math0703407/DMC.tex
1: \documentclass{m2an}
2: 
3: \usepackage{amsmath,amssymb}
4: \usepackage[latin1]{inputenc}
5: %\usepackage[french]{babel}
6: \usepackage{epsfig,float,psfrag}
7: \usepackage{color}
8: %\usepackage[notref,notcite]{showkeys}
9: \textwidth=16.0cm
10: \textheight=24.0cm
11: \setlength{\topmargin}{-1cm}
12: \setlength{\evensidemargin}{-0.04cm}
13: \setlength{\oddsidemargin}{-0.04cm}
14: \parindent=0cm
15: \parskip 0.3cm
16: %\numberwithin{equation}{section}
17: \newtheorem{thm}{Theorem}
18: \newtheorem{aadef}[thm]{Definition}
19: \newtheorem{alem}[thm]{Lemma}
20: \newtheorem{aprop}[thm]{Proposition}
21: \newtheorem{acor}[thm]{Corollary}
22: \newtheorem{arem}[thm]{Remark}
23: \newenvironment{adem}[1][]%
24:    {\ \\ {\bf Proof #1~: }}%
25:    {\hfill\mbox{\rule{2 true mm}{3 true mm}}\vskip 2 ex\noindent}
26: \newcommand{\E}{{\mathbb E}}
27: \renewcommand{\P}{{\mathbb P}}
28: \newcommand{\R}{{\mathbb R}}
29: \newcommand{\N}{{\mathbb N}}
30: \renewcommand{\t}{{\theta}}
31: \newcommand{\dt}{{\delta t}}
32: \newcommand{\Dt}{{\Delta t}}
33: \newcommand{\uu}[1]  {{\boldsymbol{#1}} }
34: \newcommand{\ds}[1]{\displaystyle{#1}}
35: 
36: 
37: 
38: \begin{document}
39: 
40: \title{Diffusion Monte Carlo method : Numerical Analysis in a Simple
41:   Case}
42: 
43: \thanks{We thank Eric Cancès (CERMICS) and Mathias Rousset (Université
44:   Paul Sabatier, Toulouse) for many fruitful discussions and
45:   Michel Caffarel (IRSAMC, Toulouse) for introducing us to the DMC
46:   method and suggesting the toy model studied in this paper. We also
47:   thank the referees for useful suggestions which helped us to improve the first draft
48:   of this paper.}
49: \author{M.~El Makrini}\author{B.~Jourdain}\author{T.~Leli\`evre}\address{ENPC-CERMICS, 6-8 av Blaise Pascal, Cit\'e
50:     Descartes, Champs sur Marne, 77455 Marne-la-Vall\'ee Cedex 2, France
51:     -
52:     e-mail: \{makrimo,jourdain,lelievre\}@cermics.enpc.fr}
53: 
54: %\date{}
55: 
56: \begin{abstract}
57: The Diffusion Monte Carlo method is devoted to the computation of
58: electronic ground-state energies of molecules. In this paper, we focus on
59: implementations of this method which consist in exploring the
60: configuration space with a {\bf fixed} number of random walkers evolving
61: according to a Stochastic Differential Equation discretized in time. We
62: allow stochastic reconfigurations of the walkers to reduce the
63: discrepancy between the weights that they carry.  On a simple
64: one-dimensional example, we prove the convergence of the method for a
65: fixed number of reconfigurations when the number of walkers tends to
66: $+\infty$ while the timestep tends to $0$. We confirm our theoretical
67: rates of convergence by numerical experiments. Various resampling
68: algorithms are investigated, both theoretically and numerically.
69: \end{abstract}
70: 
71: % \begin{resume} ... 
72: % \end{resume}
73: 
74: \subjclass{81Q05, 65C35, 60K35, 35P15}
75: \keywords{Diffusion Monte Carlo method, interacting particle systems,
76:   ground state, Schr\"odinger operator, Feynman-Kac formula}
77: 
78: \maketitle
79: 
80: \section*{Introduction}
81: 
82: %\subsection*{Electronic structures computations}
83: 
84: 
85: The computation of electronic structures of atoms, molecules and solids is a
86: central problem in chemistry and physics. We focus here on electronic
87: ground state calculations where the objective is the computation of
88:  the lowest eigenvalue (the so-called ground-state
89: energy) $E_0$ of a self-adjoint Hamiltonian $H=-\frac{1}{2}\Delta +V$ with domain $D_\mathcal{H}(H)$ on a
90: Hilbert space $\mathcal{H} \subset L^2(\R^{3N})$ where $N$ is the number
91: of electrons (see~\cite{handbook} for a general introduction):
92: \begin{equation}\label{eq:E0}
93: E_0=\inf \{  \langle \psi,H\psi \rangle,\, \psi \in D_\mathcal{H}(H), \|\psi\|=1
94: \},
95: \end{equation}
96: where $\langle\cdot,\cdot\rangle$ denotes the duality bracket on $L^2(\R^{3N})$
97: and $\|\cdot\|$ the $L^2(\R^{3N})$-norm. For simplicity, we omit the spin variables. The function $V$ describes the interaction between the electrons, and between the
98: electrons and the nuclei, which are supposed to be fixed point-like particles. The functions $\psi$ are square
99: integrable, their normalized square modulus $|\psi|^2$ being interpreted as the
100: probability density of the particles positions in space, and they satisfy an antisymmetry condition with respect to the
101: numbering of the electrons, due to the fermionic nature of the
102: electrons (Pauli principle): $\mathcal{H}=\bigwedge_{i=1}^N L^2(\R^3)$. We suppose that the potential $V$ is such that $E_0$ is an isolated
103: eigenvalue of $H$ (see~\cite{CJL} for sufficient conditions), and we
104: denote by $\psi_0$ a normalized eigenfunction
105: associated with $E_0$. 
106: 
107: %\subsection*{The Diffusion Monte Carlo (DMC) method}
108: 
109: Due to the
110: high dimensionality of the problem, stochastic methods are particularly
111: well suited to compute~$E_0$. The first particle approximation
112: scheme of such spectral quantities was introduced in~\cite{H84} for
113: finite state space models. Convergence analysis for such
114: interacting particle systems (both continuous
115: or discrete in time) first appeared in~\cite{DM03,Delmoral,Delmoralmiclo,DD04}. The Diffusion Monte Carlo (DMC) method is
116: widely used in chemistry (see~\cite{Caffarel,Umrigar}), but has been  only
117: recently considered from a mathematical viewpoint
118: (see~\cite{CJL,Rousset}). This method gives an estimate of~$E_0$ in
119: terms of the long-time limit of the expectation of a functional of
120: a drift-diffusion process with a source term. It requires an importance
121: sampling function~$\psi_I$ which approximates the ground-state $\psi_0$
122: of~$H$. Let us define the drift function $\uu{b}=\nabla \ln |\psi_I|$, the
123: so-called local energy $\ds{E_L=\frac{H \psi_I}{\psi_I}}$ and the DMC energy:
124: \begin{equation}\label{eq:EDMC}
125: E_{\rm DMC}(t)= \frac{\E\left(E_L(\uu{X}_t)\exp\left(-\int_0^t
126:       E_L(\uu{X}_s)ds\right)\right)}{\E\left(\exp\left(-\int_0^t
127:       E_L(\uu{X}_s)ds\right)\right)},
128: \end{equation}
129: where the $3N$-dimensional process $\uu{X}_t$ satisfies the stochastic
130: differential equation:
131: \begin{equation}\label{eq:SDE}
132: \left\lbrace
133: \begin{array}{l}
134: \ds{\uu{X}_t=\uu{X}_0 + \int_0^t \uu{b}(\uu{X}_s) \, ds + d \uu{W}_t,}\\
135: \uu{X}_0 \sim |\psi_I|^2(\uu{x}) \, d \uu{x}.
136: \end{array}
137: \right.
138: \end{equation}
139: The stochastic process $(\uu{W}_t)_{t \geq 0}$ is a standard $3N$-dimensional Brownian
140: motion. One can then show that (see~\cite{CJL})
141: \begin{equation}\label{eq:CV_E_DMC}
142: \lim_{t \to \infty} E_{\rm DMC}(t)=E_{{\rm DMC},0},
143: \end{equation}
144: where 
145: \begin{equation}\label{eq:EDMC0}
146: E_{{\rm DMC},0}=\inf \{  \langle
147: \psi,H\psi \rangle,\, \psi \in D_\mathcal{H}(H), \|\psi\|=1, \, \psi=0\mbox{ on }\psi_I^{-1}(0)
148: \}.
149: \end{equation}
150: We have proved in~\cite{CJL} that $E_{{\rm DMC},0} \geq E_0$, with equality if and
151: only if the nodal surfaces of $\psi_I$ coincide with those of a ground
152: state $\psi_0$ of $H$. In other words, if there exists a ground state
153: $\psi_0$ such that $\psi_I^{-1}(0)=\psi_0^{-1}(0)$, then $\lim_{t \to
154:   \infty} E_{\rm DMC}(t)=E_0$. The error $|E_0-E_{{\rm DMC},0}|$ is
155: related to the so-called fixed-node approximation, which is
156: well known by practitioners of the field (see~\cite{handbook}).
157: 
158: In this paper, we complement the theoretical results obtained in~\cite{CJL} with a numerical analysis in a simple case.
159: In practice, the longtime limit $E_{{\rm DMC},0}$ in (\ref{eq:CV_E_DMC}) is approximated by taking the
160: value of $E_{\rm DMC}$ at a (large) time $T>0$. Then $E_{\rm DMC}(T)$ is
161: approximated by using a
162: discretization in time of the stochastic differential
163: equation~(\ref{eq:SDE}) and of the integral in the exponential factor in
164: (\ref{eq:EDMC}), and an approximation of the expectation
165: values in~(\ref{eq:EDMC}) by an empirical mean over a large number $N$
166: of 
167: trajectories. These trajectories $(\uu{X}^i)_{1 \leq i \leq N}$, also called walkers in the
168:    physical literature or particles in the mathematical literature, satisfy a discretized version of~(\ref{eq:SDE}), and
169:    interact at times $n\Delta t$ for $n\in\{1,\hdots,\nu-1\}$ where
170:    $\Dt=T/\nu$ for $\nu\in{\mathbb N}^*$ through a stochastic
171:    reconfiguration step aimed at
172:    reducing the discrepancy between their exponential weights. We thus obtain an interacting
173:    particle system. The number of
174:    reconfiguration steps is $\nu-1$. The
175: stochastic differential equation \eqref{eq:SDE} is discretized
176: with a possibly smaller timestep $\dt=\Dt/\kappa=T/(\nu\kappa)$ with $\kappa\in\N^*$. The total
177: number of steps for the discretization of \eqref{eq:SDE} is then $K=\nu\kappa$.
178: 
179: In the following, we consider the following adapted version of the DMC
180: scheme with a fixed number of walkers (see~\cite{Caffarel}):
181: \begin{itemize}
182: \item[{$ \quad \bullet \;$}] {\bf Initialization} of an ensemble of $N$
183:   walkers $\left(\uu{X}^j_{0 \Dt} \right)_{1 \le j \le N}$ i.i.d. according to
184:   $|\psi_I|^2(\uu{x}) \, d\uu{x}$. 
185: \item[{$ \quad \bullet \;$}] {\bf Iterations in time:} let us be given
186:   the particle positions $\left( \uu{X}^j_{n \Dt} \right)_{1 \le j \le N}$
187:   at time $n \Delta t$, for $n \in \{0,\hdots,\nu-1\}$. The new particle positions at time $(n+1) \Delta t$ are obtained in two steps:
188: \begin{enumerate}
189: \item {\bf Walkers displacement:} for all $1 \le j \le N$, the
190:   successive positions $\left( \uu{X}^j_{n \Dt + \dt}, \, \hdots ,\,
191:     \uu{X}^j_{n \Dt + \kappa \dt}\right)$ over the time interval $(n \Delta t,(n+1) \Delta t)$ are obtained by an
192: appropriate discretization of~(\ref{eq:SDE}). In the field of
193: interacting particles system for Feynman-Kac formulae (see~\cite{Delmoral,Delmoralmiclo}), this step is called {\bf the mutation step}.
194: \item {\bf Stochastic reconfiguration:} The new positions\footnote{With a slight
195:   abuse of notation and though $n
196:     \Dt + \kappa \dt=(n+1)\Dt$, we distinguish between the particle positions $\uu{X}^j_{n
197:     \Dt + \kappa \dt}$ at the end of the walkers displacement on time interval
198:   $(n \Dt,(n+1) \Dt)$, and the new
199:   particle positions $\uu{X}^j_{(n+1)
200:     \Dt}$ obtained after the reconfiguration step, and
201:   which are used as the initial position for the next walkers displacement on time interval
202:   $((n+1) \Dt,(n+2) \Dt)$. We will use a more precise notation for the
203:   analysis of the numerical scheme in Section~\ref{sec:num_anal}, but this is not required
204: at this stage.} $\left(
205:     \uu{X}^j_{(n+1) \Dt} \right)_{1 \le j \le N}$ which will be used as
206:   the initial particle positions on the time interval $((n+1) \Dt,(n+2)
207:   \Dt)$ are obtained from independent sampling of the measure
208: \begin{equation}
209: \frac{\sum_{j=1}^N \exp\left( - \dt \sum_{k=1}^\kappa E_L(\uu{X}^j_{n
210:     \Dt + k \dt}) \right) \delta_{\uu{X}^j_{n
211:     \Dt + \kappa \dt}}}{ \sum_{j=1}^N \exp\left( - \dt \sum_{k=1}^\kappa E_L(\uu{X}^j_{n
212:     \Dt + k \dt}) \right)}.
213: \end{equation}
214: In words, the new particle positions $\left(
215:     \uu{X}^j_{(n+1) \Dt} \right)_{1 \le j \le N}$ are randomly chosen among the
216:   final particle positions $\left(
217:     \uu{X}^j_{n \Dt + \kappa \dt} \right)_{1 \le j \le N}$, each of them
218:   being weighted with the coefficient $\exp\left( - \dt \sum_{k=1}^\kappa E_L(\uu{X}^j_{n
219:     \Dt + k \dt}) \right)$ (accordingly to the exponential factor in~(\ref{eq:EDMC})).
220:  In the field of
221: interacting particles system for Feynman-Kac formulae, this step is called {\bf the
222:  selection step}.
223: \end{enumerate}
224: \end{itemize}
225: An estimate of $E_{\rm DMC}(t_{n+1})$ is then given by:
226: \begin{equation}\label{eq:EDMCapprox}
227: E_{\rm DMC}(t_{n+1}) \simeq \frac{1}{N} \sum_{j=1}^{N} \,
228:   E_L \left(\uu{X}^j_{(n+1) \Dt}\right).
229: \end{equation}
230: There are other possible estimations of $E_{\rm
231:   DMC}(t_{n+1})$. In~\cite{Caffarel}, the authors propose to use Cesaro
232: or weighted Cesaro means of the expression~(\ref{eq:EDMCapprox}). In Section~\ref{sec:num_anal}, we will use the
233: following expression: 
234: \begin{equation} \label{eq:EDMCapprox_prime}
235: E_{\rm DMC}(t_{n+1}) \simeq \frac{\sum_{j=1}^{N} \,
236:   E_L(\uu{X}^j_{n \Dt + \kappa \dt}) \exp\left( - \dt
237:     \sum_{k=1}^{\kappa} E_L(\uu{X}^j_{n \Dt + k \dt}) \right)}{\sum_{j=1}^{N}
238:   \exp\left( - \dt \sum_{k=1}^{\kappa} E_L(\uu{X}^j_{n \Dt + k \dt}) \right)},
239: \end{equation}
240: in an intermediate step to prove the convergence result. 
241: 
242: We would like to mention that a continuous in time version of the DMC scheme with stochastic
243: reconfiguration has been proposed in~\cite{Rousset}. The author
244: analyzes the longtime behavior of the interacting particle system and
245: proves in particular a uniform in time control of the variance of the
246: estimated energy.
247: 
248: The DMC algorithm presented above is prototypical. Many refinements
249: are used in practice. For example, an acception-rejection step is generally used in the
250: walkers displacement step (see~\cite{RCAL82}). This will not be
251: discussed here. Likewise, the selection
252: step can be done in many ways (see~\cite{CDM05,C04} for general
253: algorithms, and~\cite{Caffarel,Umrigar,Sorella} for algorithms used in
254: the context of DMC computations). In this paper, we restrict ourselves
255: to resampling methods with a fixed number of particles, and such that
256: the weights of the particles after resampling are equal to $1$. Then, the basic
257: consistency requirement of the selection step is that, conditionally on
258: the former positions $\left( \uu{X}^j_{n \Dt + k \dt} \right)_{1 \le
259:   j \le N, 1 \leq k \leq \kappa}$, the $i$-th particle~$\uu{X}^i_{n
260:     \Dt + \kappa \dt}$ is replicated $N \rho^i_n$ times in mean, where $\rho^i_n=\exp\left( - \dt \sum_{k=1}^\kappa E_L(\uu{X}^i_{n
261:     \Dt + k \dt}) \right) \Big/ \sum_{j=1}^N \exp\left( - \dt \sum_{k=1}^\kappa E_L(\uu{X}^j_{n
262:     \Dt + k \dt}) \right)$ denotes the (normalized) weight of the $i$-th particle. There are of course many ways to satisfy this requirement.
263: We presented above the so-called
264: {\it multinomial resampling} method. We will also discuss below {\it residual
265: resampling} (also called stochastic remainder resampling), {\it
266: stratified resampling} and {\it
267: systematic resampling}, which may also be used for DMC computations. Let us briefly describe these three resampling
268: methods. Residual resampling consists in reproducing
269: $\lfloor N \rho^i_n  \rfloor$ times the $i$-th particle, and then completing
270: the set of particles by using multinomial resampling to draw the
271: $N^R=N-\sum_{l=1}^N \lfloor N \rho^l_n  \rfloor$ remaining particles, the
272: $i$-th particle being assigned the weight $\rho^{R,i}_n=\{ N \rho^i_n \} /
273: N^R$. Here and in the following,
274: $\lfloor x\rfloor$ and $\{x\}$ respectively denote the
275:   integer and the fractional part of $x\in\R$. In the stratified
276:   resampling method, the interval $(0,1)$ is divided into $N$ intervals $((i-1)/N,i/N)$
277:   ($1 \leq i \leq N$), $N$ random variables are then drawn independently and
278:   uniformly in each interval, and the new particle positions are then
279:   obtained by the inversion method: $\uu{X}^i_{(n+1)\Dt} = \sum_{j=1}^N
280:   1_{\{\sum_{l=1}^{j-1} \rho^l_n < (i-U^i_n)/N \leq \sum_{l=1}^{j}
281:     \rho^l_n \}} \uu{X}^j_{n\dt + \kappa \dt} $, where $U^i_n$ are
282:   i.i.d. random variables uniformly distributed over $[0,1]$. Here and
283:   in the following, we use the convention $\sum_{l=1}^0 \cdot =0$.
284: Systematic resampling
285:   consists in replicating the $i$-th particle $\Big\lfloor N \sum_{l=1}^i
286:   \rho^l_n + U_n \Big\rfloor - \Big\lfloor N \sum_{l=1}^{i-1} \rho^l_n + U_n \Big\rfloor$
287:   times\footnote{The consistency of this resampling method follows
288:     from the following easy computation $$\E\left(\lfloor
289:       x+U\rfloor\right)=\lfloor x\rfloor\P(U<1-\{x\})+(\lfloor
290:   x\rfloor+1)\P(U\geq 1-\{x\})=\lfloor x\rfloor(1-\{x\})+(\lfloor
291:   x\rfloor+1)\{x\}=x.$$},  where $(U_n)_{n \geq 1}$ are independent random variables uniformly distributed in
292:   $[0,1]$. Notice that systematic resampling can be seen as the
293:   stratified resampling method, with $U^1_n= \ldots=U^N_n=U_n$.
294: Contrary to the three other resampling methods, after a systematic
295: resampling step,
296:  the new particle positions are not independent,
297:   conditionally on the former positions. This makes
298:   systematic resampling much
299:   more difficult to study mathematically. To our knowledge, its
300:   convergence even in a discrete time setting is still an open question. We will therefore restrict
301:   ourselves to a numerical study of its performance.
302: 
303: Notice that practitioners often use branching algorithms with an evolving number
304: of walkers during the computation (see~\cite{RCAL82,Umrigar}): the particles with low local
305: energy are replicated and the particles with high local energy are
306: killed, without keeping the total number of particles constant. This may
307: lead to a smaller Monte Carlo error (fourth contribution to the
308: error in the classification just below).
309: 
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311: % Autre ecriture de systematic resampling
312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
313: % Systematic resampling
314: %  consists in replicating the $i$-th particle
315: % $F^{-1}[ ((i-1)+U)/N ]$ times,
316: %   where $U$ denotes a random variable uniformly distributed in $(0,1)$,
317: %   and $F^{-1}$ is the inverse of the cumulative distribution function associated with
318: %   the normalized weights $\rho^i_n$: $\forall i \in \{1,\ldots,N\}$,
319: %   $F^{-1}(u)=i$ if and only if $u \in \left(\sum_{l=1}^{i-1}
320: %   \rho^l_n,\sum_{l=1}^{i} \rho^l_n \right]$.
321: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
322: 
323: 
324: %\subsection*{Presentation of our result}
325: 
326: We can distinguish between four sources of errors in the approximation
327: of $E_0$ by $\ds{ \frac{1}{N} \sum_{j=1}^{N} \,
328:   E_L \left(\uu{X}^j_{\nu \Dt}\right)}$:
329: \begin{enumerate}
330: \item the error due to the fixed node approximation $|E_0-E_{{\rm
331:     DMC},0}|$,
332: \item the error due to finite time approximation of the limit: $\lim_{t
333:     \to \infty} E_{\rm DMC}(t) \simeq E_{\rm DMC}(T)$,
334: \item the error due to the time discretization of the stochastic differential
335: equation~(\ref{eq:SDE}) and of the integral in the exponential factor in
336: $E_{\rm DMC}(t)$ (see~(\ref{eq:EDMC})),
337: \item the error introduced by the interacting particle system, due to
338:   the approximation of the expectation value in~(\ref{eq:EDMC}) by an empirical mean.
339: \end{enumerate}
340: 
341: The error (1) due to the fixed node approximation has been analyzed
342: theoretically in~\cite{CJL}. 
343: 
344: Concerning the error (2) due to finite time approximation of the limit, the rate of convergence in time is
345: typically exponential. Indeed if $H$ admits a spectral gap (namely if the
346: distance between $E_0$ and the remaining of the spectrum of $H$ is strictly
347: positive), and if $\psi_I$ is such that $\langle \psi_I, H \psi_I \rangle <
348: \inf \sigma_{\rm ess}(H)$, then one can show that the operator $H$ with
349: domain $D_\mathcal{H}(H) \cap \{\psi, \, \psi=0\mbox{ on }\psi_I^{-1}(0) \}$ (whose
350: lowest eigenvalue is $E_{{\rm DMC},0}$, see~(\ref{eq:EDMC0})) also admits a
351: spectral gap $\gamma>0$. Then, by standard spectral decomposition methods,
352: we have:
353: $$0 \leq \left| E_{\rm DMC}(t) - E_{{\rm DMC},0} \right| \leq C \exp(-\gamma t).$$
354: 
355: Our aim in this paper is to provide some theoretical and numerical
356: results related to the errors~(3) and~(4), in the framework
357: of a simple one-dimensional case. We therefore consider in the
358: following that the final time of simulation $T$ is fixed and we analyze the error introduced by the
359: numerical scheme on the estimate of $E_{\rm DMC}(T)$. Our convergence
360: result is of the form:
361: \begin{equation}\label{eq:est_err_general}
362: \E \left|E_{\rm DMC}(T)- \frac{1}{N} \sum_{j=1}^{N} \,
363:   E_L \left(\uu{X}^j_{\nu \kappa \dt}\right)\right| \leq C(T) \, \dt +
364: \frac{C(T,\nu)}{\sqrt{N}},
365: \end{equation}
366: where $C(T)$ (resp. $C(T,\nu)$) denotes a constant which only depends on
367: $T$ (resp. on $T$ and $\nu$) (see Theorem~\ref{th:est_err} and
368: Corollary~\ref{cor:est_err} below).
369: 
370: %\subsection*{Presentation of our toy model}
371: 
372: Let us now present the toy model we consider in the following. We consider the Hamiltonian 
373: \begin{equation}\label{eq:H}
374: H=-\frac{1}{2}\frac{d^2}{dx^2}+V, \mbox{ with }
375: V=\frac{\omega^2}{2}x^2+\theta x^4,
376: \end{equation}
377: where $\omega,\theta>0$ are two constants. The ground
378: state energy $E_0$ is defined by~(\ref{eq:E0}), with
379: \begin{equation}\label{eq:cal_H}
380: \mathcal{H}=\left\{ \psi \in L^2(\R), \, \psi(x)=-\psi(-x)
381: \right\}.
382: \end{equation}
383: We restrict the functional spaces to odd functions in order to mimic the antisymmetry
384: constraint on $\psi$ for fermionic systems. The importance sampling
385: $\psi_I$ is chosen to be the ground state of
386: $H_0=-\frac{1}{2}\frac{d^2}{dx^2} + \frac{\omega^2}{2}x^2$ on
387: $\mathcal{H}$:
388: \begin{equation}\label{eq:psi_I}
389: \psi_I(x)=\sqrt{2\omega}\left(\frac{\omega}{\pi}\right)^{1/4}xe^{-\frac{\omega
390:     }{2}x^2}.
391: \end{equation}
392: It is associated with the energy $\frac{3}{2}\omega$: $H_0 \psi_I=
393: \frac{3}{2}\omega \psi_I$.
394: The drift function $b$ and the local energy $E_L$ are then defined by:
395: \begin{equation}
396: b(x)=\frac{\psi_I'}{\psi_I}(x)=\frac{1}{x}-\omega x, \mbox{ and } E_L(x)=V(x)-\frac{1}{2}\;\frac{\psi_I''}{\psi_I}(x)=\frac{3}{2}\omega+\theta x^4.
397: \end{equation}
398: Thus, using equation~(\ref{eq:EDMC}), the DMC energy is:
399: \begin{equation}E_{\rm DMC}(t)=\frac{3}{2}\omega+\theta\frac{\E\left(X^4_t\exp\left(-\theta\int_0^t
400:       X_s^4ds\right)\right)}{\E\left(\exp\left(-\theta\int_0^t
401:       X_s^4ds\right)\right)},
402: \label{defedmc}\end{equation}
403: where
404: \begin{equation}
405: X_t=X_0 + \int_0^t \left(\frac{1}{X_s}-\omega X_s\right)ds+ W_t,\label{eds}
406: \end{equation}
407: with $(W_t)_{t \geq 0}$ a Brownian motion independent from the initial variable $X_0$
408: which is distributed according to the invariant measure $2\psi_I^2(x)1_{\{x>0\}}dx$.
409: We recall that due to the explosive part in the drift function~$b$, the stochastic
410: process cannot cross $0$, which is the zero point of $\psi_I$
411: (see~\cite{CJL}): $\P(\exists t>0,\, X_t=0)=0$. This explains why the restriction of $\psi_I^2$ to
412: $\R_+^*$ is indeed an invariant measure for~(\ref{eds}). For $\theta>0$,
413: the longtime
414: limit $E_{{\rm DMC},0}$ of $E_{\rm DMC}(t)$ is not analytically known,
415: but can be very accurately computed by a spectral method (see
416: Section~\ref{sec:num_res_spec}). Let us finally make precise that for the
417: numerical analysis, we use a special feature of our simple model, namely
418: the fact that for $s \leq t$, it is possible to simulate the conditional
419: law of $X_t$ given $X_s$ (see Appendix). The time discretization error is thus only related to
420: the discretization of the integral in the exponential factor in the DMC
421: energy~(\ref{eq:EDMC}). We however indicate some possible ways to
422: prove~(\ref{eq:E0}) with a convenient time discretization of the SDE (see Equation~(\ref{schemaurel}),
423: Remark~\ref{rem:weak} and Proposition~\ref{prop:Markov}).
424: 
425:  Though our model
426: is one-dimensional (and therefore still far from the real problem~(\ref{eq:E0})), it contains one of the main difficulties related to the
427: approximation of the ground state energy for fermionic systems, namely
428: the explosive behavior of the drift in the stochastic differential
429: equation. However, two characteristics of practical problems are missing
430: in the toy model considered here. First, since we consider a
431: one-particle model, we do not treat difficulties related to
432: singularities of the drift and of the local energy at points where two
433: particles (either two electrons or one electron and one nucleus) coincide. Second, the local energy $E_L$ generally explodes at the nodes
434: of the trial wave function, and this is not the case on the simple example
435: we study
436: since the trial wave function is closely related to the exact ground
437: state. For an adaptation of the DMC algorithm to take care of these
438: singularities, we refer to~\cite{Umrigar}.
439: Despite the simplicity of the model studied in this paper, we think that the
440: convergence results we obtain and the mathematical tools we use
441: are prototypical for generalization to more complicated systems.
442: 
443: Compared to previous mathematical analysis of convergence for
444: interacting particle systems with stochastic
445: reconfiguration~\cite{DM03,Delmoral,Delmoralmiclo,DD04,Rousset}, our
446: study concentrates on the limit $\dt \to 0$ and $N \to \infty$ for a
447: fixed time~$T$, and on
448: the {\em influence of the time discretization error} in the
449: estimate~(\ref{eq:est_err_general}), where the test function~$E_L$ is
450: {\em unbounded}. It is actually important in our analysis that this
451: unbounded function $E_L$ also appears in the weights of the particles,
452: since it allows for specific estimates (see Lemmas~\ref{majomom} and~\ref{decroiss} below).
453: 
454: %\subsection*{Outline}
455: 
456: The paper is organized as follows. In Section~\ref{sec:num_anal}, we
457: prove the convergence result, by adapting
458: the methods of~\cite{Delmoral,Delmoralmiclo} to analyze the dependence
459: of the error on $\dt$. We then check the optimality of this theoretical result
460: by numerical experiments in Section~\ref{sec:num_res}, where we also
461: analyze numerically the dependence of the results on
462: various numerical parameters, including the number $(\nu-1)$ of
463: reconfiguration steps. From these numerical experiments, we propose a
464: simple heuristic method to choose the optimal number of reconfiguration steps.
465: 
466: {\bf Notation:} For any set of
467: random variables $(Y_i)_{i \in I}$, we denote by $\sigma((Y_i)_{i \in
468:   I})$ the sigma-field generated by these random variables. The parameters $\omega$ and $\theta$ are fixed positive constants.
469: By convention, any sum from one to zero is equal to zero: $\sum_{k=1}^0
470: \cdot =0$. Likewise, the subset $\{1,2, \ldots ,0\}$ of $\N$ is by convention the
471: empty set. For any real $x$, $\lfloor x\rfloor$ and $\{x\}$ respectively denote the
472:   integer and the fractional part of $x$.
473: \section{Numerical Analysis in a Simple Case}\label{sec:num_anal}
474: 
475: We perform the numerical analysis in two steps: time discretization and
476: then particle approximation.
477: 
478: \subsection{Time discretization}\label{sec:time_disc}
479: 
480: We recall that $T>0$ denotes the final simulation time, and that
481: $\dt=\frac{T}{K}$ is the smallest time-step. Since $Y_t=X_t^2$ is a
482: square root process solving $dY_t=(3-2\omega Y_t)dt+2\sqrt{Y_t}dW_t$,
483: it is possible to simulate the increments $Y_{(k+1)\dt}-Y_{k\dt}$ and
484: therefore 
485: $X_{(k+1)\dt}-X_{k\dt}$ (see Appendix or \cite{Glasserman} p.120). We
486: can thus simulate exactly in law the vector
487: $(X_0,X_{\dt},\hdots,X_{K \dt})$. That is why we are first going to study the
488: error related to the time discretization of the integral which appears
489: in the exponential factors in \eqref{defedmc}.\par
490: Let us define the corresponding approximation of $E_{\rm
491:    DMC}(T)$:
492: \begin{equation}\label{eq:EDMCdt}
493: E_{\rm
494:    DMC}^{\dt}(T)=\frac{\E\left(E_L(X_{T})\exp\left(-\dt\sum_{k=1}^KE_L(X_{k\dt})\right)\right)}{\E\left(\exp\left(-\dt\sum_{k=1}^KE_L(X_{k\dt})\right)\right)}=\frac{3}{2}\omega+\theta\frac{\E\left(X^4_T\exp\left(-\t\dt\sum_{k=1}^KX^4_{k\dt})\right)\right)}{\E\left(\exp\left(-\t\dt\sum_{k=1}^KX^4_{k\dt}\right)\right)}.
495: \end{equation}
496: \begin{aprop}
497: $$
498:   \forall K\in\N^*,\left|E_{\rm
499:    DMC}(T)-E_{\rm
500:    DMC}^{\dt}(T)\right|\leq C_T\dt.
501: $$
502: \label{erreurdt}\end{aprop}
503: \begin{adem}
504: Using H\"older inequality, we have:
505: \begin{align*}
506:    \left|E_{\rm DMC}(T)-E_{\rm DMC}^{\dt}(T)\right|\leq&\frac{\t}{\E\left(\exp\left(-\t\dt\sum_{k=1}^KX^4_{k\dt}\right)\right)}\left(\sqrt{\E(X_T^8)}+\frac{\E\left(X^4_{T}\exp\left(-\theta\int_0^T
507:       X_s^4ds\right)\right)}{\E\left(\exp\left(-\theta\int_0^T
508:       X_s^4ds\right)\right)}\right)\\
509: &\left(\E\left(\left(\exp\left(-\t\int_0^T
510:       X_s^4ds\right)-\exp\left(-\t\dt\sum_{k=1}^KX^4_{k\dt}\right)\right)^2\right)\right)^{1/2}.
511: \end{align*}
512: The conclusion is now a consequence of Lemma~\ref{vitfort} and the fact that the function $x\in\R_+\rightarrow e^{-\t x}$ is Lipschitz
513: continuous with constant $\t$.
514: \end{adem}
515:       
516: \begin{alem}
517: For any $K\in\N^*$,
518: $$\E\left(\left(\int_0^T
519:       X_s^4ds-\dt\sum_{k=1}^KX^4_{k\dt}\right)^2\right)\leq
520:   C\dt^2(T^2+T),$$
521: where $\dt=\frac{T}{K}$.
522: \label{vitfort}\end{alem}
523: 
524: \begin{adem}[of Lemma \ref{vitfort}]
525:    By Itô's formula, $dX^4_t=(10X^2_t-4\omega X^4_t)dt+4X_t^3dW_t$. With
526:    the integration by parts formula, one deduces that for any $k\in\{1,\hdots,K\}$,
527: $$\int_{(k-1)\dt}^{k\dt}(X^4_{k\dt}-X^4_s)ds=\int_{(k-1)\dt}^{k\dt}(s-(k-1)\dt)\left((10X^2_s-4\omega
528:   X^4_s)ds+4X_s^3dW_s\right).$$
529: Therefore denoting $\tau_s=\lfloor\frac{s}{\dt}\rfloor\dt$ the
530:   discretization time just before $s$, one obtains
531: $$\dt\sum_{k=1}^KX^4_{k\dt}-\int_0^T
532:       X_s^4ds=\int_0^T(s-\tau_s)(10X^2_s-4\omega
533:   X^4_s)ds+\int_0^T(s-\tau_s)4X_s^3dW_s.$$
534: Hence 
535: \begin{align*}
536:    \E\left(\left(\dt\sum_{k=1}^KX^4_{k\dt}-\int_0^T
537:       X_s^4ds\right)^2\right)\leq 2\int_0^T(s-\tau_s)^2\E\left(T(10X^2_s-4\omega
538:   X^4_s)^2+16X_s^6)\right)ds.
539: \end{align*}
540: Since $X_0$ is distributed according to the invariant measure
541: $2\psi_I^2(x)1_{\{x>0\}}dx$, so is $X_s$. As a consequence, for any $p\in\N$, $\E(X_s^p)$ does not
542: depend on $s$ and is finite and the conclusion follows readily.
543: \end{adem} 
544: 
545: In realistic situations, exact simulation of the increments
546: $X_{(k+1)\dt}-X_{k\dt}$ is not possible and one has to resort to
547: discretization schemes. The singularity of the drift coefficient
548: prevents the process $X_t$ from crossing the nodal surfaces of the
549: importance sampling function $\psi_I$. The standard explicit Euler
550: scheme does not preserve this property at the discretized level. For
551: that purpose, we suggest to use the following explicit scheme proposed by
552: \cite{Alfonsi}
553: \begin{equation}
554:    \begin{cases}
555:       \bar{X}_0=X_0,\\
556: \ds{\forall k\in\N,\;\bar{X}_{(k+1)\delta t}=\left(\left({\bar{X}_{k\delta t}}(1-\omega
557:     \delta t)+\frac{\Delta W_{k+1}}{1-\omega
558:     \delta t}\right)^2+2\delta t\right)^{1/2}}\;\;\mbox{with}\;\;\Delta
559:     W_{k+1}=W_{(k+1)\delta t}-W_{k\delta t}.
560:    \end{cases}\label{schemaurel}
561: \end{equation}
562: Because of the singularity at the origin of the drift coefficient in
563: \eqref{eds}, we have not been able so far to prove the following weak error
564: bound (see Remark~\ref{rem:weak} below):
565: \begin{equation}
566:    \left|E\left(f(X^4_{T})\exp\left(-\theta\int_0^{T}
567:       X_s^4ds\right)\right)-\E\left(f(\bar{X}^4_{T})\exp\left(-\theta \delta t\sum_{k=1}^{K}
568:       \bar{X}_{k\delta t}^4\right)\right)\right|\leq C_T\dt\mbox{ for
569:       }f(x)\equiv 1\mbox{ and }x^4.\label{weakaurel}
570: \end{equation}
571: Such a bound is expected according to~\cite{Talay-Tubaro} and would imply that
572: \begin{equation}
573:    \left|E_{\rm DMC}(T)-\frac{\E\left(E_L(\bar{X}_{T})\exp\left(-\dt\sum_{k=1}^KE_L(\bar{X}_{k\dt})\right)\right)}{\E\left(\exp\left(-\dt\sum_{k=1}^KE_L(\bar{X}_{k\dt})\right)\right)}\right|\leq C_T\dt.\label{discretaurel}
574: \end{equation}
575: \begin{arem}\label{rem:weak}
576: We would like to sketch a possible way to
577: prove~\eqref{weakaurel}. Because the square root in \eqref{schemaurel}
578: makes expansions with respect to $\delta t$ and $\Delta W_{k+1}$
579: complicated, it is easier to work with $Y_t=X_t^2$ and $\bar{Y}_{k\delta t}=\bar{X}^2_{k\delta
580:     t}$ which satisfy
581: $$dY_t=(3-2\omega Y_t)dt+2\sqrt{Y_t}\;dW_t\;\;\mbox{and}\;\;\bar{Y}_{(k+1)\delta t}=\left(\sqrt{\bar{Y}_{k\delta t}}(1-\omega
582:     \delta t)+\frac{\Delta W_{k+1}}{1-\omega
583:     \delta t}\right)^2+2\delta t.$$
584: The standard approach to analyze the time discretization error of the
585: numerator and denominator of the left hand side of~\eqref{discretaurel} is then to introduce some
586: functions $v$ and $w$ solutions to the partial differential equation:
587: \begin{equation}\label{eq:EDP}
588: \partial_t v=(3-2y)\partial_y
589: v+2y\partial_{yy}v-\t y^2v,\;(t,y)\in\R_+\times (0,+\infty)
590: \end{equation}
591: with initial conditions $v(0,y)=y^2$ and $w(0,y)=1$. Now, we write (for
592: the numerator, for example):
593: \begin{align*}
594: &\E\left(X^4_{T}\exp\left(-\theta\int_0^{T}
595:       X_s^4ds\right)\right)-\E\left(\bar{X}^4_{T}\exp\left(-\theta \delta t\sum_{k=1}^{K}
596:       \bar{X}_{k\delta t}^4\right)\right)\\
597: &=\sum_{k=0}^{K-1}\E\left(\left(v(T-k\delta t,\bar{Y}_{k\delta
598:       t})-e^{-\theta\delta t\bar{Y}_{(k+1)\delta t}^2}v(T-(k+1)\delta t,\bar{Y}_{(k+1)\delta
599:       t})\right)\exp\left(-\theta \delta t\sum_{j=0}^{k-1}
600:       \bar{Y}_{j\delta t}^2\right)\right).
601: \end{align*}
602: An error bound of the form $C_T \dt$ can now be proved by some Taylor
603: expansions as in \cite{Talay-Tubaro,Alfonsi}, provided
604: the existence of a sufficiently smooth solution $v$
605: to~(\ref{eq:EDP}). We have not been able to prove existence of such a
606: solution so far.
607: \end{arem}
608: 
609: % {\small A cause de la racine carrée, il semble difficile d'analyser directement l'erreur de
610: %     ce schéma et il est sans doute préférable de passer par les
611: %     processus $Y_t=X_t^2$ et $\bar{Y}_{k\delta t}=\bar{X}^2_{k\delta
612: %     t}$ qui vérifient
613: % $$dY_t=(3-2\omega Y_t)dt+2\sqrt{Y_t}\;dW_t\;\;\mbox{et}\;\;\bar{Y}_{(k+1)\delta t}=\left((1-\omega
614: %     \delta t)\sqrt{\bar{Y}_{k\delta t}}+\frac{\Delta W_{k+1}}{1-\omega
615: %     \delta t}\right)^2+2\delta t.$$
616: % L'approche standard pour analyser l'erreur de discrétisation du
617: %     numérateur et du dénominateur de \eqref{defedmc} consiste à
618: %     introduire les solutions $v$ et $w$ de l'EDP 
619: % $$\partial_t v=(3-2y)\partial_y
620: % v+2y\partial_{yy}v-\t y^2v,\;(t,y)\in\R_+\times (0,+\infty)$$
621: % pour les conditions initiales $v(0,y)=y^2$ et $w(0,y)=1$. Notons que
622: % $\partial_t w=-\t v$. On écrit par exemple
623: % \begin{align*}&\E\left(X^4_{T}\exp\left(-\theta\int_0^{T}
624: %       X_s^4ds\right)\right)-\E\left(\bar{X}^4_{T}\exp\left(-\theta \delta t\sum_{k=1}^{K}
625: %       \bar{X}_{k\delta t}^4\right)\right)\\
626: % &=\sum_{k=0}^{K-1}\E\left[\left(v(T-k\delta t,\bar{Y}_{k\delta
627: %       t})-e^{-\theta\delta t\bar{Y}_{(k+1)\delta t}^2}v(T-(k+1)\delta t,\bar{Y}_{(k+1)\delta
628: %       t})\right)\exp\left(-\theta \delta t\sum_{j=0}^{k-1}
629: %       \bar{Y}_{j\delta t}^2\right)\right].\end{align*}
630: % Et on analyse les termes de la forme $v(T-k\delta t,\bar{Y}_{k\delta
631: %       t})-e^{-\theta\delta t\bar{Y}_{(k+1)\delta t}^2}v(T-(k+1)\delta t,\bar{Y}_{(k+1)\delta
632: %       t})$ en effectuant des développements de Taylor. Pour cela, il
633: %       faut des contrôles sur les dérivées de $v$ que l'on peut essayer
634: %       d'obtenir sur l'interprétation probabiliste
635: % $$v(t,y)=\E\left((Y^y_t)^2\exp\left(-\t\int_0^t(Y^y_s)^2ds\right)\right)$$
636: % où $Y^y_t$ désigne la solution de l'EDS issue de $Y^y_0=y$. Les dérivées
637: % temporelles ne posent aucun problème. Par exemple, 
638: % $$\partial_t v(t,y)=\E\left((10Y^y_t-4\omega
639: %   (Y^y_t)^2-\t(Y^y_t)^4)\exp\left(-\t\int_0^t(Y^y_s)^2ds\right)\right)$$
640: % et les dérivées successives font apparaître l'espérance de polynômes en
641: %   $Y^y_t$ multipliés par le facteur exponentiel. Pour étudier les
642: %   dérivées en $y$, on introduit le flot de l'EDS en posant
643: %   $Y^{(k),y}_t=\partial^k_y Y^y_t$. Formellement,
644: % $$\partial_y
645: % v(t,y)=\E\left[\left((2Y^y_tY^{(1),y}_t-2\t Y^y_t\int_0^tY^y_sY^{(1),y}_sds\right)\exp\left(-\t\int_0^t(Y^y_s)^2ds\right)\right]$$
646: % où
647: % $$dY^{(1),y}_t=-3Y^{(1),y}_tdt+\frac{Y^{(1),y}_t}{\sqrt{Y^y_t}}\;dW_t\mbox{
648: %   i.e.
649: %   }Y^{(1),y}_t=\exp\left(\int_0^t\frac{1}{\sqrt{Y^y_s}}\;dW_s-3t-\frac{1}{2}\int_0^t\frac{1}{Y^y_s}ds\right).$$
650: % Malheureusement les moments de cette exponentielle ne sont pas
651: %   agréables... Pour l'EDS de départ, on a
652: % $$X^{(1),x}_t=\exp\left(-\int_0^t\frac{1}{(X^x_s)^2}+\omega
653: %   ds\right)\leq 1$$
654: % ce qui permet de contrôler $\partial_x u(t,x)$ où $u(t,x)=v(t,x^2)$ mais
655: %   $\partial_yv(t,y)=\frac{1}{2\sqrt{y}}\partial_xu(t,\sqrt{y})$ et on
656: %   retrouve des problèmes liés à la singularité à l'origine.}
657: 
658: % \begin{arem}
659: % We would like to mention that, in practice, some people use a
660: % Metropolis step after each integration in time on a timestep $\dt$, to
661: % ensure both that the process do not cross the nodal surface of $\psi_I$ and that
662: % the law of $\uu{X}^j_{n\Dt + k \dt}$ is the stationary law associated
663: % with~(\ref{eq:SDE}).
664: % \end{arem}
665: 
666: \subsection{Particle approximation}
667: % We consider a discrete time system with $N$-particles (also called walkers in the
668: %    physical literature) $(\xi_n=(\xi^1_n,\hdots,\xi^N_n))_{1\leq n\leq \nu}$ with $\nu\in\N^*$ and set $\Dt=T/\nu$. The
669: % stochastic differential equation \eqref{eds} is possibly discretized
670: % with a smaller step $\dt=\Dt/\kappa=T/(\nu\kappa)$ with $\kappa\in\N^*$. The total
671: % number of steps for the discretization of \eqref{eds} is then
672: % $K=\nu\kappa$.\\
673: 
674: % Let $(W^1,\hdots,W^N)$ be a $N$-dimensional Brownian motion independent
675: % from initial random variables $(\xi^{i,1}_0)_{1\leq i\leq N}$ i.i.d. according
676: % to $2\psi_I^2(x)1_{\{x>0\}}dx$. We initialize the particles positions $\left(\xi_1^i=(\xi_{1,1}^i,\hdots,\xi_{1,\kappa}^i)\right)_{1\leq
677: %   i\leq N}$ by setting $\xi^i_{1}=(X^{i,1}_{\dt},\hdots, X^{i,1}_{\kappa\dt})$ where
678: %     $(X^{i,1}_t)_{0\leq t\leq \kappa\dt}$ solves
679: % $$X^{i,1}_t=\xi^{i,1}_0+\int_0^tb(X^{i,1}_s)ds+W^i_{t}.$$ And the particles positions at
680: %   discrete time $n\geq 2$
681: %   are obtained inductively by alternation of a selection/reconfiguration
682: %   step and a
683: %   mutation step. Given $\xi_n=(\xi_n^1,\hdots,\xi_n^N)$ with
684: %   $\xi^i_n=(\xi_{n,1}^i,\hdots,\xi_{n,\kappa}^i)$ for $1\leq
685: % i\leq N$, the selection step consists in choosing
686: % $\xi^{1,n+1}_0,\hdots,\xi^{N,n+1}_0$ conditionally independent with $\xi^{i,n+1}_0$
687: % distributed according to the measure
688: % \begin{equation}
689: %    \epsilon_n g(\xi_n^i)\delta_{\xi_{n,\kappa}^i}+(1-\epsilon_ng(\xi_n^i))\frac{\sum_{j=1}^Ng(\xi_n^j)\delta_{\xi_{n,\kappa}^j}}{\sum_{j=1}^Ng(\xi_n^j)},\label{messel}
690: % \end{equation}
691: % where $g(y_1,\hdots,y_\kappa)=\exp(-\t\dt\sum_{k=1}^\kappa y_k^4)$. We consider the following possible choices for $\epsilon_n$ :
692: % \begin{equation}
693: %    \epsilon_n=0,\;\epsilon_n=1\;\mbox{and}\;\epsilon_n=\frac{1}{\max_{1\leq
694: %     i\leq N}g(\xi_n^i)}.\label{choix_eps}
695: % \end{equation}
696: % The mutation step then consists in setting
697: %     $\xi^i_{n+1}=(X^{i,n+1}_{\dt},\hdots, X^{i,n+1}_{\kappa\dt})$ where
698: %     $(X^{i,n+1}_t)_{0\leq t\leq \kappa\dt}$ solves
699: % \begin{equation}
700: %    X^{i,n+1}_t=\xi^{i,n+1}_0+\int_0^tb(X^{i,n+1}_s)ds+W^i_{t+n\Dt}-W^i_{n\Dt}.\label{sden+1}
701: % \end{equation}
702: 
703: We now introduce some notation to study the particle approximation. We
704: recall that $\nu$ denotes the number of large timesteps (the
705: number of reconfiguration steps is $\nu-1$), and
706: $\Dt=\kappa \dt$ the time period between two reconfiguration steps. Let
707: us suppose that we know the initial positions $(X^i_{n,0})_{1 \leq i \leq N}$ of
708: the $N$ walkers at time $(n-1)\Dt$, for a time index $n \in \{1,\hdots,\nu\}$. The successive positions of the walkers
709: over the time interval $((n-1) \Dt,n\Dt)$ are then given by
710: $(X^i_{n,\dt},\hdots,X^i_{n,\kappa\dt})$, where $(X^i_{n,t})_{0 \leq t \leq
711: \Dt}$ satisfies:
712: \begin{equation}
713: X^i_{n,t}=X^i_{n,0}+\int_0^t b(X^i_{n,s}) \, ds+ \left(W^i_{t+(n-1) \Dt} -
714:   W^i_{(n-1) \Dt}\right).\label{sden+1}
715: \end{equation}
716: Here $(W^1,\hdots,W^N)$ denotes a $N$-dimensional Brownian motion independent
717: from the initial positions of the walkers $(X^i_{1,0})_{1\leq i\leq N}$
718: which are i.i.d. according
719: to $2 \psi_I^2(x) 1_{\{x>0\}} dx$. We recall that in our framework, it
720: is possible to simulate exactly in law all these random variables (see
721: Appendix). We store the successive positions
722: $(X^i_{n,\dt},\hdots,X^i_{n,\kappa\dt})$ of the $i$-th walker over the
723: time interval $((n-1) \Dt,n \Dt)$ in a so-called particle $\xi^i_{n}
724: \in (\R_+^*)^\kappa$
725: (see Figure~\ref{fig:particle}): $\forall i \in \{1,\hdots,N\}, \forall  n
726: \in \{1,\hdots, \nu\}$, 
727: \begin{equation}
728: \xi^i_{n}=(X^i_{n,\dt},\hdots,X^i_{n,\kappa\dt}).
729: \end{equation}
730: In the following, we will denote by $\xi_{n}=(\xi^1_{n},\hdots,\xi^N_{n})$ the configuration of the
731: ensemble of particles at time index $n$. We have here described {\bf the mutation step}.
732: 
733: \begin{figure}[htbp]
734: \input{particle.pstex_t}
735: \caption{The $i$-th particle $\xi^i_n$ at time index $n$ is composed of the
736:   successive positions $(X^i_{n,\dt},\hdots,X^i_{n,\kappa \dt})$ of the
737:   $i$-th walker on time interval $((n-1) \Dt,n \Dt)$.}\label{fig:particle}
738: \end{figure}
739: 
740: 
741: For a given configuration of the particles $\xi_{n}$ at a time index $n \in \{1,\hdots, \nu\}$, {\bf the selection step}
742: now consists in choosing the initial positions
743: $(X^i_{n+1,0})_{1 \leq i \leq N}$ of the $N$ walkers at time $n \Dt$
744: using one of the following resampling algorithm:
745: \begin{itemize}
746:    \item[{\bf(S1)}]
747: The $(X^i_{n+1,0})_{1 \leq i \leq N}$ are conditionally
748: independent w.r.t.~$\xi_{n}$ and for $1 \leq i \leq N$, $X^i_{n+1,0}$ is distributed according to the measure
749:  \begin{equation}
750:    \epsilon_n
751:    g(\xi_n^i)\delta_{\xi_{n,\kappa}^i}+(1-\epsilon_n
752:    g(\xi_n^i)) \sum_{j=1}^N \rho_n^j \delta_{\xi_{n,\kappa}^j},\label{messel}
753: \end{equation}
754: where $g$ is defined by, for $y=(y_1,\hdots,y_\kappa)\in(\R_+^*)^\kappa$,
755: \begin{equation}\label{eq:g}
756: g(y)=\exp\left(-\t\dt\sum_{k=1}^\kappa y_k^4\right),
757: \end{equation}
758: $\rho_n^j$ denotes the weight of the $j$-th particle
759: \begin{equation}\label{eq:rho}
760: \rho_n^j=\frac{g(\xi_n^j)}{\sum_{j=1}^Ng(\xi_n^j)}
761: \end{equation}
762: and $\epsilon_n$ is a non negative function of $\xi_{n}$ such that
763: $\epsilon_n \leq 1 \big/ \max_{1\leq
764:     i\leq N}g(\xi_n^i)$. In particular the following choices are possible for $\epsilon_n$:
765: \begin{equation}
766:    \epsilon_n=0,\;\epsilon_n=1\;\mbox{ or }\;\epsilon_n=\frac{1}{\max_{1\leq
767:     i\leq N}g(\xi_n^i)}.\label{choix_eps}
768: \end{equation}
769: The so-called {\bf multinomial resampling} method which corresponds to the
770: choice $\epsilon_n=0$ gives rise to a maximum decorrelation with the former
771: position of the particles, while with growing
772: $\epsilon_n$, more and more correlation is introduced.
773: \item[{\bf (S2)}] The $(X^i_{n+1,0})_{1 \leq i \leq N}$ are such that
774: \begin{equation}
775:    \begin{cases}
776:       \forall j\in\{1,\hdots,N\},\;\forall
777: i\in\left\{ \left(1+ \sum_{l=1}^{j-1} a^l_n\right),\hdots,\left(\sum_{l=1}^{j} a^l_n\right)\right\}, \\
778: \qquad X^i_{n+1,0}=\xi^j_{n,\kappa},\\
779: \mbox{and the variables $(X^i_{n+1,0})_{1+ \sum_{l=1}^{N} a^l_n \leq i
780:     \leq N}$ are conditionally
781: independent w.r.t.~$\xi_{n}$,}\\
782: \qquad \mbox{with } X^i_{n+1,0}\mbox{ distributed according to
783: }\sum_{j=1}^N\left\{ N \rho^j_n\right\}\delta_{\xi_{n,\kappa}^j}
784: \Big/ \left(N- \sum_{l=1}^{N} a^l_n\right),
785:    \end{cases}
786: \label{messel2}\end{equation}
787: where
788:   \begin{equation}
789:    a^j_n=\big\lfloor N \rho_n^j \big\rfloor,\;j\in\{1,\hdots,N\}.\label{defpoidsa}
790:   \end{equation} 
791: Notice that the $(X^i_{n+1,0})_{1 \leq i \leq N}$ are conditionally
792: independent w.r.t.~$\xi_{n}$.
793: This is the so-called {\bf residual resampling} method.
794: \item[{\bf (S3)}] The $(X^i_{n+1,0})_{1 \leq i \leq N}$ are such that, for $1 \leq i \leq N$,
795: \begin{equation}\label{eq:strat}
796: X^i_{(n+1),0} = \sum_{j=1}^N
797:   1_{\{\sum_{l=1}^{j-1} \rho^l_n < (i-U^i_n)/N \leq \sum_{l=1}^{j}
798:     \rho^l_n \}} \xi^j_{n,\kappa},
799: \end{equation}
800: where $(U^i_n)_{1 \leq i \leq N}$ are random variables
801: i.i.d. according to the uniform law on $[0,1]$, independently of $\xi_{n}$. 
802: Notice that the $(X^i_{n+1,0})_{1 \leq i \leq N}$ are conditionally
803: independent w.r.t.~$\xi_{n}$. This is the so-called {\bf stratified resampling} method.
804: \end{itemize}
805: 
806: For $n\in \{1,\hdots,\nu\}$, let us denote by 
807: \begin{equation}\label{eq:eta_n_}
808: \eta_n^N=\frac{1}{N}\sum_{i=1}^N\delta_{\xi^i_n}
809: \end{equation} 
810: the particle approximation of the measure $\eta_n$ defined by: $\forall
811: f:(\R_+^*)^\kappa\rightarrow \R\mbox{ bounded }$,
812: \begin{equation}\label{eq:eta_n}
813: \eta_n(f)=\frac{\E\left(f\left(X_{(n-1)\Dt+\dt},\hdots,X_{(n-1)\Dt+\kappa\dt}\right)\exp\left(-\t\dt\sum_{k=1}^{(n-1)\kappa}(X_{k\dt})^4\right)\right)}{\E\left(\exp\left(-\t\dt\sum_{k=1}^{(n-1)\kappa}(X_{k\dt})^4\right)\right)},
814: \end{equation}
815: where the process $(X_t)_{0 \leq t \leq T}$ is defined by~\eqref{eds}.
816: 
817: For $y=(y_1,\hdots,y_\kappa)\in(\R_+^*)^\kappa$ and
818: $f:(\R_+^*)^\kappa \rightarrow \R$, we set 
819: \begin{equation}\label{eq:P}
820: Pf(y)=\E\left(f(X^{y_\kappa}_\dt,\hdots,X^{y_\kappa}_{\kappa \dt})\right)
821: \end{equation} 
822: where for $x\in\R^*_+$, 
823: \begin{equation}\label{eq:sdex}
824: X^x_t=x+\int_0^tb(X^x_s)ds+W_t
825: \end{equation}
826:  denotes the solution of
827: the stochastic differential equation~\eqref{eds} starting from~$x$.
828: By the Markov property, the measures $(\eta_n)_{1 \leq n \leq \nu}$
829: satisfy the inductive relations, for any function $f:(\R_+^*)^\kappa
830: \rightarrow \R$ bounded, $\forall n\in \{1,\hdots,\nu-1\}$,
831: \begin{align}
832:  \eta_{n+1}(f)&=\frac{\E\left(\exp\left(-\t\dt\sum_{k=1}^{n\kappa}(X_{k\dt})^4\right)\E\left(f\left(X_{n\Dt+\dt},\hdots,X_{n\Dt+\kappa\dt}\right)\bigg|(X_{j\dt})_{0\leq
833:  j\leq
834:  n\kappa}\right)\right)}{\eta_n(g)\E\left(\exp\left(-\t\dt\sum_{k=1}^{(n-1)\kappa}(X_{k\dt})^4\right)\right)}\\
835: &=\frac{1}{\eta_n(g)}\times\frac{\E\left(gPf\left(X_{(n-1)\Dt+\dt},\hdots,X_{(n-1)\Dt+\kappa\dt}\right)\exp\left(-\t\dt\sum_{k=1}^{(n-1)\kappa}(X_{k\dt})^4\right)\right)}{\E\left(\exp\left(-\t\dt\sum_{k=1}^{(n-1)\kappa}(X_{k\dt})^4\right)\right)}=\frac{\eta_n(gPf)}{\eta_n(g)},
836:    \label{inducetan}
837: \end{align}
838: where $g$ is defined by~(\ref{eq:g}). Moreover, we can express $E_{\rm
839:   DMC}^{\dt}(T)$ defined by~(\ref{eq:EDMCdt}) as:
840: \begin{equation}
841:    E_{\rm DMC}^{\dt}(T)=\frac{3}{2}\omega
842: +\theta\frac{\eta_\nu(gy_\kappa^4)}{\eta_\nu(g)}.\label{expr1}
843: \end{equation}
844: Therefore the particle approximation of $E_{\rm DMC}(T)$ is given by
845: \begin{equation}
846:    E_{\rm DMC}^{N,\nu,\kappa}(T)=\frac{3}{2}\omega
847: +\theta\frac{\eta^N_\nu(gy_\kappa^4)}{\eta^N_\nu(g)}.\label{expr2}
848: \end{equation}
849: This approximation of $E_{\rm DMC}(T)$ corresponds to the expression~(\ref{eq:EDMCapprox_prime})
850: given in the introduction. We will also prove in Corollary~\ref{cor:est_err} below the convergence of the
851: approximation which corresponds to the expression~(\ref{eq:EDMCapprox})
852: given in the introduction (see Equation~(\ref{expr3}) below).
853: 
854: The convergence of the approximation $E_{\rm DMC}^{N,\nu,\kappa}(T)$ is ensured by our main result :
855: \begin{thm}\label{th:est_err}
856: \begin{equation}\label{eq:est_err}
857:    \E\left|E_{\rm DMC}(T)-E_{\rm DMC}^{N,\nu,\kappa}(T)\right|\leq
858:    \frac{C}{\nu\kappa}+\frac{C_\nu}{\sqrt{N}},
859: \end{equation}
860: where the constant $C$ only depends on $T$ and the constant $C_\nu$ on $T$ and $\nu$.
861: \end{thm}
862: \begin{arem}\label{choinu}
863: The number of selection steps is $\nu-1$. For instance, when $\nu=1$,
864:    there is no selection involved in the expression of $E_{\rm
865:      DMC}^{N,\nu,\kappa}(T)$ and the particles remain independent. In
866:    this case, the first term in the right hand side
867:    of~(\ref{eq:est_err}) corresponds to the time discretization error
868:    proved in Proposition~\ref{erreurdt}, while the second term is
869:    the classical error estimate related to the law of large numbers. For a fixed number of
870:    selection steps, the theorem ensures the convergence of the particle
871:    approximation $E_{\rm DMC}^{N,\nu,\kappa}(T)$ as the time-step
872:    $\dt=T/(\nu \kappa)$ used for the discretization of the stochastic
873:    differential equation \eqref{eds} tends to~$0$ while the number $N$ of
874:    particles tends to $+\infty$. But this result does not specify the dependence of $C_\nu$ on $\nu$ and
875:    gives no hint on the optimal choice of the number of selection steps
876:    in terms of error minimization. We are going to deal with this
877:    important issue
878:    in the numerical study (see Section~\ref{sec:num_res}).
879: \end{arem}
880: According to the above expressions
881:    \eqref{expr1} and \eqref{expr2} of $E_{\rm DMC}^{\dt}(T)$ and $E_{\rm
882:    DMC}^{N,\nu,\kappa}(T)$, this theorem is easily proved by combining
883:    Proposition \ref{erreurdt} and the following result :
884: \begin{aprop}\label{propvitpart}
885:    \begin{equation}
886:    \E\left|\frac{\eta^N_\nu(gy_\kappa^4)}{\eta^N_\nu(g)}-\frac{\eta_\nu(gy_{\kappa}^4)}{\eta_\nu(g)}\right|\leq \frac{C_\nu}{\sqrt{N}}.\label{vitpart}
887: \end{equation}\end{aprop}
888: 
889: \begin{adem}[ of Proposition \ref{propvitpart}]
890: One has
891: \begin{align*}
892:    \E\left|\frac{\eta^N_\nu(gy_\kappa^4)}{\eta^N_\nu(g)}-\frac{\eta_\nu(gy_\kappa^4)}{\eta_\nu(g)}\right|\leq &\frac{\E|\eta_\nu^N(gy_\kappa^4)-\eta_\nu(gy_\kappa^4)|}{\eta_\nu(g)}\\&+\left(\E\left(\frac{\eta_\nu^N(gy_\kappa^4)}{\eta_\nu^N(g)}\right)^2\right)^{1/2}\frac{\left(\E\left(\eta_\nu^N(g)-\eta_\nu(g)\right)^2\right)^{1/2}}{\eta_\nu(g)}.
893: \end{align*}
894: According to Proposition \ref{vitmc} and Lemma \ref{contetag} below, the
895: first term of the right-hand-side and the quotient in the second term
896: are smaller than $C_\nu/\sqrt{N}$. Since by Jensen's inequality,
897: $\left(\frac{\eta_\nu^N(gy_\kappa^4)}{\eta_\nu^N(g)}\right)^2\leq
898: \frac{\eta_\nu^N(gy_\kappa^8)}{\eta_\nu^N(g)}$, the boundedness of 
899: $\E\left(\frac{\eta_\nu^N(gy_\kappa^4)}{\eta_\nu^N(g)}\right)^2$
900: follows from Lemma \ref{majomom} below. 
901: \end{adem}
902: \begin{aprop}\label{vitmc}
903:    For any bounded function $f:(\R_+^*)^\kappa\rightarrow \R$, 
904: \begin{equation}\label{eq:f_bounded}
905: \forall
906: n\in\{1,\hdots,\nu\},\;\E((\eta_n^N(f)-\eta_n(f))^2)\leq\frac{C_n}{N}\|f\|^2_\infty,
907: \end{equation}
908: where the constant $C_n$ does not depend on $\kappa$.
909: 
910: For any function $f:(\R_+^*)^\kappa\rightarrow \R$ such that for some $p\geq 2$,
911: $\ds{\|f\|_{\kappa,p}=\sup_{y\in\R_+^\kappa}\frac{|f(y)|}{1+y_\kappa^p}}$ is
912: finite,
913: \begin{equation}\label{eq:fkp_bounded}
914:  \forall
915: n\in\{1,\hdots,\nu\},\;\E|\eta_n^N(f)-\eta_n(f)|\leq\frac{C_n}{\sqrt{N}}\|f\|_{\kappa,p},
916: \end{equation}
917: where the constant $C_n$ does not depend on $\kappa$.
918: \end{aprop}
919: For $f$ bounded, the first estimate~(\ref{eq:f_bounded}) is proved in
920: \cite{Delmoralmiclo}. In order to prove Proposition~\ref{propvitpart}, we need to
921: apply Proposition~\ref{vitmc} with $f(y)=g(y)$ and $f(y)=g(y) y_\kappa^4$, which are
922: bounded functions with $L^\infty$ norm respectively equal to $1$ and
923: $\frac{C}{\dt}$ where $C$ is a constant not depending on $\dt$. But we
924: want to obtain the convergence when $\dt$ tends to $0$. This is why we need the
925: second estimate~(\ref{eq:fkp_bounded}), that we use with $f(y)=g(y)
926: y_\kappa^4$ for which $\|f\|_{\kappa,p}$ is bounded and does not depend on $\dt$.
927: 
928: Notice that for $f$ bounded, Corollary 2.20 in~\cite{Delmoralmiclo} states the convergence in law of
929: $\sqrt{N}(\eta^N_n(f)-\eta_n(f))$ to a centered Gaussian variable and
930: gives an expression of the variance of this limit variable. Because of
931: the complexity of this expression, using this result with $f(y)=g(y)y_\kappa^4$ did not really help us to
932: understand the dependence of $C_\nu$ on $\nu$ (see Remark \ref{choinu}
933: above).
934: \begin{adem}
935: For $f$ bounded, the first estimate~(\ref{eq:f_bounded}) is proved by
936: induction on $n$ in
937: \cite{Delmoralmiclo} (see Proposition 2.9). Since we follow the same
938: inductive reasoning to deal with $f$ such that
939: $\|f\|_{\kappa,p}<+\infty$, we give at the same time the proof for $f$ bounded.  
940: 
941: Since the initial positions $(\xi^i_1)_{1\leq i\leq N}$ are independent and identically
942: distributed with $\xi^i_{1,\kappa}$ distributed according to
943: $2\psi_I^2(x)1_{\{x>0\}}dx$, the
944: statement holds for $n=1$.\\
945: To deduce the statement at rank $n+1$ from the statement at rank $n$,
946: we remark that according to \eqref{inducetan}, 
947: \begin{equation}
948:    \eta_{n+1}^N(f)-\eta_{n+1}(f)=T_{n+1}+\frac{1}{\eta_n(g)}\left((\eta_n^N(gPf)-\eta_n(gPf))+\frac{\eta_n^N(gPf)}{\eta_n^N(g)}(\eta_n(g)-\eta_n^N(g))\right)\label{decomperr}
949: \end{equation}
950: where we recall that $P$ is defined by~(\ref{eq:P}), and 
951: $$T_{n+1}=\eta_{n+1}^N(f)-\frac{\eta_n^N(gPf)}{\eta_n^N(g)}.$$
952: To deal with this term $T_{n+1}$, one remarks that for the first type of
953: selection step (S1), all the possible choices
954: of $\epsilon_n$ given in \eqref{choix_eps} are $\sigma(\xi_n)$-measurable. As a consequence, for $i\in\{1,\hdots,N\}$,
955: $$\E(f(\xi^i_{n+1})|\xi_n)=\epsilon_n
956: g(\xi^i_n)Pf(\xi^i_n)+(1-\epsilon_n
957: g(\xi^i_n))\sum_{j=1}^N \rho^j_n Pf(\xi_{n}^j),$$
958: where $\rho^j_n$ is defined by~(\ref{eq:rho}). Multiplying
959: this equality by $\frac{1}{N}$ and summing over $i$, one deduces
960: \begin{equation}
961:    \E(\eta_{n+1}^N(f)|\xi_n)=\sum_{j=1}^N \rho_n^j  Pf (\xi^j_n)=\frac{\sum_{j=1}^Ng(\xi^j_n)Pf(\xi_n^j)}{\sum_{j=1}^N
962:    g(\xi^j_n)}=\frac{\eta_n^N(gPf)}{\eta_n^N(g)}.\label{condit}
963: \end{equation}
964: Now, for the stochastic remainder resampling algorithm (S2), by~\eqref{messel2},
965: $\E(\eta_{n+1}^N(f)|\xi_n)$ is equal to
966: $$\frac{1}{N}\sum_{j=1}^N\Bigg\lfloor\frac{Ng(\xi^j_n)}{\sum_{l=1}^N
967:   g(\xi^l_n)}\Bigg\rfloor
968:   Pf(\xi^j_n)+\sum_{i=1+\sum_{l=1}^N a_n^l}^N\frac{1}{N- \sum_{l=1}^N
969:     a_n^l} \, \sum_{j=1}^N\left\{\frac{Ng(\xi^j_n)}{\sum_{l=1}^N
970:   g(\xi^l_n)}\right\}Pf(\xi^j_n)$$
971: and \eqref{condit} still holds.  Finally, for the stratified resampling
972: method (S3), by~\eqref{eq:strat}, we have (using the footnote\footnotemark[2])
973: \begin{align*}
974: \E(\eta_{n+1}^N(f)|\xi_n)
975: &=\frac{1}{N}\sum_{i=1}^N\sum_{j=1}^N \E \left( 1_{\{\sum_{l=1}^{j-1} \rho^l_n < (i-U^i_n)/N \leq \sum_{l=1}^{j}
976:     \rho^l_n \}} \Big|\xi_n \right) Pf (\xi^j_n),\\
977: &=\frac{1}{N}\sum_{j=1}^N \E \left( \sum_{i=1}^N 1_{\{\sum_{l=1}^{j-1} \rho^l_n < (i-U^1_n)/N \leq \sum_{l=1}^{j}
978:     \rho^l_n \}} \Big|\xi_n \right) Pf (\xi^j_n),\\
979: &=\frac{1}{N}\sum_{j=1}^N \E \left( \left\lfloor   N \sum_{l=1}^{j}
980:     \rho^l_n  + U^1_n  \right\rfloor - \left\lfloor   N \sum_{l=1}^{j-1}
981:     \rho^l_n  + U^1_n  \right\rfloor \Big|\xi_n \right) Pf (\xi^j_n),\\
982: &=\sum_{j=1}^N \rho_n^j  Pf (\xi^j_n),
983: \end{align*}
984: which yields again~\eqref{condit}.
985: Since for all three possible selection steps, the variables $(\xi^i_{n+1})_{1\leq i\leq N}$ are independent
986: conditionally on $\xi_n$, one deduces that
987: \begin{align*}
988:    \E((T_{n+1})^2|\xi_n)=\frac{1}{N^2}\sum_{i=1}^N\E\left(\left(f(\xi_{n+1}^i)-\E(f(\xi^i_{n+1})|\xi_n)\right)^2|\xi_n\right)\leq \frac{1}{N}\E\left(\eta_{n+1}^N(f^2)|\xi_n\right).
989: \end{align*}
990: Therefore 
991: \begin{equation}
992:    \E((T_{n+1})^2)\leq \frac{1}{N}\E(\eta_{n+1}^N(f^2)).\label{majot}
993: \end{equation}
994: When $f$ is bounded, $\eta_{n+1}^N(f^2)\leq \|f\|_\infty^2$, $\left|\frac{\eta_n^N(gPf)}{\eta_n^N(g)}\right|\leq
995: \|Pf\|_\infty$, and $\|Pf\|_\infty\leq \|f\|_\infty$. Hence by \eqref{decomperr},
996: \begin{align*}
997:    \E((\eta_{n+1}^N(f)-\eta_{n+1}(f))^2)\leq 3\left(\frac{\|f\|_\infty^2}{N}+\frac{\E((\eta_n^N(gPf)-\eta_n(gPf))^2)+\|f\|_\infty^2\E((\eta_n^N(g)-\eta_n(g))^2)}{(\eta_n(g))^2}\right)
998: \end{align*}
999: with the second term of the right-hand-side smaller than $C\|f\|_\infty^2/N$ by the
1000: induction hypothesis and Lemma \ref{contetag} below.\par
1001: When $\|f\|_{\kappa,p}<+\infty$, combining \eqref{decomperr} and
1002: \eqref{majot}, one obtains
1003: \begin{align*}
1004:    \E\left|\eta_{n+1}^N(f)-\eta_{n+1}(f)\right|\leq
1005:    &\frac{\left(\E(\eta_{n+1}^N(f^2))\right)^{1/2}}{\sqrt{N}}+\frac{\E\left|\eta_n^N(gPf)-\eta_n(gPf)\right|}{\eta_n(g)}\\&+\left(\E\left(\frac{\eta_n^N(gPf)}{\eta_n^N(g)}\right)^2\right)^{1/2}\frac{ \left( \E(\eta_n^N(g)-\eta_n(g))^2 \right)^{1/2}}{\eta_n(g)}.
1006: \end{align*}
1007: Since $\|f^2\|_{k,2p} \leq 2 \|f\|_{k,p}^2$ (by using the inequality $f^2(y)\leq 2\|f\|_{\kappa,p}^2(1+y_\kappa^{2p})$), the first term of the right-hand-side is smaller than
1008: $C_n\|f\|_{\kappa,p}/\sqrt{N}$ by Lemma \ref{majomom} below. Since, according to
1009: Lemma \ref{contmom} below, $\|Pf\|_{\kappa,p}\leq e^{C_p\Delta t}\|f\|_{\kappa,p}$, the second term is smaller than
1010: $C_n\|f\|_{\kappa,p}/\sqrt{N}$ by the induction hypothesis and Lemma
1011: \ref{contetag}. Last, by using successively Cauchy Schwartz
1012: inequalities, \eqref{condit} for $f^2$ and Lemma \ref{majomom}, one obtains that 
1013: $\E\left(\frac{\eta_n^N(gPf)}{\eta_n^N(g)}\right)^2 \leq
1014: \E\left(\frac{\eta_n^N(g (Pf)^2)}{\eta_n^N(g)}\right) \leq
1015: \E\left(\frac{\eta_n^N(g P f^2)}{\eta_n^N(g)}\right) =
1016: \E(\eta_{n+1}^N(f^2)) \leq C_n\|f\|^2_{\kappa,p}$. And it follows
1017: from the Proposition statement for $f$ bounded and Lemma
1018: \ref{contetag} that $\frac{\left(\E(\eta_n^N(g)-\eta_n(g))^2\right)^{1/2}}{\eta_n(g)}$ is smaller than
1019: $C_n/\sqrt{N}$.
1020: \end{adem}
1021: 
1022: \begin{arem}\label{rem:SRR}
1023: Proposition~\ref{vitmc} (and therefore Theorem~\ref{th:est_err}) also
1024: hold for the stratified remainder resampling algorithm, which consists
1025: in combining the stochastic remainder resampling and the stratified
1026: resampling. More precisely, it consists in replicating
1027: $\lfloor N \rho^i_n  \rfloor$ times the $i$-th particle, and then completing
1028: the set of particles by using stratified resampling to draw the
1029: $N^R=N-\sum_{l=1}^N \lfloor N \rho^l_n  \rfloor$ remaining particles, the
1030: $i$-th particle being assigned the weight $\rho^{R,i}_n=\{ N \rho^i_n \} /
1031: N^R$.
1032: \end{arem}
1033: 
1034: \begin{alem}\label{majomom} Let $h:(\R_+^*)^\kappa\rightarrow \R_+$ be such that for some
1035:   $p\geq 2$, $\|h\|_{\kappa,p}<+\infty$. Then,
1036: $$\forall n \in \{1,\hdots,\nu\},\; \max\left( \E(\eta_n^N(h)) ,
1037:   \E\left(\frac{\eta_n^N(gh)}{\eta_n^N(g)}\right) \right)\leq
1038: e^{C_pn\Dt}\|h\|_{\kappa,p}(1+\E(X_0)^p),$$
1039: where $X_0$ is distributed according to the measure
1040: $2\psi_I^2(x)1_{\{x>0\}}dx$ (see~(\ref{eds})).
1041: \end{alem}
1042: \begin{adem}
1043: As the variables $\xi^i_{1,\kappa}, 1\leq i\leq N$ are distributed
1044: according to the invariant measure $2\psi_I^2(x)1_{\{x>0\}}dx$, one has
1045: $\E(\eta^N_1(h))\leq \|h\|_{\kappa,p}(1+\E(X_0)^p)$. In addition for $n\geq
1046: 1$, according to \eqref{condit},
1047: $\E(\eta^N_{n+1}(h))=\E\left(\frac{\eta_n^N(gPh)}{\eta_n^N(g)}\right)$
1048: where $\|Ph\|_{\kappa,p}\leq e^{C_p\Dt}\|h\|_{k,p}$  by Lemma \ref{contmom}. Therefore it is enough to
1049: check the bound for $\E\left(\frac{\eta_n^N(gh)}{\eta_n^N(g)}\right)$.\par
1050: For $n\geq 0$, one has
1051: \begin{align}
1052:  \E\left(\frac{\eta_{n+1}^N(gh)}{\eta_{n+1}^N(g)}\right) \leq
1053:   \|h\|_{\kappa,p}\left(1+\E\left(\frac{\sum_{i=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa }(\xi^i_{n+1,k})^4\right)(\xi^i_{n+1,\kappa})^p}{\sum_{j=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa}(\xi^j_{n+1,k})^4\right)}\right)\right).\label{decompmom}
1054: \end{align}
1055: Let us denote in this proof $\xi^i_{n+1,0}=X^i_{n+1,0}$, where $0\leq n
1056: \leq \nu-1$ and $1\leq i \leq N$. Let us set ${\mathcal
1057:   F}=\sigma(\xi^i_{n+1,k},\;1\leq i\leq N,\;0\leq k\leq
1058:   \kappa-1)$. By Lemma \ref{decroiss} below,
1059: \begin{align*}
1060:    \E\Bigg(&\frac{\sum_{i=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa}(\xi^i_{n+1,k})^4\right)(\xi^i_{n+1,\kappa})^p}{\sum_{j=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa}(\xi^j_{n+1,k})^4\right)}\bigg| {\mathcal
1061:   F}\Bigg)\leq \frac{\sum_{i=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1062:    -1}(\xi^i_{n+1,k})^4\right)\E((\xi^i_{n+1,\kappa})^p|{\mathcal
1063:   F})}{\sum_{j=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1064:    -1}(\xi^j_{n+1,k})^4\right)},\\
1065: &\phantom{xxxxxxxxxxxxxxxxxxxxxx}=\frac{\sum_{i=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1066:    -1}(\xi^i_{n+1,k})^4\right)\E((X^x_\dt)^p)|_{x=\xi^i_{n+1,\kappa-1}}}{\sum_{j=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1067:    -1}(\xi^j_{n+1,k})^4\right)},\\
1068: &\phantom{xxxxxxxxxxxxxxxxxxxxxx}\leq e^{C_p\dt}\frac{\sum_{i=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1069:    -1}(\xi^i_{n+1,k})^4\right)(\xi^i_{n+1,\kappa-1})^p}{\sum_{j=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1070:    -1}(\xi^j_{n+1,k})^4\right)}+e^{C_p\dt}-1,
1071: \end{align*}
1072: where we have used the definition of the mutation step (see~\eqref{sden+1}) and the Markov
1073:    property for the stochastic differential equation \eqref{eq:sdex} to
1074:    obtain the equality, and then Lemma \ref{contmom} for the last
1075:    inequality. Notice that this estimate also holds for $\kappa=1$, in
1076:    which case the right hand side reduces to $ \ds{\frac{e^{C_p\dt}}{N}
1077:    (\xi^i_{n+1,0})^p +e^{C_p\dt}-1}$.
1078: 
1079: Taking expectations and iterating the reasoning, one deduces that
1080: $$\E\left(\frac{\sum_{i=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa
1081:         }(\xi^i_{n+1,k})^4\right)(\xi^i_{n+1,\kappa})^p}{\sum_{j=1}^N\exp\left(-\t\dt\sum_{k=1}^{\kappa }(\xi^j_{n+1,k})^4\right)}\right)\leq \frac{e^{C_p\Dt}}{N}\sum_{i=1}^N\E((\xi^i_{n+1,0})^p)+(e^{C_p\dt}-1)\sum_{k=0}^{\kappa-1}e^{C_pk\dt}.$$
1082: Inserting this bound in \eqref{decompmom}, one concludes that
1083: \begin{align*}
1084:  \E\left(\frac{\eta_{n+1}^N(gh)}{\eta_{n+1}^N(g)}\right) &\leq
1085:  e^{C_p\Dt}\|h\|_{\kappa,p}\left(1+\E\left(\frac{1}{N}\sum_{i=1}^N(\xi^{i}_{n+1,0})^p\right)\right).
1086: \end{align*}
1087: For $n=0$, one deduces that $\E\left(\frac{\eta_{1}^N(gh)}{\eta_{1}^N(g)}\right)\leq
1088:  e^{C_p\Dt}\|h\|_{\kappa,p}(1+\E(X_0^p))$, where $X_0$ is distributed according to the measure
1089: $2\psi_I^2(x)1_{\{x>0\}}dx$.
1090: 
1091: For $n\geq 1$, since by a reasoning similar to the one made to obtain \eqref{condit},
1092: $\ds{\E\left(\frac{1}{N}\sum_{i=1}^N(\xi^{i}_{n+1,0})^p\right)=\E\left(\frac{\eta_{n}^N(g(y)
1093:       y_\kappa^p)}{\eta_n^N(g(y))}\right)}$,
1094:  one also deduces that
1095: $$\E\left(\frac{\eta_{n+1}^N(gh)}{\eta_{n+1}^N(g)}\right)\leq
1096: e^{C_p\Dt}\|h\|_{\kappa,p}\E\left(\frac{\eta_{n}^N(g(1+y_\kappa^p))}{\eta^N_n(g)}\right).$$
1097: The proof is completed by an obvious inductive reasoning.
1098: \end{adem}
1099: 
1100: \begin{alem}\label{contmom}
1101: For any $p\geq 2$, there is a constant $C_p$ such that 
1102: $$\forall x\in\R_+^*,\;\forall t\geq 0,\;\E((X^x_t)^p)\leq
1103: (1+x^p)e^{C_pt}-1,$$ 
1104: where $X_t^x$ is defined by~(\ref{eq:sdex}). Therefore, if
1105: $h:(\R_+^*)^\kappa\rightarrow \R$ is such that
1106: $\|h\|_{\kappa,p}<+\infty$ then $\|Ph\|_{\kappa,p}\leq
1107: e^{C_p\Dt}\|h\|_{\kappa,p}$, where the operator $P$ is defined by~(\ref{eq:P}).
1108: \end{alem}
1109: \begin{adem}
1110: By Itô's formula, $d(X^x_t)^p=\left(\frac{p(p+1)}{2}(X^x_t)^{p-2}-\omega
1111:   p(X^x_t)^{p}\right)dt+p(X^x_t)^{p-1}dW_t$. Hence
1112: $$(X^x_t)^p\leq
1113: x^p+\int_0^t\left(\frac{p(p+1)}{2}+\frac{p(p+1-2\omega)}{2}(X^x_s)^{p}\right)ds+p\int_0^t(X^x_s)^{p-1}dW_s.$$
1114: Formally, taking expectations in this inequality, one obtains
1115: \begin{align*}
1116:    \E((X^x_t)^p)\leq x^p+\int_0^t\frac{p(p+1)}{2}+\frac{p(p+1-2\omega)}{2}\E((X^x_s)^{p})ds,
1117: \end{align*}
1118: and check by Gronwall's lemma that the conclusion holds with
1119: $C_p=\frac{p(p+1)}{2}$. This formal argument can be made rigorous by a
1120: standard localization procedure.\\
1121: For $h:\R_+^\kappa\rightarrow \R$ such that $\|h\|_{\kappa,p}<+\infty$
1122: one deduces that
1123: $$\forall y\in\R_+^\kappa,\;|Ph(y)|\leq
1124: \E|h(X^{y_\kappa}_\dt,\hdots,X^{y_\kappa}_{\kappa\dt})|\leq
1125: C\|h\|_{\kappa,p}(1+\E((X^{y_\kappa}_{\kappa\dt})^p))\leq e^{C_p\Dt}\|h\|_{\kappa,p}(1+y_\kappa^p).$$\end{adem}
1126: \begin{alem}\label{decroiss}
1127: $$\forall(z_1,\hdots,z_N),(a_1,\hdots,a_N)\in\R_+^N\mbox{ with }\sum_{i=1}^Na_i>0,\;\forall p\geq 0,\;\forall c\geq
1128: 0,\;\frac{\sum_{i=1}^Na_iz_i^{p}e^{-cz_i^4}}{\sum_{i=1}^Na_ie^{-cz_i^4}}\leq
1129: \frac{\sum_{i=1}^Na_iz_i^{p}}{\sum_{i=1}^Na_i}.$$
1130: \end{alem}
1131: \begin{adem}Let us set
1132:    $f(c)=\frac{\sum_{i=1}^Na_iz_i^{p}e^{-cz_i^4}}{\sum_{i=1}^Na_ie^{-cz_i^4}}$. 
1133: By Hölder's inequality, the derivative
1134: $$f'(c)=\left(
1135:   \frac{\sum_{i=1}^Na_iz_i^{p}e^{-cz_i^4}}{\sum_{i=1}^Na_ie^{-cz_i^4}}
1136:   \frac{\sum_{i=1}^Na_iz_i^{4}e^{-cz_i^4}}{\sum_{i=1}^Na_ie^{-cz_i^4}} \right)-\frac{\sum_{i=1}^Na_iz_i^{p+4}e^{-cz_i^4}}{\sum_{i=1}^Na_ie^{-cz_i^4}}$$ 
1137: is non positive. Hence for any 
1138: $c\geq 0$, $f(c)\leq f(0)=\frac{\sum_{i=1}^Na_iz_i^{p}}{\sum_{i=1}^Na_i}$.
1139: \end{adem}
1140: \begin{alem}\label{contetag}
1141: The sequence $(\eta_n(g))_{1\leq n\leq\nu}$ is bounded from below by a
1142: positive constant non depending on~$\kappa$.\end{alem}
1143: \begin{adem}
1144: Since
1145: $$\eta_n(g)=\frac{\E\left(\exp\left(-\t\dt\sum_{k=1}^{n\kappa}X_{k\dt}^4\right)\right)}{\E\left(\exp\left(-\t\dt\sum_{k=1}^{(n-1)\kappa}X_{k\dt}^4\right)\right)}\leq
1146: 1$$
1147: the sequence $(\eta_n(g))_{1\leq n\leq\nu}$ is bounded from below by 
1148: $$\prod_{n=1}^\nu
1149: \eta_n(g)=\E\left(\exp\left(-\t\dt\sum_{k=1}^{\nu\kappa}X_{k\dt}^4\right)\right).$$
1150: According to Lemma \ref{vitfort}, this expectation converges to
1151: $\E\left(\exp\left(-\t\int_0^TX_{s}^4ds\right)\right)>0$ when $\kappa$
1152: tends to $+\infty$, which concludes the proof.\end{adem}
1153: 
1154: We can now prove, as a corollary of Theorem~\ref{th:est_err}, the convergence of the
1155: approximation $\overline{E_{\rm DMC}^{N,\nu,\kappa}}(T)$ of $E_{\rm
1156:   DMC}(T)$, defined by:
1157: \begin{equation}
1158: \overline{E_{\rm DMC}^{N,\nu,\kappa}}(T)=\frac{3}{2}\omega
1159: +\frac{\theta}{N}\sum_{i=1}^N (X^i_{\nu+1,0})^4.\label{expr3}
1160: \end{equation}
1161: \begin{acor}\label{cor:est_err}
1162: $$
1163:    \E\left|E_{\rm DMC}(T)-\overline{E_{\rm DMC}^{N,\nu,\kappa}}(T)\right|\leq
1164:    \frac{C}{\nu\kappa}+\frac{C_\nu}{\sqrt{N}},
1165: $$
1166: where the constant $C$ only depends on $T$ and the constant $C_\nu$ on $T$ and $\nu$.
1167: \end{acor}
1168: \begin{adem}
1169: By using the result of Theorem~\ref{th:est_err} and Cauchy Schwartz inequality, it is sufficient to prove
1170: the estimate $\ds{\E\left(E_{\rm DMC}^{N,\nu,\kappa}(T)-\overline{E_{\rm
1171:       DMC}^{N,\nu,\kappa}}(T)\right)^2\leq \frac{C_\nu}{N}}$. Let us
1172: denote in this proof $\xi^i_{\nu+1,0}=X^i_{\nu+1,0}$ for $1\leq i \leq N$. We have:
1173: \begin{equation*}
1174: E_{\rm DMC}^{N,\nu,\kappa}(T)-\overline{E_{\rm DMC}^{N,\nu,\kappa}}(T)
1175: = \theta\left(\frac{\eta_\nu^N(g \, y_\kappa^4)}{\eta_\nu^N(g)} -
1176:   \frac{1}{N} \sum_{i=1}^N (\xi^i_{\nu+1,0})^4\right)
1177: = \theta\left( \E\left( \frac{1}{N} \sum_{i=1}^N (\xi^i_{\nu+1,0})^4 \bigg|
1178:     \xi_\nu \right) -
1179:   \frac{1}{N} \sum_{i=1}^N (\xi^i_{\nu+1,0})^4\right)
1180: \end{equation*}
1181: by using the fact that, for any function $f:\R_+^* \to \R_+$, 
1182: \begin{equation}\label{eq:rec}
1183: \E\left( \frac{1}{N}
1184:   \sum_{i=1}^N f(\xi^i_{\nu+1,0}) \bigg| \xi_\nu \right)=\frac{\eta_\nu^N(g(y)
1185:   \,f(y_\kappa))}{\eta_\nu^N(g(y))},
1186: \end{equation} which is obtained by a reasoning similar to the
1187: one made to prove~\eqref{condit}. Now, using the same method as to
1188: obtain~\eqref{majot}, one easily gets the estimate:
1189: $$\E\left( E_{\rm DMC}^{N,\nu,\kappa}(T)-\overline{E_{\rm
1190:       DMC}^{N,\nu,\kappa}}(T) \right)^2 \leq \frac{\theta^2}{N}
1191: \E\left(\frac{1}{N}\sum_{i=1}^N (\xi^i_{\nu+1,0})^8\right)= \frac{\theta^2}{N} \E\left(
1192: \frac{\eta_\nu^N(g(y)
1193:   \,(y_\kappa)^8)}{\eta_\nu^N(g(y))}\right),
1194: $$
1195: by using again~(\ref{eq:rec}). Lemma~\ref{majomom} completes the proof.
1196: \end{adem}
1197: 
1198: We end this Section by proving that Proposition~\ref{propvitpart} also
1199: holds for the numerical scheme~\eqref{schemaurel}.
1200: \begin{aprop}\label{prop:Markov}
1201: Let us consider the Markov chain $(\bar{X}_{j\delta t})_{0\leq j\leq K}$
1202: generated by the explicit scheme \eqref{schemaurel} and denote by $Q$ its
1203: transition kernel. We now define the measure
1204: $\eta_n$ by replacing $(X_{j\delta t})_{0\leq j\leq
1205:   K}$ with $(\bar{X}_{j\delta t})_{0\leq j\leq K}$ in~(\ref{eq:eta_n}), and we define accordingly the
1206:    evolution of the particle system: conditionally on $\xi_n$,
1207:  the vectors $(X^i_{n+1,0},X^i_{n+1,\dt},\hdots,X^i_{n+1,\kappa\dt})_{1\leq i\leq N}$
1208:    are independent, with $(X^i_{n+1,0})_{1 \le i \le N}$ distributed according to
1209:  the selection algorithm (S1) (see~\eqref{messel}), (S2)
1210:  (see~\eqref{messel2}) or (S3) (see~\eqref{eq:strat}), and $(X^i_{n+1,j\dt})_{0\leq j\leq \kappa}$ a Markov
1211:    chain with transition kernel $Q$. Then, we have:
1212: $$
1213:    \E\left|\frac{\eta^N_\nu(gy_\kappa^4)}{\eta^N_\nu(g)}-\frac{\eta_\nu(gy_{\kappa}^4)}{\eta_\nu(g)}\right|\leq \frac{C_\nu}{\sqrt{N}}.
1214: $$
1215: \end{aprop} 
1216: \begin{adem}
1217: %    In the definition~(\ref{eq:eta_n}) of $\eta_n$, one may suppose more generally that
1218: %    $(X_{j\delta t})_{0\leq j\leq K}$ is an homogeneous Markov chain on
1219: %    $\R_+$ with trandition kernel $Q$. We think here for example of a
1220: %    Markov chain defined by a numerical scheme such as~\eqref{schemaurel}. Then, accordingly, in the
1221: %    evolution of the particle system, one assumes that conditionally on $\xi_n$,
1222: %  the vectors $(\xi^i_{n+1,0},\hdots,\xi^i_{n+1,\kappa})_{1\leq i\leq N}$
1223: %    are independent with $\xi^i_{n+1,0}$ distributed according to
1224: %    \eqref{messel} and $(\xi^i_{n+1,j})_{0\leq j\leq \kappa}$ a Markov
1225: %    chain with transition kernel $Q$. Notice that in this remark, we use the notation
1226: %    $\xi^i_{n+1,0}=X^i_{n+1,0}$, for $0\leq n
1227: % \leq \nu-1$ and $1\leq i \leq N$.
1228: Looking carefully at the proof
1229:    of Proposition \ref{propvitpart} above, one remarks that \eqref{vitpart}
1230:    holds in this framework as soon as
1231:    Lemma~\ref{contetag} holds, and the following property, which
1232:    replaces Lemma \ref{contmom}, is satisfied:
1233: \begin{equation}\label{eq:hyp_schemaurel}
1234: \exists C>0,\;\forall x\in\R_+,\;Qf(x)\leq e^{C\delta
1235:   t}(1+f(x))-1\mbox{ for }f(x)\equiv x^4\mbox{ and }f(x)\equiv x^8.
1236: \end{equation}
1237: Let us first prove~(\ref{eq:hyp_schemaurel}). We have:
1238: $Qf(x)=\E\left(f\left(\bar{X}^x_{\delta t}\right)\right)$ where $\bar{X}^x_\dt=\left((1-\omega
1239:     \delta t)^2x^2+2xW_\dt+\frac{W_\dt^2}{(1-\omega
1240:     \delta t)^2}+2\delta t\right)^{1/2}$. Now, for $q\in\N^*$,
1241: $$(\bar{X}^x_{\delta t})^{2q}=\sum_{j_1+j_2+j_3=q}\frac{q!}{j_1!j_2!j_3!}\,(1-\omega\dt)^{2j_1}\,2^{j_2}\,x^{2j_1+j_2}\,W_\dt^{j_2}\left(\frac{W_\dt^2}{(1-\omega
1242:     \delta t)^2}+2\delta t\right)^{j_3},$$
1243: where the indices $(j_1,j_2,j_3)$ are non negative integers. Remarking that the expectation of the terms with $j_2$ odd vanishes and
1244:     then using Young's inequality, one deduces that for $\dt\leq
1245:     \frac{1}{2\omega}$,
1246: \begin{eqnarray}
1247:    \E\left((\bar{X}^x_{\delta t})^{2q}\right)&\leq& (1-\omega\dt)^{2q}x^{2q}+\E\left(\left(\frac{W_\dt^2}{(1-\omega
1248:     \delta t)^2}+2\delta
1249:     t\right)^{q}\right)+C_q \!\!\!\!\!\!\!\!\!\sum_{\stackrel{j_1+j_2+j_3=q}{j_1<q,j_2\;even\;,j_3<q}}\!\!\!\!\!\!\!\!\!x^{2\left(q-\frac{j_2+2j_3}{2}\right)}\dt^{\frac{j_2+2j_3}{2}},\nonumber\\
1250: &\leq& x^{2q}+C_q\dt+C_q\!\!\!\!\!\!\!\!\!\sum_{\stackrel{j_1+j_2+j_3=q}{j_1<q,j_2
1251:     \;even\;,
1252:     j_3<q}}\!\!\!\!\!\!\!\!\!\left(x^{2q}\dt+\dt^{1+q\left(1-\frac{2}{j_2+2j_3}\right)}\right),\nonumber\\
1253: &\leq& (1+C_q\dt)x^{2q}+C_q\dt\leq e^{C_q\dt}(1+x^{2q})-1. \label{eq:Xbar_estim}
1254: \end{eqnarray}
1255: Let us now prove Lemma~\ref{contetag} for the scheme
1256: \eqref{schemaurel}. As noticed in the proof of  Lemma~\ref{contetag}
1257: above, it is sufficient to bound from below
1258: $\E\left(\exp\left(-\t\dt\sum_{k=1}^{\nu\kappa}\bar{X}_{k\dt}^4\right)\right)$.
1259: By Jensen inequality, we have
1260: $\E\left(\exp\left(-\t\dt\sum_{k=1}^{\nu\kappa}\bar{X}_{k\dt}^4\right)\right)
1261:  \geq \exp\left(-\t\frac{T}{\nu \kappa}\sum_{k=1}^{\nu\kappa}
1262:    \E\left(\bar{X}_{k\dt}^4\right)\right)$. By
1263:  using~(\ref{eq:Xbar_estim}), it is easy to prove by induction that
1264:  $\E\left(\bar{X}_{k\dt}^4\right)\leq e^{C_2 k \dt}
1265:  (1+\E\left(\bar{X}_{0}^4\right))-1$ and this concludes the proof of
1266:  Lemma~\ref{contetag} in this framework.
1267: \end{adem}
1268: In order to obtain a complete convergence result of the
1269: form~(\ref{eq:est_err}) for the scheme \eqref{schemaurel}, it remains to
1270: prove the complementary bound \eqref{discretaurel}, that we have not
1271: obtained so far. However, we will check by numerical simulations that (\ref{eq:est_err}) still holds.
1272: 
1273: \section{Numerical results}\label{sec:num_res}
1274: 
1275: \subsection{Computation of a reference solution by a spectral method}\label{sec:num_res_spec}
1276: 
1277: In this section, we would like to explain how we can obtain a very
1278: precise reference solution by using a partial differential equation approach to compute
1279: $E_{\rm DMC}(T)$ (see~\cite{CJL}).
1280: 
1281: \subsubsection{A partial differential equation approach to compute $E_{\rm DMC}(T)$}
1282: 
1283: Let us introduce the solution $\phi$ to the following partial
1284: differential equation for :
1285: \begin{equation}\label{eq:phi}
1286: \left\{
1287: \begin{array}{l}
1288: \displaystyle{\frac{\partial \phi}{\partial t}=-H \phi},\;(t,x) \in \R_+ \times \R\\
1289: \phi(0,x)=\psi_I(x),\;x\in\R
1290: \end{array}
1291: \right.
1292: \end{equation}
1293: where $H$ (resp. $\psi_I$) is defined by~(\ref{eq:H})
1294: (resp.~(\ref{eq:psi_I})). Since $\psi_I \in {\mathcal H}$, it
1295: is a standard result that this problem admits a unique solution
1296: $\phi  \in C^0(\R_+,{\mathcal H}) \cap C^0(\R_+^*,D_{\mathcal H}(H))
1297: \cap C^1(\R_+^*,{\mathcal H})$. The function $\phi$ is
1298: regular and odd, and therefore is such that $\phi(t,0)=0$ for all $t
1299: \geq 0$. Therefore the function $\phi$ is also solution to the following
1300: partial differential equation:
1301: \begin{equation}\label{eq:phi_prime}
1302: \left\{
1303: \begin{array}{l}
1304: \displaystyle{\frac{\partial \phi}{\partial t}=-H \phi},\;(t,x) \in \R_+ \times \R\\
1305: \phi(t,0)=0,\; t\geq 0\\
1306: \phi(0,x)=\psi_I(x),\;x\in\R.
1307: \end{array}
1308: \right.
1309: \end{equation}
1310: In~\cite{CJL}, we have shown that since $\phi$ satisfies~(\ref{eq:phi_prime}), we can express $E_{\rm DMC}(t)$
1311: (defined by~(\ref{eq:EDMC}))
1312: using the function $\phi$ (see Proposition 11 in~\cite{CJL}):
1313: \begin{equation}\label{eq:EDMC_phi}
1314: E_{\rm DMC}(t)=\frac{\langle H \psi_I,\phi(t)\rangle}{\langle
1315:   \psi_I,\phi(t)\rangle}.
1316: \end{equation}
1317: Our reference solution $E_{\rm DMC}(T)$ will rely on formula
1318: \eqref{eq:EDMC_phi} after discretization of (\ref{eq:phi}) by a spectral
1319: method.
1320: 
1321: % \begin{arem}
1322: % We would like to emphasize one important point. In general (namely in
1323: % the situation described at the beginning of the introduction), it is false
1324: % that~(\ref{eq:phi}) and~(\ref{eq:phi_prime}) admit the same
1325: % solution. Let us make this precise: in general,
1326: % \begin{itemize}
1327: % \item Equation~(\ref{eq:EDMC_phi}) holds with $\phi$ solution
1328: %   to~(\ref{eq:phi_prime}) (see Proposition 11 in~\cite{CJL}),
1329: % \item The quantity $E(t)=\frac{\langle H \psi_I,\phi(t)\rangle}{\langle
1330: %   \psi_I,\phi(t)\rangle}$ where $\phi$ is solution to~(\ref{eq:phi}) is
1331: % different from $E_{\rm DMC}(t)$, and is such that $\lim_{t \to \infty}
1332: % E(t)=E_0$ where $E_0$ is defined by~(\ref{eq:E0}) (see Section 3 in~\cite{CJL}),
1333: % \item However, even if $E(t) \neq E_{{\rm
1334: %       DMC}}(t)$, in the case when the nodal surfaces of $\psi_I$ coincide with
1335: %   those of a ground state of $H$, their longtime limits are the same: $E_0=E_{{\rm DMC},0}$, where $E_{{\rm
1336: %       DMC},0}$ is defined by~(\ref{eq:CV_E_DMC}) (see Theorem 12 in~\cite{CJL}).
1337: % \end{itemize}
1338: % On our toy problem, we have $E(t)=E_{\rm DMC}(t)$ for two reasons: first, the function $\psi_I$ we have chosen has indeed a zero set
1339: % which coincides with the zero set of a ground state of $H$ (namely
1340: % $\{0\}$), and second, the
1341: % antisymmetry property of the solution $\psi$ to~(\ref{eq:phi_prime})
1342: % constrains the function $\phi$ to be zero on this set:
1343: % $\forall t \geq 0,\, \phi(t,0)=0$.
1344: % \end{arem}
1345: 
1346: \subsubsection{Computation of the wave function $\phi$}
1347: 
1348: We will briefly present the spectral method developed to compute an
1349: approximation of $\phi$.
1350: We recall that the Hermite polynomials are defined by  :
1351: $$
1352: \forall n \in \N,\; h_n(x)=(-1)^n e^{x^2}\displaystyle\frac{d^n}{dx^n}(e^{-x^2}).
1353: $$ 
1354: We introduce the eigenfunctions 
1355: of the operator $H_0$, 
1356: normalized for the $L^2(\R)$ norm associated with the eigenvalues
1357: $E_n=\omega(n+1/2)$ for $n \geq 0$,
1358: $$
1359: \varphi_n(x)=\displaystyle h_n(\sqrt \omega x) \exp({-\frac{1}{2}\omega x^2})
1360: \left(\frac{(\omega /\pi)^{1/4}}{\sqrt{2^n n!}}\right).
1361: $$
1362: It is well known
1363: that the vector space spanned by the set of functions
1364: $\{\varphi_{2k+1}\}_{k \geq 0}$ is dense in ${\mathcal V}_0=\{\varphi\in
1365: H^1(\R) \cap {\mathcal H} \;| \;  x\varphi \in L^2\}$, which is the
1366: domain of the quadratic form associated with $H_0$.
1367: 
1368: Let us now introduce the functional space  
1369:  ${\mathcal V}=\{\varphi\in H^1(\R) \cap {\mathcal H} \;| \;
1370:  x^2\varphi\in L^2\}$, which is the
1371: domain of the quadratic form associated with $H$. The set of functions $\{\varphi_{2k+1}\}_{k \geq 0}$ is also a
1372:  basis of ${\mathcal V}$.
1373: 
1374: % Since the functions $\varphi_{2k+1}$ are also in
1375: %  ${\mathcal V}$ and that ${\mathcal V}$ is a closed subset of ${\mathcal
1376: %    V}_0$, the set of functions $\{\varphi_{2k+1}\}_{k \geq 0}$ is also a
1377: %  basis of ${\mathcal V}$.
1378:  Let ${\mathcal V}_n
1379:  =Span({\varphi_1},{\varphi_3},\dots ,{\varphi}_{2n-1})$. We use this
1380:  approximation space to build the following Galerkin scheme
1381:  for~(\ref{eq:phi}): find $\phi_n \in C^0(\R_+,{\mathcal V}_{n})$ such
1382:  that\footnote{Notice that $\psi_I=\varphi_1\in {\mathcal V}_n$.}  
1383: $\phi_n(0,x)=\psi_I$, and $\forall \varphi \in {\mathcal V}_n$
1384: \begin{equation}\label{eq:1}
1385: \displaystyle\left\langle\frac {\partial \phi_n(t)}{\partial t},\varphi\right\rangle=
1386: \displaystyle-\left\langle H \phi_n(x,t),\varphi\right\rangle.
1387: \end{equation}
1388: 
1389: We diagonalize the operator $H$ restricted to ${\mathcal V}_n$. We denote 
1390: $({\varphi_0^n},{\varphi}_2^n,\dots ,{\varphi}_{n-1}^n)$ 
1391: the eigenfunctions and ${E_0^n},{E_2^n},\dots, { E_{n-1}^n}$
1392:  the associated eigenvalues. Because of the symmetry of $H$, it is easy
1393:  to check that ${\mathcal V}_n$ can also be spanned by $({\varphi_0^n},{\varphi}_2^n,\dots ,{\varphi}_{n-1}^n)$:
1394: \begin{equation}\label{eq:span}
1395: {\mathcal V}_n=Span({\varphi_0^n},{\varphi}_2^n,\dots ,{\varphi}_{n-1}^n).
1396: \end{equation}
1397: Since for $t\geq 0$, $\phi_n(t,.) \in {\mathcal V}_n$, there exists $u_{k}(t), \;k=0,\dots,n-1$, such that 
1398: \begin{equation}\label{eq:psi}
1399: \displaystyle\phi_n=\sum_{k=0}^{n-1} u_{k}(t){\varphi_{k}^n}.
1400: \end{equation}
1401: In view of (\ref{eq:span}) and (\ref{eq:psi}), (\ref{eq:1}) is equivalent
1402:  to the  equations: $\forall i=0,\dots,n-1$,
1403: $$
1404: \begin{array}{ccl}
1405: \displaystyle\sum_{k=0}^{n-1}\frac {\partial u_{k}(t)}{\partial t} 
1406: \left\langle{\varphi}_{k}^n,{\varphi}_{i}^n\right\rangle&=&
1407: \displaystyle-\left\langle H\sum_{k=0}^{n-1} u_{k}(t) 
1408: {\varphi}_{k}^n,{\varphi}_{i}^n\right\rangle, \\
1409: &=&\displaystyle-\sum_{k=0}^{n-1}{ E}_{k}^nu_{k}(t)\left\langle{\varphi}_{k}^n,  
1410: {\varphi}_{i}^n\right\rangle.
1411: \end{array}
1412: $$
1413: We deduce that  $\forall k =0,\dots,n-1,$
1414: $$
1415: \displaystyle\frac {\partial u_{k}(t)}{\partial t} = -{E}_{k}^n u_{k}(t),
1416: $$
1417: so that
1418: \begin{equation}\label{eq:psi2}
1419: \displaystyle\phi_n(t,x)=\sum_{k=0}^{n-1}u_{k}(0)\exp(-{E}_{k}^n t){\varphi_{k}^n}(x),
1420: \end{equation}
1421: where $u_{k}(0)=\left\langle\psi_I,{\varphi}_k^n\right\rangle$.
1422: 
1423: \begin{arem}
1424: The eigenfunctions of $H$ are obtained by diagonalization of the matrix
1425: $A=(a_{ij})_{i,j=0,\dots,n-1}$ with $\:\:\forall i,j=0,\dots,n-1$ :
1426: $$
1427: \begin{array}{ccl}
1428: a_{ij}&=&\left\langle H\varphi_{2i+1},\varphi_{2j+1}\right\rangle, \\
1429:            &=&\left\langle H_0\varphi_{2i+1},\varphi_{2j+1}\right\rangle+
1430:              \theta \left\langle x^4\varphi_{2i+1},\varphi_{2j+1}\right\rangle,\\
1431:   &=&\delta_{ij} \, \omega \, (2i+\frac{3}{2}) + \theta\, \left\langle  x^4\varphi_{2i+1},
1432: \varphi_{2j+1}\right\rangle.
1433: \end{array}
1434: $$
1435: We can use the  n--point Gauss-Hermite formula to  deal with the integration of the
1436:  second term on the right. We recall that this method  provides an exact
1437:  result for
1438: $\int_{-\infty}^{+\infty}p(x) \exp(-x^2)dx $ as long as $p$ is a
1439: polynomial of degree $2n-1$ or less.
1440:  
1441: \end{arem}
1442: \subsubsection{Approximation of  $E_{\rm DMC}(T)$}
1443: %+++++++++++++++++++++++++++++++++++++++++
1444: 
1445: We now use formula~(\ref{eq:EDMC_phi}) to approximate $E_{\rm DMC}(T)$.
1446: By an elementary calculation, we obtain the following approximation:
1447: \begin{equation}\label{eq:E2}
1448: E_{\rm DMC}(T) \simeq \displaystyle\frac{{E}_0^n+\displaystyle\sum_{i=1}^{n-1}\displaystyle\frac{u_{i}(0)\left\langle{\varphi}_i^n,\varphi_1\right\rangle}
1449: {u_{0}(0)\left\langle{\varphi}_1^n,\varphi_1\right\rangle}{E}_{i}^n\exp(-({E}_{i}^n-{E}_0^n)T)}
1450: {1+\displaystyle\sum_{i=1}^{n-1}\frac{u_{i}(0)\left\langle{\varphi}_i^n,\varphi_1\right\rangle}
1451: {u_{0}(0)\left\langle{\varphi}_1^n,\varphi_1\right\rangle}\exp(-({E}_{i}^n-{E}_0^n)T)}.
1452: \end{equation}
1453: In our test cases, we have observed that $n=40$ is enough to reach
1454: convergence.
1455: 
1456: Notice that for a given $n$, the convergence in time to the lowest eigenvalue
1457: ${E}_0^n$ is exponentially fast, with an exponent equal to the 
1458: spectral gap ${E}_1^n-{E}_0^n$. 
1459: 
1460: 
1461: \subsection{Numerical results of Monte Carlo simulations}
1462: 
1463: In this section, we perform various numerical experiments to validate
1464: our theoretical results, and to explore some features of DMC
1465: computation. In particular, we propose in Section~\ref{sec:optim_var} an
1466: empirical method to determine the optimal number of reconfigurations. In
1467: all the computations, the final time is $T=5$, which appears to be
1468: sufficiently large for the convergence $t \to \infty$ to be achieved
1469: with enough accuracy.
1470: 
1471: \subsubsection{Error and variance as a function of the numerical parameters}\label{sec:err_var}
1472: 
1473: \begin{figure}
1474: \begin{center}
1475: \psfrag {e(dt)}{$e(\delta t)$}
1476: \psfrag{v(dt)}{$v(\delta t)$}
1477: \psfrag{e(N)}{$e(N)$}
1478: \psfrag{v(N)}{$v(N)$}
1479: \psfrag{e(Nu)}{$e(\nu)$}
1480: \psfrag{v(Nu)}{$v(\nu)$}
1481: \begin{tabular}{cc}
1482:       \includegraphics[width=6cm]{erreurEnNombre2Marcheur.eps} &
1483:       \includegraphics[width=6cm]{varianceEnNombre2Marcheur.eps} \\
1484:       \includegraphics[width=6cm]{erreurEndt.eps} &
1485:       \includegraphics[width=6cm]{varianceEndt.eps} \\
1486:       \includegraphics[width=6cm]{erreurEnNombre2reconf.eps} &
1487:       \includegraphics[width=6cm]{varianceEnNombre2reconf.eps} \\
1488:     \end{tabular}
1489: \caption{Expectation and variance of the error when \eqref{eds} is
1490:   discretized according to the method 
1491:   described in Appendix (dotted curve) and according to the
1492:   scheme~(\ref{schemaurel}) (solid curve).}\label{fig}
1493:   \end{center}
1494: \end{figure}
1495: 
1496: % \begin{figure}[ht]
1497: % \begin{center}
1498: % \psfrag {e(dt)}{$e(\delta t)$}
1499: % \psfrag{v(dt)}{$v(\delta t)$}
1500: % \psfrag{e(N)}{$e(N)$}
1501: % \psfrag{v(N)}{$v(N)$}
1502: % \psfrag{e(Nu)}{$e(\nu)$}
1503: % \psfrag{v(Nu)}{$v(\nu)$}
1504: % \includegraphics[height=10cm]{fig.eps}
1505: % \caption{Expectation and variance of the error when \eqref{eds} is
1506: %   discretized according to the method 
1507: %   described in Appendix (solid curve) and according to the
1508: %   scheme~(\ref{schemaurel}) (dotted curve).}\label{fig}
1509: % \end{center}
1510: % \end{figure}
1511: 
1512: 
1513: We represent on Figure \ref{fig}, the expectation $e$ and the variance
1514: $v$ of the error : $\left|E_{\rm DMC}^{N,\nu,T / (\nu \delta t)}(T)-E_{\rm DMC}(T) \right|$ as a function of the  
1515: number of walkers $N$, the time step $\delta t$ and the number of
1516: reconfigurations $\nu-1$, where $E_{\rm DMC}(T)$ is approximated
1517: using~(\ref{eq:E2}) and $E_{\rm DMC}^{N,\nu,T / (\nu \delta t)}(T)$ is defined
1518: by~(\ref{expr2}). The multinomial resampling method (which is (S1) with
1519:   $\epsilon_n=0$) was used.
1520: 
1521: The top figures represent the expectation of the error and its  variance 
1522: according to the number of walkers. 
1523: To compute these quantities, we perform $2000$ independent realizations,
1524:  with the  number of reconfigurations 
1525: $\nu -1=50$, a small time step $\delta t=5.10^{-3}$ and  $\theta=0.5$.
1526: The simulations confirm the theoretical result : the
1527: error decreases as $C/\sqrt{N}$.
1528: 
1529: The effect of the time step is shown on the two figures in the center.
1530: The numerical parameters are: a large number of particles $N=5000$, number of configurations
1531:  $\nu -1=30$, $\theta=2$ and $300$ independent realizations. We can  see on the figure on the left that the  error 
1532: decreases linearly as the time step decreases. We also remark that the
1533: error is smaller with the approximate scheme~(\ref{schemaurel}) than when using
1534: the exact simulation of the SDE~(\ref{eds}) proposed in the
1535: Appendix. This rather amazing result can be interpreted as follows. When
1536: using the exact simulation of the SDE, there is only one source of error
1537: related to the time discretization, namely the approximation of the
1538: integral in the exponential factor in~(\ref{eq:EDMC}). When using the
1539: scheme~(\ref{schemaurel}), we add a weak error term which seems to
1540: partly compensate the previous one.
1541: 
1542: The last figures represent the effect of the number of
1543:   reconfiguration steps. The numerical parameters are: time step $\delta
1544:   t=5.10^{-3}$, number of particles $N=5000$, $\theta=2$ and $300$ independent realizations.
1545: The curve representing the variation of the error according to the number of
1546:  reconfigurations has the shape of a basin.
1547: We deduce that on the one hand a small number of reconfigurations has 
1548: the disadvantage that walkers with increasingly differing weights are kept. On the other hand 
1549: a large number of reconfigurations introduces much noise. An optimal
1550: number of reconfiguration seems to lie between 20 and 50.
1551: 
1552: \subsubsection{Optimal number of reconfigurations}\label{sec:optim_var}
1553: 
1554: On Figure~\ref{fig:var_EDMC}, we check that the
1555: optimal number of reconfigurations in terms of the variance $\tilde{v}$
1556: of $E_{\rm DMC}^{N,\nu,T / (\nu \delta t)}(T)$ (and not of the error as
1557: in Section~\ref{sec:err_var}) is also obtained
1558: for a number of reconfiguration which seems to lie between 20 and 50
1559: (using again the multinomial resampling method). The numerical
1560: parameters are those considered for the figures below in
1561: Figure~\ref{fig}: time step $\delta
1562:   t=5.10^{-3}$, number of particles $N=5000$, $\theta=2$ and $300$
1563:   independent realizations. We have not studied how the optimal number of reconfigurations varies according to the other
1564: numerical parameters.
1565: 
1566: \begin{figure}[ht]
1567: \begin{center}
1568: \psfrag{var(EDMC)}{$\tilde{v}(\nu)$}
1569: \includegraphics[height=6cm]{var_edmc.eps}
1570: \caption{Variance of $E_{\rm DMC}^{N,\nu,T / (\nu \delta t)}(T)$ in function of the number of
1571:   reconfigurations when \eqref{eds} is
1572:   discretized according to the method 
1573:   described in Appendix (solid curve) and according to the
1574:   scheme~(\ref{schemaurel}) (dashed curve).}\label{fig:var_EDMC}
1575: \end{center}
1576: \end{figure}
1577: 
1578: We have investigated a practical method to estimate numerically the optimal number of 
1579: reconfigurations. On Figure~\ref{Nreconf} we represent the variance of
1580: $E_{\rm DMC}^{N,1,t / \delta t}(t)$ according 
1581: to time $t$, without any reconfiguration step (which corresponds to $\nu=1$). The other numerical
1582: parameters are again those considered for the figures below in
1583: Figure~\ref{fig}. We observe that the variance is minimal at $t^*
1584: \approx 0.25 $. We remark that $\nu =
1585: T/t^*= 20$ is close to the optimal number of reconfigurations 
1586: obtained on the previous figures. We have checked this empirical result
1587: for various sets of the parameters. 
1588: It seems that the optimal
1589: number of reconfigurations is related to $T/t^*$ where $t^*$ minimizes
1590: the variance of $E_{\rm DMC}^{N,1,t / \delta t}(t)$. Since $\nu=1$, no
1591: selection step occurs and the particles are thus
1592: independent. According to the multidimensional
1593: central limit theorem, the variance of $E_{\rm DMC}^{N,1,t / \delta
1594:   t}(t)$ can be approximated by
1595: $$\frac{1}{N}\left( \frac{{\rm Var}(Y_t)}{ (\E(Z_t))^2} - 2 \E(Y_t) \frac{{\rm
1596:        Covar}(Y_t,Z_t)}{(\E(Z_t))^3} + (\E(Y_t))^2 \frac{{\rm Var}(Z_t)}{(\E(Z_t))^4}\right)$$
1597: where
1598: $$Y_t=E_L(X_t) \exp\left(-\dt \sum_{k=1}^{t /\dt} E_L(X_{k \dt}) \right)$$
1599: and
1600: $$Z_t=\exp\left(-\dt \sum_{k=1}^{t /\dt} E_L(X_{k \dt}) \right).$$
1601: 
1602:  Therefore, the optimal number of reconfiguration steps
1603: could be estimated by this method, through a precomputation over a
1604: few independent trajectories.
1605: 
1606: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1607: % Details pour le calcul de la variance asymptotique
1608: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1609: % \begin{alem}\label{lem:var}
1610: % Let us consider two sequences of random variables $(X_i)_{i
1611: %   \ge 1}$ and $(Y_i)_{i \ge 1}$ such that the $X_i$ (resp. the $Y_i$)
1612: % are i.i.d. and square integrable. Then,
1613: % $$\lim_{N \to \infty} N \, {\rm Var}\left(\frac{\sum_{i=1}^N X_i}{\sum_{i=1}^N
1614: %     Y_i} \right) = \frac{{\rm Var}(X_1)}{ \E(Y_1)^2} - 2 \E(X_1) \frac{{\rm
1615: %       Covar}(X_1,Y_1)}{\E(Y_1)^3} + \E(X_1)^2 \frac{{\rm Var}(Y_1)}{\E(Y_1)^4}.$$
1616: % \end{alem}
1617: % \begin{proof}
1618: % Observe that
1619: % \begin{equation}\label{eq:TCL1}
1620: % \sqrt{N} \left( \frac{\sum_{i=1}^N X_i}{\sum_{i=1}^N Y_i} -
1621: %   \frac{\E(X_1)}{\E(Y_1)}\right) 
1622: %  =\sqrt{N} \left( \frac{\frac{1}{N}\sum_{i=1}^N X_i -
1623: %     \E(X_1)}{\frac{1}{N} \sum_{i=1}^N Y_i} \right) + \E(X_1) \sqrt{N}
1624: % \frac{\E(Y_1)-\frac{1}{N} \sum_{i=1}^N Y_i}{\E(Y_1) \frac{1}{N} \sum_{i=1}^N Y_i}
1625: % \end{equation}
1626: % and then apply the multi-dimensional central limit theorem and Slutsky's
1627: % Lemma to obtain that
1628: % \begin{align}
1629: % &\lim_{N \to \infty}\frac{\sqrt{N}}{\frac{1}{N} \sum_{i=1}^N Y_i}
1630: %  \left(\begin{array}{c} \frac{1}{N}\sum_{i=1}^N X_i -
1631: %      \E(X_1) \\ \left(\E(Y_1)-\frac{1}{N} \sum_{i=1}^N Y_i\right) \Big/
1632: %      \E(Y_1) \end{array}\right) \\
1633: % &\qquad \stackrel{{\mathcal
1634: %   L}}{=} {\mathcal N}\left(
1635: % \left(\begin{array}{c}0\\0\end{array}\right),
1636: % \left[\begin{array}{cc}{\rm Var}(X_1) / \E(Y_1)^2 & - {\rm
1637: %       Covar}(X_1,Y_1)/ \E(Y_1)^3\\- {\rm
1638: %       Covar}(X_1,Y_1)/ \E(Y_1)^3 & {\rm Var}(Y_1) / \E(Y_1)^4
1639: %   \end{array}\right] \right).\label{eq:TCL2}
1640: % \end{align}
1641: % The result can now be deduced from~(\ref{eq:TCL1})--(\ref{eq:TCL2}) and the fact that $$N \, {\rm Var}\left(\frac{\sum_{i=1}^N X_i}{\sum_{i=1}^N
1642: %     Y_i} \right) = {\rm Var} \left( \sqrt{N} \left( \frac{\sum_{i=1}^N
1643: %       X_i}{\sum_{i=1}^N Y_i} - \frac{\E(X_1)}{\E(Y_1)}\right)\right).$$
1644: % \end{proof}
1645: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1646: 
1647: \begin{figure}[ht]
1648: \begin{center}
1649: \psfrag{var EDMC(t)}{$\tilde{v}(t)$}
1650: \includegraphics[height=6cm]{var.eps}
1651: \caption{Variance of $E_{\rm DMC}^{N,1,t / \delta t}(t)$ as a function of time $t$.}\label{Nreconf}
1652: \end{center}
1653: \end{figure}  
1654: 
1655:  \subsubsection{Comparison of the resampling algorithms}
1656: 
1657: \begin{figure}[ht]
1658: \begin{center}
1659: \begin{tabular}{cc}
1660:       \includegraphics[width=8cm]{methode2reconfiguration.eps} &
1661:       \includegraphics[width=8cm]{zoomSur2_4.eps}
1662: \end{tabular}
1663: \caption{Variance of $E_{\rm DMC}^{N,\nu,T / (\nu \delta t)}(t)$ as a
1664:   function of time $t$, for various resampling algorithms: Without =
1665:   without resampling, Mult = multinomial resampling, CMult = correlated
1666:   multinomial resampling, Res = residual resampling, Strat = stratified
1667:   resampling, StratRem = stratified remainder resampling, Syst = systematic resampling.}\label{fig:var_resamp}
1668:   \end{center}
1669: \end{figure}
1670: 
1671: We finally compare various resampling algorithms on
1672: Figure~\ref{fig:var_resamp}, where the variance of $E_{\rm DMC}^{N,\nu,T / (\nu \delta t)}(t)$ as a
1673:   function of time is represented. The numerical parameters are:
1674:   $N=1000$, $\dt=5.10^{-3}$, $\nu - 1 =20$, $\theta=2$ and 200
1675:   independent realizations.
1676: 
1677: We first observe on the figure on the
1678:   left that without any resampling, the variance of the results explodes
1679:   with increasing time. This shows the necessity to use resampling
1680:   algorithms. We compare the following resampling algorithms: multinomial resampling (which is (S1) with
1681:   $\epsilon_n=0$), correlated multinomial resampling (which is (S1) with
1682:   $\epsilon_n=1 / \max_{1\leq i\leq N}g(\xi_n^i)$), residual resampling
1683:   (which is (S2)), stratified resampling (which is (S3)),
1684:   stratified remainder resampling (which combines residual and
1685:   stratified resampling, see Remark~\ref{rem:SRR}) and
1686:   systematic resampling (which corresponds to stratified resampling with
1687:   $U^1_n= \ldots=U^N_n=U_n$, see the Introduction). We observe that, as
1688:   expected, when more correlation is introduced, the variance due to the
1689:   resampling is reduced. The multinomial resampling method is generally
1690:   the worse, while the best resampling methods seem to be systematic
1691:   resampling or stratified remainder resampling.
1692: 
1693: \section*{Conclusion}
1694: 
1695: In this paper, we have proved on a simple example convergence of
1696: numerical implementations of the DMC method with a fixed number of
1697: walkers. The theoretical rates of convergence are confirmed by
1698: numerical experiments and are likely to hold in more general situations. We
1699: have also checked numerically the existence of an optimal number of
1700: reconfiguration steps. Various resampling algorithms have been
1701: considered, both theoretically and numerically. For future work, we plan to investigate criteria
1702: devoted to the choice of the number of reconfiguration steps. One
1703: interesting direction is the use of automatic criteria based on a
1704: measure of the discrepancy between the weights carried by the walkers to
1705: decide when to perform a reconfiguration step.
1706: 
1707: % Future works: proof of \eqref{discretaurel}, reconfiguration en fonction de
1708: % discrepancy between weights, comparison with the continuous in time
1709: % version~\cite{Rousset}.
1710: \section*{Appendix : Simulation of the stochastic differential equation
1711:   \eqref{eds}}
1712: In this appendix, we show that it is possible to simulate exactly in law
1713: the $(K+1)$-plet
1714: $(X_0,X_{\dt},\hdots,X_{K \dt})$, where $X_t$ is defined by~(\ref{eds}). Let $(G,U)$ denote a couple of independent random variables with $G$
1715: normal and
1716: $U$ uniformly distributed on the interval $[0,1]$.
1717: \subsection*{Simulation of the increment $X_t-X_s$, for $t \geq s$.}
1718: The square $R_t$ of the norm of a $3$-dimensional Brownian motion $\uu{W}_t=\left(\uu{W}^1_t,\uu{W}^2_t,\uu{W}^3_t\right)$ solves
1719: $dR_t=3dt+2\sqrt{R_t}dB_t$ where $\ds{B_t=\int_0^t \frac{\uu{W}_s \cdot
1720:   d\uu{W}_s}{\|\uu{W}_s \|}}$ is a one-dimensional Brownian
1721: motion. Hence $\ds{\rho_t=\frac{R_t}{1+2\omega t}}$ solves
1722: \begin{equation}
1723:    d\rho_t=(3-2\omega \rho_t)\frac{dt}{1+2\omega
1724:   t}+2\sqrt{\rho_t}\frac{dB_t}{\sqrt{1+2\omega t}}.\label{sqbes}
1725: \end{equation}
1726: It is easy to check that $\left(\int_0^{\frac{1}{2\omega}(e^{2\omega
1727:   t}-1)}\frac{dB_s}{\sqrt{1+2\omega s}}\right)_t$ is a Brownian
1728:   motion. Hence, performing a time-change in \eqref{sqbes}, one obtains that $\rho_{\frac{1}{2\omega}(e^{2\omega
1729:   t}-1)}=e^{-2\omega t}R_{\frac{1}{2\omega}(e^{2\omega
1730:   t}-1)}$ is a weak solution of the equation $dY_t=(3-2\omega
1731:   Y_t)dt+2\sqrt{Y_t}\;dW_t$ satisfied by $Y_t=X_t^2$. Therefore $e^{-\omega t}\sqrt{R_{\frac{1}{2\omega}(e^{2\omega
1732:   t}-1)}}$ is a weak solution of \eqref{eds}.\\
1733: For $v\geq u$, $R_v$ has the same distribution as $\left(\sqrt{R_u}
1734:   +\uu{W}^1_v -\uu{W}^1_u  \right)^2 + (\uu{W}^2_v -\uu{W}^2_u)^2 +
1735: (\uu{W}^3_v -\uu{W}^3_u)^2$, and therefore as
1736:   $(\sqrt{R_u}+G\sqrt{v-u})^2-2(v-u)\log(U)$ with
1737:   $(G,U)$ independent from $R_u$. Hence for $t\geq s$, $X_{t}$ has the same
1738:   distribution as
1739: \begin{align*}&\left(e^{-2\omega t}\left(\left(e^{\omega
1740:       s}X_s+\frac{G}{\sqrt{2\omega}}(e^{2\omega t}-e^{2\omega
1741:       s})^{1/2}\right)^2-2\frac{1}{2\omega}(e^{2\omega t}-e^{2\omega
1742:       s})\log(U)\right)\right)^{1/2}\\
1743: &=\left(\left(e^{-\omega
1744:       (t-s)}X_s+\frac{G}{\sqrt{2\omega}}(1-e^{-2\omega(t-s)})^{1/2}\right)^2-\frac{1}{\omega}(1-e^{-2\omega
1745:       (t-s)})\log(U)\right)^{1/2}\end{align*}
1746: where the couple $(G,U)$ is
1747:       independent from $X_s$.
1748: \subsection*{Simulation of $X_0$ with distribution $2\psi_I^2(x)1_{\{x>0\}}dx$.}
1749: The random variable
1750: $\frac{1}{\sqrt{2\omega}}\left(G^2-2\log(U)\right)^{1/2}$ is
1751: distributed according to the invariant measure
1752: $2\psi_I^2(x)1_{\{x>0\}}dx$, as suggested by letting the time increment $t-s$ tend to $+\infty$ in the previous
1753: simulation. Indeed, $G^2-2\log(U)$ is a Gamma random
1754: variable with density $\frac{1}{2^{3/2}\Gamma(3/2)}1_{\{z>0\}}
1755: \sqrt{z}e^{-z/2}$. And one deduces the
1756: density of $\frac{1}{\sqrt{2\omega}}\left(G^2-2\log(U)\right)^{1/2}$ by an easy change of variables.
1757: 
1758: \begin{thebibliography}{99}
1759: 
1760:    \bibitem{Alfonsi} A. Alfonsi, {\em  On the discretization schemes for
1761:     the CIR (and Bessel squared) processes}, Monte Carlo Methods
1762:   Appl. 11(4), 355--384, (2005).
1763: 
1764: \bibitem{Caffarel} R. Assaraf, M. Caffarel and A. Khelif, {\em Diffusion
1765:     Monte Carlo with a fixed number of walkers}, Phys. Rev. E 61,
1766:   4566--4575, (2000). 
1767: 
1768: \bibitem{CJL} E. Cancès, B. Jourdain and T. Lelièvre, {\em Quantum Monte
1769:     Carlo simulations of fermions. A mathematical analysis of the
1770:     fixed-node approximation}, Mathematical Models and Methods in
1771:   Applied Sciences 16(9), 1403--1440, (2006).
1772: 
1773: 
1774: \bibitem{handbook} E.  Canc\`es, M. Defranceschi, 
1775: W. Kutzelnigg, C. Le Bris, and Y. Maday,  {\em Computational Quantum 
1776:   Chemistry: a Primer}, in: Handbook of 
1777: Numerical Analysis, Special volume, Computational Chemistry, volume X, 
1778:  Ph. G. Ciarlet and C. Le Bris (eds), 
1779:  North-Holland, 3--270, (2003). 
1780: 
1781: \bibitem{CDM05} O. Cappé, R. Douc and E. Moulines, {\it Comparison of Resampling Schemes for Particle Filtering}, In 4th International Symposium on Image and Signal Processing and Analysis (ISPA), Zagreb, Croatia, (2005).
1782:   
1783: \bibitem{C04} N. Chopin, {\em Central limit theorem for sequential Monte Carlo
1784: methods and its application to Bayesian inference}, Ann.
1785: Statist. 32(6), 2385--2411, (2004).
1786: 
1787: \bibitem{Delmoral} P. Del Moral, {\em Feynman-Kac Formulae :
1788:     Genealogical and Interacting Particle Systems with Applications},
1789:     Springer-Verlag (2004).
1790: 
1791: \bibitem{DD04} P. Del Moral and A. Doucet, {\em Particle motions in
1792:     absorbing medium with hard and soft obstacles}, Stochastic
1793:   Anal. Appl. 22(5), 1175--1207, (2004).
1794: 
1795: \bibitem{Delmoralmiclo} P. Del Moral and L. Miclo, {\em
1796:     Branching and Interacting Particle Systems. Approximation of
1797:     Feynman-Kac Formulae with Applications to Non-Linear Filtering}, in
1798:     S\'eminaire de Probabilit\'es XXXIV, Lecture Notes in Mathematics 1729,
1799:     1--145, Springer-Verlag (2000).
1800: 
1801: \bibitem{DM03} P. Del Moral and L. Miclo, {\em Particle approximations
1802:     of {L}yapunov exponents connected to {S}chr\"odinger operators and
1803:     {F}eynman-{K}ac semigroups}, ESAIM Probab. Stat. 7, 171--208, (2003).
1804: 
1805: \bibitem{Glasserman} P. Glasserman, {\em Monte Carlo methods in
1806:   financial engineering}, Springer-Verlag (2004).
1807: 
1808: \bibitem{H84} J. H. Hetherington, {\em Observations on the statistical
1809:     iteration of matrices}, Phys. Rev. A 30(5), 2713--2719, (1984).
1810: 
1811: \bibitem{RCAL82}  P.J. Reynolds, D.M. Ceperley, B.J. Alder and
1812:   W.A. Lester,  {\em Fixed-node quantum Monte Carlo for molecules},
1813:   J. Chem. Phys. 77(11), 5593--5603, (1982).
1814:  
1815: \bibitem{Rousset} M. Rousset, {\em On the Control of an Interacting
1816:     Particle Approximation of Schr\"odinger Groundstates},
1817: SIAM J. Math. Anal., 38(3), 824--844, (2006).
1818: 
1819: 
1820: \bibitem{Sorella} S. Sorella, {\em Green Function Monte Carlo with
1821:     Stochastic Reconfiguration}, Phys. Rev. Lett. 80(20), 4558--4561,
1822:   (1998).
1823: 
1824: 
1825: \bibitem{Talay-Tubaro} D. Talay and L. Tubaro, {\em Expansion of the
1826:     global error for numerical schemes solving stochastic differential
1827:     equations}, Stochastic Analysis and Applications 8(4), 94--120, (1990).
1828: 
1829: \bibitem{Umrigar} C.J. Umrigar, M.P. Nightingale and K.J. Runge, {\em A
1830:     Diffusion Monte Carlo algorithm with very small time-step errors},
1831:   J. Chem. Phys. 99(4), 2865--2890, (1993).
1832: 
1833: \end{thebibliography}
1834: \end{document}
1835: 
1836: