage{amsmath, amsthm, amsfonts, amssymb}
7: 
8: %\usepackage[active]{/usr/share/doc/HTML/en/kdvi/srcltx}
9: 
10: \usepackage[dvips,final]{graphicx}
11: %\pagestyle{headings}
12: 
13: \usepackage{graphicx}
14: 
15: %\usepackage{subfigure}
16: \usepackage{float}
17: \newfloat{figure}{H}{lof}
18: \floatname{figure}{\figurename}
19: 
20: \numberwithin{figure}{section}
21: 
22: \DeclareMathAlphabet{\eufrak}{U}{}{}{}  % Euler fraktur math
23: \SetMathAlphabet\eufrak{normal}{U}{euf}{m}{n}
24: \SetMathAlphabet\eufrak{bold}{U}{euf}{b}{n}
25: 
26: \numberwithin{equation}{section}
27: 
28: \newenvironment{Proof}{\removelastskip\par\medskip
29: \noindent{\em Proof.}
30: \rm}{\hfill$\square$\\\par\medbreak}
31: 
32: \def\div{{\mathord{{\rm div}}}}
33: \def\Ad{{\mathord{{\rm Ad}}}}
34: \def\ad{{\mathord{{\rm ad}}}}
35: \def\trace{{\mathord{{\rm trace ~}}}}
36: \def\ric{{\mathord{{\rm ric}}}}
37: \def\so{{\mathord{{\rm so}}}}
38: \def\realic{{\mathord{{\rm Ric}}}}
39: \def\Id{{\mathord{{\rm Id}}}}
40: \def\diff{{\mathord{{\rm Diff}}}}
41: \def\real{{\mathord{{\rm I\kern-2.8pt R}}}}        % Fake blackboard bold R.
42: \def\inte{{\mathord{{\rm I\kern-2.8pt N}}}}
43: \def\PP{{\mathord{{\rm I\kern-2.8pt P}}}}
44: \def\ii{{\mathord{{\rm {\Im}}}}}
45: \def\G{{\mathord{{\sl {\sf G}}}}}
46: 
47: \def\real{{\mathord{\mathbb R}}}
48: \def\bbr{{\mathord{\mathbb R}}}
49: \def\inte{{\mathord{\mathbb N}}}
50: \def\z{{\mathord{\mathbb Z}}}
51: \def\qu{{\mathord{\mathbb Z}}}
52: \def\H{{\mathord{\mathbb H}}}
53: \def\HB{{\mathord{\mathbb{\bf H}}}}
54: \def\calf{{\cal F}}
55: \newcommand{\dee}{\mbox{$I  \! \! \! \, D$}}
56: \def\Cov{{\mathrm{{\rm Cov}}}}
57: \def\Dom{{\mathrm{{\rm Dom}}}}
58: \def\Var{{\mathrm{{\rm Var}}}}
59: \def\grad{{\mathrm{ \ \! {\rm grad}}}}
60: \def\crb{{\mathrm{{\rm R}}}}
61: \def\R{\right}
62: \def\L{\left}
63: \def\realef#1{(\ref{#1})}
64: 
65: 
66: %%%%%%%%%%%%%%%%%%%%%   def    JCB
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: 
69: \newcommand{\simgeq}{\;
70: \raisebox{-0.4ex}{\tiny$\stackrel{{\textstyle>}}{\sim}$}\;}
71: \newcommand{\simleq}{\;
72: \raisebox{-0.4ex}{\tiny$\stackrel{{\textstyle<}}{\sim}$}\;}
73: \newcommand{\theor}{{\bf Theorem}}
74: \newcommand{\etape}{{\bf Step}}
75: \newcommand{\rem}{{\bf Remark:}}
76: \newcommand{\rems}{{\bf Remarks:}}
77: \newcommand{\ex}{{\bf Example:}}
78: \newcommand{\demo}{{\bf Proof:}}
79: \newcommand{\CQFD}{\nolinebreak\hfill\rule{2mm}{2mm}\medbreak\par}
80: \newcommand{\lem}{{\bf Lemma }}
81: \newcommand{\ind}{\mathbf{1}}
82: \newcommand{\sgn}{\mathrm{sign }\:}
83: \newcommand{\disp}{\displaystyle}
84: \newcommand{\cvar}{\stackrel{var}{\longrightarrow}}
85: \newcommand{\lto}{\longrightarrow}
86: \newcommand{\e}{\varepsilon}
87: \newcommand{\s}{\sigma}
88: \newcommand{\om}{\omega}
89: \newcommand{\vide}{{}}
90: 
91: 
92: 
93: \def\rit{\mathbb{R}}
94: \def\cit{\mathbb{C}}
95: \def\nit{\mathbb{N}}
96: \def\zit{\mathbb{Z}}
97: \def\qit{\mathbb{Q}}
98: \def\dit{\mathbb{D}}
99: \def\Eit{\mathbb{E}}
100: \def\pit{\mathbb{P}}\def\P{\mathbb{P}}
101: \def\E{\mathop{\hbox{\rm I\kern-0.20em E}}\nolimits}
102: \def\Var{\mathop{\hbox{\rm Var}}\nolimits}
103: \def\Cov{\mathop{\hbox{\rm Cov}}\nolimits}
104: 
105: \def\cqfd{\hbox{\rule{2.5mm}{2.5mm}}}
106: 
107: \def\court {\hskip 5pt}
108: \def\med{\hskip 10pt}
109: \def\lng{\hskip 20pt}
110: 
111: 
112: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
113: %
114: 
115: 
116: \newtheorem{prop}{Proposition}[section]
117: \newtheorem{assumption}{Assumption}[section]
118: \newtheorem{lemma}[prop]{Lemma}
119: \newtheorem{definition}[prop]{Definition}
120: \newtheorem{corollary}[prop]{Corollary}
121: \newtheorem{theorem}[prop]{Theorem}
122: \newtheorem{remark}[prop]{Remark}
123: \def\Dom{\rm Dom \ \! }
124: \def\trace{{\mathrm{{\rm trace}}}}
125: \def\Ent{{\mathrm{{\rm Ent}}}}
126: \def\var{{\mathrm{{\rm Var}}}}
127: \def\div{{\mathrm{{\rm div}}}}
128: \def\argmax{{\mathrm{{\rm argmax}}}}
129: 
130: \textwidth15.3cm
131: \textheight21.5cm
132: 
133: \oddsidemargin0.5cm
134: \evensidemargin0.5cm
135: \topmargin1cm
136: 
137: %\oddsidemargin1.5in
138: %\evensidemargin1.5in
139: %\topmargin1cm
140: 
141: \headheight0cm
142: \headsep0cm
143: \baselineskip1in
144: \parindent0.2in
145: 
146: 
147: \title{
148: \Huge
149:  Stochastic analysis on Gaussian space applied to drift estimation
150: }
151: 
152: \author
153: {
154: Nicolas Privault\footnote{nicolas.privault@math.univ-poitiers.fr}
155: \\
156: Laboratoire de Math\'ematiques
157: \\
158: Universit\'e de Poitiers
159: \\
160: T\'el\'eport 2 - BP 30179
161: \\
162: 86962 Chasseneuil Cedex
163: \\
164: France
165: \and
166: Anthony R\'eveillac\footnote{anthony.reveillac@univ-lr.fr}
167: \\
168: Laboratoire de Math\'ematiques
169: \\
170: Universit\'e de La Rochelle
171: \\
172: Avenue Michel Cr\'epeau
173: \\
174: 17042 La Rochelle Cedex
175: \\
176: France
177: }
178: 
179: %\date{}
180: 
181: \date{august 29, 2007}
182: 
183: \allowdisplaybreaks
184: 
185: \begin{document}
186: \hyphenation{func-tio-nals}
187: \maketitle
188: 
189: \begin{abstract}
190:  In this paper we consider the nonparametric
191:  functional estimation of the drift
192:  of Gaussian processes using Paley-Wiener
193:  and Karhunen-Lo\`eve expansions.
194:  We construct efficient estimators for
195:  the drift of such processes,
196:  and prove their minimaxity using Bayes estimators.
197:  We also construct superefficient estimators of Stein type
198:  for such drifts using the Malliavin integration by parts formula
199:  and stochastic analysis on Gaussian space, in which superharmonic
200:  functionals of the process paths play a particular role.
201:  Our results  are illustrated by numerical simulations
202:  and extend the construction of James-Stein type estimators
203:  for Gaussian processes by Berger and Wolpert \cite{berger1}.
204: \end{abstract}
205: 
206: \normalsize
207: 
208: \vspace{0.5cm}
209: 
210: \small \noindent {\bf Key words:}
211:  Nonparametric drift estimation, Stein estimation, Gaussian space,
212:  Malliavin calculus, harmonic analysis.
213: \\
214: {\em Mathematics Subject Classification:} 62G05, 60H07, 31B05.
215: 
216: %\tableofcontents
217: 
218: \normalsize
219: 
220: \baselineskip0.7cm
221: 
222: \section{Introduction}
223: 
224: \noindent
225:  The maximum likelihood estimator $\hat{\mu}$ of
226:  the mean $\mu\in \real^d$ of a Gaussian random vector $X$ in $\real^d$
227:  with covariance $\sigma^2 {\rm I}_{\real^d}$ under a probability $\P_\mu$
228:  is well-known to be equal to $X$ itself,
229:  and can be computed by maximizing the likelihood ratio
230: $$
231: \frac{1}{(2\pi\sigma^2)^{d/2}} e^{-\frac{\Vert X-m \Vert^2_d}{2\sigma^2}}
232: $$
233:  with respect to $m$,
234:  where $\Vert \cdot \Vert_d$ denotes the Euclidean norm on $\real^d$.
235:  It is efficient in the sense that it attains the Cramer-Rao bound
236: $$
237:  \sigma^2 d
238:  =
239:  \E_\mu [ \Vert X - \mu \Vert_d^2 ]
240:  = \inf_Z
241:  \E_\mu [ \Vert Z - \mu \Vert_d^2 ]
242: , \qquad
243:  \mu \in \real^d,
244: $$
245:  over all unbiased estimators $Z$ satisfying
246:  $\E_\mu [Z] = \mu$, for all $\mu \in \real^d$.
247: \\
248: 
249: \noindent
250:  In \cite{jamesstein}, James and Stein have constructed
251:  superefficient estimators for the mean of $X \in \real^d$,
252:  of the form
253: $$
254:  \left(
255:  1- \frac{d-2}{\Vert X \Vert_d^2}
256:  \right)
257:  X
258: $$
259:  whose risk is lower than the Cramer Rao bound $\sigma^2 d$
260:  in dimension $d\geq 3$.
261: \\
262: 
263: \noindent
264:  Drift estimation for Gaussian processes is of interest in several fields of
265:  application.
266:  For example in the decomposition
267: $$
268:  X_t=X^u_t+u_t, \qquad t \in [0,T],
269: $$
270:  the process $(X_t)_{t\in [0,T]}$ is interpreted as
271:  an observed output signal, the drift $(u_t)_{t \in [0,T]}$
272:  is viewed as an input signal to be estimated and perturbed by a
273:  centered Gaussian noise $(X^u_t)_{t \in [0,T]}$,
274:  cf. e.g. \cite{IbragimovRozanov}, Ch.~VII.
275:  Such results find applications in e.g. telecommunication
276:  (additive Gaussian channels) and finance (identification of market trends).
277: \\
278: 
279: \noindent
280:  Berger and Wolpert \cite{berger1}, \cite{berger2}, have constructed estimators
281:  of James-Stein type for the drift of a Gaussian process $(X_t)_{t\in [0,T]}$
282:  by applying the James-Stein procedure to the independent Gaussian random variables
283:  appearing in the Karhunen-Lo\`eve expansion of the process.
284:  In this context, $\hat{u} : = (X_t)_{t\in \real_+}$ is seen as a minimax
285:  estimator of its own drift $(u_t)_{t\in \real_+}$.
286: \\
287: 
288: \noindent
289:  Stein~\cite{stein} has shown that the James-Stein estimators on $\real^d$
290:  could be extended to a wider family of estimators, using integration by parts for
291:  Gaussian measures.
292:  Let us briefly recall Stein's argument, which relies on integration by parts
293:  with respect to the Gaussian density
294:  and on the properties of superharmonic functionals for the Laplacian on $\real^d$.
295:  Given an estimator of $\mu \in \real^d$ of the form $X+g(X)$, where
296:  $g: \real^d \to \real^d$ is sufficiently smooth,
297:  and applying the integration by parts formula
298: \begin{equation}
299: \label{ibp2}
300:  \E_\mu [(X_i - \mu_i ) g_i (X)] =
301:  \sigma^2
302:  \E_\mu [\partial_i g_i (X)],
303: \end{equation}
304:  $g = \sigma^2 \grad \log f = \sigma^2 ( \partial_1 \log f , \ldots , \partial_d \log f )$,
305:  one obtains
306: $$
307: \E_\mu [\Vert X +
308:  \sigma^2 \grad \log f (X)
309:  -
310:  \mu \Vert_d^2] = \sigma^2 d
311:  +
312:  4 \sigma^4
313:  \sum_{i=1}^d
314:  \E_\mu \left[ \frac{
315:  \partial^2_i \sqrt{f} (X)}{\sqrt{f} (X)} \right]
316: ,
317: $$
318:  i.e. $X+ \sigma^2 \grad \log f (X)$ is a superefficient estimator if
319: $$
320:  \sum_{i=1}^d \partial^2_i \sqrt{f} (x) < 0, \qquad dx-a.e.
321: ,
322: $$
323:  which is possible if $d\geq 3$.
324:  In this case, $X+ \sigma^2 \grad \log f (X)$ improves in the mean square sense
325:  over the efficient estimator $\hat{u}$ which
326:  attains the Cramer-Rao bound $\sigma^2 d$ on unbiased estimators of $\mu$.
327: \\
328: 
329: \noindent
330:  In this paper we present an extension of Stein's argument to an
331:  infinite-dimensional setting using the Malliavin integration by
332:  parts formula, with application to the construction of Stein type
333:  estimators for the drift of a Gaussian process $(X_t)_{t\in [0,T]}$.
334:  Our approach applies to Gaussian processes such as Volterra processes
335:  and fractional Brownian motions.
336:  It also extends the results of Berger and Wolpert \cite{berger1}
337:  in the same way that the construction of Stein~\cite{stein} extends that
338:  of James and Stein~\cite{jamesstein},
339:  and this allows us to recover the estimators of James-Stein type
340:  introduced by Berger and Wolpert \cite{berger1} as particular cases.
341:  Here we replace the Stein equation
342:  \eqref{ibp2} with the integration by parts formula of the Malliavin
343:  calculus on Gaussian space. Our estimators are given by processes of the
344:  form
345: $$
346:  X_t + D_t \log F, \qquad
347:  t\in [0,T]
348: ,
349: $$
350:  where $F$ is a positive superharmonic random variable on
351:  Gaussian space and $D_t$ is the Malliavin derivative indexed by $t\in [0,T]$.
352:  In contrast to the minimax estimator $\hat{u}$, such estimators
353:  are not only biased but also anticipating with respect to the Brownian filtration
354:  $({\cal F}_t)_{t\in [0,T]}$.
355:  This however poses no problem when one has access to complete paths
356:  from time $0$ to $T$.
357: \\
358: 
359: \noindent
360:  For large values of $\sigma$ it can be shown that the percentage
361:  gain of this estimator is at least equal to the universal constant
362: \begin{equation}
363: \label{gsint1}
364: \frac{16}{\pi^4}
365:  \int_{\real^4}
366:  e^{-\frac{x^2+y^2+z^2+r^2}{2}}
367:  \frac{ dx dy dz dr}{x^2+9y^2+25z^2+49r^2}
368: \end{equation}
369:  which approximately represents $11.38\%$, see \eqref{gsint} below.
370: \\
371: 
372: \noindent
373:  We proceed as follows.
374:  In Section~\ref{2} we use stochastic calculus in
375:  the independent increment case to derive a Cramer-Rao bound over
376:  all unbiased drift estimators.
377:  This bound is attained by the process $\hat{u}:= (X_t)_{t\in [0,T]}$,
378:  which will be considered as an efficient drift estimator.
379:  In Section~\ref{2.0} we compute the Bayes estimators obtained
380:  under prior Gaussian distributions.
381:  We show that these Bayes estimators are admissible, and use
382:  them to prove that the drift estimator $\hat{u}$ is minimax.
383:  The tools and results presented in Sections~\ref{2} and \ref{2.0}
384:  are not surprising, but we did not find any source covering them in the literature.
385:  In Section~\ref{2.1} we recall the elements of
386:  analysis and integration by parts on Gaussian space
387:  which will be needed in Section~\ref{3}
388:  to construct superefficient drift estimators for Gaussian processes
389:  using superharmonic random functionals on Gaussian space.
390:  The superefficiency of these estimators will show, as in the
391:  classical case, that the minimax estimator $\hat{u}$ is not admissible.
392:  In Section~\ref{app} we give examples of nonnegative superharmonic
393:  functionals using cylindrical functionals and potential theory on Gaussian space.
394:  Examples are considered in Section~\ref{4} in case $u$ is deterministic.
395:  We show that the James-Stein estimators of Berger and Wolpert \cite{berger1}
396:  can be recovered as particular cases in our approach,
397:  and we provide numerical simulations for the gain of such estimators.
398:  It turns out that in those examples,
399:  the gain obtained in comparison with the minimax
400:  estimator $(X_t)_{t\in [0,T]}$
401:  is a function of $\sigma^2/T$, thus making $\sigma$ and $T$ play
402:  inverse roles, unlike in the usual setting of Brownian rescaling.
403: \\
404: 
405: \noindent
406:  This paper is an extended version of \cite{pr3} and provides
407:  proofs of the results presented in \cite{p-r-c}.
408: \section*{Notation}
409: %\label{2}
410:  Let $T>0$.
411:  Consider a real-valued centered Gaussian process $(X_t)_{t\in [0,T]}$
412:  with covariance function
413: $$
414:  \gamma (s,t) = \E [ X_s X_t ], \qquad s,t\in [0,T],
415: $$
416:  on a probability space $(\Omega , {\cal F} , \P )$,
417:  where ${\cal F}$ is the $\sigma$-algebra generated by $X$.
418:  Recall that $(X_t)_{t\in [0,T]}$ can be represented in different ways
419:  as an isonormal Gaussian process
420:  on a real separable Hilbert space $H$, i.e. as an isometry
421:  $X : H \to L^2 (\Omega , {\cal F} , P)$ such that
422:  $\{ X (h) \ : \ h\in H\}$ is a family of centered
423:  Gaussian random variables satisfying
424: $$\E
425:  [ X (h) X (g) ]
426:  = \langle h , g \rangle_H,
427:  \qquad
428:  h,g \in H
429: ,
430: $$
431:  where $\langle \cdot , \cdot \rangle_H$ and $\Vert \cdot \Vert_H$
432:  denote the scalar product and norm on $H$.
433: \\
434: 
435: \noindent
436:  One can distinguish two main types of such isonormal representations
437:  of $X_t$, see e.g. \cite{amn} and \cite{berger1} respectively for details.
438: \begin{description}
439: \item{\bf (A)}
440:  Paley-Wiener expansions.
441:  In this case, $H$ is the completion
442:  of the linear space generated by the functions
443:  $\stackrel{}{\chi}_t \! \! (s) =\min (s,t)$, $s,t\in [0,T]$,
444:  with respect to the norm
445: $$
446:  \langle
447:  \stackrel{}{\chi}_t
448:  ,
449:  \stackrel{}{\chi}_s \rangle_H : = \gamma (s,t)
450: ,
451:  \qquad
452:  s, t \in [0,T]
453: ,
454: $$
455:  and $ X (\cdot )$ is constructed on $H$ from
456:  $X ( \stackrel{}{\chi}_t ) : = X_t$, $t\in [0,T]$,
457:  i.e. we have
458: $$
459:  X ( \stackrel{}{\chi}_t )
460:  =
461:  \sum_{k=0}^\infty
462:  \langle
463:  \stackrel{}{\chi}_t
464:  ,
465:  h_k (t)
466:  \rangle_H
467:  X (h_k)
468: , \qquad
469:  t\in [0,T],
470: $$
471:  for any orthonormal basis $(h_k)_{k\in\inte}$ of $H$.
472:  Assume in addition $\gamma (s,t)$ has the form
473: $$
474:  \gamma (s,t) = \int_0^{s\wedge t} K(t,r)K(s,r) dr
475: ,
476:  \qquad
477:  s,t\in [0,T]
478: ,
479: $$
480:  where $K(\cdot , \cdot )$ is a deterministic kernel and
481: $$
482:  (Kh)(t)
483:  : =
484:  \int_0^t K(t,s) \dot{h} (s) ds
485: $$
486:  is differentiable in $t\in [0,T]$,
487:  and let $K^*$ denote the adjoint of $K$ with respect to
488: $$
489:  \langle h , g \rangle
490:  : = \langle \dot{h} , \dot{g} \rangle_{L^2([0,T] , dt )}.
491: $$
492:  The scalar product in $H$ then satisfies
493: $$
494:  \langle h , g \rangle_H
495:  = \langle K^* h , K^* g \rangle
496:  = \langle h , \Gamma g \rangle
497: ,
498: $$
499:  where $\Gamma = K K^*$, and we have the decomposition
500: $$
501:  X_t
502:  =
503:  \sum_{k=0}^\infty
504:  \langle
505:  \stackrel{}{\chi}_t
506:  ,
507:  h_k
508:  \rangle_H
509:  X (h_k)
510:  = \sum_{k=0}^\infty
511:  \langle 1_{[0,t]}
512:  ,
513:  \dot{\Gamma} h_k
514:  \rangle_{L^2([0,T] , dt )}
515:  X (h_k)
516:  = \sum_{k=0}^\infty
517:  \Gamma h_k (t)
518:  X (h_k)
519: ,
520: $$
521:  $t\in [0,T]$.
522:  In this case we also have the representation
523: $$
524:  X_t
525:  = \int_0^t K(t,s) dW_s,
526:  \qquad
527:  t\in [0,T],
528: $$
529:  where $(W_s)_{s\in [0,T]}$ is a standard Brownian
530:  motion, cf. \cite{amn}.
531: \item{\bf (B)}
532:  Karhunen-Lo\`eve expansions. This framework is used in \cite{berger1}.
533:  In this case, $\mu$ is a finite Borel measure on $[0,T]$ and
534:  $H$ is defined from
535: $$
536:  \langle
537:  h
538:  ,
539:  g
540:  \rangle_H
541:  =
542:  \langle
543:  h
544:  ,
545:  \Gamma
546:  g
547:  \rangle
548: ,
549: $$
550:  where
551: $$\langle h , g \rangle
552:  : =
553:  \langle h , g \rangle_{L^2([0,T],d\mu )}
554: ,
555: $$
556:  and
557: $$
558:  ( \Gamma g ) (t)
559:  =
560:  \int_0^T g(s) \gamma (s,t) \mu (ds ),
561: \qquad
562:  t\in [0,T]
563: ,
564: $$
565:  with
566: $$
567:  X (h) = \int_0^T X_s h(s) \mu (ds), \qquad h\in H
568: .
569: $$
570:  Given $(h_k)_{k\in\inte}$ an orthonormal basis of
571:  $L^2([0,T],d\mu)$, we have the expansion
572: $$X_t = \sum_{k=0}^\infty h_k (t) X (h_k)
573: ,
574:  \qquad
575:  t\in [0,T]
576: .
577: $$
578: \end{description}
579:  In the sequel we will use mainly the framework $(A)$ with
580:  $\langle h , g \rangle = \langle \dot{h} , \dot{g} \rangle_{L^2([0,T] , dt )}$,
581:  which is
582:  better adapted to our approach, although some results valid
583:  in the general framework of Gaussian processes will be valid
584:  for $(B)$ as well.
585: \\
586: 
587: \noindent
588:  The Girsanov theorem for Gaussian processes,
589:  cf. e.g. \cite{nualartm2}, states that $X^u$ defined as
590: %correction ci-dessous: gamma^{-1} au lieu de gamma:
591: $$
592:  X^u(g) : = X(g) - \langle g , u \rangle
593:  = X(g) - \langle g , \Gamma^{-1} u \rangle_H,
594:  \qquad g\in H,
595: $$
596:  where $u\in H$ is deterministic,
597:  has same law as $X$ under the probability $\P_u$ defined by
598: $$
599:  \frac{d\P_u}{d\P}
600:  = \exp \left( X( \Gamma^{-1} u ) - \frac{1}{2}
601:  \Vert \Gamma^{-1} u \Vert_H^2
602:  \right)
603: .
604: $$
605:  In other terms, in case $(A)$ we have
606: \begin{eqnarray*}
607:  X_t - u(t)
608:  & = &
609:  X_t - \Gamma \Gamma^{-1} u(t)
610: \\
611:  & = &
612:  \sum_{k=0}^\infty
613:  \Gamma h_k (t)
614:  ( X(h_k) - \langle h_k, \Gamma^{-1} u\rangle_H )
615: \\
616:  & = &
617:  \sum_{k=0}^\infty
618:  \Gamma h_k (t) X^u (h_k)
619: , \qquad
620:  t\in [0,T]
621: ,
622: \end{eqnarray*}
623:  where $(h_k)_{k\in\inte}$ is orthonormal basis of $H$,
624:  and in case $(B)$,
625: \begin{eqnarray*}
626:  X_t - u(t)
627:  & = &
628:  \sum_{k=0}^\infty
629:  h_k (t)
630:  (
631:  X( h_k) - \langle h_k, u \rangle_{L^2([0,T],d\mu)}
632:  )
633: \\
634:  & = &
635:  \sum_{k=0}^\infty
636:  h_k (t)
637:  (
638:  X(h_k) - \langle h_k, \Gamma^{-1} u \rangle_H
639:  )
640: \\
641:  & = &
642:  \sum_{k=0}^\infty
643:  h_k (t) X^u (h_k)
644: , \qquad
645:  t\in [0,T]
646: ,
647: \end{eqnarray*}
648:  where $(h_k)_{k\in\inte}$ is orthonormal basis of
649:  $L^2([0,T],d\mu)$.
650: \section{Efficient drift estimator}
651: \label{2}
652:  Here we work in the framework of $(A)$, in the particular case where
653:  $(X_t)_{t\in [0,T]}$ has independent increments,
654:  i.e.
655: $$\gamma (s,t) = \int_0^{s\wedge t} \sigma^2_u du
656: ,
657: $$
658:  where $\sigma \in L^2([0,T] , dt )$ is an a.e. non-vanishing function,
659: $$(\dot{K} h)(t) = (\dot{K}^* h )(t) = \sigma_t \dot{h} (t), \qquad
660:  t\in [0,T],
661: $$
662:  with $K(t,r) = 1_{[0,t]} (r) \sigma_r$ and
663: $$
664:  \Gamma h (t) = \int_0^t \dot{h}_s \sigma_s^2 ds,
665:  \qquad
666:  t\in [0,T].
667: $$
668:  In other terms, $(X_t )_{t \in [0,T]}$ is a continuous Gaussian martingale
669:  with
670:  quadratic variation $\sigma^2_t dt$,
671:  which can be represented as the time change
672: $$X_t = W_{\int_0^t \sigma^2_s ds}, \qquad
673:  t\in [0,T],
674: $$
675:  of the standard Brownian motion $(W_t)_{t\in \real_+}$,
676:  or as the stochastic integral process $\displaystyle
677:  X_t = \int_0^t \sigma_s dW_s$, $t\in [0,T]$, and we have
678:  $\displaystyle X (h) = \int_0^T \dot{h}(s) dX_s$,
679:  $h \in H$, where
680: $$H
681:  = \left\{
682:  v : [0,T] \to \real
683:  \ : \
684:  v (t) = \int_0^t \dot{v}(s) ds,
685:  \ t\in [0,T],
686:  \
687:  \dot{v} \in L^2 ([0,T], \sigma^2_t dt)
688:  \right\}$$
689:  is the Cameron-Martin space with inner product
690: $$
691:  \langle v_1 , v_2 \rangle_H
692:  =
693:  \int_0^T \dot{v}_1 (s) \dot{v}_2 (s) \sigma^2_s ds,
694:  \qquad v_1, v_2 \in H
695: .
696: $$
697: 
698: \noindent
699:  Let $({\cal F}_t)_{t\in [0,T]}$ denote the filtration generated
700:  by $(X_t)_{t\in [0,T]}$, and for $u$ an ${\cal F}_t$-adapted process,
701:  let $\P^\sigma_u$ denote the translation of the Wiener measure
702:  on $\Omega$ by $u$, i.e. $\P^\sigma_u$ is the measure on $\Omega$ under which
703: $$X^u_t : = X_t - u_t
704:  , \qquad t \in [0,T],
705: $$
706:  is a continuous Gaussian martingale with quadratic variation
707: $$d\langle X^u , X^u \rangle_t = \sigma^2_t dt.
708: $$
709:  Consider $u$ an ${\cal F}_t$-adapted processes of the form
710: $$
711:  u_t = \int_0^t \dot{u}_s ds, \qquad t\in [0,T]
712: ,
713: $$
714:  with
715: $$
716:  \E^\sigma
717:  \left[
718:  \int_0^T \frac{\dot{u}^2_s}{\sigma^2_s} ds
719:  \right]
720:  < \infty
721: .
722: $$
723:  By the Girsanov theorem, $\P^\sigma_u$ is absolutely continuous with
724:  respect to $\P^\sigma$, with
725: $$d\P^\sigma_u = \Lambda ( u ) d\P^\sigma,
726: $$
727:  where
728: $$
729:  \Lambda ( u )
730:  : =
731:  \exp \left(
732:  \int_0^T
733:  \frac{\dot{u}_s}{\sigma^2_s}
734:  dX_s
735:  -
736:  \frac{1}{2} \int_0^T \frac{\dot{u}_s^2}{\sigma^2_s} ds \right)
737: $$
738:  denotes the Girsanov-Cameron-Martin density,
739:  the canonical process
740:  $(X_t)_{t\in [0,T]}$ becomes a continuous Gaussian semimartingale
741:  under $\P^\sigma_u$,
742:  with quadratic variation $\sigma^2_t dt$ and  drift $\dot{u}_t dt$.
743:  The expectation under $\P_u$ will be denoted by $\E_u$.
744: \begin{definition}
745:  A drift estimator $\xi$ is called unbiased if
746: $$
747:  \E_u [\xi_t]= \E_u [u_t], \quad t \in [0,T],
748: $$
749:  for all square-integrable ${\cal F}_t$-adapted process
750:  $(u_t)_{t\in [0,T]}$.
751:  It is called adapted if the process $(\xi_t)_{t\in [0,T]}$ is
752:  ${\cal F}_t$-adapted.
753:  %  generated by $(X^u_t)_{t\in [0,T]}$.
754: \end{definition}
755: \noindent
756:  Here, the canonical process $(X_t)_{t\in [0,T]}$ will be considered
757:  as an unbiased estimator of own its drift $(u_t)_{t\in [0,T]}$
758:  under $\P^\sigma_u$, with risk defined as
759: $$
760:  \E^\sigma_u \left[
761:  \Vert X - u \Vert_{L^2([0,T] , d\mu )}^2
762:  \right]
763:  =
764:  \int_0^T \E^\sigma_u
765:  \left[
766:  | X^u_t |^2 \right]
767:  \mu ( dt )
768:  =
769:  \int_0^T \int_0^t \sigma^2_s ds
770:  \mu ( dt )
771: ,
772: $$
773:  where $\mu$ is a finite Borel measure on $[0,T]$.
774:  Clearly this estimator is consistent as $\sigma$ or $T$ tend to $0$:
775:  precisely, given $N$ independent samples
776: $$(X_t^1 )_{t\in [0,T]}, \ldots , (X_t^N )_{t\in [0,T]},
777: $$
778:  of $(X_t)_{t\in [0,T]}$, the process
779: \begin{equation}
780: \label{xbar}
781:  \bar{X}_t
782:  : = \frac{
783:  X_t^1 + \cdots + X_t^N
784:  }{N}
785: ,
786:  \qquad
787:  t\in [0,T],
788: \end{equation}
789:  is an unbiased estimator of $(u_t)_{t\in [0,T]}$ whose risk
790: $$
791:  \E^{\sigma/\sqrt{N}}_u
792:  \left[
793:  \Vert \bar{X} - u \Vert_{L^2([0,T] , d\mu )}^2
794:  \right]
795:  =
796:  \frac{1}{N}
797:  \int_0^T \int_0^t \sigma^2_s ds  \mu ( dt )
798: $$
799:  converges to zero as $N$ goes to infinity.
800: \\
801: 
802: \noindent
803:  The justification of the use of $\hat{u}=(X_t)_{t\in [0,T]}$ as an
804:  efficient estimator
805:  comes from the following proposition which
806:  allows us to compute a Cramer-Rao bound attained by $\hat{u}$.
807:  Here the parameter space is restricted to the space
808:  of adapted processes in $L^2 (\Omega \times [0,T] , \P \otimes \mu )$,
809:  which corresponds in a sense to a parametric estimation.
810: \begin{prop}
811: \label{prop:CramerRaoBound}
812:  {Cramer-Rao inequality.}
813:  For any unbiased and adapted estimator $\xi$ of
814:  $u$ we have
815: \begin{equation}
816: \label{11}
817:  \E^\sigma_u \left[ \int_0^T |\xi_t - u_t|^2  \mu ( dt ) \right]
818:  \geq
819:  \crb (\sigma,\mu, \hat{u} )
820: ,
821: \end{equation}
822:  where $u \in L^2 (\Omega \times [0,T] , \P_u^\sigma \otimes \mu )$
823:  is adapted and the Cramer-Rao type bound
824: $$
825:  \crb (\sigma,\mu, \hat{u} )
826:  := \int_0^T \int_0^t \sigma^2_s ds \mu ( dt )
827: $$
828:  is independent of $u$ and attained by the
829:  efficient estimator $\hat{u}=X$.
830: \end{prop}
831: \begin{Proof}
832:  Since $\xi$ is unbiased, for all $\zeta\in H$ we have
833: \begin{eqnarray*}
834:  \E^\sigma_{u+\varepsilon \zeta}[\xi_t]
835:  & = &
836:  \E^\sigma_{u+\varepsilon \zeta}[u_t
837:  +
838:  \varepsilon
839:  \zeta_t ]
840: \\
841:  & = &
842:  \E^\sigma_{u+\varepsilon \zeta}[u_t]
843:  +
844:  \varepsilon
845:  \E^\sigma_{u+\varepsilon \zeta}[\zeta_t ]
846: \\
847:  & = &
848:  \E^\sigma_{u+\varepsilon \zeta}[u_t] +\varepsilon \zeta_t,
849:  \quad t\in [0,T], \quad \varepsilon \in \real,
850: \end{eqnarray*}
851:  hence
852: \begin{eqnarray*}
853:  \zeta_t
854:  &=& \frac{d}{d\varepsilon} \E^\sigma_{u+\varepsilon \zeta}[
855:  \xi_t - u_t
856:  ]_{|\varepsilon=0}
857: \\
858: &=& \frac{d}{d\varepsilon} \E^\sigma
859:  [
860:  ( \xi_t - u_t )
861:  \Lambda ( u+\varepsilon \zeta ) ]_{| \varepsilon=0}
862: \\
863:  &=&
864:  \E^\sigma
865:  \left[
866:  (
867:  \xi_t - u_t
868:  )
869:  \frac{d}{d\varepsilon}
870:  \Lambda ( u+\varepsilon \zeta )_{| \varepsilon=0}
871:  \right]
872: \\
873: &=&
874:  \E^\sigma_u \left[
875:  ( \xi_t - u_t )
876:  \frac{d}{d\varepsilon}
877:  \log \Lambda ( u+\varepsilon \zeta )_{| \varepsilon=0}
878: \right]
879: \\
880:  &=&
881:  \E^\sigma_u \left[
882:  (
883:  \xi_t - u_t
884:  )
885:  \left(
886:  \int_0^T \frac{\dot{\zeta}_s}{\sigma^2_s }
887:  dX_s
888:  -
889:  \int_0^T \frac{\dot{\zeta}_s\dot{u}_s}{\sigma^2_s}
890:  ds
891:  \right)
892:  \right]
893: \\
894:  &=&
895:  \E^\sigma_u \left[
896:  (
897:  \xi_t - u_t
898:  )
899:  \int_0^T
900:  \frac{\dot{\zeta}_s}{\sigma^2_s}
901:  dX^u_s
902:  \right]
903: \\
904:  &=&
905:  \E^\sigma_u \left[
906:  (
907:  \xi_t - u_t
908:  )
909:  \int_0^t
910:  \frac{\dot{\zeta}_s}{\sigma^2_s}
911:  dX^u_s
912:  \right]
913: ,
914: \end{eqnarray*}
915:  where the exchange between expectation and derivative is justified by
916:  classical uniform integrability arguments.
917:  Thus, by the Cauchy-Schwarz inequality and the It\^o isometry we have
918: $$
919:  \zeta_t^2
920:  \leq
921:  \E^\sigma_u \left[ \left(
922:  \int_0^t
923:  \frac{\dot{\zeta}_s}{\sigma^2_s}
924:  dX^u_s \right)^2 \right] \E^\sigma_u [ | \xi_t - u_t|^2 ]
925:  =
926:  \int_0^t
927:  \frac{\dot{\zeta}_s^2}{\sigma^2_s}
928:   ds \E^\sigma_u [ | \xi_t - u_t|^2 ],
929:  \qquad
930:  t\in [0,T].
931: $$
932:  It then suffices to take
933: $$\zeta_t = \int_0^t
934:  \sigma^2_s ds,
935:  \qquad
936:  t\in [0,T],
937: $$
938:  to get
939: \begin{equation}
940: \label{eqv}
941:  \var^\sigma_u [\xi_t]
942:  =
943:  \E^\sigma_u [ |\xi_t - u_t|^2 ] \geq
944:  \int_0^t \sigma^2_s ds, \qquad t\in [0,T]
945: ,
946: \end{equation}
947:  which leads to \eqref{11} after integration with respect to
948:  $\mu (dt)$.
949:  As noted above, $\hat{u} = (X_t)_{t\in [0,T]}$
950:  is clearly unbiased under $\P^\sigma_u$ and it attains
951:  the lower bound $\crb (\sigma,\mu, \hat{u} ) $.
952: \end{Proof}
953: \noindent
954:  Recall that the classical linear
955:  parametric estimation problem for the drift of a diffusion
956:  consists in estimating the coefficient $\theta$ appearing
957:  in
958: $$d\xi_t = \theta a_t(\xi_t) dt + dY_t, \qquad
959:  \xi_0 = 0
960: ,
961: $$
962:  with a maximum likelihood estimator $\hat{\theta}_T$ given by
963: \begin{equation}
964: \label{mlet}
965: \hat{\theta}_T = \frac{\int_0^T a_t (\xi_t)d\xi_t}{\int_0^T a_t^2(\xi_t)dt}
966: ,
967: \end{equation}
968:  cf. \cite{liptser}, \cite{prakasarao} for Brownian motion and
969:  \cite{tudorviens} for an extension to fractional Brownian motions.
970: \\
971: 
972: \noindent
973:  Here we consider the nonparametric functional estimation of the drift of
974:  a one-dimensional drifted Brownian motion
975:  $(X_t)_{t\in \real_+}$ with decomposition
976: \begin{equation}
977: \label{jk}
978:  dX_t = \dot{u}_t dt + dX^u_t,
979: \end{equation}
980:  where $(\dot{u}_t)_{t\in [0,T]} \in L^2 (\Omega \times [0,T])$ is
981:  an adapted process and
982:  $(X^u_t)_{t\in \real_+}$ is a standard Brownian motion with
983:  quadratic variation $\sigma^2_t$
984:  under a probability $\P^\sigma_u$.
985: \\
986: 
987: \noindent
988:  In case $u$ is constrained to have the form $u_t = \theta t$, $t\in [0,T]$,
989:  $\theta \in \real$, our efficient estimator
990:  $\hat{u}$ satisfies $\hat{u}_T = \hat{\theta}_T T$,
991:  where $\hat{\theta}_T$ is given by \eqref{mlet}, $T>0$, with
992:  the asymptotics $\hat{\theta}_T\to \theta$ in probability
993:  as $T$ tends to infinity.
994:  The asymptotics is not in large time since $T$ can be a fixed parameter,
995:  but the efficient estimator
996:  $\hat{u} = X$ converges to $u$ as $\sigma$ tends to $0$,
997:  or equivalently as $T$ tends to $0$ by rescaling.
998: \\
999: 
1000: \noindent
1001:  To close this section we note that, at least informally,
1002:  $\hat{u} = (X_t)_{t\in [0,T]}$ can be viewed as a
1003:  maximum likelihood estimator of its own adapted drift $(u_t)_{t\in [0,T]}$
1004:  under $\P_u^\sigma$.
1005:  Indeed the functional differentiation of the Cameron-Martin
1006:  density
1007: $$\frac{d}{d\varepsilon} \Lambda ( \hat{u}+\varepsilon \zeta )_{\mid \varepsilon = 0} = 0
1008: , \qquad \zeta \in H ,
1009: $$
1010:  implies
1011: $$\int_0^T \frac{\dot{\zeta}_s}{\sigma^2_s } d X_s
1012:  -
1013:  \int_0^T \frac{\dot{\zeta}_s}{\sigma^2_s } d\hat{u}_s
1014:  = 0
1015: , \qquad
1016:  \zeta \in H ,
1017: $$
1018:  which leads to $X = \hat{u}$.
1019: \section{Bayes estimators}
1020: \label{2.0}
1021: \noindent
1022:  In this section we
1023:  consider Bayes estimators which will be useful in proving
1024:  the minimaxity of the estimator $\hat{u} = X$ in the framework of $(A)$ for Gaussian
1025:  processes with non-necessarily independent increments.
1026:  We will make use of the next lemma which is classical in the framework of
1027:  Gaussian filtering and is proved in the Appendix.
1028: \begin{lemma}
1029: \label{ljk}
1030:  Let $Z$ be a Gaussian process with covariance operator $\Gamma_\tau$
1031:  and drift $v\in H$,
1032:  and assume that $X$ is a Gaussian process with drift $Z$
1033:  and covariance operator $\Gamma$ given $Z$.
1034:  Then, conditionally to $X$, $Z$ has drift
1035: $$
1036:  f
1037:  \mapsto
1038:  \langle
1039:  f
1040:  ,
1041:  ( \Gamma+\Gamma_\tau)^{-1}\Gamma
1042:  v
1043:  \rangle
1044:  +
1045:  X (
1046:  ( \Gamma+\Gamma_\tau)^{-1}\Gamma_\tau
1047:  f
1048:  )
1049:  \quad
1050:  \mbox{and covariance}
1051:  \quad
1052:  \Gamma_\tau ( \Gamma + \Gamma_\tau )^{-1} \Gamma
1053: .
1054: $$
1055: \end{lemma}
1056: 
1057: \noindent
1058:  Note that unlike in Proposition~\ref{prop:CramerRaoBound},
1059:  no adaptedness or unbiasedness restriction is made on
1060:  $\xi$ in the infimum taken in \eqref{infr} below.
1061: \begin{prop}
1062: \label{best}
1063:  {Bayes estimator.}
1064:  Let $P^\tau_v$ denote the Gaussian distribution on $\Omega$
1065:  with covariance operator $\Gamma_\tau$ and drift $v\in H$.
1066:  The Bayes risk
1067: \begin{equation}
1068: \label{br}
1069:  \int_\Omega
1070:  \E_z \left[ \int_0^T | \xi_t - z_t|^2  \mu ( dt )
1071:  \right]
1072:  d\P^\tau_v (z)
1073: \end{equation}
1074:  of any estimator $(\xi_t)_{t\in [0,T]}$
1075:  on $\Omega$ under the prior distribution $\P^{\tau}_v$
1076:  is uniquely minimized by
1077: $$
1078:  \xi^{\tau,v}_t
1079:  : =
1080:  \langle
1081:  \stackrel{}{\chi}_t ,
1082:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma
1083:  v \rangle
1084:  +
1085:  X (
1086:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1087:  \stackrel{}{\chi}_t
1088:  )
1089: ,
1090:  \qquad
1091:  t\in [0,T]
1092: ,
1093: $$
1094:  which has risk
1095: \begin{equation}
1096: \label{infr}
1097:  \int_0^T
1098:  \langle
1099:  \stackrel{}{\chi}_t
1100:  ,
1101:  \Gamma
1102:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1103:  \stackrel{}{\chi}_t
1104:  \rangle
1105:  \mu ( dt )
1106:  =
1107:  \inf_\xi
1108:  \int_\Omega
1109:  \E_z \left[ \int_0^T | \xi_t - z_t|^2
1110:  \mu ( dt )
1111:  \right]
1112:  d\P^\tau_v (z)
1113: .
1114: \end{equation}
1115: \end{prop}
1116: \begin{Proof}
1117:  Let $Z$ denote a Gaussian process
1118:  with drift $v\in H$ and covariance $\Gamma_\tau$.
1119:  Recall (cf. Lemma~\ref{ljk}) that
1120:  if $X$ has drift $Z$ and covariance $\Gamma$
1121:  then, conditionally to $X$, $(Z_t)_{t\in [0,T]}$ has drift
1122: $$
1123:  t \mapsto
1124:  \langle
1125:  \stackrel{}{\chi}_t
1126:  ,
1127:  (\Gamma_\tau + \Gamma )^{-1}
1128:  \Gamma
1129:  v
1130:  \rangle
1131:  +
1132:  X (
1133:  (\Gamma_\tau + \Gamma )^{-1}
1134:  \Gamma_\tau
1135:  \stackrel{}{\chi}_t
1136:  )
1137: $$
1138:  and covariance
1139:  $
1140:  \Gamma_\tau
1141:  (\Gamma_\tau + \Gamma )^{-1}
1142:  \Gamma$.
1143:  Hence the Bayes risk of an estimator $\xi$
1144:  under the prior distribution $\P^\tau_v$ is given by
1145: \begin{eqnarray*}
1146: \lefteqn{
1147:  \int_\Omega
1148:  \E_z \left[ \int_0^T | \xi_t - z_t|^2 \mu ( dt )
1149:  \right]
1150:  d\P^\tau_v (z)
1151:  =
1152:  \E \left[
1153:  \E \left[ \int_0^T | \xi_t - Z_t|^2 \mu ( dt ) \Big| X
1154:  \right]
1155:  \right]
1156: }
1157: \\
1158:  & = &
1159:  \E \left[
1160:  \int_0^T | \xi_t - \E[Z_t \mid X ] |^2 \mu ( dt )
1161:  \right]
1162:  +
1163:  \E \left[
1164:  \int_0^T
1165:  \var ( Z_t | X )
1166:  \mu ( dt )
1167:  \right]
1168: \\
1169:  & = &
1170:  \E \left[
1171:  \int_0^T
1172:  \left| \xi_t -
1173:  \langle
1174:  \stackrel{}{\chi}_t ,
1175:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma
1176:  v \rangle
1177:  -
1178:  X (
1179:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1180:  \stackrel{}{\chi}_t
1181:  )
1182:  \right|^2
1183:  \mu ( dt )
1184:  \right]
1185: \\
1186:  & &
1187:  +
1188:  \int_0^T
1189:  \langle
1190:  \stackrel{}{\chi}_t
1191:  ,
1192:  \Gamma
1193:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1194:  \stackrel{}{\chi}_t
1195:  \rangle
1196:  \mu ( dt )
1197: ,
1198: \end{eqnarray*}
1199:  which is minimized by
1200: $$
1201:  \xi^{\tau,v}_t
1202:  : =
1203:  \E[Z_t \mid X ]
1204:  =
1205:  \langle
1206:  \stackrel{}{\chi}_t ,
1207:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma
1208:  v \rangle
1209:  -
1210:  X (
1211:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1212:  \stackrel{}{\chi}_t
1213:  )
1214: ,
1215:  \qquad t\in [0,T]
1216: .
1217: $$
1218: \end{Proof}
1219: \noindent
1220:  Clearly $\xi^{\tau,v}$ is unique in the sense that it is the only
1221:  estimator to minimize the Bayes risk \eqref{br}.
1222:  This shows in particular that every $\xi^{\tau,v}$ is admissible
1223:  in the sense that if an estimator $\xi$ satisfies
1224: $$
1225:  \E_z \left[
1226:  \Vert \xi - z \Vert_{L^2 ([0,T] , d\mu )}^2
1227:  \right]
1228:  \leq
1229:  \E_z \left[
1230:  \Vert \xi^{\tau,v} - z \Vert_{L^2 ([0,T] , d\mu )}^2
1231:  \right], \qquad
1232:  z\in \Omega
1233: ,
1234: $$
1235:  then
1236: \begin{eqnarray*}
1237:  \int_\Omega
1238:  \E_z \left[
1239:  \Vert \xi - z \Vert_{L^2 ([0,T] , d\mu )}^2
1240:  \right]
1241:  d\P^\tau_v
1242:  & \leq &
1243:  \int_\Omega
1244:  \E_z \left[
1245:  \Vert \xi^{\tau,v} - z \Vert_{L^2 ([0,T] , d\mu )}^2
1246:  \right]
1247:  d\P^\tau_v
1248: \\
1249:  & = &
1250:  \int_0^T
1251:  \langle
1252:  \stackrel{}{\chi}_t
1253:  ,
1254:  \Gamma
1255:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1256:  \stackrel{}{\chi}_t
1257:  \rangle
1258:  \mu ( dt )
1259: ,
1260: \end{eqnarray*}
1261:  hence
1262: \begin{equation}
1263: \label{mrsk}
1264:  \int_\Omega
1265:  \E_z \left[
1266:  \Vert \xi - z \Vert_{L^2 ([0,T] , d\mu )}^2
1267:  \right]
1268:  d\P^\tau_v
1269:  =
1270:  \int_0^T
1271:  \langle
1272:  \stackrel{}{\chi}_t
1273:  ,
1274:  \Gamma
1275:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1276:  \stackrel{}{\chi}_t
1277:  \rangle
1278:  \mu ( dt )
1279: ,
1280: \end{equation}
1281:  and $\xi = \xi^{\tau,v}$ by Proposition~\ref{best}.
1282: \\
1283: 
1284: \noindent
1285:  The Bayes estimator $\xi^{\tau , v}$ is biased in general,
1286:  and for deterministic $u\in H$ its mean square error under
1287:  $\P_u$ is equal to
1288: \begin{eqnarray}
1289: \label{abid}
1290: \lefteqn{
1291: % \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \!
1292:  \E_u \left[ \int_0^T | \xi^{\tau , v}_t - u_t|^2  \mu ( dt )
1293:  \right]
1294: }
1295: \\
1296: \nonumber
1297:  & = &
1298:  \E_u \left[ \int_0^T \Big|
1299:  \xi^{\tau , v}_t - \E_u [ \xi^{\tau , v}_t ]
1300:  \Big|^2  \mu ( dt )
1301:  \right]
1302:  +
1303:  \E_u \left[
1304:  \int_0^T |
1305:  \E_u [ \xi^{\tau , v}_t ]
1306:  - u_t
1307:  |^2
1308:  \mu ( dt )
1309:  \right]
1310: \\
1311: \nonumber
1312:  & = &
1313:  \E_u \left[ \int_0^T \Big|
1314:  X^u (
1315:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma_\tau
1316:  \stackrel{}{\chi}_t
1317:  )
1318:  \Big|^2  \mu ( dt )
1319:  \right]
1320:  +
1321:  \int_0^T \Big|
1322:  \langle
1323:  \stackrel{}{\chi}_t ,
1324:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma
1325:  ( v - u )
1326:  \rangle
1327:  \Big|^2
1328:  \mu ( dt )
1329: \\
1330: \nonumber
1331:  & = &
1332:  \int_0^T
1333:  \langle
1334:  ( \Gamma_\tau+\Gamma )^{-1}
1335:  \Gamma_\tau
1336:  \stackrel{}{\chi}_t
1337:  ,
1338:  \Gamma ( \Gamma_\tau+\Gamma )^{-1}
1339:  \Gamma_\tau
1340:  \stackrel{}{\chi}_t
1341:  \rangle
1342:  \mu ( dt )
1343: \\
1344: \nonumber
1345: & &
1346:  +
1347:  \int_0^T \Big|
1348:  \langle
1349:  \stackrel{}{\chi}_t ,
1350:  ( \Gamma_\tau+\Gamma )^{-1} \Gamma
1351:  ( v - u )
1352:  \rangle
1353:  \Big|^2
1354:  \mu ( dt )
1355: ,
1356: \end{eqnarray}
1357:  which shows that
1358: $$
1359:  \sup_{u\in H}
1360:  \E_u \left[ \int_0^T | \xi^{\tau , v}_t - u_t|^2 \mu ( dt )
1361:  \right]
1362:  =
1363:  + \infty
1364: ,
1365: $$
1366:  hence $\xi^{\tau , v}$ is not minimax.
1367: \\
1368: 
1369: \noindent In the independent increment case of Section~\ref{2} we have,
1370:  if
1371:  $\Gamma_\tau f(s) = \tau_s^2 f(s)$, $s\in [0,T]$:
1372: $$
1373:  \xi^{\tau,v}_t
1374:  : =
1375:  \int_0^t \frac{\sigma^2_s\dot{v}_s}{\tau^2_s+\sigma^2_s} ds
1376:  +
1377:  \int_0^t \frac{\tau^2_s}{\tau^2_s+\sigma^2_s} dX_s
1378: ,
1379:  \qquad
1380:  t\in [0,T]
1381: ,
1382: $$
1383:  with risk
1384: \begin{equation}
1385: \label{infr2}
1386:  \int_0^T
1387:  \int_0^t \frac{\tau^2_s\sigma^2_s}{\tau^2_s+\sigma^2_s} ds
1388:  \mu ( dt )
1389:  =
1390:  \inf_\xi
1391:  \int_\Omega
1392:  \E_z \left[ \int_0^T | \xi_t - z_t|^2
1393:  \mu ( dt )
1394:  \right]
1395:  d\P^\tau_v (z)
1396: .
1397: \end{equation}
1398: \noindent
1399:  Assuming now that $\Gamma_\tau f (t) = \tau^2 f(t)$, $t\in [0,T]$,
1400:  the Bayes risk
1401: $$
1402:  \int_0^T
1403:  \langle
1404:  \stackrel{}{\chi}_t ,
1405:  \Gamma_\tau
1406:  ( \Gamma_\tau+\Gamma )^{-1}
1407:  \Gamma
1408:  \stackrel{}{\chi}_t
1409:  \rangle
1410:  \mu ( dt )
1411:  =
1412:  \int_0^T
1413:  \langle
1414:  \stackrel{}{\chi}_t ,
1415:  ( I + \Gamma /\tau^2 )^{-1}
1416:  \Gamma \stackrel{}{\chi}_t
1417:  \rangle
1418:  \mu ( dt )
1419: ,
1420: $$
1421:  of $\xi^{\tau , v}$, $\tau \in \real$,
1422:  converges as $\tau \to \infty$ to the bound
1423: $$
1424:  \crb (\sigma,\mu, \hat{u} )
1425:  =
1426:  \int_0^T
1427:  \langle
1428:  \stackrel{}{\chi}_t ,
1429:  \Gamma
1430:  \stackrel{}{\chi}_t
1431:  \rangle
1432:  \mu ( dt )
1433: ,
1434: $$
1435:  hence it follows in the next proposition that, as in
1436:  the finite dimensional Gaussian case, the
1437:  estimator $\hat{u} = (X_t)_{t\in [0,T]}$ is minimax.
1438:  Note again that unlike in Proposition~\ref{prop:CramerRaoBound},
1439:  no adaptedness condition is imposed on $\xi$ in the infima
1440:  \eqref{infr} and \eqref{nmx}.
1441: \begin{prop}
1442: \label{minimax}
1443:  The estimator $\hat{u}=X$ is minimax.
1444:  For all $u\in \Omega$ we have
1445: \begin{equation}
1446: \label{nmx}
1447:  \crb ( \gamma ,\mu, \hat{u} )
1448:  =
1449:  \E_u \left[ \int_0^T | X_t - u_t|^2 \mu ( dt ) \right]
1450:  = \inf_\xi
1451:  \sup_{v \in \Omega }
1452:  \E_v \left[ \int_0^T | \xi_t - v_t|^2 \mu ( dt ) \right]
1453: .
1454: \end{equation}
1455: \end{prop}
1456: \begin{Proof}
1457:  Clearly, taking $\xi=0$ yields
1458: $$
1459:  \crb (\gamma ,\mu, \hat{u} )
1460:  =
1461:  \sup_{u\in \Omega}
1462:  \E_u \left[ \int_0^T | X_t - u_t|^2 \mu ( dt ) \right]
1463:  \geq \inf_\xi
1464:  \sup_{u\in \Omega}
1465:  \E_u \left[ \int_0^T | \xi_t - u_t|^2 \mu ( dt ) \right]
1466: .
1467: $$
1468:  On the other hand, from Proposition~\ref{best}, for all processes
1469:  $\xi$ we have
1470: \begin{eqnarray*}
1471:  \sup_{u \in \Omega}
1472:  \E_u \left[ \int_0^T | \xi_t - u_t|^2 \mu ( dt ) \right]
1473:  & \geq &
1474:  \int_\Omega
1475:  \E_z \left[ \int_0^T | \xi_t - z_t|^2 \mu ( dt )
1476:  \right]
1477:  d\P^\tau_0 (z)
1478: \\
1479:  & \geq &
1480:  \int_0^T
1481:  \langle
1482:  \stackrel{}{\chi}_t ,
1483:  ( I + \Gamma /\tau^2 )^{-1}
1484:  \Gamma
1485:  \stackrel{}{\chi}_t
1486:  \rangle
1487:  \mu ( dt )
1488: ,
1489: \end{eqnarray*}
1490:  for all $\tau > 0$, hence
1491: $$
1492:  \inf_\xi
1493:  \sup_{u\in H}
1494:  \E_u \left[ \int_0^T | \xi_t - u_t|^2 \mu ( dt ) \right]
1495:  \geq
1496:  \int_0^T
1497:  \langle
1498:  \stackrel{}{\chi}_t ,
1499:  \Gamma
1500:  \stackrel{}{\chi}_t
1501:  \rangle
1502:  \mu ( dt )
1503:  =
1504:  \crb ( \gamma ,\mu, \hat{u} )
1505: .
1506: $$
1507: \end{Proof}
1508: \section{Malliavin calculus on Gaussian space}
1509: \label{2.1}
1510:  Before proceeding to the construction
1511:  of Stein type estimators, we need to introduce
1512:  some elements of analysis on Gaussian space, see e.g. \cite{nualartm2}.
1513:  This construction is valid in both frameworks $(A)$ and $(B)$.
1514:  Given $u \in H$, let
1515: $$
1516:  X^u = X - u.
1517: $$
1518: % be an ${\cal F}_t$-adapted process and let $X^u = X - u$.
1519: 
1520: \noindent
1521:  We fix $(h_n)_{n \geq 1}$ a total subset of $H$
1522:  and let ${\cal S}$ denote the space of cylindrical functionals of the
1523:  form
1524: \begin{equation}
1525: \label{e1}
1526:  F= f_n
1527:  \left(
1528:  X^u ( h_1 )
1529:  ,
1530:  \ldots
1531:  ,
1532:  X^u ( h_n)
1533:  \right)
1534: ,
1535: \end{equation}
1536:  where $f_n$ is in the space of infinitely differentiable rapidly decreasing
1537:  functions on $\real^n$, $n\geq 1$.
1538: \noindent
1539: \begin{definition}
1540:  The $H $-valued Malliavin derivative is defined as
1541: $$
1542:  \nabla_t F =
1543:  \sum_{i=1}^{n}
1544:  h_i ( t )
1545:  \partial_i f_n
1546:  \left(
1547:  X^u (h_1 )
1548:  ,
1549:  \ldots
1550:  ,
1551:  X^u (h_n)
1552:  \right)
1553: ,
1554: $$
1555:  for $F \in {\cal S}$ of the form \eqref{e1}.
1556: \end{definition}
1557: \noindent
1558:  It is known that $\nabla$ is closable, cf. Proposition~1.2.1 of
1559:  \cite{nualartm2}, and its closed domain will be denoted by $\Dom (\nabla )$.
1560: \begin{definition}
1561:  Let $D$ be defined on $\Dom ( \nabla )$ as
1562: $$D_t F : = ( \Gamma \nabla F ) (t)
1563: , \quad t\in [0,T], \quad F\in \Dom (\nabla )
1564: .
1565: $$
1566: \end{definition}
1567: \noindent
1568:  Let $\delta : L^2_u (\Omega ; H ) \to L^2 (\Omega , \P_u )$
1569:  denote the closable adjoint of $\nabla$, i.e. the
1570:  divergence operator under $\P_u$,
1571:  which satisfies the integration by parts formula
1572: \begin{equation}
1573: \label{intparpartiesvar1/N}
1574:  \E_u [ F \delta ( v ) ]
1575:  = \E_u [ \langle v , \nabla F \rangle_{H} ]
1576: ,
1577:  \qquad
1578:  F \in \Dom ( \nabla )
1579:  ,
1580:  \quad
1581:  v \in \Dom (\delta )
1582: ,
1583: \end{equation}
1584:  with the relation
1585: $$\delta (h F ) = F X (h) - \langle h , \nabla F\rangle_H,
1586: $$
1587:  cf. \cite{nualartm2}, for $F\in \Dom (\nabla )$ and $h\in H$ such
1588:  that $hF\in \Dom (\delta )$.
1589:  Note that \eqref{intparpartiesvar1/N} is an infinite-dimensional version
1590:  of the integration by parts \eqref{ibp2}, which can be proved e.g. using
1591:  the countable Gaussian random variables constructed from $X$.
1592: \begin{lemma}
1593: \label{prt}
1594:  We have
1595: $$
1596:  \E_u [ F X^u_t ]
1597:  = \E_u [
1598:  D_t F
1599:  ]
1600: ,
1601:  \qquad
1602:  t\in [0,T]
1603:  ,
1604:  \quad
1605:  F \in \Dom ( \nabla )
1606: .
1607: $$
1608: \end{lemma}
1609: \begin{Proof}
1610: \begin{description}
1611: \item{(A)}
1612:  In the case of Paley-Wiener expansions we have
1613: \begin{eqnarray*}
1614:  \E_u [ F X^u_t ]
1615:  & = &
1616:  \E_u [ F X^u  ( \stackrel{}{\chi}_t ) ]
1617: \\
1618:  & = &
1619:  \E_u [ F \delta ( \stackrel{}{\chi}_t ) ]
1620: \\
1621:  & = &
1622:  \E_u [ \langle \stackrel{}{\chi}_t , \nabla F\rangle_H ]
1623: \\
1624:  & = &
1625:  \E_u [ \langle 1_{[0,t]} , \dot{\Gamma} \nabla F\rangle_{L^2([0,T] , dt )} ]
1626: \\
1627:  & = &
1628:  \E_u [ ( \Gamma \nabla F) ( t ) ]
1629: ,
1630:  \qquad
1631:  F \in \Dom ( \nabla )
1632:  ,
1633:  \quad
1634:  t\in [0,T]
1635: .
1636: \end{eqnarray*}
1637: \item{(B)}
1638:  In the case of Karhunen-Lo\`eve expansions we have
1639: \begin{eqnarray*}
1640:  \E_u [ F X^u_t ]
1641:  & = &
1642:  \sum_{k=0}^\infty
1643:  h_k(t)
1644:  \E_u [ F X^u (h_k) ]
1645: \\
1646:  & = &
1647:  \sum_{k=0}^\infty
1648:  h_k(t)
1649:  \E_u [ F \delta (h_k) ]
1650: \\
1651:  & = &
1652:  \sum_{k=0}^\infty
1653:  h_k(t)
1654:  \E_u [ \langle h_k , \nabla F\rangle_H ]
1655: \\
1656:  & = &
1657:  \sum_{k=0}^\infty
1658:  h_k(t)
1659:  \E_u [ \langle h_k , \Gamma \nabla F\rangle_{L^2([0,T],\mu )} ]
1660: \\
1661:  & = &
1662:  \E_u [
1663:  ( \Gamma \nabla F ) (t)
1664:  ]
1665: ,
1666:  \qquad
1667:  F \in \Dom ( \nabla )
1668:  ,
1669:  \quad
1670:  t\in [0,T]
1671: .
1672: \end{eqnarray*}
1673: \end{description}
1674: \end{Proof}
1675: \noindent
1676: %
1677: %\noindent
1678: \begin{definition}
1679:  We define the Laplacian $\Delta$ by
1680: $$
1681:  \Delta F = \trace_{L^2([0,T] , d\mu )^{\otimes 2}} D D F
1682:  = \int_0^T D_t D_t F \mu (dt)
1683: $$
1684:  on the space $\Dom (\Delta )$ made
1685:  of all $F\in \Dom (\nabla )$
1686:  such that $D_tF \in \Dom (\nabla)$, $t\in [0,T]$, and
1687:  $(D_tD_tF)_{t\in [0,T]}\in L^2([0,T],\mu)$, $\P$-a.s.
1688: \end{definition}
1689: \noindent
1690:  If $F\in {\cal S}$ has the form \eqref{e1} we have
1691: $$
1692:  \Delta F
1693:  =
1694:  \sum_{i,j=1}^n
1695:  \langle
1696:  \Gamma h_i
1697:  ,
1698:  \Gamma h_j
1699:  \rangle_{L^2([0,T],\mu )}
1700:  \partial_i \partial_j f_n
1701:  \left(
1702:  X^u ( h_1 )
1703:  ,
1704:  \ldots
1705:  ,
1706:  X^u ( h_n )
1707:  \right)
1708: .
1709: $$
1710:  Unlike the Gross Laplacian $\Delta_G$ defined by
1711: $$\Delta_G F = \trace_{H^{\otimes 2}} \nabla \nabla F
1712: ,
1713: $$
1714:  the operator $\Delta$ is closable, as shown in the following
1715:  proposition.
1716: \begin{prop}
1717: \label{prop:closability}
1718:  Closability of $\Delta$.
1719:  For any sequence $(F_n)_{n \in \inte}$ of random variables converging
1720:  to $0$ in $L^2 (\Omega , \P_u )$ and such that
1721:  $(\Delta F_n)_{n \in \inte}$ converges in $L^2 (\Omega , \P_u )$, we have
1722: $$\lim_{n\to \infty }
1723:  \Delta F_n = 0
1724: .
1725: $$
1726: \end{prop}
1727: \begin{proof}
1728:  Let $(G_n)_{n \in \inte}$ a sequence in $\mathcal{S}$ converging to $0$ in
1729:  $L^2(\Omega, \P_u )$, and such that $(\Delta G_n)_{n \in \inte}$ converges to $F$
1730:  in $L^2(\Omega, \P_u )$.
1731:  For all $G\in\mathcal{S}$ we have, in the notation of $(A)$:
1732: \begin{eqnarray*}
1733: \disp{\vert \langle \Delta G_n, G\rangle_{L^2 (\Omega , \P_u )} \vert} &=& \disp{ \left\vert
1734:  \E_u \left[ G \int_0^T D_tD_t G_n \mu(dt) \right] \right\vert } \\
1735: &=& \disp{ \left\vert \int_0^T \E_u [\langle\nabla D_t G_n ,
1736:  \stackrel{}{\chi}_t G \rangle_H] \, \mu(dt)
1737:  \right\vert }
1738: \\
1739: &=& \disp{ \left\vert \int_0^T \E_u [D_t G_n \, \delta(
1740:  \stackrel{}{\chi}_t
1741:  G) ] \, \mu(dt)  \right\vert } \\
1742: &=& \disp{ \left\vert \int_0^T \E_u [\langle \nabla G_n,
1743:  \stackrel{}{\chi}_t
1744:  \delta(
1745:  \stackrel{}{\chi}_t
1746:  G) \rangle_H ] \, \mu(dt)  \right\vert } \\
1747: &=& \disp{ \left\vert \int_0^T \E_u [G_n \delta(
1748:  \stackrel{}{\chi}_t
1749:  \delta(
1750:  \stackrel{}{\chi}_t
1751:  G)) ] \, \mu(dt)  \right\vert } \\
1752: &\leq& \disp{ \|G_n\|_{L^2 (\Omega, \P_u )} \int_0^T \|\delta(
1753:  \stackrel{}{\chi}_t
1754:  \delta(
1755:  \stackrel{}{\chi}_t
1756:  G))\|_{L^2(\Omega, \P_u )} \, \mu(dt), }
1757: \end{eqnarray*}
1758: hence $\langle F, G \rangle_{L^2(\Omega, \P_u )} =0, \; G \in \mathcal{S}$,
1759:  which implies $F=0$.
1760: \end{proof}
1761: \noindent
1762:  We will say that a random variable $F$ in $\Dom (\Delta )$
1763:  is $\Delta$-superharmonic on $\Omega$ if
1764: \begin{equation}
1765: \label{sph}
1766: \Delta F (\omega ) \leq 0, \qquad \P (d\omega)-a.s.
1767: \end{equation}
1768: \begin{remark}
1769:  In the independent increment case where $\gamma (s,t)$ is given by
1770: $$\gamma (s,t) = \int_0^{s\wedge t}
1771:  \sigma_u^2 dt
1772: ,
1773:  \qquad
1774:  s,t\in [0,T]
1775: ,
1776: $$
1777:  we have
1778: \begin{equation}
1779: \label{intparpartiesvar1/N.1}
1780: \delta ( v ) = \int_0^T \dot{v}_t dX^u_t,
1781: \end{equation}
1782:  for every ${\cal F}_t$-adapted process $v \in L^2 (\Omega ; H , \P_u )$.
1783: \end{remark}
1784: \section{Superefficient drift estimators}
1785: \label{3}
1786:  Our aim is to construct a superefficient estimator of $u$ of the form
1787:  $X+\xi$, whose mean square error is strictly smaller than the
1788:  minimax risk $\crb (\gamma , \mu , \hat{u})$ of Proposition~\ref{minimax}
1789:  when $\xi \in L^2 ([0,T]\times \Omega , \P_u \otimes \mu )$
1790:  is a suitably chosen stochastic process.
1791:  This estimator will be biased and anticipating with respect
1792:  to the Brownian filtration.
1793:  In the next lemma we follow Stein's argument which
1794:  uses integration by parts but we replace \eqref{ibp2} by
1795:  the duality relation \eqref{intparpartiesvar1/N}
1796:  between the gradient and divergence operators on Gaussian space.
1797:  The results of this section are valid in both frameworks $(A)$ and $(B)$.
1798: \begin{lemma}
1799: \label{lemma1}
1800:  {Unbiased risk estimate}.
1801:  For any $\xi \in L^2 (\Omega \times [0,T] , \P_u \otimes \mu )$ such that
1802:  $\xi_t \in \Dom (\nabla)$, $t\in [0,T]$, and
1803:  $(D_t\xi_t)_{t\in [0,T]}\in L^1 (\Omega \times [0,T] , \P_u \otimes \mu )$, we have
1804: \begin{equation}
1805: \label{eq:erreur}
1806:  \E_u
1807:  \left[
1808:  \Vert X + \xi - u \Vert_{L^2([0,T], \mu )}^2
1809:  \right]
1810:  =
1811:  \crb (\gamma ,\mu, \hat{u} )
1812:  +
1813:  \Vert \xi \Vert_{L^2 (\Omega \times [0,T] , \P_u \otimes \mu )}^2
1814:  + 2
1815:  \E_u \left[
1816:  \int_0^T
1817:  D_t \xi_t
1818:  \mu ( dt )
1819:  \right]
1820: .
1821: \end{equation}
1822: \end{lemma}
1823: \begin{Proof}
1824:  We have
1825: \begin{eqnarray*}
1826: \lefteqn{
1827:  \E_u \left[
1828:  \Vert X + \xi - u \Vert_{L^2([0,T] , d\mu )}^2 \right]
1829:  =
1830:  { \E_u \left[ \int_0^T \Big{|} X^u_t + \xi_t \Big{|}^2
1831:  \mu ( dt )
1832:  \right] }
1833: }
1834: \\
1835: &=& { \E_u \left[ \int_0^T | X^u_t |^2
1836:  \mu ( dt )
1837:  \right] +
1838:  \Vert \xi \Vert_{L^2 ( \Omega \times [0,T] , \P_u \otimes \mu )}^2  +
1839:  2 \E_u \left[ \int_0^T X^u_t \xi_t
1840:  \mu ( dt )
1841:  \right] } \\
1842:  &=&
1843:  \crb (\gamma ,\mu, \hat{u} )
1844:  +
1845:  \Vert \xi \Vert_{L^2 ( \Omega \times [0,T] , \P_u \otimes \mu )}^2
1846:  + 2 \E_u \left[ \int_0^T X^u_t \xi_t
1847:  \mu ( dt )
1848:  \right]
1849: ,
1850: \end{eqnarray*}
1851:  and apply Lemma~\ref{prt} to obtain \eqref{eq:erreur}.
1852: \end{Proof}
1853: \noindent
1854:  The next proposition specializes the above lemma to
1855:  processes $\xi$ of the form
1856: $$\xi_t = D_t \log F,
1857:  \qquad t\in [0,T],
1858: $$
1859:  where $F$ is an a.s. strictly positive and sufficiently
1860:  smooth random variable.
1861: \begin{prop}
1862: \label{prop:erreurcylindricalfunctions.1}
1863:  Logarithmic gradient.
1864:  Stein-type estimator.
1865:  For any $\P$-a.s. positive random variable
1866:  $F\in \Dom (\nabla)$ such that $D_t F\in \Dom (\nabla)$, $t\in [0,T]$,
1867:  and
1868:  $(D_tD_tF )_{t\in [0,T]}\in L^1 (\Omega \times [0,T] , \P_u \otimes \mu )$, we have
1869: $$
1870:  \E_u
1871:  \left[
1872:  \Vert X +
1873:  D \log F - u \Vert_{L^2 ([0,T] , d\mu )}^2
1874:  \right]
1875:  =
1876:  \crb (\gamma ,\mu, \hat{u} )
1877:  -
1878:  \E_u \left[
1879:  \Vert
1880:  D \log F \Vert_{L^2 ([0,T], \mu )}^2
1881:  \right]
1882:  +
1883:  2
1884:  \E_u \left[
1885:  \frac{ \Delta F }{F}
1886:  \right]
1887: .
1888: $$
1889: \end{prop}
1890: \begin{Proof}
1891:  From \eqref{eq:erreur} we have
1892: \begin{eqnarray*}
1893: \lefteqn{
1894:  \! \! \! \! \! \! \! \! \! \! \!
1895:  \E_u
1896:  \left[
1897:  \Vert X + D \log F - u \Vert_{L^2 ([0,T],\mu )}^2
1898:  \right]
1899: }
1900: \\
1901:  & = &
1902:  \crb (\gamma ,\mu, \hat{u} )
1903:  +
1904:  \Vert D \log F \Vert_{L^2 (\Omega \times [0,T] , \P_u \otimes \mu )}^2
1905:  +
1906:  2 \E_u \left[ \int_0^T D_t D_t \log F
1907:  \mu ( dt )
1908:  \right]
1909: \\
1910: & = &
1911:  \crb (\gamma ,\mu, \hat{u} )
1912:  +
1913:  \E_u \left[
1914:  \int_0^T \left(
1915:  \bigg| \frac{D_t F }{F} \bigg|^2
1916:  + 2 D_t D_t \log F
1917:  \right)
1918:  \mu ( dt )
1919:  \right]
1920: ,
1921: \end{eqnarray*}
1922:  and we use the relation
1923: $$
1924:  \bigg| \frac{D_t F }{F} \bigg|^2
1925:  + 2 D_t D_t \log F
1926:  =
1927:  2 \frac{D_tD_t F}{F}
1928:  - \bigg| \frac{D_t F }{F} \bigg|^2
1929:  ,
1930:  \qquad
1931:  t\in [0,T].
1932: $$
1933: \end{Proof}
1934: \noindent
1935:  From the above proposition it suffices that $F$ be $\Delta$-superharmonic for
1936:  $X + D \log F$ to be superefficient.
1937:  In this case we have
1938: \begin{equation}
1939: \label{sh}
1940:  \E_u
1941:  \left[
1942:  \Vert X +
1943:  D \log F - u \Vert_{L^2([0,T] , d\mu )}^2
1944:  \right]
1945:  \leq
1946:  \crb (\gamma ,\mu, \hat{u} )
1947:  -
1948:  \E_u \left[
1949:  \Vert
1950:  D \log F \Vert_{L^2([0,T] , d\mu )}^2
1951:  \right]
1952: ,
1953: \end{equation}
1954:  with equality in \eqref{sh} when $F$ is $\Delta$-harmonic.
1955: \\
1956: 
1957: \noindent
1958:  In the next proposition we show
1959:  that the $\Delta$-superharmonicity of $F$ is not necessary
1960:  for $X + D \log F$ to be superefficient, namely
1961:  the $\Delta$-superharmonicity
1962:  of $F$ can be replaced by the $\Delta$-superharmonicity
1963:  of $\sqrt{F}$, which is a weaker assumption, see \cite{fourdrinier}
1964:  in the finite dimensional case.
1965:  In particular, $X+D \log F$ is a superefficient estimator of $u$
1966:  if $\Delta \sqrt{F} < 0$ on a set of strictly positive $\P$-measure.
1967: \begin{prop}
1968: \label{prop:erreurcylindricalfunctions}
1969:  Stein-type estimator.
1970:  For any $\P$-a.s. positive random variable
1971:  $F\in \Dom (\nabla)$ such that $D_t F\in \Dom (\nabla)$, $t\in [0,T]$,
1972:  and
1973:  $(D_tD_tF )_{t\in [0,T]}\in L^1 (\Omega \times [0,T] , \P_u \otimes \mu )$, we have
1974: \begin{equation}
1975: \label{hjk}
1976:  \E_u
1977:  \left[
1978:  \Vert X +
1979:  D \log F - u \Vert_{L^2 ([0,T] , d\mu )}^2
1980:  \right]
1981:  =
1982:  \crb (\gamma ,\mu, \hat{u} )
1983:  +
1984:  4
1985:  \E_u
1986:  \left[ \frac{\Delta \sqrt{F}}{\sqrt{F}} \right]
1987: .
1988: \end{equation}
1989: \end{prop}
1990: \begin{Proof}
1991:  For any $F\in \Dom ( \nabla )$ such that $F>0$, $\P$-a.s.,
1992:  and $\sqrt{F} \in \Dom (\Delta^\vide)$, we have
1993: $$
1994:  2 \frac{D_tD_t F}{F} - \bigg| \frac{D_t F }{F} \bigg|^2
1995:  =
1996:  \frac{2}{\sqrt{F}}
1997:  D_t \left( \frac{D_t F}{\sqrt{F}} \right)
1998:  =
1999:  \frac{4}{\sqrt{F}}
2000:  D_t D_t \sqrt{F},
2001:  \qquad t\in [0,T]
2002: ,
2003: $$
2004:  which implies
2005: \begin{equation}
2006: \label{qsd}
2007:  4
2008:  \frac{\Delta \sqrt{F}}{\sqrt{F}}
2009:  =
2010:  2 \frac{ \Delta F }{F}
2011:  - \int_0^T
2012:  | D_t \log F |^2
2013:  \mu ( dt )
2014: ,
2015: \end{equation}
2016:  and allows us to conclude from Lemma~\ref{lemma1}.
2017: \end{Proof}
2018: \noindent
2019:  Relation \eqref{hjk} extends to any $F\in \Dom (\nabla )$ such that
2020:  $\sqrt{F} \in \Dom (\Delta )$, and $F \geq 0$, $\Delta \sqrt{F} \leq 0$, $\P$-a.s.
2021: \\
2022: 
2023: \noindent
2024:  In case $(X_t)_{t\in [0,T]}$ is a Brownian motion with constant
2025:  variance $\sigma_t = \sigma$, $t\in [0,T]$, we have
2026: \begin{equation}
2027: \label{sh2}
2028:  \E_u
2029:  \left[
2030:  \Vert X +
2031:  D \log F - u \Vert_{L^2([0,T])}^2
2032:  \right]
2033:  \leq
2034:  \frac{\sigma^2T^2}{2}
2035:  +
2036:  4
2037:  \E_u \left[
2038:  \frac{\Delta \sqrt{F}}{\sqrt{F}}
2039:  \right]
2040: .
2041: \end{equation}
2042:  Given $(X_t^{\sigma,1})_{t\in [0,T]}, \ldots , (X_t^{\sigma,N})_{t\in [0,T]}$ are
2043:  $N$ independent samples of $(X_t)_{t\in [0,T]}$, the process
2044:  $\bar{X}$ defined in \eqref{xbar} satisfies
2045: $$
2046:  \E^{\sigma/N}_u
2047:  \left[
2048:  \Vert \bar{X} +
2049:  D \log F - u \Vert_{L^2([0,T])}^2
2050:  \right]
2051:  =
2052:  \frac{1}{N}
2053:  \crb (\sigma,\mu, \hat{u} )
2054:  +
2055:  \frac{4}{N^2}
2056:  \E_u \left[
2057:  \frac{\Delta \sqrt{F}}{\sqrt{F}}
2058:  \right]
2059: .
2060: $$
2061: 
2062: \noindent
2063:  As in \cite{stein}, the superefficient estimators constructed in this way
2064:  are minimax in the sense that from Proposition~\ref{minimax}
2065:  and Proposition~\ref{prop:erreurcylindricalfunctions.1},
2066:  for all $u\in H$
2067:  we have
2068: $$\E_u
2069:  \left[
2070:  \Vert X + D \log F - u \Vert_{L^2([0,T], \mu )}^2
2071:  \right]
2072:  <
2073:  \crb (\gamma ,\mu, \hat{u} )
2074:   = \inf_\xi
2075:  \sup_{v \in \Omega }
2076:  \E_v \left[ \int_0^T | \xi_t - v_t|^2 dt \right]
2077: ,
2078: $$
2079:  provided $\Delta \sqrt{F} < 0$ on a set of strictly positive $\P$-measure,
2080:  thus showing that the minimax estimator
2081:  $\hat{u} = (X_t)_{t\in [0,T]}$ is inadmissible.
2082: \\
2083: 
2084: \noindent
2085: 
2086: \noindent
2087: % However the computation of $X_t +\E_u [D_t \log F \mid {\cal F}_t]$ requires an a priori knowledge of $u$.
2088: %\noindent
2089: % Note that these properties hold in $L^2([0,T], \mu )$-norm but not
2090: % pointwise, as e.g. for the estimation of $u_T$ from $X_T$ by
2091: % the standard Stein method.
2092:  Both estimators $X_t+D_t \log F$ and
2093:  $X_t +\E_u [D_t \log F \mid {\cal F}_t]$
2094:  have bias
2095: $$b_t = \E_u [X_t + D_t \log F -u_t]
2096:  = \E_u [ D_t \log F ]
2097: , \qquad
2098:  t\in [0,T],
2099: $$
2100:  which can be bounded as follows from  \eqref{qsd}:
2101: \begin{eqnarray*}
2102: \Vert b \Vert_{L^2([0,T], \mu )}^2
2103:  & = &
2104:  \int_0^T
2105:  |
2106:  \E_u [ D_t \log F ]
2107:  |^2 dt
2108: \\
2109:  & \leq &
2110:  \E_u \left[ \int_0^T
2111:  | D_t \log F |^2 dt
2112:  \right]
2113: \\
2114:  & =&
2115:  2
2116:  \E_u \left[
2117:  \frac{ \Delta F }{F}
2118:  \right]
2119:  -
2120:  4
2121:  \E_u \left[
2122:  \frac{\Delta \sqrt{F}}{\sqrt{F}}
2123:  \right]
2124: .
2125: \end{eqnarray*}
2126: \begin{remark}
2127:  In the independent increment case of Section~\ref{2},
2128:  the formulas obtained in this section also hold for $u$ an adapted
2129:  process in $L^2 (\Omega \times [0,T] )$.
2130:  However, in this case the computation of the gradient $D \log F$
2131:  requires in principle the knowledge of $X^u$, except when $u$ is
2132:  deterministic, in which case the knowledge of $X$ is sufficient.
2133:  Thus, assuming $u$ to be deterministic will be necessary
2134:  for the applications of Section~\ref{4}.
2135: \end{remark}
2136: \section{Superharmonic functionals}
2137: \label{app}
2138:  In this section we give examples of nonnegative superharmonic
2139:  functionals with respect to the Laplacian $\Delta$.
2140:  We start by reviewing the construction of such functionals
2141:  using potential theory on the Gaussian space $(\Omega , H  , \P )$,
2142:  and next we turn to cylindrical functionals which will be used
2143:  in the numerical applications of Section~\ref{4}.
2144:  We assume that $( \Gamma h_k )_{k \geq 1}$ is orthogonal in $L^2 ([0,T ] , d\mu )$,
2145:  and we let
2146: $$
2147:  \lambda_k = \Vert \Gamma h_k \Vert_{L^2([0,T],\mu)},
2148: \qquad
2149:  k \geq 1.
2150: $$
2151:  The sequence $(h_k)_{k\geq 1}$ can be realized as the solution of the eigenvalue problem
2152: \begin{equation}
2153: \label{*1}
2154:  \Gamma h_k = - \lambda^2_k \ddot{h}_k,
2155:  \qquad
2156:  \dot{h}_k (T) = 0, \qquad
2157:  k\geq 1,
2158: \end{equation}
2159:  in case $(A)$, provided $\mu (dt)=dt$, and
2160: $$\Gamma h_k = \lambda^2_k h_k, \qquad
2161:  k\geq 1,
2162: $$
2163:  in case $(B)$ for general $\mu$.
2164: %\end{assumption}
2165: 
2166: \subsubsection*{Potentials}
2167:  We refer to \cite{grossp} and \cite{Goodman} for
2168:  the notion of harmonicity on the Wiener space with respect to
2169:  the Gross Laplacian.
2170: \noindent From our orthonormality assumption on $(h_k)_{k\geq 1}$,
2171:  the Laplacian $\Delta$ is written as
2172: $$
2173:  \Delta^\vide F
2174:  =
2175:  {\sum_{i=1}^n
2176:  \partial_i^2 f_n
2177:  \left(
2178:   \int_0^T \dot{h}_1 ( s ) dX^u_s
2179:  ,
2180:  \ldots
2181:  ,
2182:  \int_0^T \dot{h}_n( s ) dX^u_s
2183:  \right)}
2184: $$
2185:  on cylindrical functionals.
2186:  Let $(W^\Omega_t)_{t \geq 0}$ denote the standard $\Omega$-valued Wiener
2187:  process with generator $\frac{1}{2} \Delta_G$
2188:  on $(\tilde{\Omega} , \tilde{\cal F} , \tilde{\P})$, represented as
2189: \begin{equation}
2190: \label{cv}
2191:  W^\Omega_t = \sum_{n=1}^\infty
2192:  \int_0^\cdot \dot{h}_n(s) \sigma^2_s ds
2193:  \frac{\beta_n ( t ) }{\Vert h_n \Vert_{H^\vide}}
2194: ,
2195:  \qquad t\in \real_+,
2196: \end{equation}
2197:  where $(\beta_n(t))_{t\in \real_+}$, $n\geq 1$,
2198:  are independent standard Brownian motions
2199:  on $(\tilde{\Omega} , \tilde{\cal F} , \tilde{\P})$,
2200:  given as
2201: $$\beta_n ( t ) = \int_0^T
2202:  \frac{\dot{h}_n (r)}{\Vert h_n \Vert_{H^\vide}}
2203: % \sigma^2_s
2204:  dW^\Omega_t( r )
2205: ,
2206:  \qquad t \in \real_+,
2207:  \quad
2208:  n\geq 1
2209: .
2210: $$
2211:  We have the covariance relation
2212: $$\tilde{\E}
2213:  \left[
2214:  \int_0^T
2215:  \dot{v}_1 (r) dW^\Omega_s (r)
2216:  \int_0^T
2217:  \dot{v}_2 (r) dW^\Omega_t (r)
2218:  \right] =
2219:  ( s \wedge t )
2220:  \langle
2221:  v_1 ,
2222:  v_2
2223:  \rangle_{H^\vide}
2224: ,
2225:  \qquad
2226:  s,t \in \real_+
2227: ,
2228:  \quad
2229:  v_1, v_2 \in H
2230: .
2231: $$
2232:  In other terms we have
2233: \begin{eqnarray*}
2234: \lefteqn{
2235:  \tilde{\E}
2236:  [ W^\Omega_t ( a ) W^\Omega_s ( b ) ]
2237:  =
2238:  ( s\wedge t )
2239:  \sum_{n=1}^\infty
2240:  \frac{
2241:  \int_0^a \dot{h}_n(s) \sigma^2_s ds
2242:  \int_0^b \dot{h}_n(s) \sigma^2_s ds
2243:  }{\Vert h_n \Vert_{H^\vide}^2}
2244: }
2245: \\
2246:  & = &
2247:  ( s\wedge t )
2248: \left<
2249:  \sum_{n=1}^\infty
2250:  \frac{\dot{h}_n}{\Vert h_n \Vert_{H^\vide}^2}
2251:  \int_0^a \dot{h}_n(s) \sigma^2_s ds
2252:  ,
2253:  \sum_{n=1}^\infty
2254:  \frac{\dot{h}_n}{\Vert h_n \Vert_{H^\vide}^2}
2255:  \int_0^b \dot{h}_n(s) \sigma^2_s ds
2256:  \right>_{L^2([0,T],\sigma^2_t dt)}
2257: \\
2258:  & = &
2259:  ( s\wedge t )
2260:  \langle {\bf 1}_{[0,a]}
2261:  ,
2262:  {\bf 1}_{[0,b]}
2263:  \rangle_{L^2([0,T],\sigma^2_t dt)}
2264: \\
2265:  & = &
2266:  ( s \wedge t )
2267:  \int_0^{a \wedge b}
2268:  \sigma^2_r dr
2269: , \qquad 0 \leq a, b \leq T,
2270:  \quad s,t \in \real_+,
2271: \end{eqnarray*}
2272:  which shows that $(W^\Omega_t(a))_{a\in [0,T]}$ is a continuous Gaussian
2273:  martingale with quadratic variation $\sigma^2_a da$
2274:  for fixed $t\in \real_+$.
2275: \\
2276: 
2277: \noindent
2278:  Denote by $(B_t)_{t\in \real_+}$ the $H^\vide $-valued
2279:  Wiener process represented as
2280: $$
2281:  B_t
2282:  =
2283:  \sum_{n=0}^\infty
2284:  \frac{h_n}{\Vert h_n \Vert_{H^\vide}^2}
2285:  \beta_n ( t )
2286: ,
2287: $$
2288:  with $\beta_n ( t ) = \langle B_t , h_n \rangle_H$,
2289:  $n\geq 1$, and covariance
2290: $$\tilde{\E} [\langle B_s , h_n \rangle_{H^\vide }
2291:  \langle B_t , h_m \rangle_{H^\vide } ] =
2292:  {\bf 1}_{\{ n = m \}} ( s \wedge t ),
2293:  \qquad
2294:  s,t \in \real_+,
2295: $$
2296:  i.e.
2297: $$
2298:  \tilde{\E}
2299:  [\langle B_t , v_1 \rangle_{H^\vide }
2300:  \langle B_s ,v_2 \rangle_{H^\vide } ]
2301:  = ( s \wedge t )
2302:  \langle Q v_1, v_2\rangle_{H^\vide}
2303: ,
2304:  \qquad
2305:  s,t \in \real_+,
2306:  \quad
2307:  v_1,v_2\in H^\vide ,
2308: $$
2309:  where $Q:H^\vide  \to H^\vide $ is the operator with eigenvalues
2310:  $\{ \Vert h_n \Vert_{H^\vide}^{-2} \ : \ n \geq 1 \}$
2311:  in the Hilbert basis $(h_n)_{n\geq 1}$.
2312:  It\^o's formula for Hilbert-valued Wiener processes, cf.
2313:  Theorem~4.17 of \cite{daprato}, shows that
2314: $$F(B_t)
2315:  = F(B_0)
2316:  +
2317:  \int_0^t
2318:  \langle D F (B_s) , dB_s \rangle_{H^\vide}
2319:  +
2320:  \frac{1}{2}
2321:  \int_0^t
2322:  \Delta^\vide F (B_s)
2323:  ds
2324: , \qquad
2325:  F\in {\cal S},
2326: $$
2327:  hence $(B_t)_{t\in \real_+}$ has generator $\frac{1}{2} \Delta^\vide$.
2328: \\
2329: 
2330: \noindent
2331:  Dynkin's formula, cf. \cite{Dynkin}, Theorem~5.1, shows that
2332:  for all stopping time $\tau$ such that
2333:  $\tilde{\E} [\tau \mid B_0 = \omega ] < \infty$ we have,
2334:  $\P^\sigma (d\omega )$-a.s.:
2335: $$
2336:  \tilde{\E}
2337:  [
2338:  F ( B_\tau )
2339:  \mid B_0 = \omega
2340:  ]
2341:  -
2342:  F ( \omega )
2343:  =
2344:  \frac{1}{2}
2345:  \tilde{\E}
2346:  \left[
2347:  \int_0^\tau
2348:  \Delta^\vide F ( B_s )
2349:  ds
2350:  \mid B_0 = \omega
2351:  \right]
2352: ,
2353: $$
2354:  hence $\Delta^\vide F \leq 0$ implies
2355: $$
2356:  F(\omega )
2357:  \geq
2358:  \tilde{\E} [
2359:  F( B_\tau ) \mid B_0 = \omega
2360:  ]
2361: .
2362: $$
2363:  For $r>0$, let
2364: $$\tau_{r}
2365:  = \inf \{ t \in \real_+ \ : \ B_t \notin {\eufrak B}_r (B_0)\}
2366: $$
2367:  denotes the first exit time of $(B_t)_{t\in [0,T]}$ from the open ball
2368:  ${\eufrak B}_r (\omega )$ of radius $r>0$, centered at $B_0 = \omega \in \Omega$.
2369:  We have the following converse.
2370: \begin{prop}
2371:  Let $F\in \Dom (\Delta^\vide )$ be such that $\Delta^\vide F$ is continuous
2372:  on $\Omega$, and assume that there exists $r_0>0$ such that
2373: \begin{equation}
2374: \label{jkl}
2375:  F(\omega )
2376:  \geq
2377:   \tilde{\E} [
2378:  F( B_{\tau_r} ) \mid B_0 = \omega
2379:  ]
2380: , \qquad
2381:  \P^\sigma_u (d\omega )-a.s.,
2382:  \quad
2383:  0 < r < r_0
2384: .
2385: \end{equation}
2386:  Then $F$ is $\Delta^\vide$-superharmonic on $\Omega$ in the sense of
2387:  Relation \eqref{sph}.
2388: \end{prop}
2389: \begin{Proof}
2390:  From Remark~3, page 134 of \cite{Dynkin}, we have
2391: \begin{equation}
2392: \label{dk}
2393:  \frac{1}{2}
2394:  \Delta^\vide F ( \omega )
2395:  =
2396:  \lim_{n\to \infty}
2397:  \frac{ \tilde{\E}
2398:  [
2399:  F ( B_{\tau_{1/n}} )
2400:  \mid B_0 = \omega
2401:  ]
2402:  -
2403:  F ( \omega )
2404: }{\tilde{\E} [\tau_{1/n} \mid B_0 = \omega ]} ,
2405: \end{equation}
2406:  which shows that $\Delta^\vide F \leq 0$ when \eqref{jkl}
2407:  is satisfied.
2408: \end{Proof}
2409: \noindent
2410:  This yields in particular the following class of
2411:  $\Delta^\vide$-superharmonic functionals.
2412: \begin{prop}
2413:  Let the potential of $F\geq 0$ be defined by
2414: \begin{equation}
2415: \label{hj}
2416:  G (\omega ) = \int_0^{+\infty}
2417:  \tilde{\E}
2418:  [
2419:  F ( B_t ) \mid B_0 = \omega
2420:  ]
2421:  dt
2422: , \qquad
2423:  \P^\sigma_u (d\omega )-a.s.
2424: ,
2425: \end{equation}
2426:  assume that $G\in \Dom ( \Delta^\vide ) $ and
2427:  that $\Delta^\vide G$ is continuous on $\Omega$.
2428:  Then $G$ is a $\Delta^\vide$-superharmonic on $\Omega$.
2429: \end{prop}
2430: \begin{Proof}
2431:  For all $r>0$ we have
2432: \begin{eqnarray*}
2433:  G (\omega )
2434:  & = &
2435:  \tilde{\E}
2436:  \left[
2437:  \int_0^{\tau_r} F ( B_t )  dt
2438:  \Big|
2439:  B_0 = \omega
2440:  \right]
2441:  +
2442:  \tilde{\E}
2443:  [
2444:  G ( B_{\tau_r } )
2445:  \mid
2446:  B_0 = \omega
2447:  ]
2448: \\
2449: & \geq &
2450:  \tilde{\E}
2451:  [
2452:  G ( B_{\tau_r } )
2453:  \mid
2454:  B_0 = \omega
2455:  ]
2456: ,
2457: \end{eqnarray*}
2458:  which shows that $G$ is $\Delta^\vide$-superharmonic.
2459: \end{Proof}
2460: \noindent
2461:  Note that if $F$ is bounded with bounded support
2462:  in $\Omega$ then $G$ is bounded on $\Omega$,
2463:  see e.g. Remark~3.5 of \cite{grossp}.
2464: \\
2465: 
2466: \subsubsection*{Convolution}
2467: \noindent
2468:  Positive superharmonic functionals can also be obtained
2469:  by convolution, i.e. if
2470:  $F$ is $\Delta^\vide$-superharmonic and $G$ is positive and
2471:  sufficiently integrable, then
2472: $$
2473:  \omega \mapsto
2474:  \int_\Omega
2475:  G ( \tilde{\omega} )
2476:  F ( \omega - \tilde{\omega} )
2477:  \P^\sigma ( d \tilde{\omega} )
2478: $$
2479:  is positive and $\Delta^\vide$-superharmonic.
2480: \subsubsection*{Cylindrical functionals}
2481: \noindent
2482:  Superharmonic functionals on Gaussian space can also
2483:  be constructed as cylindrical functionals,
2484:  by composition with finite-dimensional functions.
2485:  Here we use the expansions of case $(A)$.
2486:  From the expression of $\Delta$ on cylindrical functionals
2487: $$
2488:  \Delta F
2489:  =
2490:  \sum_{i=1}^n
2491:  \partial_i^2 f_n
2492:  \left(
2493:  \lambda_1^{-1}
2494:  X^u ( h_1 )
2495:  ,
2496:  \ldots
2497:  ,
2498:  \lambda_n^{-1}
2499:  X^u ( h_n )
2500:  \right)
2501: ,
2502: $$
2503:  we check that
2504: $$F
2505:  =
2506:  f_n
2507:  \left(
2508:  \lambda_1^{-1}
2509:  X^u ( h_1 )
2510:  ,
2511:  \ldots
2512:  ,
2513:  \lambda_n^{-1}
2514:  X^u ( h_n )
2515:  \right)
2516: $$
2517:  is superharmonic on $\Omega$ if and only if $f_n$ is
2518:  superharmonic on $\real^n$.
2519:  Given $a\in \real$ and $b\in \real^n$, let
2520:  $f_{n,a,b} : \real^n \to \real$ be defined as
2521: $$
2522:  f_{n,a,b} (x_1,\ldots ,x_n) =
2523:  \Vert x + b \Vert^a
2524:  =
2525:  ((x_1+b_1)^2+\cdots + (x_n+b_n)^2)^{a/2}
2526: ,
2527: $$
2528:  then $\sqrt{f_{n,a,b}}$ is superharmonic on $\real^n$,
2529:  $n\geq 3$, if and only if $a\in [4-2n,0]$.
2530:  Let
2531: $$F_{n,a,b} = f_{n,a,b} \left(
2532:  \lambda_1^{-1} X^u ( h_1 )
2533:  ,
2534:  \ldots
2535:  ,
2536:  \lambda_n^{-1} X^u ( h_n )
2537:  \right)
2538: .
2539: $$
2540:  We have
2541: $$
2542:  D_t \log F_{n,a,b}
2543:  =
2544:  a
2545:  \sum_{i=1}^n
2546:  \frac{
2547:  \lambda_i^{-1}
2548:  \Gamma h_i (t)
2549:  \left(
2550:  b_i
2551:  +
2552:  \lambda_i^{-1} X^u ( h_i )
2553:  \right)
2554: }{
2555:  \left|
2556:   b_1 +
2557:  \lambda_1^{-1}
2558:  X^u ( h_1 )
2559:  \right|^2
2560:  +
2561:  \cdots
2562:  +
2563:  \left|
2564:  b_n
2565:  +
2566:  \lambda_n^{-1}
2567:  X^u ( h_n )
2568:  \right|^2
2569: }
2570: ,
2571: $$
2572:  and
2573: \begin{eqnarray*}
2574:  \Delta \sqrt{F_{n,a,b}}
2575:  & = &
2576: % \sigma^4
2577:  \sum_{i=1}^n
2578:  \partial_i^2
2579:  \sqrt{f_{n,a,b}}
2580:  \left(
2581:  \lambda_1^{-1} X^u ( h_1 )
2582:  ,
2583:  \ldots
2584:  ,
2585:  \lambda_n^{-1} X^u ( h_n )
2586:  \right)
2587: ,
2588: \end{eqnarray*}
2589:  since $( \Gamma h_k )_{k \geq 1}$ is orthogonal in
2590:  $L^2([0,T] , dt )$, hence
2591: \begin{eqnarray*}
2592: \label{eq:laplacienracineFsurracineF}
2593:  \frac{\Delta
2594:  \sqrt{F_{n,a,b}}}{\sqrt{F_{n,a,b}}}
2595:  & = &
2596:  \frac{
2597: % \sigma^4
2598:  a ( n -2 + a/2 )/2}{
2599:  \left|
2600:  b_1
2601:  +
2602:  \lambda_1^{-1} X^u ( h_1 )
2603:  \right|^2
2604:  +
2605:  \cdots
2606:  +
2607:  \left|
2608:  b_n
2609:  +
2610:  \lambda_n^{-1}
2611:  X^u ( h_n )
2612:  \right|^2
2613: }
2614: ,
2615: \end{eqnarray*}
2616:  is negative
2617:  if $4-2n \leq a \leq 0$, which is minimal for $a=2-n$.
2618:  We also have
2619: \begin{eqnarray*}
2620: %\label{eq:laplacienracineFsurracineF}
2621:  \frac{\Delta
2622:  F_{n,a,b}}{F_{n,a,b}}
2623:  & = &
2624:  \frac{
2625: % \sigma^4
2626:  a ( n+ a - 2 )}{
2627:  \left|
2628:  b_1
2629:  +
2630:  \lambda_1^{-1}
2631:  X^u ( h_1 )
2632:  \right|^2
2633:  +
2634:  \cdots
2635:  +
2636:  \left|
2637:  b_n
2638:  +
2639:  \lambda_n^{-1}
2640:  X^u ( h_n )
2641:  \right|^2
2642: }
2643: ,
2644: \end{eqnarray*}
2645:  which is negative for $a\in [2-n,0]$ and vanishes for $a=2-n$.
2646:  In this case the estimator is given by
2647: $$
2648:  D_t \log F_{n,2-n,b}
2649:  =
2650:  -
2651:  (n-2)
2652:  \sum_{i=1}^n
2653:  \frac{
2654:  \lambda_i^{-1}
2655:  \left(
2656:  b_i
2657:  +
2658:  \lambda_i^{-1}
2659:  X^u ( h_i )
2660:  \right)
2661:  \Gamma h_i ( t )
2662: }{
2663:  \left|
2664:  b_1 +
2665:  \lambda_1^{-1}
2666:  X^u ( h_1 )
2667:  \right|^2
2668:  +
2669:  \cdots
2670:  +
2671:  \left|
2672:  b_n
2673:  +
2674:  \lambda_n^{-1}
2675:  X^u ( h_n )
2676:  \right|^2
2677: }
2678: ,
2679: $$
2680:  and from Proposition~\ref{prop:erreurcylindricalfunctions.1},
2681:  inequality \eqref{sh} actually also holds as an equality:
2682: \begin{equation}
2683: \label{hldseq}
2684:  \E_u
2685:  \left[
2686:  \Vert X +
2687: % \sigma^2
2688:  D \log F_{n,2-n,b} - u \Vert_{L^2([0,T] , dt )}^2
2689:  \right]
2690:  =
2691:  \crb (\sigma,\mu, \hat{u} )
2692:  -
2693: % \sigma^4
2694:  \E_u \left[
2695:  \int_0^T
2696:  | D_t \log F_{n,2-n,b} |^2
2697:  dt \right]
2698: ,
2699: \end{equation}
2700:  with
2701: \begin{equation}
2702: \label{eq:est2}
2703:  \Vert D \log F_{n,2-n,b} \Vert_{L^2([0,T] , dt )}^2
2704:  =
2705:  \frac{ (n-2)^2 }{ \left|
2706:  b_1 +
2707:  \lambda_1^{-1}
2708:  X^u ( h_1 )
2709:  \right|^2
2710:  +
2711:  \cdots
2712:  +
2713:  \left|
2714:  b_n
2715:  +
2716:  \lambda_n^{-1}
2717:  X^u ( h_n )
2718:  \right|^2
2719: }
2720: .
2721: \end{equation}
2722: \noindent
2723:  Note that when $u$ is deterministic, any superharmonic functional of the form
2724: $$
2725:  f_n
2726:  \left(
2727:  \lambda_1^{-1}
2728:  X^u ( h_1 )
2729:  ,
2730:  \ldots
2731:  ,
2732:  \lambda_n^{-1}
2733:  X^u ( h_n )
2734:  \right)
2735: ,
2736: $$
2737:  can be replaced with
2738: $$
2739:  f_n
2740:  \left(
2741:  \lambda_1^{-1}
2742:  X ( h_1 )
2743:  ,
2744:  \ldots
2745:  ,
2746:  \lambda_n^{-1}
2747:  X ( h_n )
2748:  \right)
2749: ,
2750: $$
2751:  which retains the same harmonicity property,
2752:  and can be directly computed from an observation of $X$.
2753: \\
2754: 
2755: \noindent
2756:  The Stein type estimator of $u$ is given by
2757: $$
2758:  X_t + D_t \log F_{n,2-n,b}, \qquad t \in [0,T],
2759: $$
2760:  with
2761: $$
2762:  b_i = \lambda_i^{-1} \langle u , h_i\rangle,
2763:  \qquad i=1,\ldots , n,
2764: $$
2765:  i.e.
2766: $$
2767:  D_t \log F_{n,2-n,b}
2768:  =
2769:  -
2770:  (n-2)
2771:  \frac{[\Pi_n X]_t }{\Vert \Pi_n X \Vert_{L^2([0,T] , dt )}^2}
2772: ,
2773: $$
2774:  where $\Pi_n$ denotes the orthogonal projection
2775: $$
2776:  \Pi_n X (t)
2777:  : =
2778:  \sum_{k=1}^n
2779:  \lambda_k^{-1}
2780:  X (h_k)
2781:  \Gamma h_k (t)
2782:  =
2783:  \sum_{k=1}^n
2784:  \lambda_k^{-1}
2785:  \left(
2786:  b_k
2787:  +
2788:  \lambda_k^{-1}
2789:  X^u( h_k )
2790:  \right)
2791:  \Gamma h_k (t)
2792: .
2793: $$
2794:  We have
2795: \begin{eqnarray*}
2796:  \Vert D \log F_{n,2-n,b} \Vert_{L^2 ([0,T]\times \Omega , \P_u \otimes dt )}^2
2797:  & = &
2798:  - 4
2799:  \E_u \left[
2800:  \frac{\Delta \sqrt{F_{n,2-n,b}}}{\sqrt{F_{n,2-n,b}}} \right]
2801: \\
2802:  & = &
2803:  (n-2)^2
2804:  \E_u \left[
2805:  \frac{ 1 }{ \left|
2806:  \lambda_1^{-1} X ( h_1 )
2807:  \right|^2
2808:  +
2809:  \cdots
2810:  +
2811:  \left|
2812:  \lambda_n^{-1}  X ( h_n )
2813:  \right|^2
2814: }
2815:  \right]
2816: \\
2817:  & = &
2818:  (n-2)^2
2819:  \E_u \left[
2820:  \Vert \Pi_n X \Vert_{L^2([0,T] , dt )}^{-2}
2821:  \right]
2822: ,
2823: \end{eqnarray*}
2824:  and
2825: $$
2826:  \E_u
2827:  \left[
2828:  \Vert X
2829:  +
2830:  D \log F_{n,2-n,b} - u \Vert_{L^2([0,T] , dt )}^2
2831:  \right]
2832:  =
2833:  \crb (\sigma,\mu, \hat{u} )
2834:  -
2835:  (n-2)^2
2836:  \E_u \left[
2837:  \Vert \Pi_n X \Vert_{L^2([0,T] , dt )}^{-2}
2838:  \right]
2839: .
2840: $$
2841:  Note that the estimator
2842: $$
2843:  X_t - (n-2)
2844:  \frac{[\Pi_n X]_t }{\Vert \Pi_n X \Vert_{L^2([0,T] , dt )}^2},
2845:  \qquad
2846:  t\in [0,T],
2847: $$
2848:  is of James-Stein type, but it is not a shrinkage operator.
2849:  Another difference with James-Stein estimators is that here
2850:  the denominator consists in a sum of squared Gaussians with different
2851:  variances.
2852: \\
2853: 
2854: \noindent
2855:  Given $(X_t^1)_{t\in [0,T]}, \ldots , (X_t^N )_{t\in [0,T]}$,
2856:  $N$ independent samples of $(X_t)_{t\in [0,T]}$, the process
2857: $$\bar{X}_t = \frac{1}{N}
2858:  \left(
2859:  X_t^1 + \cdots + X_t^N \right)
2860: $$
2861:  is a Brownian motion with drift $u$ and
2862:  quadratic variation
2863:  $\sigma^2_t dt /N$ under $\P_u$, and
2864:  can be used for both efficient and Stein type
2865:  estimation.
2866: \section{Numerical application}
2867: \label{4}
2868: \noindent
2869:  In this section we present numerical simulations which
2870:  allow us to measure the efficiency of our estimators.
2871:  We use the framework of case $(A)$ and the
2872:  superharmonic functionals constructed
2873:  as cylindrical functionals in the previous section,
2874:  and we assume that $u \in H$ is deterministic.
2875: \\
2876: 
2877: \noindent
2878:  We work in the independent increment framework of Section~\ref{2}
2879:  and we additionally assume that $\sigma_t = \sigma$ is constant, $t\in [0,T]$,
2880:  i.e. $(X_t)_{t\in [0,T]}$ is a Brownian motion with variance $\sigma^2$,
2881:  $\Gamma h (t)= \sigma^2h(t)$, $t\in [0,T]$, and
2882: $$\crb (\sigma,\mu, \hat{u} ) = \frac{\sigma^2}{2} T^2.
2883: $$
2884:  Letting
2885: $$
2886:  h_n (t) =
2887: \frac{\sqrt{2T}}{
2888:  \sigma
2889:  \pi
2890:  \left( n- 1/ 2 \right)
2891: }
2892:  \sin \left( \left(
2893:  n-\frac{1}{2} \right) \frac{ \pi t}{T}
2894:  \right),
2895:  \qquad
2896:  t \in [0,T]
2897: ,
2898:  \quad
2899:   n \geq 1,
2900: $$
2901:  i.e.
2902: $$\dot{h}_n (t) =
2903:  \frac{1}{\sigma}
2904:  \sqrt{\frac{2}{T}}
2905:  \cos \left(
2906:  \left(
2907:  n-\frac{1}{2} \right) \frac{ \pi t}{T}
2908:  \right),
2909:  \qquad
2910:  t \in [0,T],
2911:  \quad n \geq 1,
2912: $$
2913:  provides an orthonormal basis $(h_n)_{n\geq 1}$ of $H$
2914:  such that $(\Gamma h_k)_{k\geq 1}$ is orthogonal in $L^2([0,T] , dt )$, with
2915: $$
2916:  \lambda_n
2917:  =
2918:  \frac{\sigma T}{
2919:  \pi
2920:  ( n-1/2 ) }, \qquad n\geq 1
2921: ,
2922: $$
2923:  solution of \eqref{*1}.
2924:  The estimator of $u$ will be given by
2925: $$
2926:  D_t \log F_{n,2-n,b}
2927:  =
2928:  -
2929:  (n-2)
2930:  \sqrt{\frac{2}{T}}
2931:  \sum_{k=1}^n
2932:  \frac{
2933:  \displaystyle
2934:  X (
2935:  h_k
2936:  )
2937:  }{
2938:  \left|
2939:  \lambda_1^{-1}
2940:  X ( h_1 )
2941:  \right|^2
2942:  +
2943:  \cdots
2944:  +
2945:  \left|
2946:  \lambda_n^{-1} X ( h_n )
2947:  \right|^2
2948: }
2949:  \sin \left(
2950:  \left(
2951:  k-\frac{1}{2}
2952:  \right)
2953:  \frac{ \pi t}{T} \right)
2954: ,
2955: $$
2956:  For simulation purposes we will use $X+D\log F$, and
2957:  construct the (nondrifted) Brownian motion $(X^u_t)_{t\in [0,T]}$
2958:  via the Paley-Wiener expansion
2959: \begin{equation}
2960: \label{pw}
2961:  X^u_t
2962:  = \sigma^2
2963:  \sum_{n=1}^\infty \eta_n h_n (t)
2964:  =
2965:  \sigma
2966:  \frac{\sqrt{2T}}{\pi}
2967:  \sum_{n=1}^\infty \eta_n
2968:  \frac{
2969:  \sin \left(
2970:  \left(
2971:  n-\frac{1}{2} \right) \frac{ \pi t}{T}
2972:  \right)}{
2973:  \left(
2974:  n-\frac{1}{2}
2975:  \right)
2976:  }
2977: ,
2978: \end{equation}
2979:  where $(\eta_n)_{n\geq 1}$ are independent standard Gaussian
2980:  random variables
2981:  with unit variance under $\P_u$ and
2982: $$
2983:  \eta_n
2984:  =
2985:  \int_0^T \dot{h}_n (s) dX^u_s
2986: ,
2987:  \qquad
2988:  n\geq 1
2989: .
2990: $$
2991:  In this case we have
2992: \begin{eqnarray}
2993: \label{hldseq1}
2994:  \lefteqn{
2995:  D_t \log F_{n,2-n,b}
2996: }
2997: \\
2998: \nonumber
2999:  & = &
3000:  -
3001:  (n-2)
3002:  \sqrt{\frac{2}{T}}
3003:  \sum_{k=1}^n
3004:  \frac{
3005:  \displaystyle
3006: % \left(
3007:  \eta_k
3008:  +
3009:  \langle u , h_k \rangle
3010: % \right)
3011:  }{
3012:  \displaystyle
3013:  \sum_{l=1}^n
3014:  \lambda_l^{-2}
3015:  \left(
3016:  \eta_l
3017:  +
3018:  \langle u , h_l \rangle
3019:  \right)^2
3020: }
3021:  \sin \left(
3022:  \left(k-\frac{1}{2} \right) \frac{ \pi t}{T}
3023:  \right)
3024: .
3025: \end{eqnarray}
3026: \noindent
3027: %\subsubsection*{Numerical Simulations}
3028: \noindent
3029:  Recall that the improvement obtained in comparison with the
3030:  efficient estimator $\hat{u}$ is not obtained
3031:  pathwise, but in expectation.
3032:  The gain of the superefficient estimator
3033:  $X + D \log F_{n,2-n,b}$ compared to the
3034:  efficient estimator $\hat{u}$ is given by
3035: $$
3036:  G ( u , \sigma , T , n )
3037:  :
3038:  =
3039:  - \frac{4}{ \crb (\sigma,\mu, \hat{u} ) }
3040:  \E_u \left[ \frac{\Delta \sqrt{F_{n,2-n,b}}}{\sqrt{F_{n,2-n,b}}} \right]
3041: $$
3042:  as a function of $n \geq 3$.
3043:  From \eqref{hldseq} and \eqref{hldseq1} we have
3044: \begin{equation}
3045: \label{jkl2}
3046:  G ( u , \sigma , T , n )
3047:  =
3048:  2
3049:  (n-2)^2
3050:  \E \left[
3051:  \left(
3052:  \displaystyle
3053:  \sum_{l=1}^n
3054:  \left(
3055:  \pi
3056:  \left(
3057:  l
3058:  -
3059:  \frac{1}{2}
3060:  \right)
3061:  \left(
3062:  \eta_l
3063:  +
3064:  \langle u , h_l \rangle
3065:  \right)
3066:  \right)^2
3067:  \right)^{-1}
3068:  \right]
3069: ,
3070: \end{equation}
3071:  hence $G(u,\sigma , T , n)$ converges to
3072: \begin{equation}
3073: \label{gsint}
3074:  (n-2)^2
3075:  \frac{8}{\pi^2}
3076:  \E \left[
3077:  \left(
3078:  \displaystyle
3079:  \sum_{l=1}^n
3080:  \left(
3081:  2 l
3082:  -
3083:  1
3084:  \right)^2
3085:  \eta_l^2
3086:  \right)^{-1}
3087:  \right]
3088: ,
3089: \end{equation}
3090:  as $\sigma$ tends to infinity.
3091:  The quantity \eqref{gsint} can be evaluated as a Gaussian integral
3092:  to yield \eqref{gsint1}.
3093:  Unlike in the classical Stein method, we stress that here $n$ becomes
3094:  a free parameter and there is some interest in determining the values
3095:  of $n$ which yield the best performance.
3096: %\newpage
3097: \begin{prop}
3098: \label{prop:limitesestimateur}
3099:  For all $\sigma , T>0$, and $u\in H$ we have
3100: $$
3101:  G( u , \sigma , T, n )
3102:  \simeq
3103:  \frac{6}{ n \pi^2}
3104: $$
3105:  as $n$ goes to infinity.
3106: \end{prop}
3107: \begin{Proof}
3108:  Let
3109: $$S_n =
3110:  \sum_{l=1}^n
3111:  \left(
3112:  \pi
3113:  \left(
3114:  n - l
3115:  +
3116:  \frac{1}{2}
3117:  \right)
3118:  \left(
3119:  \eta_l
3120:  +
3121:  \langle u , h_l \rangle
3122:  \right)
3123:  \right)^2
3124: ,
3125:  \qquad
3126:  n \geq 1
3127: .
3128: $$
3129:  We have
3130: $$
3131:  G ( \alpha , \sigma , T , n )
3132:  =
3133:  2 (n-2)^2
3134:  \E \left[
3135:  \frac{1}{
3136:  S_n}
3137:  \right]
3138: ,
3139: $$
3140:  and by the strong law of large numbers, $n (n-2)^2{S_n}^{-1}$ converges to
3141:  $3/\pi^2$ as $n$ goes to infinity, since
3142: $$
3143:  \lim_{n\to \infty}
3144:  \frac{\E [S_n]}{n^3}
3145:  =
3146:  \frac{\pi^2}{4}
3147:  \lim_{n\to \infty}
3148:  \frac{1}{n^3}
3149:  \sum_{i=1}^n (2i-1)^2
3150:  =
3151:  \frac{\pi^2}{3}
3152: .
3153: $$
3154:  Now for all $n >10$ we have
3155: \begin{eqnarray*}
3156: \E_u
3157: \left[
3158:  \left(
3159:  \frac{(n-2)^3}{S_n}
3160:  \right)^2
3161: \right]
3162:  & = &
3163: \E
3164: \left[
3165: \Lambda (u)
3166:  \left(
3167:  \frac{(n-2)^3}{S_n}
3168:  \right)^2
3169: \right]
3170: \\
3171:  & \leq &
3172:  n^2 \pi^2
3173: \E
3174: \left[
3175: \Lambda (u)^2
3176: \right]^{1/2}
3177: \E
3178: \left[
3179:  \left(
3180:  \sum_{l=1}^{[n/2]}
3181:  \left(
3182:  1 - \frac{l}{n}
3183:  +
3184:  \frac{1}{2n}
3185:  \right)^2
3186:  \eta_l^2
3187:  \right)^{-4}
3188: \right]^{1/2}
3189: \\
3190:  & \leq &
3191:  n^2
3192:  \frac{4}{\pi^4}
3193: \E
3194: \left[
3195: \Lambda (u)^2
3196: \right]^{1/2}
3197: \E
3198: \left[
3199:  \left(
3200:  \sum_{l=1}^{[n/2]}
3201:  \eta_l^2
3202:  \right)^{-4}
3203: \right]^{1/2}
3204: \\
3205:  & \leq &
3206:  \frac{4n^2}{\pi^4}
3207: \E
3208: \left[
3209: \Lambda (u)^2
3210: \right]^{1/2}
3211:  \left(
3212:  \prod_{k=1}^4
3213:  \left(
3214:  [n/2] - 2 k
3215:  \right)
3216:  \right)^{-1/4}
3217: ,
3218: \end{eqnarray*}
3219:  hence $n^3/S_n$ is uniformly integrable in $n > 16$,
3220:  where $[n/2]$ denotes the integer part of $n/2$.
3221:  This concludes the proof.
3222: \end{Proof}
3223: %\newpage
3224: \noindent
3225:  In the sequel we choose $u_t=\alpha t$, $t\in [0,T]$, $\alpha \in  \real$.
3226:  Figure~\ref{g1} gives a sample path representation of the process $X+D\log F$.
3227: \vskip0cm
3228: \begin{figure}[!ht]
3229: \centering \caption{\small $u(t)=t, \; t \in [0,T]; \; n=5$.}
3230: \label{g1}
3231: \resizebox*{13cm}{7cm}{\rotatebox{0}{\includegraphics[scale=0.9]{drifttraceNB.eps}}}
3232: \end{figure}
3233: \noindent
3234:  In this case, from \eqref{jkl2} we have
3235: $$
3236:  G ( \alpha , \sigma , T , n )
3237:  = 2
3238:  (n-2)^2
3239:  \E \left[
3240:  \left(
3241:  \displaystyle
3242:  \sum_{l=1}^n
3243:  \left(
3244:  \pi
3245:  \left(
3246:  l
3247:  -
3248:  \frac{1}{2}
3249:  \right)
3250:  \eta_l
3251:  -
3252:  \alpha
3253:  \frac{\sqrt{2T}}{\sigma}
3254:  (-1)^l
3255:  \right)^2
3256:  \right)^{-1}
3257:  \right]
3258: ,
3259: $$
3260:  from which it follows that $G(\alpha , \sigma , T , n)$ converges to
3261: $$
3262:  (n-2)^2
3263:  \frac{8}{\pi^2}
3264:  \E \left[
3265:  \left(
3266:  \displaystyle
3267:  \sum_{l=1}^n
3268:  \left(
3269:  2 l
3270:  -
3271:  1
3272:  \right)^2
3273:  \eta_l^2
3274:  \right)^{-1}
3275:  \right]
3276: ,
3277: $$
3278:  when $\alpha^{-2}\sigma^2/T$ tends to infinity, and is equivalent to
3279: $$
3280:  \left( 1 -\frac{2}{n} \right)^2
3281:  \frac{\sigma^2}{\alpha^2 T}
3282: $$
3283:  as $\alpha^{-2}\sigma^2/T$ tends to $0$.
3284: %\newpage
3285: \noindent
3286:  Figure~\ref{g3} represents the gain in percentage of the
3287:  superefficient estimator
3288:  $X + \sigma^2 D \log F_{n,2-n,b}$
3289:  compared to the efficient estimator $\hat{u}$ using Monte-Carlo simulations, i.e.
3290:  we represent $100\times G(\alpha , \sigma , T , n)$ as a function of $n\geq 3$.
3291: \vskip0cm
3292: \begin{figure}[!ht]
3293: \centering
3294: \caption{\small Percentage gain as a function of $n$ for 10000 samples and $\alpha = \sigma = T=1$.}
3295: \label{g3}
3296: \resizebox*{13cm}{7cm}{\rotatebox{0}{\includegraphics[scale=0.9]{gain0-12-1.eps}}}
3297: \end{figure}
3298: %\newpage
3299: %
3300: \noindent
3301:  An optimal value
3302: $$
3303:  n_{\rm opt}
3304:  = \argmax
3305:  \left\{
3306:  G ( \alpha , \sigma , T , n )
3307:  ~ : ~
3308:  n\geq 3
3309:  \right\}
3310: $$
3311:  of $n$ exists in general and is equal to $4$ when $\alpha = \sigma = T = 1$.
3312: %\newpage
3313: \noindent
3314:  Figure~\ref{g4} shows the variation of the gain as a function
3315:  of $n$ and $T$ for $\alpha = \sigma = 1$.
3316: \vskip-0.5cm
3317: \begin{figure}[!ht]
3318: \centering \caption{\small Gain as a function of $n$ and $T$.}
3319: \label{g4}
3320: \resizebox*{13cm}{7cm}{\rotatebox{0}{\includegraphics[scale=0.9]{gainnetTbisNB.eps}}}
3321: \end{figure}
3322: \noindent
3323:  Figure~\ref{g5} represents the variation of the gain as a function
3324:  of $n$ and $\sigma$.
3325: \vskip0cm
3326: \begin{figure}%[!ht]
3327: \centering \caption{\small Gain as a function of $n$ and $\sigma$.}
3328: \label{g5}
3329: \resizebox*{13cm}{7cm}{\rotatebox{0}{\includegraphics[scale=0.9]{gainnetsigmaNB.eps}}}
3330: \\
3331: \end{figure}
3332: %\subsubsection*{Acknowledgement} We thank the editors and referees for suggestions which led to several improvements of this paper, and in particular for communicating to us the references \cite{berger1} and \cite{berger2}.
3333: \section{Appendix}
3334:  The next Proposition is classical in the framework of Gaussian filtering
3335:  and is needed in Section~\ref{2} for Bayes estimation.
3336:  Its proof is stated for completeness since we did not find it in the literature.
3337: \begin{prop}
3338: %\label{ljk}
3339:  Let $Z$ be a Gaussian process with covariance operator $\Gamma_\tau$
3340:  and drift $v\in H$,
3341:  and assume that $X$ is a Gaussian process with drift $Z$
3342:  and quadratic covariance operator $\Gamma$ given $Z$.
3343:  Then, conditionally to $X$, $Z$ has drift
3344: $$
3345:  f
3346:  \mapsto
3347:  \langle
3348:  \stackrel{}{\chi}_t
3349:  ,
3350:  ( \Gamma+\Gamma_\tau)^{-1}\Gamma
3351:  v
3352:  \rangle
3353:  +
3354:  X (
3355:  ( \Gamma+\Gamma_\tau)^{-1}\Gamma_\tau f
3356:  \stackrel{}{\chi}_t
3357:  )
3358:  \quad
3359:  \mbox{and covariance}
3360:  \quad
3361:  \Gamma_\tau ( \Gamma + \Gamma_\tau )^{-1} \Gamma
3362: .
3363: $$
3364: \end{prop}
3365: \begin{Proof}
3366:  For convenience of notation, let
3367: $$
3368:  V(f)
3369:  =
3370:  \langle
3371:  f
3372:  ,
3373:  ( \Gamma+\Gamma_\tau)^{-1}\Gamma
3374:  v
3375:  \rangle
3376: ,
3377:  \qquad
3378:  f\in H
3379: .
3380: $$
3381:  For all $f, g \in H$ we have:
3382: \begin{eqnarray*}
3383:  \E\left[
3384:  \exp
3385:  \left(
3386:  i
3387:  X (
3388:  f
3389:  )
3390:  \right)
3391:  \right]
3392:  & = &
3393:  \E\left[
3394:  \E\left[
3395:  \exp
3396:  \left(
3397:  i
3398:  X ( f )
3399:  \right)
3400:  \Big |
3401:  Z
3402:  \right]
3403:  \right]
3404: \\
3405:  & = &
3406:  \E\left[
3407:  \exp
3408:  \left(
3409:  i
3410:  Z ( f )
3411:  -
3412:  \frac{1}{2}
3413:  \langle
3414:  f
3415:  ,
3416:  \Gamma
3417:  f \rangle
3418:  \right)
3419: \right]
3420: \\
3421:  & = &
3422:  \exp
3423:  \left(
3424:  i
3425:  V ( f )
3426:  -
3427:  \frac{1}{2}
3428:  \langle
3429:  f
3430:  ,
3431:  (
3432:  \Gamma_\tau
3433:  +
3434:  \Gamma
3435:  )
3436:  f \rangle
3437: \right)
3438: ,
3439: \end{eqnarray*}
3440:  and
3441: \begin{eqnarray*}
3442: \lefteqn{
3443:  \E\left[
3444:  \exp
3445:  \left(
3446:  i
3447:  X (
3448:  g
3449:  )
3450:  \right)
3451:  \exp
3452:  \left(
3453:  i
3454:  Z ( f  )
3455:  \right)
3456:  \right]
3457:  }
3458: \\
3459:  & = &
3460:  \E\left[
3461:  \exp
3462:  \left(
3463:  i
3464:  Z ( f )
3465:  \right)
3466:  \E\left[
3467:  \exp
3468:  \left(
3469:  i
3470:  X ( g )
3471:  \right)
3472:  \Big |
3473:  Z
3474:  \right]
3475:  \right]
3476: \\
3477:  & = &
3478:  \E\left[
3479:  \exp
3480:  \left(
3481:  i
3482:  Z (
3483:  f +
3484:  g
3485:  )
3486:  -
3487:  \frac{1}{2}
3488:  \langle
3489:  g , \Gamma g
3490:  \rangle
3491:  \right)
3492: \right]
3493: \\
3494:  & = &
3495:  \exp
3496:  \left(
3497:  -
3498:  \frac{1}{2}
3499:  \langle
3500:  f+g,
3501:  \Gamma_\tau (f+g)
3502:  \rangle
3503:  -
3504:  \frac{1}{2}
3505:  \langle
3506:  g
3507:  ,
3508:  \Gamma
3509:  g
3510:  \rangle
3511:  +
3512:  i
3513:  V ( f+g )
3514: \right)
3515: \\
3516:  & = &
3517:  \exp
3518:  \left(
3519:  i
3520:  V (
3521:  (\Gamma + \Gamma_\tau )^{-1} \Gamma_\tau
3522:  f
3523:  )
3524:  +
3525:  i
3526:  V (
3527:  g
3528:  +
3529:  (\Gamma + \Gamma_\tau )^{-1}
3530:  \Gamma
3531:  f
3532:  )
3533:  -
3534:  \frac{1}{2}
3535:  \langle
3536:  \Gamma_\tau f , (\Gamma +\Gamma_\tau )^{-1}
3537:  \Gamma f
3538:  \rangle
3539: \right.
3540: \\
3541: & & \left.
3542:  -
3543:  \frac{1}{2}
3544:  \langle
3545:  g
3546:  +
3547:  (\Gamma + \Gamma_\tau )^{-1}
3548:  \Gamma_\tau
3549:  f
3550:  ,
3551:  (\Gamma + \Gamma_\tau)
3552:  (
3553:  g
3554:  +
3555:  (\Gamma + \Gamma_\tau )^{-1}
3556:  \Gamma_\tau
3557:  f
3558:  )
3559:  \rangle
3560:  \right)
3561: \\
3562:  & = &
3563:  \E\left[
3564:  \exp
3565:  \left(
3566:  i
3567:  X ( g )
3568:  \right)
3569: \right.
3570: \\
3571: & &
3572: \left.
3573:  \exp
3574:  \left(
3575:  i
3576:  X (
3577:  (\Gamma + \Gamma_\tau )^{-1}
3578:  \Gamma_\tau f )
3579:  +
3580:  i
3581:  V((\Gamma +\Gamma_\tau)^{-1}\Gamma f)
3582:  -
3583:  \frac{1}{2}
3584:  \langle
3585:  \Gamma_\tau f
3586:  ,
3587:  (\Gamma_\tau + \Gamma )^{-1}
3588:  \Gamma
3589:  f
3590:  \rangle
3591:  \right)
3592:  \right]
3593: ,
3594: \end{eqnarray*}
3595:  which shows that
3596: \begin{eqnarray*}
3597: \lefteqn{
3598:  \E\left[
3599:  \exp
3600:  \left(
3601:  i
3602:  Z (f)
3603:  \right)
3604:  \Big |
3605:  X
3606:  \right]
3607: }
3608: \\
3609:  & = &
3610:  \exp
3611:  \left(
3612:  i
3613:  V (
3614:  (\Gamma + \Gamma_\tau )^{-1}
3615:  \Gamma
3616:  f )
3617:  +
3618:  i
3619:  X (
3620:  (\Gamma +\Gamma_\tau)^{-1}\Gamma_\tau
3621:  f )
3622:  -
3623:  \frac{1}{2}
3624:  \langle
3625:  \Gamma_\tau f
3626:  ,
3627:  (\Gamma + \Gamma_\tau )^{-1}
3628:  \Gamma
3629:  f \rangle
3630:  \right)
3631: .
3632: \end{eqnarray*}
3633: \end{Proof}
3634: \noindent
3635:  In particular we get the following corollary
3636:  which is classical in the framework of Gaussian filtering.
3637: \begin{prop}
3638: \label{ljk2}
3639:  Let $(Z_t)_{t\in [0,T]}$ be a Brownian motion with quadratic variation $\tau^2_tdt$,
3640:  $\tau \in L^2 ([0,T] , dt )$, and drift $(v_t)_{t\in [0,T]}$, $v\in H$,
3641:  and let $(X_t)_{t\in [0,T]}$ have drift $(Z_t)_{t\in [0,T]}$
3642:  and quadratic variation $(\sigma^2_t)_{t\in [0,T]}$,
3643:  given $Z$.
3644:  Then, conditionally to $X$, the process $(Z_t)_{t\in [0,T]}$ has drift
3645: $$
3646:  \int_0^t \frac{\sigma^2_s}{\tau^2_s+\sigma^2_s} dv_s
3647:  +
3648:  \int_0^t \frac{\tau^2_s}{\tau^2_s+\sigma^2_s} dX_s
3649:  \quad
3650:  \mbox{and variance}
3651:  \quad
3652: \int_0^t \frac{\tau^2_s\sigma^2_s}{\tau^2_s+\sigma^2_s} ds,
3653:  \qquad t\in [0,T]
3654: .
3655: $$
3656: \end{prop}
3657: 
3658: ~
3659: %\\
3660: 
3661: %\footnotesize
3662: 
3663: \small
3664: 
3665: \def\cprime{$'$} \def\polhk#1{\setbox0=\hbox{#1}{\ooalign{\hidewidth
3666:   \lower1.5ex\hbox{`}\hidewidth\crcr\unhbox0}}}
3667:   \def\polhk#1{\setbox0=\hbox{#1}{\ooalign{\hidewidth
3668:   \lower1.5ex\hbox{`}\hidewidth\crcr\unhbox0}}} \def\cprime{$'$}
3669: \begin{thebibliography}{10}
3670: 
3671: \bibitem{amn}
3672: E.~Al{\`o}s, O.~Mazet, and D.~Nualart.
3673: \newblock Stochastic calculus with respect to {G}aussian processes.
3674: \newblock {\em Ann. Probab.}, 29(2):766--801, 2001.
3675: 
3676: \bibitem{berger1}
3677: J.~Berger and R.~Wolpert.
3678: \newblock Estimating the mean function of a {G}aussian process and the {S}tein
3679:   effect.
3680: \newblock {\em J. Multivariate Anal.}, 13(3):401--424, 1983.
3681: 
3682: \bibitem{daprato}
3683: G.~Da~Prato and J.~Zabczyk.
3684: \newblock {\em Stochastic equations in infinite dimensions}.
3685: \newblock Encyclopedia of Mathematics and its Applications. Cambridge
3686:   University Press, Cambridge, 1992.
3687: 
3688: \bibitem{Dynkin}
3689: E.B. Dynkin.
3690: \newblock {\em Markov processes. {V}ols. {I}, {II}}.
3691: \newblock Die Grundlehren der Mathematischen Wissenschaften. Academic Press
3692:   Inc., New York, 1965.
3693: 
3694: \bibitem{fourdrinier}
3695: D.~Fourdrinier, W.E. Strawderman, and M.T. Wells.
3696: \newblock On the construction of {Bayes} minimax estimators.
3697: \newblock {\em Ann. Stat.}, 26(2):660--671, 1998.
3698: 
3699: \bibitem{Goodman}
3700: V.~Goodman.
3701: \newblock Harmonic functions on {H}ilbert space.
3702: \newblock {\em J. Funct. Anal.}, 10:451--470, 1972.
3703: 
3704: \bibitem{grossp}
3705: L.~Gross.
3706: \newblock Potential theory on {H}ilbert space.
3707: \newblock {\em J. Funct. Anal.}, 1:123--181, 1967.
3708: 
3709: \bibitem{IbragimovRozanov}
3710: I.A. Ibragimov and Y.A. Rozanov.
3711: \newblock {\em Gaussian random processes}, volume~9 of {\em Applications of
3712:   Mathematics}.
3713: \newblock Springer-Verlag, New York, 1978.
3714: 
3715: \bibitem{jamesstein}
3716: W.~James and C.~Stein.
3717: \newblock Estimation with quadratic loss.
3718: \newblock In {\em Proc. 4th Berkeley Sympos. Math. Statist. and Prob., Vol. I},
3719:   pages 361--379. Univ. California Press, Berkeley, Calif., 1961.
3720: 
3721: \bibitem{liptser}
3722: R.S. Liptser and A.N. Shiryaev.
3723: \newblock {\em Statistics of random processes. {II}}, volume~6 of {\em
3724:   Applications of Mathematics (New York)}.
3725: \newblock Springer-Verlag, Berlin, 2001.
3726: 
3727: \bibitem{nualartm2}
3728: D.~Nualart.
3729: \newblock {\em The {M}alliavin calculus and related topics}.
3730: \newblock Probability and its Applications (New York). Springer-Verlag, Berlin,
3731:   second edition, 2006.
3732: 
3733: \bibitem{pr3}
3734: N.~Privault and A.~R\'eveillac.
3735: \newblock Stein estimation for the drift of {G}aussian processes using the
3736:   {M}alliavin calculus.
3737: \newblock Preprint, 2006, to appear in the Annals of Statistics.
3738: 
3739: \bibitem{p-r-c}
3740: N.~Privault and A.~R\'eveillac.
3741: \newblock Superefficient drift estimation on the {W}iener space.
3742: \newblock {\em C. R. Acad. Sci. Paris S\'er. I Math.}, 343:607--612, 2006.
3743: 
3744: \bibitem{prakasarao}
3745: B.L.S.~Prakasa Rao.
3746: \newblock {\em Statistical inference for diffusion type processes}, volume~8 of
3747:   {\em Kendall's Library of Statistics}.
3748: \newblock Edward Arnold, London, 1999.
3749: 
3750: \bibitem{stein}
3751: C.~Stein.
3752: \newblock Estimation of the mean of a multivariate normal distribution.
3753: \newblock {\em Ann. Stat.}, 9(6):1135--1151, 1981.
3754: 
3755: \bibitem{tudorviens}
3756: C.A. Tudor and F.G. Viens.
3757: \newblock Statistical aspects of the fractional stochastic calculus.
3758: \newblock {\em Ann. Stat.}, 35(3):1183--1212, 2007.
3759: 
3760: \bibitem{berger2}
3761: R.~Wolpert and J.~Berger.
3762: \newblock Incorporating prior information in minimax estimation of the mean of
3763:   a {G}aussian process.
3764: \newblock In {\em Statistical decision theory and related topics, III, Vol. 2
3765:   (West Lafayette, Ind., 1981)}, pages 451--464. Academic Press, New York,
3766:   1982.
3767: 
3768: \end{thebibliography}
3769: 
3770: 
3771: 
3772: \end{document}
3773: