1: %\documentclass[aps,preprint,prl,nofootinbib,endfloats,showpacs]{revtex4}
2: \documentclass[aps,preprint,nofootinbib,endfloats,showpacs]{revtex4}
3:
4: \usepackage{graphics}
5: \usepackage{epsfig}
6:
7: %\catcode`\@=11
8:
9:
10: \begin{document}
11:
12: \title{The Fermion Monte Carlo revisited}
13:
14: \author{Roland Assaraf$^1$, Michel Caffarel$^2$, and Anatole Khelif$^3$}
15: \affiliation{
16: $^{1}$Laboratoire de Chimie Th{\'e}orique CNRS-UMR 7616, Universit{\'e} Pierre
17: et Marie Curie, 4 Place Jussieu, 75252 Paris, France \\
18: $^{2}$Laboratoire de Chimie et Physique Quantiques CNRS-UMR 5626, IRSAMC, Universit\'e Paul
19: Sabatier, 118 route de Narbonne 31062 Toulouse Cedex, France \\
20: $^{3}$Laboratoire de Logique Math\'ematique, Universit{\'e} Denis Diderot,
21: 2 Place Jussieu, 75251 Paris, France}
22:
23: \date{\today}
24:
25: \begin{abstract}
26: In this work we present a detailed study of the Fermion Monte Carlo algorithm
27: (FMC), a recently proposed stochastic method for calculating fermionic
28: ground-state energies. A proof that the FMC method is an exact method is given.
29: In this work the stability of the method is related to the difference
30: between the lowest (bosonic-type) eigenvalue of the FMC diffusion operator and the
31: exact fermi energy. It is shown that within a FMC framework
32: the lowest eigenvalue of the new diffusion operator is no longer the bosonic
33: ground-state eigenvalue as in standard exact Diffusion Monte Carlo (DMC) schemes
34: but a modified value which is strictly greater.
35: Accordingly, FMC can be viewed as an exact DMC method
36: built from a correlated diffusion process having a {\it reduced} Bose-Fermi gap.
37: As a consequence, the FMC method is more stable than any transient method (or
38: nodal release-type approaches). It is shown that the most recent ingredient
39: of the FMC approach [M.H. Kalos and F. Pederiva, Phys. Rev. Lett.
40: {\bf 85}, 3547 (2000)], namely the introduction of non-symmetric guiding
41: functions, does not necessarily improve the stability of the algorithm.
42: We argue that the stability observed with such
43: guiding functions is in general a finite-size population effect disappearing for a
44: very large population of walkers.
45: The counterpart of this stability is a control population error which is
46: different in nature from the standard Diffusion Monte Carlo algorithm
47: and which is at the origin of an uncontrolled approximation in FMC.
48: We illustrate the various ideas presented in this work with calculations
49: performed on a very simple model having only nine states but a full ``sign problem''.
50: Already for this toy model it is clearly seen that FMC calculations are inherently
51: uncontrolled.
52: \end{abstract}
53:
54: \pacs{ 02.70Ss, 05.30.Fk }
55: \maketitle
56:
57: \section{Introduction}
58: In theory quantum Monte Carlo (QMC) techniques\cite{kalosfermion99}
59: are capable of giving an exact estimate of the energy with an evaluation of the error:
60: the statistical error. Unfortunately, such an ideal situation is not realized in practice.
61: Exact results with a controlled finite statistical error are only achieved for bosonic systems.
62: For fermionic systems, we do not have at our disposal an algorithm which is both exact
63: and stable (statistical fluctuations going to zero in the large simulation time regime).
64: This well-known problem is usually referred to as the ``sign problem''.
65: The usual solution to cope with this difficulty consists in defining a stable algorithm
66: based on an uncontrolled approximation, the so-called Fixed-Node approximation.
67: \cite{ceperleykalos79,schmittkalos84,lesterbook94,reynoldsJCP82,andersonJCP75,andersonJCP76}
68: In practice, the fixed-node error on the energy is small when one uses good trial
69: wavefunctions and, thus, QMC methods can be considered today as reference methods to
70: compute groundstate energies as shown by a large variety of
71: applications\cite{reynoldsJCP82,filippiJCP96,surf1,surf2,needs1,needs2,MitasGrossman00}.
72: However, the accuracy of the results is never known from the calculation, it is known only
73: {\it a posteriori}, for example by comparison with experimental data.
74: Exact methods, which are basically transient
75: methods\cite{ceperleyalderPRL84,bernuJCP90,bernuJCP91,caffarelJCP1992}
76: including the nodal release method\cite{ceperleyalderPRL84}, have been applied with success only
77: to very specific models (small or homogeneous systems) for which the sign instability
78: is not too severe (small Bose-Fermi energy gap).
79:
80: Recently, Kalos and coworkers \cite{kalosfermion00} have proposed a method
81: presented as curing the sign problem, the so-called Fermion Monte Carlo method (FMC).
82: This work makes use of two previously introduced ingredients, a cancellation process
83: between ``positive'' and ``negative'' walkers introduced by Arnow {\sl et al.} \cite{arnow1982}
84: and a modified process correlating explicitly the dynamics of the walkers
85: of different signs.\cite{Liu1994}
86: The new feature introduced in \cite{kalosfermion00} is the introduction of non-symmetric
87: guiding functions. The method has been tested on various simple systems including free fermions
88: and interacting systems
89: such as the $^3$He fluid.\cite{kalosfermion00,CollettiFMC2005}
90: The results are found to be compatible with the
91: assumption that the method is stable and not biased. However, this conclusion is not
92: clear at all because of the presence of large error bars.
93: The purpose of this work is to present a detailed analysis of the
94: algorithm and a definitive answer to this assumption.
95:
96: The content of this paper is as follows.
97: In section II we give a brief presentation of the ``sign problem''.
98: The sign instability in an exact DMC approach comes from the blowing
99: up in imaginary time of the undesirable bosonic component associated with the
100: lowest mathematical eigenstate of the Hamiltonian (a wavefunction which is
101: positive and symmetric with respect to the exchange of particles).
102: The fluctuations of the transient energy estimator grows like $e^{t(E_0^F-E_0^B)}$,
103: where $E_0^F$ is the fermionic ground-state energy (here and in what follows, the superscript
104: $F$ stands for ``Fermionic'') and $E_0^B$, the lowest mathematical eigenvalue
105: of the Hamiltonian (the superscript ``B'' standing for Bosonic).
106: In Sec. III we briefly recall the main elements of the standard DMC
107: method, the basic stochastic algorithm simulating the imaginary-time
108: evolution of the Hamiltonian.
109: In section IV, we describe the FMC method as a generalization of the DMC
110: method and we show that {\it FMC is an exact method}, that is, no systematic bias is introduced.
111: This section is made of two parts.
112: In a first part we introduce the notion of positive and negative walkers
113: to represent a signed wavefunction in DMC.
114: This will help us to view the FMC method as a generalization of
115: the DMC method, the FMC introducing two important modifications
116: with respect to DMC: a correlated dynamics for positive and negative walkers
117: and a cancellation process for such pairs whenever they meet.
118:
119: In section V we study the stability of the algorithm.
120: We prove that the {\it FMC method is in general not stable}, the fluctuations of the transient
121: estimator of the energy growing exponentially like $e^{t(E_0^F-\tilde{E}_0^B)}$
122: where $\tilde{E}_0^B$ is the lowest bosonic-like eigenvalue
123: of the generalized diffusion process operator
124: associated with FMC whose expression is given explicitly.
125: It is shown that {\it FMC is more stable than the standard nodal release DMC method}
126: because $\tilde{E}_0^B > E_0^B$.
127: In section VI we illustrate our theoretical results using a toy model,
128: a ``minimal'' quantum system having a genuine sign problem (two coupled
129: oscillators on a finite lattice). The different aspects of the FMC method are
130: highlighted in this application.
131: In the case of non-symmetric wavefunctions introduced recently in Ref.\cite{kalosfermion00}
132: it is shown that, in contrast with DMC where the population control
133: error decays linearly as a function of the population size, the FMC decay displays a much
134: more slower power law.
135: As a consequence, one needs a very large
136: population of walkers to remove the control population error
137: and to observe the instability of FMC.
138: Let us emphasize that observing such a subtle finite population effect
139: for a genuine many-fermion system is actually impossible. Here, to illustrate
140: this important point numerically, we have been led to consider a very simple system having
141: only a few states.
142: For this system, a large population of
143: walkers -eventually much larger than the
144: dimension of the quantum Hilbert space itself- can be considered. The results obtained
145: in that regime confirm our theoretical findings, in particular the fact that the stability of the
146: algorithm presented elsewhere \cite{kalosfermion00}
147: is only apparent. As an important conclusion, we emphasize that the FMC control
148: population error is an uncontrolled approximation for realistic fermion systems.
149:
150: \section{Fermion instability in quantum Monte Carlo}
151: We consider a Schr\"odinger operator for a system of $N$ fermions,
152: \begin{equation}
153: H = -\frac{1}{2} {\bf \nabla}^2 + V ({\bf R})
154: \label{schro}
155: \end{equation}
156: where we note ${\bf R}=({\bf r}_1 \dots {\bf r}_N)$ the $3N$ coordinates of the
157: $N$ particles in the three dimensional space. In this expression,
158: the first term is the kinetic energy (${\bf \nabla}^2$ is the Laplacian operator
159: in the space of the $3N$ coordinates). The second term, $V$, is the potential.
160: DMC techniques are based upon the evolution of the Hamiltonian
161: in imaginary time. We express this evolution using the spectral decomposition:
162: \begin{equation}
163: e^{-t(H-E_T)} = \sum_{i} e^{-t (E_i-E_T)} \mid \phi_i \rangle \langle \phi_i \mid
164: \label{evolution}
165: \end{equation}
166: where $\phi_i$ are the (normalized) eigenfunctions of $H$, $E_i$ are the
167: corresponding eigenvalues and $E_T$ is a so-called reference energy.
168: The fundamental property of operator (\ref{evolution}) is to filter
169: out the lowest eigenstate $\phi_0$.
170: To understand how it works, we consider the time evolution of a wavefunction
171: $f_0 ({\bf R})$
172: \begin{equation}
173: \mid f_t \rangle \equiv e^{-t (H-E_T)} \mid f_0 \rangle
174: \label{f_tdef}
175: \end{equation}
176: and calculate the overlap with an eigenstate $\phi_i$
177: \begin{equation}
178: \langle \phi_i \mid f_t \rangle = e^{-t(E_i-E_T)} \langle
179: \phi_i \mid f_0 \rangle.
180: \end{equation}
181: From this expression it is easily seen that the component on the lowest
182: eigenstate $\phi_0$ is growing exponentially faster than the higher components.
183: In DMC methods the function $f_t$ is generated using random
184: walks. For $t$ large enough the lowest eigenstate
185: \begin{equation}
186: f_t \sim e^{-t (E_0-E_T)} \phi_0
187: \label{ft}
188: \end{equation}
189: is produced.
190: This asymptotic behaviour makes DMC an accurate method for
191: computing the properties of $\phi_0$, in particular the energy $E_0$.
192: Unfortunately, for fermionic systems the physical groundstate $\phi_0^F$, which is
193: antisymmetric with respect to the exchange of particles, is not
194: $\phi_0$, but some ``mathematically'' excited state because
195: $\phi_0$ is positive and symmetric for a Schr\"odinger
196: operator (bosonic ground-state). For the sake of clarity,
197: we shall denote from now on this bosonic ground-state as $\phi_0^B$ and its energy, $E_0^B$.
198: The necessity of extracting an exponentially small component to evaluate
199: $E_0^F$ or any fermionic
200: property is at the origin of the fermion instability in quantum Monte Carlo.
201:
202: We now give a quantitative analysis of this instability.
203: Since the asymptotic behaviour of $f_t$ is not useful to compute the fermionic energy,
204: we consider the transient behaviour of the evolution of $f_t$.
205: Basically, exact methods \cite{ceperleyalderPRL84} filter out $\phi_0^F$ by projecting
206: the transient behaviour of $f_t$ on the antisymmetric space.
207: Introducing the antisymmetrization operator ${\cal A}$, one obtains the
208: physical groundstate for $t$ large enough
209: \begin{equation}
210: {\cal A} \mid f_t \rangle \sim e^{-t(E_0^F-E_T)} \phi_0^F
211: \label{aft}
212: \end{equation}
213: since the components of $f_t$ over the higher antisymmetric eigenstates are
214: decreasing exponentially with respect to the component on $\phi_0^F$.
215: In practice, the fermionic energy $E_0^F$ is calculated using an
216: antisymmetric function $\psi_T$ and using the fact that, at large enough $t$, one has
217: \begin{equation}
218: E_0^F = \frac{\langle \psi_T \mid H \mid f_t \rangle}{\langle \psi_T \mid f_t
219: \rangle}.
220: \label{E_Fproj}
221: \end{equation}
222: In the long-time regime the stochastic estimation of the R.H.S. of Eq.(\ref{E_Fproj})
223: is unstable.
224: In essence, the signal, the antisymmetric component of $f_t$,
225: %RA: pourquoi des - et pas des, ? MC: je ne sais pas! je trouve ca plus joli. Bon!
226: mettons des virgules....
227: decreases exponentially fast with respect to $f_t$, Eq.(\ref{ft}). The signal-over-noise
228: ratio behaves like $e^{-t(E_0^F-E_0^B)}$ and, thus, an exponential growth
229: of the fluctuations of the DMC estimator, Eq.(\ref{E_Fproj}), appears.
230:
231: Now, let us give a more quantitative analysis by writing this estimator and
232: evaluating the variance.
233: In a standard diffusion Monte Carlo calculation, the time-dependent distribution
234: $f_t$ is generated by a random walk over a population of walkers $\{{\bf R}_i\}$.
235: Formally $f_t$ is given in the calculation as an average at time $t$ over Dirac functions
236: centered on the walkers and weighted by some positive function $\psi_G$
237: \begin{equation}
238: f_t ({\bf R}) = \frac{1}{\psi_G ({\bf R})} \left\langle \frac{}{} \sum_i
239: \delta ({\bf R}-{\bf R}_i ) \right\rangle
240: \label{sampleft}
241: \end{equation}
242: In this expression, $\langle ... \rangle$ denotes the average over populations of
243: walkers $\{{\bf R}_i\}$ obtained at the given time $t$.
244: The function $\psi_G$ is usually called the importance or guiding function.
245: Replacing $f_t$ in (\ref{E_Fproj}) by its expression (\ref{sampleft}),
246: the estimator of the energy reads for large enough $t$
247: \begin{eqnarray}
248: E_0^F & = &
249: \frac{\left\langle \sum_i \frac{H\psi_T}{\psi_G} ({\bf R}_i) \right\rangle}
250: {\left\langle \sum_i \frac{\psi_T}{\psi_G} ({\bf R}_i)\right\rangle}
251: \label{energystat}
252: \end{eqnarray}
253: In practice, both numerator and denominator are computed as an average
254: over the $N_S$ walkers produced by the algorithm at time $t$.
255: As a consequence, the energy is obtained as
256: \begin{equation}
257: E_0^F = \frac{\cal{N} }{\cal{D}}
258: \equiv \frac{ \frac{1}{N_S}\sum_{i=1}^{N_S} \frac{H\psi_T}{\psi_G} ({\bf
259: R}_i)}{ \frac{1}{N_S} \sum_{i=1}^{N_S} \frac{\psi_T}{\psi_G} ({\bf R}_i)}.
260: \label{statea}
261: \end{equation}
262: The ratio $\frac{\cal{N} }{\cal{D}}$ is an estimator of the energy for $N_S$ large enough,
263: when the numerator and denominator have small fluctuations aroung their average.
264: Now, let us evaluate the fluctuations of this ratio in the limit of a large population $N_S$
265: \begin{equation}
266: \sigma^2 \left( \frac{\cal{N}}{\cal{D}} \right) = \frac{\langle
267: ({\cal N}-E_0^F{\cal D})^2 \rangle}{\langle {\cal D} \rangle^2}.
268: \end{equation}
269: Here, we have used the hypothesis that the fluctuations of the denominator and the
270: numerator are very small.
271: Using the fact that ${\cal N}$ and ${\cal D}$ are statistical averages
272: over independent random variables
273: with the same distribution one obtains
274: \begin{equation}
275: \sigma^2 \left( \frac{\cal{N}}{\cal{D}} \right) = \frac{1}{N_S} \frac{\langle
276: \left[\frac{(H-E_0^F)\psi_T}{\psi_G} ({\bf R}_i)\right]^2 \rangle}{ \langle \frac{\psi_T}{\psi_G} ({\bf R}_i) \rangle^2}.
277: \end{equation}
278:
279: We can replace these averages by integrals over the distribution $f_t \psi_G$,
280: Eq.(\ref{sampleft})
281: \begin{equation}
282: \sigma^2 \left( \frac{\cal{N}}{\cal{D}} \right) = \frac{1}{N_S}
283: \frac{\langle \frac{[(H-E_0^F)\psi_T]^2}{\psi_G} \mid \phi_0^B \rangle
284: \langle \psi_G \mid \phi_0^B \rangle }{ \langle \psi_T \mid \phi_0^F \rangle^2}
285: e^{2t(E_0^F-E_0^B)} \propto \frac{1}{N_S} e^{2t(E_0^F-E_0^B)}
286: \label{sigmaea}
287: \end{equation}
288: which confirms quantitatively that the statistical error grows exponentially
289: in time. Let us emphasize that this problem is particularly severe because the Bose-Fermi
290: energy gap, $\Delta_{B-F} \equiv E_0^F-E_0^B$,
291: usually grows faster than linearly as a function of the number of particles.
292: In practice, one has to find a trade-off between the systematic error coming from
293: short projection times $t$ and the large fluctuations arising at large projection times.
294:
295: \section{The diffusion Monte Carlo method}
296: The FMC method is a generalization of the well-known DMC
297: method. Presenting this algorithm is a useful preparation
298: for the next section about Fermion Monte Carlo.
299: The DMC method generates the function $f_t$ following the imaginary time dymamics of $H$
300: \begin{equation}
301: f_t \equiv e^{-t(H-E_T)} f_0
302: \label{evolt0}
303: \end{equation}
304: where $f_0$ is a positive function.
305: The imaginary time dynamics is produced by iterating many times the short-time
306: Green function $e^{-\tau (H-E_T)}$ where $\tau$ is a small time step.
307: The distribution $f_{t^\prime}$ at the time $t^\prime \equiv t+\tau$ is then obtained from $f_t$ as follows
308: \begin{equation}
309: f_{t^\prime} = e^{-\tau (H-E_T)} f_t
310: \label{shorttime}
311: \end{equation}
312: where the density $f_t$ is sampled by the population of walkers $\{ {\bf R}_i\}$.
313: Using the Dirac ket notation, $f_t$ given by Eq.(\ref{sampleft}) is rewritten as
314: \begin{equation}
315: f_t = \frac{1}{\psi_G} \left\langle \frac{}{} \sum_i \mid {\bf R}_i \rangle \right\rangle
316: \label{def_ft2},
317: \end{equation}
318: where $\psi_G$ is some positive function, the so-called guiding function.
319: Let us show how the density $f_{t^\prime}$ is generated from the distribution (\ref{def_ft2}).
320: Replacing in (\ref{shorttime}) the function $f_t$ by the R.H.S. of (\ref{def_ft2}) one has
321: \begin{eqnarray}
322: f_{t^\prime} & = & e^{-\tau (H-E_T)} \frac{1}{\psi_G}
323: \left\langle \frac{}{} \sum_i \mid {\bf R}_i \rangle \right\rangle \\
324: & = &
325: \left\langle \sum_i e^{-\tau (H-E_T)} \frac{1}{\psi_G} \mid
326: {\bf R}_i \rangle \right\rangle \\
327: & = & \frac{1}{\psi_G} \left\langle \frac{}{} \sum_i e^{\tau L} \mid {\bf R}_i \rangle \right\rangle
328: \label{dmcopap}
329: \end{eqnarray}
330: where we have introduced the operator
331: \begin{eqnarray}
332: L & \equiv & -\psi_G (H-E_T) \frac{1}{\psi_G}.
333: \label{dmcop}
334: \end{eqnarray}
335: For a Schr\"odinger Hamiltonian (\ref{schro}) the operator $L$ takes the form
336: \begin{eqnarray}
337: L & = & -\psi_G (H-E_L) \frac{1}{\psi_G} - (E_L-E_T) \\
338: & = &
339: \frac{1}{2} {\bf \nabla}^2 - {\bf \nabla} [{\bf b}.]
340: \label{fkp} \\
341: & & - (E_L -E_T)
342: \label{branching}
343: \end{eqnarray}
344: where we have introduced the so-called drift vector
345: \begin{equation}
346: {\bf b} \equiv \frac { {\bf \nabla} \psi_G}{\psi_G}
347: \end{equation}
348: and the local energy of the guiding function $\psi_G$,
349: \begin{equation}
350: E_L \equiv \frac{H \psi_G}{\psi_G}.
351: \end{equation}
352: The operator $L$ is the sum of the so-called Fokker Planck operator (\ref{fkp}) and a local
353: operator (\ref{branching}). Using this decomposition,
354: the vector $e^{\tau L} \mid {\bf R}_i \rangle$ appearing in the
355: average (\ref{dmcopap}) can be rewritten for small enough time step $\tau$ as follows
356: \begin{eqnarray}
357: e^{\tau L} \mid {\bf R}_i \rangle & = &
358: e^{\tau \left[ \frac{1}{2} {\bf \nabla}^2 - {\bf \nabla} [{\bf b}.] \right]}
359: \label{fkp_ap}
360: \\
361: & & \times e^{-\tau (E_L-E_T)} \mid {\bf R}_i \rangle.
362: \label{branch_ap}
363: \end{eqnarray}
364: The action of $e^{\tau L}$ on $ \mid {\bf R}_i \rangle $ is sampled as follows.
365: The short-time dynamics of the Fokker Planck operator (\ref{fkp_ap}) is performed by the way of
366: a Langevin process,
367: \begin{equation}
368: {{R_i}^\prime}^\mu = {R_i}^\mu + b_i^\mu \tau + \sqrt{\tau} \eta_i^\mu
369: \label{lang}
370: \end{equation}
371: where $\mu$ runs over the $3N$ coordinates (three space coordinate for each
372: fermion), and $\eta_i^\mu$ are independent gaussian random variables centered and normalized
373: \begin{equation}
374: \langle \eta_i^\mu \eta_i^\nu \rangle =\delta^{\mu\nu}
375: \end{equation}
376: The averaged Langevin process (\ref{lang}) is equivalent to apply
377: the short-time dynamics of the Fokker Planck operator (\ref{fkp_ap}):
378: \begin{equation}
379: \left\langle \frac{}{} \mid {\bf R}_i^\prime \rangle \right\rangle =e^{\tau \left[ \frac{1}{2}
380: {\bf \nabla}^2 - {\bf \nabla} [{\bf b}.] \right]} \mid {\bf R}_i \rangle
381: \end{equation}
382: The factor
383: \begin{equation}
384: w_i \equiv e^{-\tau (E_L ({\bf R_i})-E_T)}
385: \label{branching2}
386: \end{equation}
387: being a normalization term, called the
388: branching term. The new walker ${\bf R}_i^\prime$ is duplicated (branched)
389: a number of times equal to $w_i$ in average.
390: This process is a birth-death process since some walkers can be duplicated
391: and some can be removed. The population of walkers
392: fluctuating, one has to resort to control
393: population techniques. \cite{umrigarJCP93,sorella98rec,khelif2000}
394: With these two processes, diffusion and branching
395: a new population of walkers $\{{\bf R}_i^\prime\}$ is produced, representing in
396: average the desired result :
397: \begin{equation}
398: \left\langle \frac{}{} \sum_i \mid {\bf R}_i^\prime \rangle \right\rangle
399: = \left\langle \sum_i e^{\tau L} \mid {\bf R}_i \rangle \right\rangle
400: \label{riprim}
401: \end{equation}
402: From (\ref{dmcopap}) and (\ref{riprim}) one can see that
403: the distribution $f_{t^\prime}$ is sampled by the new population of
404: walkers $\{{\bf R}_i^\prime\}$ according to (\ref{def_ft2})
405: \begin{equation}
406: f_{t^\prime} = \left\langle \frac{1}{\psi_G} \mid {\bf R}_i^\prime) \right\rangle.
407: \label{ft_samplep}
408: \end{equation}
409: In summary, by iterating these two simple operations, namely the Langevin and
410: branching processes, the DMC method allows to simulate the imaginary time dynamics
411: of the Hamiltonian, thus producing a sample of $f_t$, Eq.(\ref{def_ft2}).
412: Various properties of the system can be computed from this sample, e.g.
413: ground-state bosonic energies,\cite{klv,ceperleykalos79,schmittkalos84}
414: excited-state energies,\cite{bernuJCP91,caffarelJCP1992} and various observables.
415: \cite{rothstein97,Rothstein2004,force00,filippi2000,caffrerat,rcaf03,casalegno2003}
416:
417: For the vast majority of the DMC simulations on fermionic systems, only an
418: approximation of the exact
419: fermionic ground-state energy is computed, namely the so-called Fixed-Node energy.
420: \cite{schmittkalos84,reynoldsJCP82,filippiJCP96,ceperleyalder86,
421: ceperleyalderPRL80,Grossman2002}
422: In a fixed-node DMC calculation the guiding function is chosen as
423: $\psi_G=|\psi_T|$ where $\psi_T$ is some fermionic antisymmetric trial wavefunction.
424: With this choice, the guiding function
425: vanishes at the nodes (zeroes) of the trial wavefunction and the drift vector diverges.
426: As a consequence, the walkers cannot cross the nodes of $\psi_T$
427: and are confined within the nodal regions of the configuration space.
428: It can be shown that the resulting DMC stationary state is the best variational
429: solution having the same nodes as $\psi_T$. In other words, the ``Fixed-Node'' energy obtained
430: from the R.H.S of (\ref{energystat}) or (\ref{statea})
431: is an upper bound of the exact fermi energy,
432: $E_0^{FN}>E_0^F$.\cite{ceperleyalderPRL80,ceperleyalder86}
433: Note that, in practice, $E_0^{FN}$ is in general a good approximation of
434: the true energy. \cite{filippiJCP96,Grossman2002}
435:
436: In the present work, we are considering ``exact'' DMC approaches for which the exact
437: fermionic energy calculated from expression (\ref{energystat}) is searched for.
438: As discussed in the previous section, such exact DMC calculations are fundamentally unstable.
439: A famous example of an exact DMC approach is the
440: nodal release method of Ceperley and Alder.\cite{ceperleyalderPRL80,ceperleyalderPRL84}
441: Basically, nodal release methods are standard DMC methods where
442: the fixed-node distribution (sampled with a standard Fixed-Node
443: DMC) is chosen as initial distribution $f_0$.
444: In exact methods the guiding function $\psi_G$ is strictly positive
445: everywhere, so that the walkers can cross the nodes of the trial wavefunction.
446: Exact fermi methods are efficient in practice only when
447: the convergence of the estimator is fast enough, that is, when it occurs before
448: the blowing up of fluctuations, Eq. (\ref{sigmaea}).
449: In practice, two conditions are to be satisfied.
450: First, the fixed-node wavefunction
451: $\phi_0^{FN}$ must be already close enough to the exact solution $\phi_0^F$. For this
452: reason the choice of the trial function $\psi_T$ (quality of the nodes of $\psi_T$)
453: is crucial.
454: Second, the Bose-Fermi gap, $\Delta_{B-F} = E_0^F-E_0^B$,
455: which drives the asymptotic behaviour of the
456: fluctuations, Eq. (\ref{sigmaea}), must be small.
457: The quantity $E_0^F-E_0^B$ depends only on the Hamiltonian at hand; there is
458: no freedom in the nodal release method to modify the asymptotic behaviour.
459: We will define in the next section the Fermion Monte Carlo method (FMC) as
460: a generalization of the DMC method and show in sections V and VI
461: that, in contrast with the nodal release method, the FMC method can improve
462: substantially the asymptotic behaviour, Eq.(\ref{sigmaea}).
463:
464: \section{The FMC method}
465: \subsection{ Preliminary: Introducing positive and negative walkers in DMC}
466: In Fermion Monte Carlo a dynamics on a signed function $f_t$ is performed.
467: In what follows, we show that DMC can be easily generalized to the case of
468: a signed distribution $f_t$ and, thus, FMC can be viewed as
469: %This formulation has usually no interest if we are interested in the
470: %asymptotic behaviour of $f_t$, which is a positive function,
471: %namely the lowest eigenstate of $H$.
472: %Here this formulation may have an interest because we are looking at the transient behaviour.
473: a simple generalization of DMC.
474: %End of RA
475: If $f_t$ carries a sign, it can be written as the difference of two positive functions
476: \begin{equation}
477: f_t = f_t^+-f_t^-
478: \label{f_tdif}
479: \end{equation}
480: both satisfying the following equations of evolution
481: \begin{eqnarray}
482: f_t^+ & = & e^{-t(H-E_T)} f_0^+ \\
483: f_t^- & = & e^{-t(H-E_T)} f_0^-
484: \end{eqnarray}
485: To sample these expressions, two independent DMC calculations
486: can be carried out. The positive part $f_t^+$ is then sampled by a population of
487: walkers $\{{\bf R_i}^+\}$ (called ``positive'' walkers). The distribution $f_t^+$
488: is related to $\{{\bf R_i}^+\}$ as in Eq.(\ref{def_ft2})
489: \begin{equation}
490: f_t^+ = \frac{1}{\psi_G^+} \left\langle \frac{}{} \sum_i \mid {\bf R}_i^+ \rangle \right\rangle
491: \label{sampleft+}
492: \end{equation}
493: and the negative part is similarly sampled by a population of
494: ``negative'' walkers $\{{\bf R_i}^-\}$
495: \begin{equation}
496: f_t^- = \frac{1}{\psi_G^-} \left\langle \frac{}{} \sum_i \mid {\bf R}_i^- \rangle \right\rangle.
497: \label{sampleft-}
498: \end{equation}
499: Note that we consider here the general case
500: where the guiding functions associated with the positive and negative walkers,
501: $\psi_G^+$ and $\psi_G^-$, are different.
502: Finally, the dynamics of positive and negative walkers is described by the DMC-like diffusion
503: operators
504: \begin{eqnarray}
505: L^\pm & \equiv & -\psi_G^{\pm} (H-E_T) \frac{1}{\psi_G^{\pm}}
506: \\
507: & = & \frac{1}{2} {\bf \nabla}^2 - \nabla [{\bf b^\pm}.]
508: \label{fkpp} \\
509: & & - (E_L^\pm -E_T) \label{branchingp}
510: \end{eqnarray}
511: where the drift vectors are given by
512: \begin{equation}
513: {\bf b}^\pm \equiv \frac { {\bf \nabla} \psi_G^\pm}{\psi_G^\pm}
514: \end{equation}
515: and the local energies of the guiding functions $\psi_G^\pm$ by
516: \begin{equation}
517: E_L^\pm \equiv \frac{H \psi_G^\pm}{\psi_G^\pm}.
518: \label{elgw}
519: \end{equation}
520: In actual calculations, two Langevin dynamics on the positive and negative walkers are
521: performed
522: \begin{equation}
523: {{{R_i}^\prime}^\mu}^\pm = {{R_i}^\mu}^\pm + b^\mu \tau + \sqrt{\tau} {\eta_i^\mu}^\pm
524: \label{lang+-}
525: \end{equation}
526: and positive and negative walkers are branched according to their respective weight
527: \begin{equation}
528: W^\pm ({\bf R}_i^\pm) \equiv e^{-\tau ({E_L}^\pm ({\bf R}^\pm_i)-E_T)}.
529: \label{FKW+-}
530: \end{equation}
531: \subsection{The detailed rules of FMC}
532: In a few words, the FMC method is similar to a DMC method on a
533: signed function, except that the positive and negative walkers are correlated and can
534: annihilate whenever they meet. The Langevin processes are correlated in such a way that
535: positive walkers and negative walkers meet as much as possible.
536: A cancellation procedure is then performed when a positive walker and a negative
537: walker meet. We will see in the next section that this cancellation procedure is
538: at the origin of an improved stability of the algorithm. In this section we give
539: a complete description of the algorithm and show that this algorithm does not
540: introduce any systematic bias.
541:
542: In FMC the guiding functions $\psi_G^+$ and $\psi_G^-$ are not arbitrary,
543: they are related under any permutation $P$ of two particles as follows
544: \begin{equation}
545: \psi_G^+ ({\bf R}) = \psi_G^- ( P{\bf R}).
546: \label{ppsig+-}
547: \end{equation}
548: Various choices are possible for the guiding functions. Here, we consider the
549: form proposed by Kalos and Pederiva,\cite{kalosfermion00} namely
550: \begin{equation}
551: \psi_G^{\pm} = \sqrt{\psi_S^2 + c^2 \psi_T^2}\pm c \psi_T
552: \label{psig+-def}
553: \end{equation}
554: where $\psi_S$ is a symmetric (Bose-like) function, $\psi_T$ an antisymmetric trial wavefunction,
555: and $c$ some positive mixing parameter allowing to introduce some antisymmetric component into
556: $\psi_G^{\pm}$.
557:
558: Having in mind this choice for the guiding functions we will show how the two
559: DMC processes over the two populations of walkers $\{{\bf R}_i^+\}$ and $\{{\bf
560: R}_i^-\}$ can be replaced by a diffusion process over a population of
561: {\it pairs} of walkers $\{({\bf R}_i^+,{\bf R}_i^-)\}$.
562: We first show how the DMC process can be modified to maintain as many
563: positive and negative walkers during the simulation.
564: In the DMC dynamics, the branching terms associated with the positive and negative walkers
565: are in general different.
566: As a consequence, the number of positive walkers $N_S^+$, can be different from the
567: number of negative walkers $N_S^-$. At time $t$ the DMC density $f_t$ reads
568: \begin{equation}
569: f_t = \left\langle \sum_{i=1}^{N_S^+} \frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle - \sum_{i=1}^{N_S^-} \frac{1}{\psi_G^-} \mid {\bf R}_i^- \rangle \right\rangle.
570: \label{dmcsample+-}
571: \end{equation}
572: This formula is obtained by replacing in equation (\ref{f_tdif}) the expressions
573: (\ref{sampleft+}) and (\ref{sampleft-}) for
574: $f_t^+$ and $f_t^-$, respectively.
575: If $N_S^+$ and $N_S^-$ are different,
576: we will replace $f_t$ (\ref{dmcsample+-}) by a
577: new function $g_t$ sampled with an equal number of positive and negative walkers.
578: Such an operation does not introduce any bias if the antisymmetric
579: components of the future evolution of $f_t$ and $g_t$ are identical.
580: Indeed, only the antisymmetric component of $f_t$ contributes to the
581: estimator of the energy, Eq. (\ref{E_Fproj}).
582: At time $t^\prime > t$ the two densities are given by
583: \begin{eqnarray}
584: f_{t^\prime} & = & e^{-(t^\prime -t) (H-E_T)} f_t \\
585: g_{t^\prime} & = & e^{-(t^\prime -t) (H-E_T)} g_t.
586: \end{eqnarray}
587: Let us write that the antisymmetric components of $f_{t^\prime}$ and $g_{t^\prime}$ must be equal
588: \begin{equation}
589: {\cal A} e^{-(t^\prime -t) (H-E_T)} f_t = {\cal A} e^{-(t^\prime -t) (H-E_T)} g_t.
590: \end{equation}
591: Using the fact that the antisymmetrisation operator ${\cal A}$ commutes with the evolution
592: operator and regrouping all the terms one finally finds
593: \begin{equation}
594: e^{-(t^\prime -t) (H-E_T)} {\cal A} (f_t-g_t) = 0.
595: \end{equation}
596: This condition is satisfied whenever ${\cal A} (f_t-g_t) = 0$.
597: The important conclusion is that one can replace $f_t$ by any function $g_t$
598: such that the difference $g_t-f_t$ is
599: orthogonal to the space of antisymmetric functions.
600: Let us now show how this property can be used to impose a common number of walkers in
601: the positive and negative populations.
602:
603: Let us consider the case where there are more positive walkers than negative walkers,
604: $N_S^+ > N_S^-$.
605: In this case one can substract from $f_t$, Eq.(\ref{dmcsample+-}), the following vector
606: \begin{equation}
607: \frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle + P \frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle
608: \label{1+P+}
609: \end{equation}
610: where ${\bf R}_i^+$ is a positive walker and $P$ is a two-particle permutation.
611: Such an operation is allowed since the application of the antisymmetrizer to the
612: vector (\ref{1+P+}) gives zero [a direct consequence of ${\cal A}(1 + P)=0$].
613: Now, because of Eq.(\ref{ppsig+-}) the vector (\ref{1+P+}) can also be written as
614: \begin{equation}
615: \frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle + \frac{1}{\psi_G^-} P \mid {\bf R}_i^+ \rangle.
616: \end{equation}
617: Substracting this vector from $f_t$ removes the contribution $\frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle$ from (\ref{dmcsample+-}) and adds the contribution $- \frac{1}{\psi_G^-} P \mid {\bf R}_i^+ \rangle$ to (\ref{dmcsample+-}).
618: In other words, the positive walker ${\bf R}_i^+$ has been removed and the negative walker
619: \begin{equation}
620: {\bf R}_i^- = P {\bf R}_i^+
621: \end{equation}
622: has been created.
623: Similarly, one can remove a negative walker ${\bf R}_i^-$ and create a positive walker
624: \begin{equation}
625: {\bf R}_i^+ = P {\bf R}_i^-.
626: \end{equation}
627: This possibility of transfering one walker from one population to the other one allows to
628: keep an identical number of walkers for the two populations at each step.
629: Now, thanks to this possibility, it is possible to
630: interpret the two populations consisting of
631: the $N_S$ positive walkers ${\bf R}_i^+$ and the $N_S$ negative walkers
632: ${\bf R}_i^- $ ($i \in [1..N_S]$)
633: as an unique population of $N_S$ pairs of walkers $\{({\bf R}_i^+,{\bf R}_i^-)\}$.
634: Following this interpretation, the density $f_t$ (\ref{dmcsample+-}) can be then rewritten
635: as an average over a population of pairs of walkers
636: \begin{equation}
637: f_t = \left\langle \sum_{i=1}^{N_S} (\frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle
638: - \frac{1}{\psi_G^-} \mid {\bf R}_i^- )\rangle \right\rangle
639: \label{f_tfmc}
640: \end{equation}
641: and the energy can be computed as a ratio of averages performed on the population of pairs
642: \begin{equation}
643: E_0^F = \frac{\left\langle \sum_i (\frac{H\psi_T}{\psi_G^+} ({\bf R}_i^+) -
644: \frac{H\psi_T}{\psi_G^-} ({\bf R}_i^-)) \right\rangle}
645: {\left\langle \sum_i (\frac{\psi_T}{\psi_G^+} ({\bf R}_i^+)-
646: \frac{\psi_T}{\psi_G^-} ({\bf R}_i^-))\right\rangle},
647: \label{eapair}
648: \end{equation}
649: where Eq.(\ref{E_Fproj}) has been rewritten by replacing $f_t$ using Eq.(\ref{f_tfmc}).
650: Now, everything is in order to detail the short-time dynamics of FMC.
651: The FMC dynamics consists of three steps (Langevin, branching, and
652: cancellation steps):
653:
654: (i) {\sl Langevin step}
655: The Langevin processes (\ref{lang+-}) are simulated as in DMC, except that the gaussian
656: random variables of the positive and negative walkers are no longer independent.
657: The positive walker ${\bf R}_i^+$ and the negative walker ${\bf R}_i^-$ are moved according
658: to Eq.(\ref{lang+-})
659: \begin{equation}
660: {{{R_i}^\prime}^\mu}^\pm = {{R_i}^\mu}^\pm + {b^\mu}^\pm \tau + \sqrt{\tau} {\eta_i^\mu}^\pm
661: \label{lang+-2}
662: \end{equation}
663: where ${\eta_i^\mu}^\pm$ are gaussian centered random variables verifying
664: \begin{equation}
665: \langle {\eta_i^\mu}^\pm {\eta_j^\mu}^\pm \rangle = \delta_{ij}.
666: \end{equation}
667: Such a move insures that the density of positive and negative walkers obey the Fokker Planck
668: equation:
669: \begin{equation}
670: \left\langle \frac{}{} \mid {{\bf R}_i^\prime}^\pm \rangle \right\rangle = e^{\tau ( \frac{1}{2}
671: {\bf \nabla}^2
672: - {\bf \nabla} [{\bf b^\pm}.])} \mid {\bf R}_i^\pm \rangle.
673: \end{equation}
674:
675: However, the gaussian random variables are no more independent,
676: they are correlated within a pair
677: \begin{equation}
678: c_i^{\mu \nu} \equiv \langle {\eta_i^\mu}^+ {\eta_i^\nu}^- \rangle \ne 0.
679: \end{equation}
680: Differents ways of correlating positive and negative walkers can be considered.
681: We shall employ here the approach used in Refs.\cite{Liu1994,kalosfermion00}
682: which consists in obtaining
683: the vector $\vec{{\bf \eta}_i}^-$, representing the $3N$
684: coordinates ${\vec{\bf \eta}_i}^-$, from ${\vec{\bf \eta}_i}^+$ by reflexion
685: with respect to the hyperplane perpendicular to the vector ${\bf R}_i^+-{\bf R}_i^-$.
686: \begin{equation}
687: \vec{\eta_i} ^- = \vec{\eta_i}^+ - 2 \frac{({\bf R}_i^+-{\bf R}_i^-).\vec{\eta_i} ^+}{({\bf R}_i^+-{\bf R}_i^-)^2} ({\bf R}_i^+-{\bf R}_i^-).
688: \label{cor+-}
689: \end{equation}
690: This relation between the gaussian random variables makes the move deterministic
691: along the direction ${\bf R}_i^+-{\bf R}_i^-$. Such a construction insures that the walkers
692: within a pair will meet each other in a finite time (even in large-dimensional spaces).
693: This aspect will be illustrated numerically in the last section.
694: Formally, the two correlated Langevin processes can be seen as one Langevin process
695: in the space of pairs of walkers
696: \begin{equation}
697: ({{\bf R}_i^+}^\prime,{{\bf R}_i^-}^\prime) = ({{\bf R}_i^+},{{\bf R}_i^-}) + ({\bf b^+}({\bf R}_i^+), {\bf b^-({\bf R}_i^-)})\tau + \sqrt{\tau} (\vec{\eta}_i^+,\vec{\eta}_i^-)
698: \label{corrlanffmc}
699: \end{equation}
700: where $\vec{\eta}_i^+$ and $\vec{\eta}_i^-$ are related via (\ref{cor+-}).
701:
702: {\sl (ii) Branching step}
703: As we have already noticed, the branching of the negative and the positive
704: walker are different within a pair:
705: \begin{equation}
706: w_i^+ \equiv e^{-\tau (E_L^+ ({\bf R}_i^+)-E_T)} \ne w_i^- \equiv e^{-\tau (E_L^- ({\bf R}_i^-)-E_T)}.
707: \end{equation}
708: Taking into account their respective weights, the two walkers of a pair give the
709: following contribution to the density $f_t$
710: \begin{equation}
711: w_i^+ \frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle -
712: w_i^- \frac{1}{\psi_G^-} \mid {\bf R}_i^- \rangle .
713: \label{contw+-}
714: \end{equation}
715: If for example $w_i^+>w_i^-$, this vector can be written as
716: \begin{eqnarray}
717: & & w_i^- \left[\frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle -\frac{1}{\psi_G^-} \mid {\bf R}_i^- \rangle \right]
718: \label{branchingpair+-} \\
719: & & + (w_i^+-w_i^-) \frac{1}{\psi_G^+} \mid {\bf R}_i^+ \rangle
720: \label{branchingsingle}
721: \end{eqnarray}
722: This density is the sum of two contributions. The first contribution, Eq.(\ref{branchingpair+-}),
723: comes from a pair of walkers $({\bf R}_i^+, {\bf R}_i^-)$ and carries the weight $w_i^-$.
724: The second, Eq.(\ref{branchingsingle}), comes from a single
725: positive walker ${\bf R}_i^+$ and carries the weight $w_i^+-w_i^-$.
726: This single walker ${\bf R}_i^+$ can be replaced by a pair as follows.
727: First, this single
728: walker can be replaced by two positive walkers ${\bf R}_i^+$ with half of the
729: weight, $\frac{1}{2}(w_i^+-w_i^-)$. One of these two positive walkers carrying
730: half of the weight can be transfered to the population of negative walkers by
731: exchanging two particles.
732: Finally, this single walker ${\bf R}_i^+$ can be replaced by a pair
733: $({\bf R}^+_i,P{\bf R}^+_i)$ carrying the weight $\frac{1}{2}(w_i^+-w_i^-)$.
734: The resulting process just described is a branching of the pair
735: $({\bf R}_i^+, {\bf R}_i^-)$ with the weight $w_i^-$ and the creation of
736: the pair $({\bf R}^+_i,P{\bf R}^+_i)$ with the weight $\frac{1}{2}(w_i^+-w_i^-)$.
737: Of course, if one has $w_i^+<w_i^-$, then, the pair $({\bf R}_i^+, {\bf R}_i^-)$ is
738: branched with the weight $w_i^+$ and the pair $(P{\bf R}^-_i,{\bf R}^-_i)$ is created
739: with the weight $\frac{1}{2}(w_i^--w_i^+)$.
740:
741: Both cases, $w_i^-<w_i^+$ or $w_i^->w_i^+$, can be summarized as follows.
742: The pair of walkers $({\bf R}_i^+,{\bf R}_i^-)$ is branched with the weight
743: \begin{equation}
744: \text{min} (w_i^+,w_i^-) = e^{-\tau (\text{max} (E_L^+,E_L^-)-E_T)}
745: \label{wminpair}
746: \end{equation}
747: and the pairs $({\bf R}^+_i,P{\bf R}^+_i)$ and $(P{\bf R}^-_i,{\bf R}^-_i)$
748: are created with their respective weights
749: \begin{equation}
750: \frac{1}{2}(w_i^+-\text{min}(w_i^+,w_i^-)) = \frac{\tau}{2} [E_L^+-\text{max}
751: (E_L^+,E_L^-)]
752: + O(\tau^2)
753: \label{wsingle1}
754: \end{equation}
755: and
756: \begin{equation}
757: \frac{1}{2}(w_i^--\text{min}(w_i^+,w_i^-)) = \frac{\tau}{2} [E_L^+-\text{max}
758: (E_L^+,E_L^-)]
759: +O(\tau^2)
760: \label{wsingle2}
761: \end{equation}
762:
763: {\sl (iii) Cancellation step}
764: The third step is a cancellation procedure performed whenever a positive and negative walker meet.
765: When ${\bf R_i}^+={\bf R_i}^-$, the contribution of the pair to the density can be simplified
766: as follows
767: \begin{equation}
768: \left[1-\frac{\psi_G^+}{\psi_G^-}({\bf R}_i^+)\right] \frac{1}{\psi_G^+}({\bf R}_i^+) \mid {\bf R}_i^+ \rangle.
769: \end{equation}
770: If the term in brackets is positive, this contribution comes from one single positive
771: walker ${\bf R}_i^+$ with multiplicity $\left[1-\frac{\psi_G^+}{\psi_G^-}({\bf R}_i^+)\right]$.
772: One can transform this single walker into a pair of positive and negative walker
773: $({\bf R}_i^+,P{\bf R}_i^+)$ with the new multiplicity
774: \begin{equation}
775: \frac{1}{2}\left[1-\frac{\psi_G^+}{\psi_G^-}({\bf R}_i^+)\right].
776: \label{multc+}
777: \end{equation}
778: If the term in brackets is negative, the pair $(P{\bf R}_i^+,{\bf R}_i^+)$ is drawn and
779: the multiplicity is given by
780: \begin{equation}
781: \frac{1}{2}\left[1-\frac{\psi_G^-}{\psi_G^+}({\bf R}_i^+)\right].
782: \label{multc-}
783: \end{equation}
784: This is a cancellation procedure because a pair
785: $({\bf R}_i^+,{\bf R}_i^+)$ with a multiplicity 1 has been transformed into a pair
786: with a multiplicity smaller than one.
787: Note that, when $\psi_G^+=\psi_G^-$, the multiplicities (\ref{multc+}) or
788: (\ref{multc-}) reduce both to zero. In other words, there is a total cancellation
789: of the pair whenever the walkers meet.
790: As we shall see in the next section, the cancellation step is
791: at the origin of the improved stability. The basic reason is that this
792: procedure removes pairs which do not contribute to the signal but only to the statistical noise.
793: A rigorous analysis of this point is provided in the next section.
794:
795: \section{Stability of the FMC method}
796: \subsection{Criterium for stability}
797: We have just seen that the FMC method is a generalization of the DMC approach and
798: we have shown that FMC preserves the evolution of the antisymmetric component of the
799: sampled density.
800: Now, having shown that FMC is an exact method, it is necessary to study the stability
801: of the method.
802: For that purpose, we consider the estimator of the energy, Eq. (\ref{eapair}), in the large-time
803: regime
804: \begin{equation}
805: E_0^F = \frac{\cal N}{\cal D} =
806: \frac{\frac{1}{N_S} \sum_i \frac{H\psi_T}{\psi_G^+} ({\bf R}_i^+) - \frac{H\psi_T}{\psi_G^-} ({\bf R}_i^-) }
807: { \frac{1}{N_S} \sum_i \frac{\psi_T}{\psi_G^+} ({\bf R}_i^+)- \frac{\psi_T}{\psi_G^-} ({\bf R}_i^-)}
808: \label{efmcN/D}
809: \end{equation}
810: where $N_S$ is the population size at time $t$.
811: In the same way as done
812: for the estimator (\ref{statea}), we can evaluate the variance of $E_0^F$ by supposing
813: that $N_S$ is large enough
814: so that both numerator and denominator have small fluctuations around their average
815: \begin{equation}
816: \sigma^2 \left( \frac{\cal{N}}{\cal{D}} \right) = \frac{\langle
817: ({\cal N}-E_0^F{\cal D})^2 \rangle}{\langle {\cal D} \rangle^2}.
818: \label{sigmaefmc}
819: \end{equation}
820: Let us begin with the denominator.
821: Using identity (\ref{f_tfmc}), the average of the random variable ${\cal
822: D}$ defined in(\ref{efmcN/D}) is nothing but
823: \begin{equation}
824: \langle {\cal D} \rangle = \frac{1}{N_S} \langle \psi_T \mid f_t \rangle.
825: \end{equation}
826: Replacing $f_t$ by its asymptotic behaviour, Eq.(\ref{aft}), we
827: finally find that the denominator of (\ref{sigmaefmc}) behaves for large $t$ as
828: \begin{equation}
829: \langle {\cal D} \rangle^2 = \frac{1}{N_S^2} e^{-2t(E_0^F-E_T)} \langle \psi_T
830: \mid \phi_0^F \rangle^2.
831: \label{denomfmcbehaviour}
832: \end{equation}
833: Now, let us compute the numerator of (\ref{sigmaefmc}). This numerator can be
834: written as the variance of a sum of random variables defined over the pairs of walkers
835: \begin{equation}
836: \langle ({\cal N}-E_0^F{\cal D})^2 \rangle =
837: \left\langle \left[ \frac{ \sum_i \Gamma({\bf R^+}_i,{\bf R}^-_i)}{N_S} \right]^2 \right\rangle
838: \end{equation}
839: where we have introduced the function $\Gamma({\bf R}^+,{\bf R}^-)$
840: \begin{equation}
841: \Gamma({\bf R}^+,{\bf R}^-)\equiv \frac{(H-E_0^F)\psi_T}{\psi_G^+} ({\bf R}^+) -
842: \frac{(H-E_0^F)\psi_T}{\psi_G^-} ({\bf R}^-).
843: \end{equation}
844: Using the fact that the pairs of walkers have the same distribution and supposing that
845: they are independent we finally find
846: \begin{equation}
847: \langle ({\cal N}-E_0^F{\cal D})^2 \rangle = \frac{1}{N_S} \sigma^2 \left(\frac{}{}\Gamma({\bf R}^+,{\bf R}^-)\right).
848: \end{equation}
849: If the pairs of walkers are not independent
850: this expression is only modified by a correlation factor
851: independent on the time and the population size $N_S$
852: provided that $N_S$ and $t$ are large enough.
853: Finally, up to a multiplicative factor, the variance
854: of the energy estimator has the following asymptotic behaviour
855: \begin{equation}
856: \sigma^2(\frac{\cal N}{\cal D}) \propto \frac{1}{N_S(t)} C(t)
857: \label{sigma2efmc}
858: \end{equation}
859: where the coefficient $C(t)$ is given by
860: \begin{equation}
861: C(t) = N_S(t)^2 e^{-2t(E_0^F-E_T)}
862: \label{C(t)}
863: \end{equation}
864: and where $N_S(t)$, the number of pairs, depends on time $t$ due to the
865: birth-death process.
866: In expression (\ref{C(t)}) we note that the behaviour of (\ref{sigmaefmc}) at
867: large times $t$ is related to the value of $E_0^F$ and the asymptotic behaviour of
868: the number of pairs of walkers $N_S(t)$.
869: We can already understand physically the interest of
870: the cancellation process: this process limits the growth of $N_S(t)$ the number of
871: pairs of walkers, thus limiting the growth of $C(t)$ (\ref{C(t)}) and the
872: variance (\ref{sigma2efmc}).
873: Let us now precise this criterium more rigorously by evaluating the
874: asymptotic behaviour of $N_S(t)$.
875: For that purpose we introduce the density of pairs $\Pi_t ({\bf R}_i^+,{\bf R}_i^-)$, this
876: density obeys a diffusion equation we write down
877: \begin{equation}
878: \frac{\partial \Pi_t}{\partial t} = - ({ D}_{\text{FMC}}-E_T) \Pi_t
879: \label{evotilde}
880: \end{equation}
881: where we have introduced the diffusion operator $-({ D}_{\text{FMC}}-E_T)$.
882: We will give its expression later; for the present purpose we just need
883: the asymptotic behaviour of $\Pi_t$ given by
884: \begin{equation}
885: \Pi_t = e^{-t(\tilde{E}_0^B-E_T)} \Pi_S
886: \label{asympit}
887: \end{equation}
888: where $\Pi_S$ is the stationary density of the process, namely the
889: lowest eigenstate of the operator $D_{\text{FMC}}$, and $\tilde{E}_0^B$ is the
890: corresponding eigenvalue.
891: The number of pairs $N_S(t)$ behaves as the normalization of $\Pi_t$, and consequently grows like
892: $e^{-t(\tilde{E}_0^B-E_T)}$.
893: Note that in practice one adjusts the reference energy $E_T$ to $\tilde{E}_0^B$ during
894: the simulation to keep a constant population of average size $\bar{N_S}$ along the dynamics.
895: Such a procedure is referred to as a control population technique\cite{khelif2000}
896: and will be discussed later.
897: Finally, the asymptotic behaviour of the variance of the FMC estimator of the energy is
898: \begin{equation}
899: \sigma^2(\frac{\cal N}{\cal D}) \propto \frac{1}{\bar{N_S}} e^{2t(E_0^F-\tilde{E}_0^B)}
900: \label{fluctfmc}
901: \end{equation}
902: This expression is analogous to (\ref{sigmaea}) except that the lowest energy of the
903: Hamiltonian operator $H$ has been replaced by the lowest energy of the operator
904: $D_{\text{FMC}}$.
905: In conclusion the stability of the algorithm is related to the lowest eigenvalue
906: of the FMC diffusion operator, $\tilde{E}_0^B$.
907: It is clear from (\ref{fluctfmc}) that the higher this eigenvalue is, the more
908: stable the simulation will be.
909: In the next section, we will discuss the allowed values of $\tilde{E}_0^B$.
910: This will prove that FMC is not a stable method in general, but
911: is more stable than any standard transient method.
912:
913: \subsection{Stability of the Fermion Monte Carlo algorithm}
914: In this section we prove that the lowest eigenvalue of the FMC operator,
915: $\tilde{E}_0^B$, has the following upper and lower bounds
916: \begin{eqnarray}
917: \tilde{E}_0^B & \leq & E_0^F
918: \label{stableineq}
919: \\
920: \tilde{E}_0^B & > & E_0^B
921: \label{improvstable}
922: \end{eqnarray}
923: >From the expression of the variance, Eq. (\ref{fluctfmc}), one can easily
924: understand the meaning of these two inequalities.
925: The first inequality indicates that FMC is not a stable
926: method, the stability being achieved only in the limit $\tilde{E}_0^B=E_0^F$.
927: Note that, even for very simple systems, this stability is in general not obtained.
928: This important point will be illustrated in the next section.
929: The second inequality shows that FMC is more stable than any standard
930: transient DMC method (nodal release method). Indeed, the exponent associated with the explosion
931: of fluctuations, Eq. (\ref{fluctfmc}), is smaller than in the standard case, Eq.(\ref{sigmaea}).
932: Before giving a mathematical proof of these two inequalities, let us first present
933: some intuitive arguments in their favor.
934: The first inequality, Eq. (\ref{stableineq}), takes its origin in the fact that the
935: signal -the antisymmetric component of $f_t$, Eq.(\ref{aft}) is extracted from the population
936: of pairs of walkers, Eq.(\ref{f_tfmc}), and, consequently, cannot grow faster than the
937: population of pairs itself, Eq.(\ref{asympit}).
938: The second inequality can be understood as follows.
939: Without the cancellation process, the FMC method reduces to
940: two correlated DMC algorithms. The number of walkers grows as in a standard
941: DMC, namely $ \sim e^{-(E_0^B-E_T)t}$. The cancellation
942: process obviously reduces the growth of the population of walkers, Eq.
943: (\ref{asympit}), and, thus, we should expect that $\tilde{E}_0^B > E_0^B$.
944: Now, let us give some more rigorous proofs. For that purpose we
945: compare the Fermion Monte Carlo operators with and without cancellation process.
946: Without the cancellation process the FMC diffusion operator reads
947: \begin{eqnarray}
948: D -E_T & \equiv & {\psi_G^+} (H^+-{E_L}^+) \frac{1}{{\psi_G}^+}
949: +{\psi_G}^- (H^--{E_L}^-) \frac{1}{{\psi_G}^-}
950: \label{langevin+-}
951: \\
952: & & - \frac{1}{2} \frac{\partial^2}{\partial {R^+}_\mu \partial {R^-}_\nu}[c_{\mu \nu} .]
953: \label{langevincorr}
954: \\
955: & & + \text{max}({E_L}^+,{E_L}^-) -E_T
956: \label{branchingpair}
957: \\
958: & & + \frac{1}{2} [{E_L}^+-\text{max}({E_L}^+,{E_L}^-)]
959: e^{(P{\bf R}^+- {\bf R}^-).{\bf \nabla}^-. }
960: \label{hoping-}
961: \\
962: & & + \frac{1}{2} [{E_L}^--\text{max}({E_L}^+,{E_L}^-)]
963: e^{ (P{\bf R}^-- {\bf R}^+). {\bf \nabla}^+. }
964: \label{hoping+}
965: \end{eqnarray}
966: where the operators $H^\pm$ are both identical to $H$, except
967: that $H^+$ and $H^-$ act on the space of positive and negative configurations, respectively.
968: \begin{equation}
969: H^{\pm} \equiv H ({\bf R}^{\pm})
970: = -\frac{1}{2} \sum_\mu \frac{\partial^2}{{\partial R^{\pm}_\mu}^2} + V ({\bf R}^\pm).
971: \end{equation}
972: The coefficients $c_{\mu\nu}$ are real coefficients and will be defined below.
973: To justify that this operator is the diffusion operator
974: corresponding to FMC with no cancellation process, we need to
975: check that the short-time dynamics described by
976: \begin{equation}
977: \frac{\partial \Pi_t}{\partial t} = -( D-E_T) \Pi_t
978: \label{fmcwcan}
979: \end{equation}
980: is indeed realized via the two first steps of the FMC algorithm (Langevin and
981: branching steps).
982: In the expression of $D$, the two operators appearing in Eq.(\ref{langevin+-})
983: define a Fokker Planck operator in analogy to Eq.(\ref{fkpp}).
984: This operator is the diffusion operator associated with the Langevin process,
985: Eq.(\ref{corrlanffmc}).
986: The term in Eq.(\ref{langevincorr}) is a coupling term between the moves of
987: positive and negative walkers taking into account the correlation of the
988: gaussian random variables $\eta_\mu^+$ and $\eta_\mu^-$.
989: The quantities $c_{\mu \nu} ({\bf R}^+,{\bf R}^-)$ introduced in Eq.(\ref{langevincorr})
990: are nothing but the covariance of these variables
991: \begin{equation}
992: c_{\mu \nu} ({\bf R}^+,{\bf R}^-) = \langle \eta_\mu^+ \eta_\nu^- \rangle.
993: \end{equation}
994: The three last contributions describe the branching processes at work in FMC.
995: One recognizes in Eq.(\ref{branchingpair}) the branching of a pair, Eq.(\ref{wminpair}).
996: The two following contributions, Eqs.(\ref{hoping-}) and
997: (\ref{hoping+}), correspond to the
998: creation of pairs $({\bf R}^+,P{\bf R}^+)$ and $(P{\bf R}^-,{\bf
999: R}^-)$, with the respective weights given by (\ref{wsingle1}) and (\ref{wsingle2}).
1000: Note that in Eqs.(\ref{hoping-},\ref{hoping+}) the operator
1001: $e^{ (P{\bf R}^-- {\bf R}^+). {\bf \nabla}^+ }$ is written in a symbolic form
1002: representing a translation of the vector ${\bf R}^+$ to $P{\bf R}^-$, the action of this operator
1003: on the pair $({\bf R}^+,{\bf R}^-)$ being indeed to create the pair $(P{\bf R}^-,{\bf R}^-)$.
1004:
1005: Now, let us prove that the lowest eigenvalue of $D$ is $E_0^B$ (bosonic ground-state).
1006: For that purpose, it is convenient to introduce the operator ${\cal R}$ which transforms a
1007: distribution of pairs of walkers into a distribution of walkers and then to define
1008: the following reduced density
1009: \begin{equation}
1010: {\cal R} {\Pi}_t ({\bf R}) \equiv \int dR^\prime \left[\frac{{\Pi}_t ({\bf
1011: R},{\bf R}^\prime)}{\psi_G^+ ({\bf R})}
1012: +\frac{{\Pi}_t ({\bf R}^\prime,P{\bf R})}{\psi_G^- ({\bf R})} \right].
1013: \label{reduceddensity}
1014: \end{equation}
1015: The density ${\cal R} {\Pi}_t$ represents the sum of the distributions sampled by each type of
1016: walkers when the contribution of the other type of walkers is integrated out.
1017: Using the explicit expression of $D$ it is a simple matter of algebra to
1018: verify that
1019: \begin{equation}
1020: {\cal R} D {\Pi_t}= H {\cal R }{\Pi}_t.
1021: \label{relevint}
1022: \end{equation}
1023: % MC: This relation may look very formal, but it has a physical meaning.
1024: % Si tu ne dis rien de plus, il faut enlever cette phrase.
1025: Using Eqs.(\ref{relevint}) and (\ref{fmcwcan}), one can also write
1026: \begin{equation}
1027: \frac{\partial {\cal R} \Pi_t}{\partial t} = - (H-E_T) {\cal R }\Pi_t
1028: \label{evof}
1029: \end{equation}
1030: which means that the reduced density evolves under the dynamics of $H$.
1031: In other words, the set of positive and negative walkers sample the same
1032: distribution as a standard Diffusion Monte Carlo algorithm.
1033: Now, suppose that $\lambda_S$ is the lowest eigenvalue of $D$ and $\Pi_S$ the corresponding
1034: eigenstate (the stationary density of the process described by Eq.(\ref{fmcwcan}))
1035: \begin{equation}
1036: D \Pi_S = \lambda_S \Pi_S.
1037: \end{equation}
1038: Applying ${\cal R}$ on both sides of this identity and using the relation (\ref{relevint})
1039: one gets
1040: \begin{equation}
1041: H {\cal R} \Pi_S = \lambda_S {\cal R} \Pi_S.
1042: \end{equation}
1043: In other words, the reduced density ${\cal R} \Pi_S$ is a positive eigenstate of
1044: $H$ with eigenvalue $\lambda_S$. The bosonic state being non-degenerate, we can conclude that
1045: $\lambda_S=E_0^B$. This ends our proof.
1046:
1047: Let us now consider the genuine FMC diffusion operator including the cancellation process,
1048: $D_{FMC}$. To simplify the notations let us suppose that we are in the symmetric case for which
1049: $\psi_G^+=\psi_G^-=\psi_G$ (the common guiding function is symmetric under permutation
1050: of particles). This particular case is much simple because when walkers meet, there is a
1051: full cancellation and no residual branching. From an operatorial point of view the
1052: cancellation step consists in introducing a projection operator, $P_c$,
1053: at each step of the dynamics
1054: \begin{eqnarray}
1055: P_c \equiv [1-\int dR \mid R^+=R, R^-=R \rangle \langle R^+=R, R^-=R\mid].
1056: \label{p_c}
1057: \end{eqnarray}
1058: where $\mid R^+, R^- \rangle$ denotes the usual tensorial product.
1059: The full FMC diffusion operator can thus be written as
1060: \begin{equation}
1061: D_{\text{FMC}} \equiv P_c D.
1062: \label{hprimedef}
1063: \end{equation}
1064: It is important to realize that the $D_{\text{FMC}}$ operator defined via $P_c$ and $D$
1065: (Eqs.(\ref{p_c}) and (\ref{hoping+})) represents indeed an equivalent operatorial description of
1066: the stochastic rules of FMC described in section IV.B (Langevin, branching and cancellation
1067: steps). Note also that using the expression
1068: (\ref{hprimedef}) of $D_{\text{FMC}}$, we have a simple alternative way of
1069: recovering the proof just presented above that FMC is a bias-free approach. Since this is
1070: an important point of this work, let us present this alternative proof.
1071: The action of the projection operator $P_c$ is to remove from the sample components
1072: of the form $\mid R^+=R, R^-=R \rangle$
1073: for which the antisymmetric component of the reduced density is zero
1074: \begin{equation}
1075: {\cal A} {\cal R} \mid R^+=R, R^-=R \rangle
1076: = {\cal A} \frac{1}{\psi_G ({\bf R})} (\mid R \rangle + \mid P R \rangle)=0.
1077: \label{p0part}
1078: \end{equation}
1079: In other words one has the following algebraic identity
1080: \begin{equation}
1081: {\cal A} {\cal R} P_c = {\cal A} {\cal R}.
1082: \label{p0gen}
1083: \end{equation}
1084: In the general case where $\psi_G^+ \ne \psi_G^-$, the cancellation procedure
1085: still corresponds to define a new operator written as in Eq.(\ref{hprimedef}) with
1086: $P_c$ satisfying the same identity as in (\ref{p0gen}).
1087: Applying ${A\cal R}$ to the L.H.S. and R.H.S. of equation (\ref{evotilde}), one has
1088: \begin{equation}
1089: \frac{ \partial {\cal A} {\cal R} \Pi_t} {\partial t}
1090: = - {{\cal A}\cal R} D_{\text{FMC}} \Pi_t = -{\cal A} (H-E_T) {\cal R} \Pi_t.
1091: \end{equation}
1092: This equation indicates that the evolutions of the antisymmetric component of the reduced
1093: density under the dynamics of ${D}_{\text {FMC}}$ and $H$ are identical.
1094: This confirms that the energy estimator, Eq.(\ref{f_tfmc}), or any observable estimator not
1095: coupling directly positive and negative walkers, is not biased. In the case of the
1096: energy, the estimator can be written as a function of the reduced density as follows
1097: \begin{equation}
1098: E_0^F = \frac{\langle \psi_T \mid H {\cal R} {\Pi}_t \rangle}
1099: {\langle \psi_T \mid {\cal R} {\Pi}_t \rangle}.
1100: \label{easymb}
1101: \end{equation}
1102:
1103: Let us now turn back to our discussion of the stability of FMC. For that, we need
1104: to compare the lowest eigenvalue $\tilde{E}_0^B$ of ${D}_{\text{FMC}}$ and, $E_0^B$, the lowest
1105: eigenvalue of $H$. Now, it is clear from the definition of ${D}_c$, Eq.(\ref{hprimedef}),
1106: that the following relation holds
1107: \begin{equation}
1108: \tilde{E}_0^B > E_0^B.
1109: \label{bettstab}
1110: \end{equation}
1111: Indeed, the action of $P_c$ present in the definition of
1112: ${D}_c$, Eq.(\ref{hprimedef}), consists in removing positive coefficients within the
1113: extradiagonal part of ${D}$. As well-known, a consequence of such a matrix (or operator)
1114: manipulation is to increase the energy of the lowest eigenvalue of the matrix.
1115: Expressed in a more physical way, the cancellation process reduces the growth of the
1116: population of pairs: $e^{-(\tilde{E}_0^B-E_T)t} < e^{-(E_0^B-E_T)t}$.
1117:
1118: To summarize, we have shown that FMC reduces the instability of fermion simulations.
1119: The signal-over-noise ratio decreases as $e^{(\tilde{E}_0^B-E_0^F) t}$,
1120: where $\tilde{E}_0^B$ is the lowest eigenvalue of the FMC operator, ${D}_{\text{FMC}}$.
1121: Because of the inequality (\ref{bettstab}), this ratio decreases slower in FMC that in any
1122: standard transient DMC or nodal release methods.
1123: We have shown that the cancellation process is at the origin of this improvement;
1124: however, as we shall see in the next section, the cancellation process is
1125: efficient (i.e., we have a small difference $\tilde{E}_0^B-E_0^F$) only if the correlation between
1126: walkers described by the coupling terms $c_{\mu\nu}$ is
1127: introduced. This feature is important, particularly in high-dimensional spaces where the
1128: probability of meeting and cancelling becomes extremely small for independent walkers.
1129: As a result, the correlation of positive and negative walkers is a fundamental
1130: feature of FMC.
1131: The quantitative effect of the correlation on the stabilization of the algorithm is not easy
1132: to study theoretically and to optimize in the general case.
1133: In the next section we will give a numerical illustration, for a simple system,
1134: of the interplay between cancellation and correlation (via the $c_{\mu\nu}$ parameters),
1135: and also of the role of the choice of the guiding functions, $\psi_G^+$ and $\psi_G^-$.
1136:
1137: \section{Numerical study}
1138:
1139: \subsection{The model: 2D-harmonic oscillator on a finite grid}
1140: In this section we study the FMC method on a very simple model on a lattice.
1141: For this model it is possible to calculate $E_0^F$ (fermionic ground-state energy),
1142: $E_0^B$ (bosonic ground-state energy), and $\tilde{E}_0^B$ the lowest eigenvalue of the FMC
1143: operator by a standard deterministic method (exact diagonalization).
1144: The results obtained for this simple model will provide us with a well-grounded
1145: framework to interpret the Fermion Monte Carlo simulations.
1146: The second motivation is that, using such a simple model, it is possible to study the limit
1147: of large number of walkers, large with respect to the dimension of the Hilbert space considered.
1148: This possibility turns out to be essential to better understand the FMC algorithm.
1149:
1150: Our model is based on the discretization of a system describing two-coupled harmonic oscillators
1151: \begin{equation}
1152: H = -\frac{1}{2} ( \frac{\partial^2}{\partial x^2}
1153: + \frac {\partial^2}{\partial y^2} ) + V(x,y)
1154: \end{equation}
1155: with
1156: \begin{equation}
1157: V(x,y) = \frac{1}{2} x^2 + \frac{1}{2}\lambda y^2 + xy
1158: \end{equation}
1159: In the following we shall take $\lambda=2$.
1160: Now, we define the discretization of this model on a $N$x$N$ regular grid
1161: ($N$ odd). A grid point ${\bf R}_i \;\; i \in (1, \dots, N^2)$ has the following
1162: coordinates
1163: \begin{equation}
1164: {\bf R}_i \equiv \left((-\frac{N}{2}+k-1)\delta_x, (-\frac{N}{2}+l-1)
1165: \delta_y\right) \;\;\; k \in [1 \dots N], \;\; l \in [1\dots N]
1166: \end{equation}
1167: where
1168: \begin{equation}
1169: \delta_x=\delta_y=x_{max}/N
1170: \label{xmax}
1171: \end{equation}
1172: On this lattice the Hamiltonian has a corresponding discrete representation given
1173: by a finite matrix.
1174: The diagonal part of the matrix reads
1175: \begin{equation}
1176: H_{ii}= \frac{1}{{\delta_x}^2} + \frac{1}{{\delta_y}^2} + V(R_i) \;\; i \in (1,\dots,N^2)
1177: \end{equation}
1178: and the off-diagonal part reads
1179: \begin{eqnarray}
1180: H_{ij} & = & -\frac{1}{2{\delta_x}^2} \text{ \ \ when $R_i$ and $R_j$ are nearest-neighbors on
1181: the lattice} \nonumber \\
1182: H_{ij} & = & 0 \text{ \ \ otherwise}
1183: \end{eqnarray}
1184:
1185: This Hamiltonian is symmetric with respect to the inversion $P$ of center $O = (0,0)$.
1186: \begin{equation}
1187: P (x,y) \equiv (-x,-y)
1188: \end{equation}
1189: As a consequence, the eigenfunctions are either symmetric or antisymmetric under
1190: $P$. We are interested in the energy $E_0^F$ of the lowest antisymmetric eigenstate,
1191: $\phi_0^F$. Even for this simple system, we are confronted with a sign
1192: instability and a genuine ``sign problem''. Indeed, the sign of $\phi_0^F$ for each grid
1193: point cannot be entirely determined by symmetry.
1194: Symmetry implies only that $\phi_0^F$ vanishes at the inversion center and that the two-dimensional
1195: pattern of positive and negative values for $\phi_0^F$ is symmetric by inversion. The precise
1196: delimitation between positive and negative zones of the wavefunction (analogous to
1197: ``nodal surfaces'' for continuous systems) is not known.
1198:
1199: Let us now introduce the trial functions $\psi_T$ and $\psi_S$. $\psi_T$
1200: has to be an (antisymmetric) approximation of $\phi_0^F$ and $\psi_S$ some symmetric and positive
1201: approximation of the lowest eigenstate, $\phi_0^B$.
1202: We have chosen them as discretizations of the exact solutions of the initial continuous model.
1203: To find these solutions we perform a diagonalization of the quadratic form of the potential
1204: \begin{equation}
1205: V(x,y) = \frac{1}{2} x^2 + \frac{1}{2}\lambda y^2 + xy
1206: = \frac{1}{2} k_1 \tilde{x}^2 + \frac{1}{2}k_2 \tilde{y}^2.
1207: \end{equation}
1208: It is trivial to verify that
1209: $$
1210: k_1=\frac{\cos^2{\theta} -\lambda \sin^2{\theta}}{cos{2\theta}}
1211: $$
1212: $$
1213: k_2=\frac{\lambda \cos^2{\theta} - \sin^2{\theta}}{cos{2\theta}}
1214: $$
1215: $$
1216: \tan{2\theta}= \frac{2}{\lambda-1}
1217: $$
1218: with
1219: \begin{equation}
1220: \tilde{x}= x \cos{\theta} - y \sin{\theta}
1221: \end{equation}
1222: and
1223: \begin{equation}
1224: \tilde{y}= x \sin{\theta} + y \cos{\theta}.
1225: \end{equation}
1226: If $k_1 < k_2$ we choose as trial wavefunction
1227: \begin{equation}
1228: \psi_T= \tilde{x}
1229: \exp{( -\frac{\sqrt{k_1}}{2} \tilde{x}^2 -\frac{\sqrt{k_2}}{2} \tilde{y}^2)},
1230: \end{equation}
1231: while, in the other case, we take
1232: \begin{equation}
1233: \psi_T= \tilde{y}
1234: \exp{( -\frac{\sqrt{k_1}}{2} \tilde{x}^2 -\frac{\sqrt{k_2}}{2} \tilde{y}^2)}
1235: \end{equation}
1236: The lowest (symmetric) eigenstate is chosen to be
1237: \begin{equation}
1238: \psi_S=
1239: \exp{(-\frac{\sqrt{k_1}}{2} \tilde{x}^2 -\frac{\sqrt{k_2}}{2} \tilde{y}^2)}.
1240: \end{equation}
1241: Note that, in the limit of a very large system the trial functions, $\psi_T$ and $\psi_S$
1242: reduce to two exact eigenstates of $H$; however, this is not the case for finite systems.
1243:
1244: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1245: \subsection{FMC on the lattice}
1246: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1247: Before presenting our results, let us say a few words about
1248: the implementation of the FMC on the lattice.
1249: The same ingredients as in the continuum case hold, except that in the lattice case
1250: the Langevin process is realized through a discrete transition probability matrix.
1251: The probability for a (positive or negative) walker $i$ to go to $j$, after a time step $\tau$ is
1252: \begin{equation}
1253: P^{\pm}(i\rightarrow j) \equiv \frac{\psi_G^{\pm}({\bf R}_j)}{\psi_G^{\pm}({\bf R}_i)}
1254: <{\bf R}_j|1-\tau(H-E_L^{\pm})|{\bf R}_i>
1255: \label{Plattdef}
1256: \end{equation}
1257: where
1258: $\tau$ is small enough to have a positive density, namely
1259: \begin{equation}
1260: \tau < \frac{1}{\text{Max}[H_{ii}-E_L^\pm(i)]}.
1261: \label{taumax}
1262: \end{equation}
1263: The local energies are defined as in the continuum, Eq.(\ref{elgw})
1264: \begin{equation}
1265: E_L^{\pm}({\bf R})= \frac{H \psi_G^{\pm}}{\psi_G^{\pm}}({\bf R})
1266: \nonumber
1267: \end{equation}
1268: with the same expression for the guiding functions $\psi_G^{\pm}$, Eq.(\ref{psig+-def})
1269: \begin{equation}
1270: \psi_G^{\pm}({\bf R}) \equiv \sqrt{\psi_S^2 + c^2 \psi_T^2} \pm c\psi_T.
1271: \label{psigpm}
1272: \end{equation}
1273: Let $(i_1,i_2)$ represents a given pair of positive and negative walkers
1274: $({\bf R}^+_{i_1},{\bf R}^-_{i_2})$.
1275: In a standard diffusion Monte Carlo (no correlation and no cancellation
1276: of positive and negative walkers), the density of pairs of positive and negative
1277: walkers $\Pi^{(k)}_{i_1i_2}$ evolves as follows in one time-step
1278: \begin{equation}
1279: \Pi^{(k+1)}_{i_1 i_2}= \sum_{j_1 j_2} \Pi^{(k)}_{j_1 j_2} P^+(j_1 \rightarrow
1280: i_1) W^+_{j_1} P^-(j_2 \rightarrow i_2) W^-_{j_2}
1281: \end{equation}
1282: where $W^\pm$ is the Feynman-Kac weight, Eq.(\ref{FKW+-}).
1283: To build the FMC algorithm, one first correlate the two stochastic processes
1284: $P^+$ and $P^-$, Eq.(\ref{Plattdef}).
1285: The way it is performed here is the counterpart of the correlation term introduced
1286: by Liu {\sl et al.} \cite{Liu1994} in the continuum case, Eq.(\ref{cor+-}).
1287: With such a choice,
1288: the positive and negative walkers of a pair tend to get closer or to move away
1289: in a concerted way.
1290: For the lattice case it is done as follows.
1291: The positive walker $j_1$ is connected by $P^+$ to a finite number of states
1292: ${j_1^c}$ (here, maximum five)
1293: with probability $P^+ (j_1 \rightarrow j_1^c)$.
1294: The negative walker $j_2$ is connected to a finite number of
1295: states $j_2^c$ with the probability $P^- (j_2 \rightarrow j_2^c)$.
1296: The states $j_1^c$ are ordered taking as criterium the distance to
1297: the negative walker $j_2$, $|{\bf R}_{j_1^c}-{\bf R}_{j_2}|$.
1298: We do the same for the states $i_2^c$ ordered by their distance
1299: with respect to the positive walker.
1300: An {\it unique} random number uniformly distributed between 0 and 1 is
1301: then drawn and the repartition functions of the two probability measures,
1302: $p (j_1^c) \equiv P^+ (j_1 \rightarrow j_1^c) $
1303: and $p (j_2^c) \equiv P^- (j_2 \rightarrow j_2^c)$ are then sampled using this common
1304: random number.
1305: The new pair $(j_1^c,j_2^c)$ is drawn accordingly.
1306: Such a procedure defines a {\it correlated} transition probability in the space of pairs
1307: \begin{equation}
1308: P_c (j_1 j_2 \rightarrow i_1 i_2) \ne P (j_1 \rightarrow i_2)
1309: P (j_1 \rightarrow i_2)
1310: \label{corrmove}
1311: \end{equation}
1312: whose role is to enhance the probability of having positive and negative walkers
1313: meeting at the same site. Note that, by construction, the correlation introduced via
1314: $P_c$ does not
1315: change the individual (reduced) densities associated with each type of walker (positive/negative).
1316: Now, let us write down explicitly the FMC rules in our lattice case, that is,
1317: the one time-step ($k\rightarrow k+1$) evolution of the density of pairs $\Pi^{(k)}_{j_1 j_2}$:
1318:
1319: (i) {\sl Correlation and branching.}
1320: The branching of an individual pair $(j_1,j_2)$, Eq.(\ref{branchingpair}),
1321: corresponds to the following evolution of the density
1322: \begin{equation}
1323: \Pi^{(k+1)}_{i_1 i_2}= \Pi^{(k)}_{i_1 i_2} + \sum_{j_1 j_2} \Pi^{(k)}_{j_1 j_2}
1324: P_c(j_1 j_2 \rightarrow i_1 i_2) \text{Min}[ W^+_{ i_1},
1325: W^-_{i_2} ]
1326: \label{eqfmc1}
1327: \end{equation}
1328: The creation of pairs, Eqs.(\ref{hoping-},\ref{hoping+}), can be written as follows
1329: \begin{equation}
1330: \Pi^{(k+1)}_{i_1 P(i_1)}= \Pi^{(k)}_{i_1 P(i_1)} + \sum_{j_1 j_2} \Pi^{(k)}_{j_1 j_2} P_c(j_1 j_2
1331: \rightarrow i_1 i_2)
1332: \theta[W^+_{i_1},W^-_{i_2}] |W^+_{i_1}-W^-_{i_2}|/2
1333: \label{eqfmc2}
1334: \end{equation}
1335: \begin{equation}
1336: \Pi^{(k+1)}_{P(i_2) i_2}= \Pi^{(k)}_{P(i_2) i_2} +
1337: \sum_{j_1 j_2} \Pi^{(k)}_{j_1 j_2} P_c(j_1 j_2 \rightarrow i_1 i_2)
1338: (1- \theta[W^+_{i_1},W^-_{i_2}]) |W^+_{i_1}-W^-_{i_2}|/2
1339: \label{eqfmc3}
1340: \end{equation}
1341: where $\theta(x,y)=1$ if $x > y$, $\theta(x,y)=0$, otherwise.
1342:
1343: (ii) {\sl Cancellation.}
1344: The cancellation process is done when a positive walker and a negative walker
1345: meet $i_1=i_2=i$
1346: \begin{equation}
1347: \Pi^{(k+1)}_{ii} = \left[ 1 - \text{min} (\frac{\psi_G^+}{\psi_G^-} (i),
1348: \frac{\psi_G^-}{\psi_G^+} (i)) \right] \Pi^{(k)}_{ii}.
1349: \label{eqfmc4}
1350: \end{equation}
1351: If $\psi_G^+(i) > \psi_G^-(i)$ we have
1352: \begin{equation}
1353: \Pi^{(k+1)}_{P(i) i}= \Pi^{(k)}_{P(i) i} + \frac{[1-\psi_G^-(i)/\psi_G^+(i)]}{2} \Pi^{(k)}_{i i}.
1354: \label{eqfmc5}
1355: \end{equation}
1356: If $\psi_G^+(i) < \psi_G^-(i)$ we have
1357: \begin{equation}
1358: \Pi^{(k+1)}_{i P(i)}= \Pi^{(k)}_{i P(i)} + \frac{[1-\psi_G^+(i)/\psi_G^-(i)]}{2} \Pi^{(k)}_{i i}.
1359: \label{eqfmc6}
1360: \end{equation}
1361: The operations (\ref{eqfmc1},\ref{eqfmc2},\ref{eqfmc3},\ref{eqfmc4},\ref{eqfmc5},\ref{eqfmc6})
1362: describe the one time-step dynamics of the simulation.
1363: At iteration $k$, the distribution of pairs $\Pi^{(k)} (i_1,i_2)$ is obtained
1364: and the transient estimator of the energy (\ref{eapair}) can be computed from
1365: \begin{equation}
1366: E(k) = \frac{ \sum_{i_1,i_2} \Pi^{(k)}_{i_1,i_2} [ \frac{H\psi_T (i_1)}{\psi_G^+(i_1)}- \frac{H\psi_T (i_2)}{\psi_G^-(i_2)}]}
1367: { \sum_{i_1,i_2} \Pi^{(k)}_{i_1,i_2}
1368: [ \frac{\psi_T (i_1)}{\psi_G^+(i_1)}- \frac{\psi_T (i_2)}{\psi_G^-(i_2)}]}.
1369: \label{energyest}
1370: \end{equation}
1371: This estimator converges to $E_0^F$ when $k \rightarrow \infty$.
1372:
1373: Now, it is important to realize that the FMC rules just presented have, in principle,
1374: no stochastic character at all. For a finite system the FMC rules can be viewed
1375: as simple deterministic matrix manipulations between finite vectors of the Hilbert
1376: space (here, the multiplications to be performed have been explicitly written).
1377: At the beginning of the simulation (iteration $k=0$)
1378: some arbitrary starting vector $\Pi^{(0)}_{i_1,i_2}$ is chosen and, then, iterations
1379: are performed up to convergence. This important remark will allow us to organize our
1380: discussion of FMC results into two parts. In a first part (Section VI.C), we perform explicitly
1381: the matrix multiplications involved and any stochastic aspect is removed from the results.
1382: Stated differently, this procedure can be viewed as performing a standard FMC simulation with an
1383: infinite number of walkers (the distribution at each point of the configuration space
1384: is exactly obtained, no statistical fluctuations are present). In a second part (Section VI.D),
1385: we implement the usual stochastic interpretation of the FMC rules using a finite number
1386: of walkers. This second part will allow to discuss the important consequences of the finiteness of
1387: the number of walkers and, in particular, the role played by the use of population
1388: control techniques.
1389:
1390: \subsection{FMC using an infinite number of walkers: the deterministic approach}
1391:
1392: \subsubsection{No systematic error: FMC is an exact method}
1393:
1394: In this section we verify on our simple example that the FMC rules do not introduce any
1395: systematic error (bias). The energy expression (\ref{energyest}) has been computed by
1396: iterating the applications of the elementary
1397: operators defined by
1398: (\ref{eqfmc1},\ref{eqfmc2},\ref{eqfmc3},\ref{eqfmc4},\ref{eqfmc5},\ref{eqfmc6}). In practice,
1399: this corresponds to iterate a matrix ${G}_{\text{FMC}} (\tau)$.
1400: The distribution $\Pi^{(k+1)}$ at iteration $k+1$ is obtained from $\Pi^{(k)}$ as follows
1401: \begin{equation}
1402: \Pi^{(k+1)} = {G}_{\text{FMC}} (\tau) \Pi^{(k)}
1403: \end{equation}
1404: The operator ${G}_{\text{FMC}} (\tau)$ has been applied a large number of times on some initial
1405: density $\Pi^{(0)}_{j_1 j_2}$
1406: (a $N^4$ component vector, $N$ being the linear size of our lattice) and
1407: the energy (\ref{energyest}) has been computed at each iteration $k$.
1408: We have checked that the energy converges to the exact
1409: value, $E_0^F$, corresponding the lowest antisymmetric state, with all decimal places.
1410: We have verified that this is true for several cases corresponding to $N$ ranging from
1411: 4 to 17. For this specific problem these results confirm numerically that FMC is an exact method.
1412:
1413: \subsubsection{Meeting time between positive and negative walkers}
1414: Here, we want to illustrate quantitatively the fact that the correlation introduced
1415: via the probability transition helps greatly to lower the meeting time between
1416: positive and negative walkers. The influence of the choice of the guiding functions
1417: (here, parameter $c$ in Eq.(\ref{psigpm})) on the meeting time is also examined.
1418: The meeting time is defined and evaluated as follows.
1419: We start with a configuration consisting of a
1420: positive walker located at a corner of the lattice and
1421: a negative walker located at the opposite corner. The positive and
1422: negative walkers are moving stochastically with the transition probability
1423: defined in (\ref{Plattdef}).
1424: We test the two cases cases corresponding to uncorrelated and correlated moves.
1425: The average time $\langle T \rangle$
1426: (number of Monte Carlo steps times $\tau$) before the walkers meet is computed.
1427: Our results are presented in Table \ref{table1} and are
1428: given for different linear sizes of the grid.
1429: In this first case the guiding functions are chosen with a large antisymmetric component, $c=4$.
1430: The results indicate clearly that more than one order of magnitude is gained by
1431: correlating the moves of the two stochastic processes.
1432: In Table \ref{table2} the same calculations are done, except that a symmetric
1433: guiding function $c=0$ ($\psi_G^+=\psi_G^-=\psi_S$) is employed.
1434: The average meeting time is found to be much lower than in the non-symmetric case, $c=4$, by
1435: nearly two orders of magnitude. This is true whether or not the stochastic processes are
1436: correlated. This behaviour of the meeting time as a function of $c$ is not surprising.
1437: When $c$ is large the two functions
1438: $\psi_G^+$ and $\psi_G^-$ are localized in the nodal pockets of $\psi_T$.
1439: In the large-$c$ limit $\psi_G^+$ is zero whenever $\psi_T$ is negative
1440: and $\psi_G^-$ is zero whenever $\psi_T$ is positive. In this limit the overlap between the two
1441: distributions $\psi_G^+$ and $\psi_G^-$ is zero and we have a similar result for the
1442: probability that walkers meet.
1443: >From these preliminary results the introduction of non-symmetric wavefunctions
1444: seems to deteriorate the stability, this property will be confirmed in the next section.
1445: \subsubsection{Stability in time of FMC}
1446: We know from section V that the stability in time is directly related to the magnitude of
1447: the reduced Bose-Fermi energy gap
1448: \begin{equation}
1449: \tilde{\Delta}_{B-F} \equiv E_0^F -\tilde{E}_0^B
1450: \label{gapdelta}
1451: \end{equation}
1452: where $\tilde{E}_0^B$ is the lowest eigenvalue of the FMC diffusion operator.
1453: The greater this gap is, the faster the signal-over-noise ratio of the Monte Carlo simulation
1454: deteriorates, the full stability being obtained only when this gap vanishes.
1455: The ultimate goal of an efficient FMC algorithm is to reduce the Bose-Fermi gap from its bare
1456: value $\Delta_{B-F}= E_0^F-E_0^B$ to a value very close to zero (ideally, zero).
1457: The energies $\tilde{E}_0^B$ and $E_0^F$ can be calculated by exact diagonalization
1458: of the Fermion Monte Carlo operator, $G_{\text{FMC}}$.
1459: In practice, we have chosen here to extract this information from large-time behaviour of the
1460: denominator of the energy, Eq. (\ref{energyest}).
1461: In this regime
1462: the denominator behaves as in Eq. (\ref{denomfmcbehaviour}) where the reference
1463: energy is adjusted to keep the number of pairs constant, $E_T=\tilde{E}_0^B$.
1464: \begin{equation}
1465: \langle {\cal D}(t=k\tau) \rangle \propto e^{-(E_0^F-\tilde{E}_0^B)t}
1466: \label{Dbehavesim}
1467: \end{equation}
1468: The gap $E_0^F-\tilde{E}_0^B$ has been extracted from the large-k values of ${\cal D}(t=k\tau)$,
1469: a quantity calculated deterministically by iterating the matrix $G_{\text{FMC}}$.
1470: The results for different values of $c$ are reported in Table \ref{gapf(c)}.
1471: For both the correlated and uncorrelated processes it is found that the gap
1472: increases with $c$.
1473: The minimal gap is obtained for $c=0$, that is when the guiding functions are symmetric
1474: $\psi_G^+=\psi_G^-=\psi_S$.
1475: This result is easily explained from the fact that there are two factors which favour
1476: the cancellation of walkers.
1477: First, as we have already seen, the average meeting time is minimal when $c=0$ since,
1478: in this case, the overlap between the functions $\psi_G^+$ and $\psi_G^-$ is maximal. Furthermore,
1479: a full cancellation between the walkers is precisely obtained when $c=0$.
1480: In conclusion, the greatest stability is obtained for a symmetric guiding function.
1481:
1482: In Table \ref{gapf(N)} the gaps obtained at $c=0$ for different linear sizes $N$ are reported.
1483: This table shows that, in the limit of large grids, that is for a system
1484: close to the continuous model, the FMC algorithm reduces the gap by a factor $\sim 20$.
1485: Such a result corresponds to a huge gain in the
1486: stability since projection times about twenty times larger than in a standard nodal
1487: release method can be used.
1488:
1489: \subsection{FMC using a finite number of walkers: the stochastic approach}
1490:
1491: In the previous section the Fermion Monte Carlo method has been discussed and implemented
1492: by manipulating the exact Fermion Monte Carlo diffusion operator without making reference to
1493: any stochastic aspect (as already mentioned it is formally equivalent to use an infinite
1494: number of walkers).
1495: Of course, for non-trivial systems it is not possible to propagate exactly
1496: the dynamics of the FMC operator. Accordingly, a finite population of walkers is introduced and
1497: specific stochastic rules allowing to simulate in average the action of the FMC operator are
1498: defined. Now, the important point is that in practice -like in any DMC method- one does not
1499: sample exactly the dynamics of the FMC operator because
1500: of the population control step needed to keep the finite number of walkers roughly constant.
1501: \cite{umrigarJCP93,sorella98rec,khelif2000}
1502: This step introduces a small modification of the sampled diffusion operator
1503: which is at the origin of
1504: a systematic error known as the population control error. For a bosonic system, the error
1505: on the ground-state energy
1506: behaves as $\frac{1}{M}$ ($M$ is the average size of the population) and an
1507: extrapolation in $\frac{1}{M}$ can be done to obtain the exact energy.
1508: For a fermionic system, as we shall see below, this behaviour is qualitatively different and,
1509: furthermore, depends
1510: on the guiding function used. To have a precise estimate of the mathematical behaviour of
1511: the population control error is fundamental since, in practice,
1512: it is essential to be able to reach the exact fermi result using a reasonable number of walkers.
1513: As we shall see later, this will not be in general possible with FMC.
1514:
1515: In this section the Fermion Monte Carlo simulations are performed
1516: using Eqs.(\ref{eqfmc1},\ref{eqfmc2},\ref{eqfmc3},\ref{eqfmc4},\ref{eqfmc5}) which
1517: allow to propagate stochastically a population of $M$ walkers.
1518: The population is kept constant during the simulation by using the
1519: stochastic reconfiguration Monte Carlo (SRMC) method.\cite{sorella98rec,khelif2000}
1520: In short, the SRMC method is a DMC method in which a reconfiguration step replaces the
1521: branching step. A configuration step
1522: consists in drawing $M$ new walkers among the $M$ previous ones
1523: according to their respective Feynman-Kac weight (for the details, see the references
1524: given above).
1525:
1526: In Figure \ref{fig1} the time-averaged energy defined as
1527: \begin{equation}
1528: E(K) \equiv \frac{\sum_{k=1}^K {\cal N} (k)}{\sum_{k=1}^K {\cal D}(k)}
1529: \end{equation}
1530: is ploted as a function of $K$. In this formula
1531: ${\cal N}(k)$ and ${\cal D}(k)$ represent the numerator and denominator at
1532: iteration $k$ of the estimator (\ref{efmcN/D}) evaluated as an average over the population of
1533: pairs.
1534: In Figure \ref{fig2} we plot the time-averaged denominator given by
1535: \begin{equation}
1536: {\cal D}_K = \frac{1}{K} \sum_{k=1}^K {\cal D}(k).
1537: \label{cumdenom}
1538: \end{equation}
1539: The time dependence of this quantity is interesting since it can be used as a measure
1540: of the stability of the algorithm.\cite{kalosfermion00} As we have shown above,
1541: the algorithm is stable only when the reduced Bose-Fermi energy gap, $\tilde{\Delta}_{B-F}
1542: =E_0^F-\tilde{E}_0^B$
1543: is equal to zero. Equivalently, the denominator (\ref{cumdenom}) must converge to a constant
1544: different from zero. In our simulations the number of walkers was chosen to be $M=100$,
1545: a value which is much larger than the total number of states of the system (here, nine states).
1546: Of course, such a study is possible only for very simple systems.
1547: As seen on the Figures \ref{fig1},\ref{fig2}, and \ref{fig3} the results
1548: obtained in the case $c=0$ (symmetric guiding function) and $c\neq 0$ are qualitatively
1549: very different.
1550: Figure \ref{fig1} shows that, within statistical error bars, there is no systematic
1551: error on the energy when a symmetric guiding function is used, $c=0$.
1552: However, the price to pay is that the statistical fluctuations are very large.
1553: This point can be easily understood by looking at the behaviour of the denominator,
1554: Fig.\ref{fig2}. Indeed, this denominator vanishes at large times, thus
1555: indicating that the simulation is not stable.
1556: In sharp contrast, for $c=4$ (non-symmetric guiding functions), the
1557: statistical fluctuations are much more smaller (by a factor of about 40)
1558: but a systematic error appears for the energy.
1559: Furthermore, the denominator ploted in Fig.\ref{fig2} is seen to converge
1560: to a finite value. The stability observed in the case $c=4$ seems to confirm
1561: the results of Kalos {\sl et al.}\cite{kalosfermion00} for non-symmetric guiding
1562: functions ($c\ne 0$). However, the situation deserves a closer look.
1563: Indeed, the existence of this finite asymptotic
1564: value seems to be in contradiction with our theoretical analysis: the denominator should converge
1565: exponentially fast to zero, and the algorithm should not be stable ($\tilde{\Delta}_{B-F} > 0$).
1566: In fact, as we shall show now, the asymptotic value obtained for $c=4$ and the
1567: corresponding stability result from a population control error.
1568: To illustrate this point, we have computed the average denominator as a function of the
1569: population size $M$. Results are reported in Fig.\ref{fig3}. On this plot
1570: we compare the population dependence of the denominator (\ref{cumdenom}) for $c=4$ and
1571: for a much smaller value of $c=0.5$. The values of $M$ range from $M=100$ to $M=6400$.
1572: A first remark is that the population control error can be quite large
1573: and is much larger for $c=4$ than for $c=0.5$.
1574: In the Appendix it is shown that the theoretical asymptotic behaviour of the
1575: error as a function of $M$ is expected to be in $1/M$.
1576: In the $c=0.5$ case, the denominator is clearly seen to extrapolate to zero like $\frac{1}{M}$
1577: In the $c=4$ case, we can just say that the data are compatible
1578: with such a behaviour but even for the largest $M$ reported in Fig.\ref{fig3} ($M=6400$)
1579: this asymptotic regime is not yet reached. Much larger populations would be necessary.
1580: This result illustrates the great difficulty in reaching
1581: the asymptotic regime, even for such a simple system having only nine states.
1582: Stated differently, the stability observed when using non-symmetric guiding
1583: functions disappears for a large number of walkers, thus confirming that the
1584: stability obtained at finite $M$ is a control population artefact.
1585: Note that a large population control error on the denominator is not surprising.
1586: Indeed, when $c\ne 0$, the local energies of the guiding functions, Eq.(\ref{elgw}),
1587: have strong fluctuations because $\psi_G^\pm$ is far from any eigenstate of $H$ ($\psi_G^\pm$
1588: contains a symmetric and an antisymmetric components).
1589: In the case of a symmetric guiding function ($c=0$), the distribution of walkers is
1590: also symmetric at large times and, thus, the average of this distribution on the antisymmetric
1591: wavefunction $\psi_T$ must necessarily be zero. Consequently, in the $c=0$ case there is no
1592: control population error on the denominator, Fig.\ref{fig2}.
1593:
1594: An example of the behaviour of the energy as a function of the population size $M$
1595: is presented in figure \ref{fig4}.
1596: Some theoretical estimates of the energy bias dependence on $M$ are derived in the Appendix.
1597: Let us summarize the results obtained.
1598: When $c=0$ (use of a symmetric guiding function),
1599: the systematic error behaves as in a standard DMC calculation for a bosonic system
1600: \begin{equation}
1601: \delta E \propto \frac{1}{M}.
1602: \end{equation}
1603: However, the statistical fluctuations are exponentially large since the calculation is
1604: no longer stable.
1605: Now, when $c > 0$ the systematic error has a radically different behaviour.
1606: For a population size $M$ the control population error grows exponentially as a function
1607: of the projection time $t$
1608: \begin{equation}
1609: \delta E \propto \frac{1}{M}e^{t \tilde{\Delta}_{B-F}},
1610: \label{formx}
1611: \end{equation}
1612: where $\tilde{\Delta}_{B-F}$ is the reduced Bose-Fermi energy gap.
1613: This dependence of the control population error as a function of the projection
1614: time is of course pathological and is a direct consequence of the use of non-symmetric
1615: guiding functions. Now, because of the form (\ref{formx}) it is clear that a population size
1616: exponentially larger than the projection time is necessary to remove the systematic population
1617: control error.
1618: In practical calculations, for a given population of walkers $M$, one has to choose a finite
1619: projection time, $t$. This time has to be small enough to have a small finite population control
1620: error but, at the same time, large enough to extract the exact fermionic groundstate
1621: from the initial distribution of walkers.
1622: The best compromise is easily calculated and leads to the following expression of the
1623: systematic error as a function of the number of walkers (for more details, see the Appendix)
1624: \begin{equation}
1625: \delta E \propto \frac{1}{M^{\gamma}}
1626: \label{ebehav}
1627: \end{equation}
1628: where
1629: \begin{equation}
1630: \gamma \equiv \frac{ \Delta_F }{ \tilde{\Delta}_{B-F} + \Delta_F} < 1
1631: \label{gammabehav}
1632: \end{equation}
1633: and $\Delta_F$ is the usual Fermi gap
1634: (energy difference between the two lowest fermionic states).
1635:
1636: In figure \ref{fig4} some numerical results for the $c=4$
1637: and $c=0.5$ cases are presented.
1638: The number of walkers considered are $M=100,200,400,800$, and $M=1600$.
1639: No data are shown for the symmetric case, $c=0$, because of the very large
1640: error bars, see Fig.\ref{fig1}. The calculations have been done for the smallest
1641: system, $N=3$ (recall that the finite configuration space consists of
1642: only nine states) and for very large numbers of Monte Carlo steps (more than $10^8$).
1643: As it should be, the systematic errors
1644: are found to be larger for the $c=4$ case than for the $c=0.5$ case (note that the
1645: data corresponding to
1646: $M=800$ and $M=1600$ must not be considered because of their large statistical noise).
1647: The concavity of both curves confirms our theoretical result,
1648: $\gamma < 1$, Eqs.(\ref{ebehav},\ref{gammabehav}). However, it is clear that to get a quantitative
1649: estimate of this exponent is hopeless because of the rapid increase of error bars
1650: as a function of $M$. The only qualitative conclusion which can be drawn by looking at the
1651: curves is that $\gamma_{c=4} < \gamma_{c=0.5}$,
1652: in agreement with our formula, Eq.(\ref{gammabehav}). Finally, let us insist on the fact
1653: that, despite these very intensive calculations for a nine-state configuration
1654: space, no controlled extrapolation to the exact energy is possible.
1655:
1656: To summarize, when $\psi_G$ has an antisymmetric component, the error is expected
1657: to decrease -for $M$ large enough- very slowly as a function of the
1658: population size [algebraically with a (very) small exponent],
1659: while in the symmetric case the bias has a much more interesting $\frac{1}{M}$-behaviour.
1660: However, in this latter case the price to pay is the presence of an exponential growth
1661: of the statistical error. In both cases, and this is the fundamental point,
1662: the number of walkers needed to get a given accuracy grows pathologically.
1663: In addition, as illustrated by our data for the very simple model problem treated
1664: here, the asymtotic regimes corresponding to the $1/M^\gamma$-behaviour appear
1665: to be very difficult to reach in practice (very large values of $M$ are needed).
1666:
1667: \section{Conclusion and perspectives}
1668: The FMC method differs from the DMC method by correlating the diffusion of the walkers
1669: and introducing a cancellation procedure between positive and negative walkers whenever they
1670: meet. In this work we have shown that the Fermion Monte Carlo approach is exact but
1671: in general not stable. FMC can be viewed as belonging to the class of transient DMC methods,
1672: the most famous one being probably the nodal release approach.
1673: \cite{ceperleyalderPRL80,ceperleyalderPRL84}
1674: However, in contrast with the standard transient methods, FMC
1675: allows to reduce in a systematic way the fermi instability.
1676: The importance of this instability is directly related to the magnitude of some
1677: ``effective'' Bose-Fermi energy gap, $\tilde{\Delta}_{B-F}=E_0^F-\tilde{E}_0^B$, where
1678: $E_0^F$ is the exact Fermi energy and $\tilde{E}_0^B$ some effective Bose energy.
1679: We have seen that this gap is intimately connected to the
1680: the cancellation rate, that is to say, to the speed at which positive and negative walkers cancel.
1681: We have shown that $E_0^B < \tilde{E}_0^B < E_0^F$, where $E_0^B$ is
1682: the standard bosonic ground-state
1683: energy. As an important consequence, the closest $\tilde{E}_0^B$ is from the exact fermionic energy,
1684: the smoother the sign problem is. For the toy model considered, the reduction obtained for
1685: the instability is very large (orders of magnitude).
1686: For large dimensional systems, there are strong indications in favor also of an
1687: important reduction. A first argument is purely theoretical. In FMC the walkers within a pair
1688: $({\bf R}_i^+,{\bf R}_i^-)$ are correlated in a such a way that ${\bf R}_i^+-{\bf R}_i^-$
1689: makes a random move only in one dimension.
1690: As a result there is a high probability that the walkers meet in a finite time
1691: even if they move in a high-dimensional space.
1692: The second argument is numerical. As shown by previous authors, the impact of
1693: correlating walkers on the average meeting time is important even for
1694: much larger systems.\cite{kalosfermion00,CollettiFMC2005}
1695:
1696: We have shown that the recent introduction of nonsymmetric guiding functions in
1697: FMC introduces a large systematic error which goes to zero very slowly
1698: as a function of the population size [$\sim 1/M^\gamma$, $\gamma =
1699: \Delta_F/(\tilde{\Delta}_{B-F} + \Delta_F)$ and $\Delta_F= E_1^F -E_0^F$ is the usual
1700: fermionic gap].
1701: For an infinite number of walkers, this systematic
1702: error is removed and the algorithm recovers the Fermi instability.
1703: Morever, we have shown that using such guiding functions does not in general improve the
1704: stability. For a large enough number of walkers, the
1705: simulation can be less stable than the simulation using a symmetric guiding function.
1706: Finally, it is important to emphasize that the conclusion of this work is that the
1707: FMC algorithm is not a solution
1708: to the sign problem. However, it is a promising way toward ``improved'' transient methods.
1709: As a transient method, FMC is expected to converge much better than a standard nodal
1710: release method. We are presently working in this direction.
1711:
1712: \section*{ACKNOWLEDGMENTS}
1713: The authors warmly thank Malvin Kalos and Francesco Pederiva for
1714: numerous useful discussions and for having inspired us the present work.
1715: This work was supported by the ``Centre National de la Recherche
1716: Scientifique'' (CNRS), Universit\'e Pierre et Marie Curie (Paris VI), Universit\'e
1717: Pauls Sabatier (Toulouse III), and Universit\'e Denis Diderot (Paris VII).
1718: Finally, we would like to acknowledge computational support from IDRIS (CNRS, Orsay) and
1719: CALMIP (Toulouse).
1720:
1721: \appendix
1722: \section*{APPENDIX: Population control error in FMC}
1723: FMC, like any Monte Carlo method using a branching process,
1724: suffers from a so-called population control bias. This systematic error appears
1725: because the branching rules (creation/annihilation of walkers) are implemented using a
1726: population consisting of a large but {\it finite} number of walkers. Nothing preventing
1727: the population size from implosing or exploding, a population control step is required
1728: to keep the average number of walkers finite.
1729: A standard strategy to cope with this difficulty consists in introducing a
1730: time-dependent reference energy whose effect is to slightly modify the elementary weights
1731: of each walker by a common multiplicative factor (close to one)
1732: so that the total weight of the population remains nearly constant during the
1733: simulation. This step, which introduces some correlation between walkers and, therefore,
1734: slightly modifies the stationary density, must be performed
1735: very smoothly to keep the population control error as small as possible.
1736: In practice, for standard DMC calculations done with accurate trial wavefunctions
1737: and population sizes large enough, the error
1738: is found to be very small, in general much smaller than the statistical fluctuations.
1739: As a consequence, the presence of a population control bias is usually not considered as critical.
1740: Here, the situation is rather different.
1741: In FMC the use of bosonic-type guiding functions introduces very large
1742: fluctuations of the local energy and the cancellation rules a very small signal-over-noise
1743: ratio for fermionic properties. In this case, it is not clear whether
1744: the bias can be kept small with a reasonable number of walkers.
1745:
1746: In this section we present an estimate of the population control bias in FMC. As we shall
1747: see our estimate shows that the sign problem is actually not solved but attenuated in FMC
1748: (an {\it exponentially} large number of walkers is needed to maintain
1749: a constant bias as the number of electrons is increased).
1750: The derivation presented in this section is very
1751: general: it is valid for any exact fermion QMC method based on the use of a nodeless bosonic-type
1752: reference process and some projection to extract the Fermi ground-state. Accordingly,
1753: we have chosen not to use the specific framework and notations of FMC but,
1754: instead, notations of a general DMC algorithm (transient method). Of course, we do not
1755: need FMC to introduce nonsymmetric guiding functions.
1756: The adaptation of what follows to FMC is straightforward.
1757:
1758: In quantum Monte Carlo we evaluate stochastically
1759: the following expression for the lowest eigenstate of energy $E_0^F$
1760: \begin{equation}
1761: E_F^t = \frac{ \langle \psi_T \mid H e^{-t(H-E_T)} f_0 \rangle}
1762: {\langle \psi_T \mid e^{-t(H-E_T)} f_0 \rangle}
1763: \label{energt}
1764: \end{equation}
1765: where $f_0$ is some positive initial distribution and $\psi_T$ an approximation of
1766: the eigenstate with energy $E_0^F$.
1767: Here, we deal with a fermionic problem so $\psi_T$ must be antisymmetric.
1768: The expression (\ref{energt}) gives the exact energy $E_0^F$ only when taking the limit $t \to \infty$. In practice for a finite $t$, there is a systematic error \begin{equation}
1769: \Delta E_F^t \equiv E_F^t-E_0^F \propto e^{-\Delta_F t} \equiv e^{-(E_1^F-E_0^F) t}
1770: \label{deltaeat}
1771: \end{equation}
1772: where $E_1^F$ is the energy of the first excited state in the antisymmetric sector and,
1773: $\Delta_F$, the fermionic energy gap.
1774: For an exact algorithm with one walker (e.g. Pure Diffusion Monte Carlo,
1775: \cite{Cafclav188,Cafclav288}) one computes the R.H.S of (\ref{energt}) by evaluating the following expression
1776: \begin{equation}
1777: E_F^t = \frac{\left\langle \frac{H \psi_T}{\psi_G} [R(t)] e^{-\int_0^t ds (E_L[R(s)]-E_T)}
1778: \right\rangle}
1779: {\left\langle \frac{ \psi_T}{\psi_G} [R(t)] e^{-\int_0^t ds (E_L[R(s)]-E_T) }
1780: \right\rangle}
1781: \label{eexact}
1782: \end{equation}
1783: where $\psi_G$ is the guiding function, strictly positive, with eventually an antisymmetric component.
1784: We have also introduced in this expression, the local energy of the guiding function
1785: \begin{equation}
1786: E_L = \frac{H\psi_G}{\psi_G}
1787: \end{equation}
1788: The integral in (\ref{eexact}) is done over the drifted random walks going from $0$ to $t$.
1789: To simplify the notations we note (\ref{eexact}) as follows
1790: \begin{equation}
1791: E_F^t = \frac{\left\langle h^t W^t \right\rangle}
1792: {\left\langle p^t W^t \right\rangle}
1793: \end{equation}
1794: where
1795: \begin{eqnarray}
1796: h^t & \equiv & \frac{H \psi_T}{\psi_G} [R(t)] \\
1797: W^t & \equiv & e^{-\int_0^t ds E_L[R(s)]} \\
1798: p^t & \equiv & \frac{\psi_T}{\psi_G} [R(t)]
1799: \end{eqnarray}
1800: For $M$ independant walkers $R_i$ one has
1801: \begin{equation}
1802: E_F^t = \frac{\left\langle \frac{1}{M} \sum_i h_i^t W_i^t \right\rangle}
1803: {\left\langle \frac{1}{M} \sum_i p_i W_i^t \right\rangle}
1804: \label{eareconf}
1805: \end{equation}
1806: In the analysis presented here based on a population of $M$ the walkers branched
1807: according to their relative multiplicities, (the $M$ walkers are therefore no longer independant), one replace the individual weight $W_i$ by a global weight
1808: \cite{khelif2000}
1809: \begin{equation}
1810: \bar{W}^t = \frac{1}{M}\sum_i W_i^t
1811: \end{equation}
1812: As a result the energy may be written as
1813: \begin{equation}
1814: E_F^t = \frac{\left\langle \bar{h}^t \bar{W}^t \right\rangle}
1815: {\left\langle \bar{p}^t \bar{W}^t \right\rangle}
1816: \label{reconf}
1817: \end{equation}
1818: where
1819: \begin{eqnarray}
1820: \bar{h}^t & \equiv & \frac{1}{M}\sum_i h_i^t \\
1821: \bar{p}^t & \equiv & \frac{1}{M}\sum_i p_i^t
1822: \end{eqnarray}
1823: Expression (\ref{reconf}) is exact when the weights $\bar{W}^t$ are included.
1824: A control population error arises when one does not take into account the weights
1825: in expression (\ref{reconf}). This population control error is thus given by
1826: \begin{eqnarray}
1827: \Delta E_F^M & = & \frac{\left\langle \bar{h}^t \right\rangle}
1828: {\left\langle \bar{p}^t \right\rangle}
1829: - \frac{\left\langle \bar{h}^t \bar{W}^t \right\rangle}
1830: {\left\langle \bar{p}^t \bar{W}^t \right\rangle}
1831: \\
1832: & = & \frac{\left\langle \bar{h}^t \right\rangle}
1833: {\left\langle \bar{p}^t \right\rangle} -
1834: \frac{\text{cov} ( \bar{h}^t, \bar{W}^t) + \left\langle \bar{h}^t\right\rangle\left\langle \bar{W}^t\right\rangle}
1835: {\text{cov}( \bar{p}^t, \bar{W}^t) + \left\langle \bar{p}^t\right\rangle \left\langle \bar{W}^t\right\rangle }
1836: \end{eqnarray}
1837: or, after normalizing $\bar{W}^t$ in such a way that $ \langle \bar{W}^t\rangle =1$
1838: (for example, by suitably adjusting the reference energy $E_T$),
1839: \begin{equation}
1840: \Delta E_F^M = \frac{\left\langle \bar{h}^t \right\rangle}
1841: {\left\langle \bar{p}^t \right\rangle}
1842: - \frac{\text{cov} ( \bar{h}^t, \bar{W}^t) + <\bar{h}^t>}
1843: {\text{cov}( \bar{p}^t, \bar{W}^t) + <\bar{p}^t> }
1844: \label{eamcov}
1845: \end{equation}
1846: This is our basic formula expressing the systematic error at time $t$
1847: resulting from the use of a finite population. Now, let us evaluate this expression
1848: in the large time $t$ and large $M$ regimes.
1849: First, we consider the two denominators appearing in the R.H.S. of Eq.(\ref{eamcov}).
1850: Let us begin with the denominator of the second ratio:
1851: \begin{equation}
1852: D_e \equiv \text{cov}( \bar{p}^t, \bar{W}^t) + <\bar{p}^t>
1853: \label{D2def}
1854: \end{equation}
1855: Because this denominator is nothing but the denominator of the R.H.S. of Eq.(\ref{eexact})
1856: we can conclude that $D_e$ does not depend on the population size $M$ and that
1857: it vanishes exponentially fast
1858: \begin{equation}
1859: D_e = K_e e^{-(E_0^F-E_T)t},
1860: \label{D2t}
1861: \end{equation}
1862: where $E_T$ is the reference energy, $E_T = E_0^B$ for a nodal release-type method,
1863: and $E_T = \tilde{E}_0^B > E_0^B$ for the FMC method.
1864: Let us now look at the other denominator of Eq.(\ref{eamcov})
1865: \begin{equation}
1866: D_a \equiv \langle \bar{p}^t\rangle.
1867: \label{D1def}
1868: \end{equation}
1869: This denominator is the usual quantity evaluated during the simulation. It is
1870: an approximate quantity since it does not include the corrective weights.
1871: The asymptotic behaviour of $D_a$ depends on $\psi_G$. We distinguish two cases:
1872:
1873: (i) If $\psi_G$ is symmetric ($c=0$), the stationary density ($t$ large enough) is symmetric.
1874: Consequently, $D_a$, which is the average of an antisymmetric function, converges to zero
1875: exponentially fast at large times. For $M$ large enough, in a regime where the dynamics
1876: is close to the exact dynamics of the Hamiltonian, we know that the convergence is
1877: given by
1878: \begin{equation}
1879: D_a = K_a(M) e^{-\tilde{\Delta}_{B-F} t},
1880: \label{D1t}
1881: \end{equation}
1882: where the coefficient $K_a$ depends on $M$ in general. This coefficient will be determined later.
1883: >From equations (\ref{D1t}) and (\ref{D2t}), one can evaluate the error on the denominator
1884: \begin{equation}
1885: D_a-D_e = (K_a (M)-K_e) e^{-\tilde{\Delta}_{B-F} t}.
1886: \label{difD1}
1887: \end{equation}
1888: We also know from the definitions of $D_a$ and $D_e$[(\ref{D1def}), (\ref{D2def})] that
1889: the difference $D_a -D_e$ is a covariance of two averages
1890: \begin{equation}
1891: D_a -D_e = \text{cov}( \bar{p}^t, \bar{W}^t)
1892: \end{equation}
1893: which, due to the central-limit theorem, behaves as
1894: \begin{equation}
1895: D_a -D_e = \frac{1}{M} {\cal C} (t)
1896: \label{difD2}
1897: \end{equation}
1898: where ${\cal C} (t)$ is some function of $t$.
1899: Identifying (\ref{difD1}) and (\ref{difD2}), one finally obtains a determination of $K_a$.
1900: Finally, we obtain the following behaviour for the systematic error on the denominator
1901: \begin{equation}
1902: D_a -D_e = \text{cov}( \bar{p}^t, \bar{W}^t) \propto \frac{1}{M} e^{-\tilde{\Delta}_{B-F} t}
1903: \end{equation}
1904: (ii) If $\psi_G$ is not symmetric ($c\ne 0$) the stationary density has an antisymmetric
1905: component and $D_a$ converges to a constant different from zero at large times.
1906: Of course, this constant depends on the number of walkers $M$.
1907: This dependence can be easily found by using the central limit theorem as before
1908: \begin{equation}
1909: D_a -D_e = \text{cov}( \bar{p}^t, \bar{W}^t) = K \frac{1}{M}.
1910: \end{equation}
1911: Finally, we have just proved that, when the guiding function is not symmetric,
1912: the asymptotic behaviour of the denominator is $D_a \propto \frac{1}{M}$.
1913: This important result is in agreement with the numerical data shown in figure (\ref{fig3}).
1914: Using exactly the same arguments, the asymptotic behaviour (large $M$, large $t$)
1915: of the difference of the two numerators in the R.H.S. of expression (\ref{eamcov})
1916: is found to be the same as $D_a-D_e$
1917: \begin{equation}
1918: N_a-N_e = \text{cov}( \bar{h}^t, \bar{W}^t) \propto D_a -D_e.
1919: \end{equation}
1920:
1921: Now, we are ready to write down the expression of the systematic population control error,
1922: $\Delta E_F^M$. For $M$ large enough, $\Delta E_F^M$ is well approximated by its first-order
1923: contribution in the $\frac{1}{M}$ expansion. Here also, we need to
1924: distinguish between the nature of the guiding function
1925:
1926: (i) If $\psi_G$ is symmetric one easily obtains
1927: \begin{equation}
1928: \Delta E_F^M \propto \frac{1}{M}
1929: \label{deamsym}
1930: \end{equation}
1931: (ii) If $\psi_G$ has an antisymmetric component
1932: \begin{equation}
1933: \Delta E_F^M \propto \frac{1}{M} e^{\tilde{\Delta}_{B-F} t}
1934: \label{deamnonsym}
1935: \end{equation}
1936:
1937: Let us now evaluate the total systematic error resulting from using a finite time $t$ and
1938: a finite population size $M$
1939: \begin{equation}
1940: \Delta E_0^F = \Delta E_F^M + \Delta E_F^t
1941: \label{deltaeatot}
1942: \end{equation}
1943: where $\Delta E_F^t$, the systematic error coming from a finite simulation time, is given by
1944: Eq.(\ref{deltaeat}) and $\Delta E_F^M$ is the error just discussed.
1945: The strategy consists in determining, for a given systematic error $\Delta E_0^F \sim \epsilon$,
1946: what is the time $t$ and the number of walkers $M$ one should consider.
1947: The condition for the total systematic error to be of order $\epsilon$ is that both terms in
1948: (\ref{deltaeatot}) are also of order $\epsilon$
1949: \begin{eqnarray}
1950: \Delta E_F^t & \sim & \epsilon \\
1951: \label{eatpsi}
1952: \Delta E_F^M & \sim & \epsilon.
1953: \label{eampsi}
1954: \end{eqnarray}
1955: This is true because no error compensation are present, $\Delta E_F^t$ and $\Delta E_F^M$
1956: being generally of the same sign (both positive). Our numerical results on
1957: the toy model give an illustration of this property.
1958: >From both equations (\ref{eatpsi}) and (\ref{deltaeat}) one can deduce the simulation time
1959: corresponding to such a systematic error
1960: \begin{equation}
1961: t \sim -\frac{\ln \epsilon}{\Delta_F}.
1962: \label{teps}
1963: \end{equation}
1964: In other words, to obtain an error of order $\epsilon$ it is sufficient to stop the simulation
1965: at a time $t$ of order (\ref{teps}).
1966: Now, let us come to the number of walkers $M$ needed.
1967: If $\psi_G$ is symmetric, we already know from expression (\ref{deamsym}) that
1968: the systematic error does not depend on the projection time and that the number
1969: of walkers $M$ and the systematic error $\epsilon$ are related as follows
1970: \begin{equation}
1971: M \propto \frac{1}{\epsilon}.
1972: \end{equation}
1973: If $\psi_G$ is not symmetric, the equation (\ref{eampsi}) can be easily solved.
1974: Replacing in Eq.(\ref{eampsi}), $\Delta E_F^M$ by its expression (\ref{deamnonsym})
1975: and using the relation (\ref{teps}) one finally finds the number of walkers required
1976: to obtain a systematic error $\epsilon$.
1977: \begin{equation}
1978: M \propto \epsilon^{-\frac{\tilde{\Delta}_{B-F}}{\Delta_F}-1}.
1979: \label{Meps}
1980: \end{equation}
1981: Let us make some important comments. First, note that in this formula the dependence on
1982: the guiding function is not included in the exponent, only in the prefactor.
1983: Second, this formula shows that the FMC algorithm reduces the systematic error by lowering
1984: the exponent. As already mentioned, the gap is equal to $E_0^F-E_0^B$ in a standard
1985: release node method and equal to $E_0^F-\tilde{E}_0^B <E_0^F-E_0^B$ in FMC.
1986: Third, in the zero-limit gap, the $\frac{1}{M}$ behaviour of the systematic error is recovered.
1987: This formula shows that the number of walkers needed for a given accuracy, $\epsilon$,
1988: grows exponentially with respect to the number of electrons. Indeed, although
1989: the gap is indeed reduced by FMC, there is no reason not to believe that it will still be
1990: proportional to the number of electrons. In consequence, the ``sign problem'' fully remains in FMC.
1991:
1992: Finally, let us write the systematic error as a function of the finite population $M$ by
1993: inverting the preceding equation (\ref{Meps})
1994: \begin{equation}
1995: \epsilon \propto M^{-\gamma}
1996: \end{equation}
1997: where
1998: \begin{equation}
1999: \gamma \equiv
2000: {\frac{\Delta_F}{\tilde{\Delta}_{B-F}+ \Delta_F}}
2001: \end{equation}
2002: This latter equation shows very clearly the respective role played by the Fermi gap, $ \Delta_F$,
2003: and the reduced Bose-Fermi gap, $\tilde{\Delta}_{B-F}$.
2004: %\bibliographystyle{physrev}
2005: %\bibliography{Assaraf}
2006:
2007: \begin{thebibliography}{99}
2008: \bibitem{kalosfermion99}
2009: M.H. Kalos and F. Pederiva, in {\it NATO-ASI Conference Quantum Monte
2010: Carlo Methods in Physics and Chemistry, Cornell}, edited by M.P.
2011: Nightingale and C.J. Umrigar, NATO-ASI (Kluwer, Dordrecht, The
2012: Netherlands, 1999).
2013: \bibitem{ceperleykalos79}
2014: D.M. Ceperley and M.H. Kalos, {\it Monte Carlo methods in statistical
2015: physics} (Springer-Verlag, Berlin, 1979).
2016: \bibitem{schmittkalos84}
2017: K.E. Schmidt and M.H. Kalos, {\it Monte Carlo Methods in Statistical
2018: Physics}, edited by K. Binder, vol. 2 (Springer, Berlin, 1984).
2019: \bibitem{lesterbook94}
2020: B.L. Hammond and W.A. Lester, {\it Monte Carlo Methods in Ab Initio
2021: Quantum Chemisty} (World Scientific, Singapore, 1994).
2022: \bibitem{reynoldsJCP82}
2023: P.J. Reynolds, D.M. Ceperley, B.J. Alder, and W. Lester, J. Chem. Phys. {\bf 77}, 5593 (1982).
2024: \bibitem{andersonJCP75}
2025: J.~B. {Anderson}, J. Chem. Phys. {\bf 63}, 1499 (1975).
2026: \bibitem{andersonJCP76}
2027: J.~B. {Anderson}, J. Chem. Phys. {\bf 65}, 4121 (1976).
2028: \bibitem{filippiJCP96}
2029: C.~{Filippi} and C.~J. {Umrigar}, J. Chem. Phys. {\bf 105}, 213 (1996).
2030: \bibitem{surf1}
2031: S.~B. {Healy}, C.~{Filippi}, P.~{Kratzer}, E.~{Penev}, and M.~{Scheffler},
2032: Phys. Rev. Lett. {\bf 87}, 016105 (2001).
2033: \bibitem{surf2}
2034: C.~{Filippi}, S.~B. {Healy}, P.~{Kratzer}, E.~{Pehlke}, and M.~{Scheffler},
2035: Phys. Rev. Lett. {\bf 89}, 166102 (2002).
2036: \bibitem{needs1}
2037: P.~R.~C. {Kent}, M.~D. {Towler}, R.~J. {Needs}, and G.~Rajagopal, Phys.Rev.B
2038: {\bf 62}, 15394 (2000).
2039: \bibitem{needs2}
2040: A.~R. {Porter}, M.~D. {Towler}, and R.~J. {Needs}, Phys.Rev.B {\bf 64}, 035320
2041: (2001).
2042: \bibitem{MitasGrossman00}
2043: L.~{Mitas}, J.~C. {Grossman}, I.~{Stich}, and J.~{Tobik}, Phys. Rev. Lett. {\bf
2044: 84}, 1479 (2000).
2045: \bibitem{ceperleyalderPRL84}
2046: D.~M. {Ceperley} and B.~J. {Alder}, J. Chem. Phys. {\bf 81}, 5833 (1984).
2047: \bibitem{bernuJCP90}
2048: B.~{Bernu}, D.~{Ceperley}, and W.~{Lester}, J. Chem. Phys. {\bf 93}, 552
2049: (1990).
2050: \bibitem{bernuJCP91}
2051: B.~{Bernu}, D.~{Ceperley}, and W.~{Lester}, J. Chem. Phys. {\bf 95}, 7782
2052: (1991).
2053: \bibitem{caffarelJCP1992}
2054: M.~{Caffarel} and D.~{Ceperley}, J. Chem. Phys. {\bf 97}, 8415 (1992).
2055: \bibitem{kalosfermion00}
2056: M.~H. {Kalos} and F.~{Pederiva}, Phys. Rev. Lett. {\bf 85}, 3547 (2000).
2057: \bibitem{arnow1982}
2058: D.~Arnow, M.~H. {Kalos}, M.~A. {Lee}, and K.~E. {Schmidt}, J. Chem. Phys. {\bf
2059: 77}, 5562 (1982).
2060: \bibitem{Liu1994}
2061: Z.~{Liu}, S.~{Zhang}, and M.~H. {Kalos}, Phys. Rev. E {\bf 50}, 3220 (1994).
2062: \bibitem{CollettiFMC2005}
2063: L.~{Colletti}, F.~{Pederiva}, and M.~H. {Kalos}, Fermion Monte Carlo
2064: Calculation of liquid-3He in {\it Exact Monte Carlo Method for Continuum
2065: Fermion Systems} (2005).
2066: \bibitem{umrigarJCP93}
2067: C.~J. {Umrigar}, M.~P. {Nightingale}, and K.~J. {Runge}, J. Chem. Phys. {\bf
2068: 99}, 2865 (1993).
2069: \bibitem{sorella98rec}
2070: M.~{Calandra Buonaura} and S.~{Sorella}, Phys. Rev. B {\bf 57}, 11446 (1998).
2071: \bibitem{khelif2000}
2072: R.~{Assaraf}, M.~{Caffarel}, and A.~{Khelif}, Phys. Rev. E {\bf 61}, 4566
2073: (2000).
2074: \bibitem{klv}
2075: M.~H. {Kalos}, D.~{Levesque}, and L.~{Verlet}, Phys. Rev. A {\bf 9}, 2178
2076: (1974).
2077: \bibitem{rothstein97}
2078: P.~{Langfelder}, S.~M. {Rothstein}, and J.~{Vrbik}, J. Chem. Phys. {\bf 107},
2079: 8525 (1997).
2080: \bibitem{Rothstein2004}
2081: I.~{Bosá} and S.~M. {Rothstein}, J. Chem. Phys. {\bf 121}, 4486 (2004).
2082: \bibitem{force00}
2083: R.~{Assaraf} and M.~{Caffarel}, J. Chem. Phys. {\bf 113}, 4028 (2000).
2084: \bibitem{filippi2000}
2085: C.~{Filippi} and C.~J. {Umrigar}, Phys. Rev. B {\bf 61}, R16291 (2000).
2086: \bibitem{caffrerat}
2087: M.~{Caffarel}, M.~{R\'erat}, and C.~{Pouchan}, Phys. Rev. A. {\bf 47}, 3704
2088: (1993).
2089: \bibitem{rcaf03}
2090: R.~{Assaraf} and M.~{Caffarel}, J. Chem. Phys. {\bf 119}, 10536 (2003).
2091: \bibitem{casalegno2003}
2092: M.~{Casalegno}, M.~{Mella}, and A.~M. {Rappe}, J. Chem. Phys. {\bf 118}, 7193
2093: (2003).
2094: \bibitem{ceperleyalder86}
2095: D.~M. {Ceperley} and B.~J. {Alder}, Science {\bf 231}, 555 (1986).
2096: \bibitem{ceperleyalderPRL80}
2097: D.~M. {Ceperley} and B.~J. {Alder}, Phys. Rev. Lett. {\bf 45}, 566 (1980).
2098: \bibitem{Grossman2002}
2099: J.~{Grossman}, J. Chem.Phys. {\bf 117}, 1434 (2002).
2100: \bibitem{Cafclav188}
2101: M.~{Caffarel} and P.~{Claverie}, J. Chem. Phys. {\bf 88}, 1088 (1988).
2102: \bibitem{Cafclav288}
2103: M.~{Caffarel} and P.~{Claverie}, J. Chem. Phys. {\bf 88}, 1100 (1988).
2104: \end{thebibliography}
2105:
2106: \newpage
2107: \begin{table}[htp]
2108: \noindent
2109: \caption{Average meeting times $\frac{\langle T\rangle}{N}$ for the correlated and
2110: uncorrelated cases in the non-symmetric case $c=4$, Eq.(\ref{psigpm}). In this example,
2111: $x_{max}=3$, Eq.(\ref{xmax}), and $\tau =0.9\tau_{max}$,
2112: where $\tau_{max}$ is the maximal time-step defined in Eq.(\ref{taumax}).}
2113: \begin{tabular}{|c|c|c|}
2114: \hline
2115: Linear size $N$ & $\frac{\langle T\rangle}{N}$ Uncorr. & $\frac{\langle T\rangle}{N}$ Corr. \\
2116: \hline
2117: \hline
2118: $N=3 $ & 2152(23) & 134(1) \\
2119: $N=5 $ & 2162(20) & 92(1) \\
2120: $N=7 $ & 2234(26) & 75(1) \\
2121: $N=9 $ & 2834(27) & 76(1) \\
2122: $N=11$ & 3546(38) & 82(1) \\
2123: $N=13$ & 4214(41) & 88(1) \\
2124: $N=15$ & & 94(1) \\
2125: $N=17$ & & 102(1) \\
2126: \hline
2127: \hline
2128: \end{tabular}
2129: \raggedright
2130: \label{table1}
2131: \end{table}
2132:
2133: \begin{table}[htp]
2134: \noindent
2135: \caption{Average meeting times $\frac{\langle T\rangle}{N}$ for the correlated and
2136: uncorrelated cases in the symmetric case ($c=0$, symmetric guiding function), Eq.(\ref{psigpm}).
2137: In this example, $x_{max}=3$,
2138: Eq.(\ref{xmax}), and $\tau =0.9\tau_{max}$,
2139: where $\tau_{max}$ is the maximal time-step defined in Eq.(\ref{taumax}).}
2140: \begin{tabular}{|c|c|c|}
2141: \hline
2142: Linear size $N$ & $\frac{\langle T\rangle}{N}$ Uncorr. & $\frac{\langle T\rangle}{N}$ Corr. \\
2143: \hline
2144: \hline
2145: $N=3 $ & 3.32(2) & 1.634(9) \\
2146: $N=5 $ & 5.64(6) & 2.27(1) \\
2147: $N=7 $ & 7.6(1) & 2.65(1.6) \\
2148: $N=9 $ & 10.3(1) & 3.32(2.5) \\
2149: $N=11$ & 12.95(8)& 4.18(4) \\
2150: $N=13$ & 15.7(1) & 4.78(4) \\
2151: $N=15$ & 18.5(2) & 5.52(6) \\
2152: $N=17$ & 21.6(2) & 6.26(7) \\
2153: \hline \hline
2154: \end{tabular}
2155: \raggedright
2156: \label{table2}
2157: \end{table}
2158:
2159: \begin{table}[htp]
2160: \noindent
2161: \caption{$N=3$. Reduced Bose-Fermi gap $\tilde{\Delta}_{B-F}$, Eq.(\ref{gapdelta}),
2162: with or without correlation for different values of $c$.
2163: The average meeting times are also indicated in parentheses. The bare
2164: Bose-Fermi gap, ${\Delta_{B-F}}$, is $\sim 0.7695$
2165: (here, $x_{max}=3.$ and $\tau=0.09\tau_{max}$).
2166: }
2167: \begin{tabular}{|c|c|c|}
2168: \hline
2169: Value of c &
2170: Correlated process &
2171: Uncorrelated process\\
2172: \hline
2173: \hline
2174: $c=0\;\;\;\;$ & $\tilde{\Delta}_{B-F}= 0.0366\;\;\;\;$ [$\frac{\langle T\rangle}{N}$= 1.634(9)]&
2175: $\tilde{\Delta}_{B-F}= 0.1629\
2176: ;\;$
2177: [$\frac{\langle T\rangle}{N}$= 3.32(2)]$\;\;\;\;$\\
2178: $c=1\;\;\;\;$ & $\tilde{\Delta}_{B-F}= 0.0917\;\;\;\;$ [$\frac{\langle T\rangle}{N}$= 6.37(7)] &
2179: $\tilde{\Delta}_{B-F}= 0.2540\
2180: ;\;$
2181: [$\frac{\langle T\rangle}{N}$= 16.5(2)]$\;\;\;\;$\\
2182: $c=2\;\;\;\;$ & $\tilde{\Delta}_{B-F}= 0.1336\;\;\;\;$ [$\frac{\langle T\rangle}{N}$= 24.6(2)] &
2183: $\tilde{\Delta}_{B-F}= 0.2277\
2184: ;\;$
2185: [$\frac{\langle T\rangle}{N}$= 130.5(5)]$\;\;\;\;$\\
2186: $c=3\;\;\;\;$ & $\tilde{\Delta}_{B-F}= 0.1026\;\;\;\;$ [$\frac{\langle T\rangle}{N}$= 62.9(3)] &
2187: $\tilde{\Delta}_{B-F}= 0.1981\
2188: ;\;$
2189: [$\frac{\langle T\rangle}{N}$= 627(3) ]$\;\;\;\;$\\
2190: $c=4\;\;\;\;$ & $\tilde{\Delta}_{B-F}= 0.1092\;\;\;\;$ [$\frac{\langle T\rangle}{N}$= 134(1)] &
2191: $\tilde{\Delta}_{B-F}= 0.1787\
2192: ;\;$
2193: [$\frac{\langle T\rangle}{N}$= 2152(23)]$\;\;\;\;$\\
2194: \hline \hline
2195: \end{tabular}
2196: \raggedright
2197: \label{gapf(c)}
2198: \end{table}
2199:
2200: \begin{table}[htp]
2201: \noindent
2202: \caption{Comparison between the reduced Bose-Fermi gap, $\tilde{\Delta}_{B-F}$, and
2203: the bare Bose-Fermi gap, ${\Delta}_{B-F}$, in the symmetric
2204: case ($c=0$, symmetric guiding function) as a function of $N$}
2205: \begin{tabular}{|c|c|c|c|}
2206: \hline
2207: Value of $N$ & $\tilde{\Delta}_{B-F}$ & ${\Delta_{B-F}}$ & Gap ratio \\
2208: \hline
2209: \hline
2210: $N=3$ & $\tilde{\Delta}_{B-F}= 0.0366$ & ${\Delta_{B-F}}= 0.7695$ &
2211: $\frac{\tilde{\Delta}_{B-F}}{{\Delta_{B-F}}}=0.0476$ \\
2212: $N=5$ & $\tilde{\Delta}_{B-F}= 0.0516$ & ${\Delta_{B-F}}= 1.0195$ &
2213: $\frac{\tilde{\Delta}_{B-F}}{{\Delta_{B-F}}}=0.0506$ \\
2214: $N=7$ & $\tilde{\Delta}_{B-F}= 0.0577$ & ${\Delta_{B-F}}= 1.1782$ &
2215: $\frac{\tilde{\Delta}_{B-F}}{{\Delta_{B-F}}}=0.0490$ \\
2216: \hline \hline
2217: \end{tabular}
2218: \raggedright
2219: \label{gapf(N)}
2220: \end{table}
2221:
2222: \begin{center}
2223: \begin{figure}[htp]
2224: \includegraphics[height=12cm,width=12cm,angle=-90]{Fig1.ps}
2225: \caption{$N=3$ Energy estimator as a function of the projection time
2226: (number of iterations $K$). Comparaison between $c=0$ (large fluctuations) and
2227: $c=4$ (small fluctuations).
2228: The exact energy is $E_0^F=1.86822..$. Number of walkers
2229: $M=100$. Number of Monte Carlo steps: $4.10^7$.
2230: }
2231: \label{fig1}
2232: \end{figure}
2233: \end{center}
2234:
2235: \begin{center}
2236: \begin{figure}[htp]
2237: \includegraphics[height=12cm,width=12cm,angle=-90]{Fig2.ps}
2238: \caption{$N=3$ Denominator as a function of the projection time (number of
2239: iterations $K$). Comparaison between $c=4$ (upper curve) and
2240: $c=0$ (lower curve). Number of walkers
2241: $M=100$. Number of Monte Carlo steps: $4.10^7$. }
2242: \label{fig2}
2243: \end{figure}
2244: \end{center}
2245:
2246: \begin{center}
2247: \begin{figure}[htp]
2248: \includegraphics[height=12cm,width=12cm,angle=-90]{Fig3.ps}
2249: \caption{$N=3$, time-averaged denominator, Eq.(\ref{cumdenom}), as a function of $1/M$. }
2250: \label{fig3}
2251: \end{figure}
2252: \end{center}
2253:
2254: \begin{center}
2255: \begin{figure}[htp]
2256: \includegraphics[height=12cm,width=8cm,angle=-90]{Fig4.ps}
2257: \caption{$N=3$ Energy, Eq.(\ref{energyest}), as a function of $1/M$ for the $c=0$ and
2258: $c=4$ cases. Exact energy: $E_0^F$=1.86822...(horizontal solid line)}
2259: \label{fig4}
2260: \end{figure}
2261: \end{center}
2262: \end{document}
2263: