physics0310035/f.tex
1: \documentstyle[manuscript,aps,epsfig]{revtex}
2: \textwidth=16truecm
3: \renewcommand{\baselinestretch} {1.7}
4: \catcode`\@=11
5: \newcommand{\fcaption}[1]{\vspace{1ex}
6:         \refstepcounter{figure}
7:         \setbox\@tempboxa = \hbox{\footnotesize {\bf ig.~\thefigure.} #1}
8:         \ifdim \wd\@tempboxa > 8cm
9:            {\begin{center}
10:         \parbox{8cm}{\footnotesize\baselineskip=8pt {\bf Fig.~\thefigure.} #1}
11: 
12:             \end{center}}
13:         \else
14:              {\begin{center}
15:              {\footnotesize {\bf Fig.~\thefigure.} #1}
16:               \end{center}}
17:         \fi}
18: \newcommand{\be}{\begin{equation}}
19: \newcommand{\ee}{\end{equation}}
20: \newcommand{\bea}{\begin{eqnarray}}
21: \newcommand{\eea}{\end{eqnarray}}
22: \newcommand{\rD}{\mbox{D}}
23: \newcommand{\reff}{\mbox{eff}}
24: \newcommand{\rR}{\mbox{R}}
25: \newcommand{\rL}{\mbox{L}}
26: \newcommand{\p}{\partial}
27: \newcommand{\s}{\sigma}
28: \newcommand{\rF}{\mbox{F}}
29: \newcommand{\rf}{\mbox{f}}
30: \newcommand{\up}{\uparrow}
31: \newcommand{\down}{\downarrow}
32: %\newcommand{\right}{\rightarrow}
33: \newcommand{\la}{\langle}
34: \newcommand{\ra}{\rangle}
35: \newcommand{\rd}{\mbox{d}}
36: \newcommand{\ri}{\mbox{i}}
37: \newcommand{\re}{\mbox{e}}
38: \newcommand{\sumnn}{\sum_{\langle jk \rangle}}
39: \newcommand{\rk}{\mbox{k}}
40: 
41: \begin{document}
42: 
43: \centerline{\large \bf Zero-Variance Zero-Bias Principle for Observables in 
44: quantum Monte Carlo:}
45: \centerline{\large \bf Application to Forces}
46: 
47: \vspace{2cm}
48: 
49: \begin{center}
50: Roland Assaraf and Michel Caffarel
51: \end{center}
52: \vspace{1cm}
53: \noindent
54: {CNRS-Laboratoire de Chimie Th\'eorique Tour 22-23, Case 137\\
55: Universit\'{e} Pierre et Marie Curie, 4 place Jussieu 75252 Paris Cedex 05 France\\
56: e-mail: ra@lct.jussieu.fr,mc@lct.jussieu.fr}
57: 
58: \begin{center}
59: {\bf Abstract}
60: \end{center}
61: A simple and stable method for computing accurate expectation values 
62: of observable with Variational Monte Carlo (VMC) or Diffusion Monte 
63: Carlo (DMC) algorithms is presented. The basic idea consists in
64: replacing the usual ``bare'' estimator associated with the observable 
65: by an improved or ``renormalized'' estimator. Using this estimator
66: more accurate averages are obtained:
67: Not only the statistical fluctuations are reduced but also 
68: the systematic error (bias) associated with the approximate VMC 
69: or (fixed-node) DMC probability densities. It is shown that improved estimators 
70: obey a Zero-Variance Zero-Bias (ZVZB) property similar 
71: to the usual Zero-Variance Zero-Bias property of the energy
72: with the local energy as improved estimator.
73: Using this property improved estimators can be optimized 
74: and the resulting accuracy on expectation values may reach
75: the remarkable accuracy obtained for total energies.
76: As an important example, we present the application of our formalism
77: to the computation of forces in molecular systems.
78: Calculations of the entire force curve of the H$_2$,LiH, and Li$_2$ molecules
79: are presented. Spectroscopic constants $R_e$ (equilibrium distance) 
80: and $\omega_e$ (harmonic frequency) are also computed. 
81: The equilibrium distances are obtained with a relative error smaller than 1$\%$, while 
82: the harmonic frequencies are computed with an error of about 10$\%$. 
83: \\
84: {PACS numbers: 71.15.-m, 31.10.+z, 31.25.Nj, 02.70.Lq}\\
85: 
86: \newpage
87: 
88: \section{Introduction}
89: 
90:  Over the recent years quantum Monte Carlo (QMC) methods have become 
91: more and more successful in computing ground-state total energies of 
92: molecular systems. For systems with large number of electrons the 
93: accuracy obtained by QMC is very good. As illustrated by a number of 
94: recent calculations,
95: \cite{wil1,wil2,wil3,mitas1,fil1,needs1,mitas2,lester1,wil4,fil2,needs2,mitas3,mitas4,anderson0} 
96: the quality of the results is comparable 
97: and, in most cases, superior to that obtained with more traditional 
98: techniques (DFT, MCSCF or coupled cluster methods). Unfortunately, 
99: for properties other than energy the situation 
100: is much less favorable and accurate results are difficult 
101: to obtain. To understand this point let us first define what 
102: we mean here by accuracy. In standard quantum Monte Carlo schemes
103: there exist essentially two types of error:
104: 
105: i) the usual statistical error resulting from the necessarily finite
106: simulation time. This error present in any Monte Carlo scheme behaves 
107: as $ \sim 1/\sqrt{N}$ where $N$ is the number of Monte Carlo steps.
108: 
109: ii) the systematic error (or ``bias'') associated with some particular 
110: choice of the trial wave function. 
111: In a Variational Monte Carlo (VMC)
112: scheme it is the systematic error resulting from the approximate trial probability 
113: density. In a fixed-node Diffusion Monte Carlo (DMC)
114: it is either the fixed-node error of energy calculations or the systematic error 
115: associated with the mixed DMC probability density 
116: for a more general observable.
117: Other types of systematic errors may also exist, e.g. the short-time 
118: error, \cite{lester2} however, such errors can be easily controlled and, therefore,
119: will not be considered here.
120: 
121: Now, to enlighten the major differences between energy and observable 
122: computations let us compute the expressions of these two errors.
123: We shall do that within the framework of the Variational Monte Carlo
124: method where, as we shall see later, all the main aspects of this work 
125: are already present.
126: 
127: In a variational Monte Carlo simulation the variational energy
128: \begin{equation}
129: E_v \equiv \frac { \langle \psi_T | H | \psi_T \rangle} 
130:                  { \langle \psi_T |     \psi_T \rangle},
131: \label{variationalenergy}
132: \end{equation}
133: where $\psi_T$ is the approximate trial wave function used, is re-expressed 
134: as the statistical average of the local energy defined as
135: \begin{equation} 
136: E_L = \frac {H\psi_T}{\psi_T} 
137: \label{localenergy}
138: \end{equation} 
139: over the probability density associated with $\psi_T^2$, namely
140: \begin{equation}
141: E_v = \langle E_L \rangle_{\psi_T^2}.
142: \label{averagenergy}
143: \end{equation}
144: An accurate calculation of the energy requires the two following conditions.
145: 
146: (i) First, the systematic (or variational) error defined as 
147: \begin{equation}
148: \Delta_E  \equiv E_v - E_0  \ge 0 ,
149: \label{deltadef}
150: \end{equation}
151: where $E_0$ is the exact energy, must be as small as possible.
152: 
153: (ii) Second, the variance of the local energy (which is directly related to the 
154: magnitude of the statistical error)
155: \begin{equation}
156: \sigma^2(E_L)= \langle (E_L-E_v)^2 \rangle_{\psi_T^2},
157: \label{vardef}
158: \end{equation}
159: must also be as small as possible.
160: 
161: To estimate both quantities we express them in terms of the 
162: trial wave function error, $\delta \psi = \psi_T-\psi_0$, where $\psi_0$
163: is the exact wave function.
164: Regarding the systematic error it is easy to check that 
165: \begin{equation} 
166: \Delta_E  = \frac {\langle \psi_T-\psi_0| H - E_0 | \psi_T-\psi_0 \rangle}
167:                   {\langle \psi_T|\psi_T \rangle }.
168: \label{delta}
169: \end{equation} 
170: In other words, $\Delta_E$ is of order two in the wave function error
171: \begin{equation}    
172: \Delta_E  \sim  O[{\delta \psi}^2].
173: \label{deltaorder}
174: \end{equation}
175: Now, regarding the variance, it is convenient to write the 
176: following equality
177: \begin{equation} 
178: E_L - E_v =  \frac { (H - E_0) (\psi_T-\psi_0) } {\psi_T} - \Delta_E,
179: \label{eqint1}
180: \end{equation}
181: from which it is directly seen that $\sigma^2(E_L)$ is also of order two 
182: \begin{equation}
183: \sigma^2(E_L) \sim  O[{\delta \psi}^2].
184: \label{varorder}
185: \end{equation} 
186: 
187: Equations (\ref{deltaorder}) and (\ref{varorder}) are at the origin of 
188: the high-quality calculations of the energy. 
189: They show that accurate energy calculations are directly related to 
190: good trial wave functions: The more accurate the
191: trial wave function is, the smaller the statistical and systematic 
192: errors are. In the limit of an exact trial wave function, both errors 
193: vanish and the energy estimator reduces to the exact energy. 
194: This most fundamental property is referred to in the literature as 
195: the ``Zero-Variance property''. Note that a much more preferable 
196: and accurate denomination should be ``Zero-Variance-Zero-Bias property" 
197: to emphasize on the existence of the {\it two} types of error. Of course, in the 
198: case of the energy this distinction is not necessary since, as just
199: seen, the two errors are not independent and vanish
200: {\it simultaneously} with the exact wave function.
201: However, as we shall see below, this peculiar aspect 
202: will be no longer true for other properties.
203: 
204: Let us now turn our attention to the computation of a general observable.
205: Defining the expectation value of some arbitrary observable $O$ 
206: as follows
207: \begin{equation}
208: O_v \equiv \frac {\langle \psi_T | O | \psi_T \rangle } {\langle \psi_T
209:   | \psi_T \rangle},
210: \label{variationalobs}
211: \end{equation}
212: its Monte Carlo expression is given by
213: \begin{equation}
214: O_v = \langle O \rangle_{\psi_T^2}.
215: \label{averageobs}
216: \end{equation}
217: It is easy to verify that the systematic error behaves as
218: \begin{equation} 
219: \Delta_O \equiv \langle O \rangle_{\psi_T^2} - 
220: \langle O \rangle_{\psi_0^2} \sim  O[\delta \psi],
221: \label{deltaobs}
222: \end{equation}
223: while the variance is given by
224: \begin{equation}
225: \sigma^2(O) \sim  O[1].
226: \label{varorderobs}
227: \end{equation}
228: Compared to the energy case we have two striking differences.
229: First, the systematic error in the averages is much larger.
230: This is a direct consequence of Eq.(\ref{deltaobs}):
231: the estimator of a general observable has only a {\it linear} zero-bias property 
232: instead of a quadratic one like in the energy case. Even worse, because
233: trial wave functions are optimized to lower the
234: systematic error in the energy (and/or its fluctuations) and not
235: the error in the observable, the prefactor associated with the linear error 
236: contribution, Eq.(\ref{deltaobs}), is usually much 
237: larger than in the energy case, Eq.(\ref{deltaorder}). 
238: In practice, this important systematic error makes in general the quality of 
239: the expectation value, Eq.(\ref{averageobs}), very poor.
240: The second important difference is that there is no zero-variance property at all 
241: for observables when Eq.(\ref{averageobs}) is used. Indeed,
242: even when the exact wave function is used as trial wave function
243: we are still left with some finite (and eventually large) statistical fluctuations,
244: Eq.(\ref{varorderobs}). Thus, statistical fluctuations are in general very large 
245: for properties.
246: A simple and popular strategy to reduce the important systematic 
247: error on properties is to mix Variational Monte Carlo (VMC) and fixed-node Diffusion 
248: Monte Carlo (DMC) calculations to build up a so-called ``hybrid'' or ``second-order'' estimator,
249: ${\langle O \rangle}_{hybrid} \equiv 2{\langle O \rangle }_{DMC}
250: - {\langle O \rangle }_{VMC}$, whose error
251: is reduced.\cite{ceperley1} An elementary calculation shows that
252: the error is now of order $O[{(\psi_T-\psi_0)}^2]$, plus a linear
253: contribution $O(\psi_0^{FN}-\psi_0)$ due to the approximate nodes
254: of the trial wave function. However, once again such a solution is not, in practice,
255: as satisfactory as it appears at first glance because of the large 
256: prefactor associated with the second-order contribution and, also, of 
257: the non-negligible linear error due to the nodes.
258: A second possible strategy to cope with the systematic error is to perform
259: an ``exact'' QMC calculation based on one of the variants of the so-called 
260: ``Forward Walking'' scheme.\cite{liu1,caffaclav,casul1} Unfortunately, such schemes are known to be
261: intrinsically unstable and, therefore, very time consuming. In practice, 
262: the possibility of getting or not a satisfactory answer depends very much on 
263: the accuracy required and on the type of observable considered. 
264: Therefore, Forward Walking is not considered as a general practical 
265: solution to the problem.
266: 
267: In this work, we propose to follow a quite different route. Our purpose is 
268: to show that it is possible to use much more efficient estimators for 
269: properties than the usual bare expression, Eq.(\ref{averageobs}). 
270: More precisely, it is shown how to construct in a simple and systematic way
271: new estimators having the same remarkable quadratic zero-variance zero-bias 
272: property as the energy case.
273: Very recently, we have made a first step in that direction by
274: showing how to generalize the zero-variance part of this property.\cite{zvp1,zvp2}
275: In short, the basic idea consists in constructing a ``renormalized'' or improved 
276: observable having the same average as the original one but a lower variance. 
277: To build the renormalized observable, an auxiliary wave function 
278: is introduced. This function plays a role analogous to the one played 
279: by the trial wave function in the case of the energy: The closer the auxiliary 
280: function is of the exact solution of some zero-variance equation 
281: (the Schroedinger equation in the case of the energy), the smaller 
282: the statistical fluctuations of the renormalized observable are. 
283: Our approach has been illustrated on some simple academic examples\cite{zvp1}
284: and also for the much more difficult case of the computation of forces
285: for some diatomic molecules.\cite{zvp2} Numerical results on these 
286: examples are very satisfactory. When suitably chosen auxiliary functions 
287: are used, statistical errors are indeed greatly reduced.
288: 
289: Here, we present the full generalization of the preceding idea: 
290: it is shown how to construct improved observables 
291: minimizing {\it both} systematic and statistical
292: errors with a quadratic behavior similar to that obtained for the energy. 
293: As a consequence, any observable is expected to be calculated, at least 
294: in principle, with the remarkable accuracy achieved by QMC for total energies.
295: The basic idea behind our approach is quite simple: it consists in 
296: making use of the relation between energy and observable calculations 
297: as expressed by the Hellmann-Feynman (HF) theorem. As well-known this 
298: theorem expresses any quantum average as a total energy derivative with 
299: respect to the magnitude of the external potential defined by the observable.
300: It is shown how the zero-variance zero-bias principle valid for 
301: each value of the energy (as a function of the external potential)
302: can be extended to the derivative and, therefore, to the observable.
303: Note that in the context of QMC simulations, the idea of using the 
304: HF theorem to compute observables, using either a finite difference scheme or 
305: the analytic derivative, is not new and has been applied by several 
306: groups\cite{wells1,traynor1,umrigar1,sun1,vrbik1,vrbik2,belohorec1,fil3,baer}.
307: In general, the results are good for very small systems but rapidly disappointing 
308: for larger systems. Indeed, only when a clear physical 
309: insight into the origin of the fluctuations of the infinitesimal difference 
310: of energy (the derivative) is available it is possible to propose an 
311: efficient solution to the problem. 
312: A very nice example of such a possibility 
313: is presented in the recent work by Filippi and Umrigar.
314: \cite{umrigarwrapped},\cite{fil3}. By using a finite representation of 
315: the energy derivative and by introducing a special coordinate transformation 
316: allowing the electrons close to a given nucleus to move almost rigidly 
317: with that nucleus, they have shown how to correlate efficiently
318: the calculation of the electronic energies associated with two slighlty 
319: different nuclear configurations of a diatomic molecule.
320: As a result they have been able to get accurate estimates of the 
321: energy derivatives (forces) for some diatomic molecules. 
322: Here, we show how the correlated sampling 
323: method of Filippi and Umrigar can be re-expressed in our framework. 
324: In addition, by generalizing their idea it is shown how coordinate 
325: transformations can be used to define a new class of improved estimators. 
326:  
327: While finishing this work, we came aware of a paper just published by Casalegno, 
328: Mella and Rappe.\cite{cmr} The idea underlying their work has some 
329: close relations with what is presented here. In short, they propose, as we do here, 
330: to compute forces using a Hellmann-Feynman-type formalism. 
331: Their expression to calculate forces is obtained by making the derivative 
332: of the VMC (or DMC) energy average with respect to nuclear positions.
333: To reduce the systematic error these authors propose to employ trial wave 
334: functions which have been very carefully optimized via energy minimization 
335: (let us recall that the HF theorem is valid when {\it fully} optimized wave functions are used). 
336: To decrease the very large statistical fluctuations associated with the infinite 
337: variance, the improved estimator introduced in our previous work 
338: on forces\cite{zvp2} is used. As we shall see below, the approach proposed by 
339: Casalegno and collaborators can be viewed as a special case of the general method presented here, except 
340: that their estimator does not obey a Zero-Variance Zero-Bias property. As we shall see below, this latter aspect 
341: has some important practical consequences when a high level of accuracy on forces is needed.
342: 
343: The organization of the paper is as follows. In Section II we present 
344: the Hellmann-Feynman theorem and the construction of improved estimators 
345: for variational Monte Carlo calculations.
346: It is also shown how the idea of Filippi and
347: Umrigar consisting in introducing a special coordinate transformation can be 
348: used to build up some more general and more efficient improved estimators.
349: In Section III we discuss the generalization of the formulae to the 
350: case of Diffusion Monte Carlo calculations.
351: In Section IV we present the application of the formalism
352: to the computation of the entire force curve 
353: for the H$_2$,LiH, and Li$_2$ molecules.
354: Calculations of the spectroscopic constants, $R_e$ 
355: and $\omega_e$, are also reported.
356: Finally, in the last section we summarize our results and present 
357: some concluding remarks.
358: 
359: \section{Improved estimators for observables} 
360: 
361: In order to make the connection between energy and observable computations 
362: we shall make use of the Hellmann-Feynman (HF) theorem which expresses
363: the expectation value of an observable as an energy derivative 
364: \begin{equation}
365: \frac{ \langle \psi_0|O|\psi_0 \rangle}
366: {\langle \psi_0|\psi_0 \rangle} 
367: = \frac{d E_0(\lambda)}{d\lambda}\big|_{\lambda=0},
368: \label{eq:exact}
369: \end{equation}
370: where $E_0(\lambda)$ is the exact ground-state energy of 
371: the ``perturbed'' Hamiltonian defined as
372: \begin{equation}
373: H(\lambda) \equiv H + \lambda O.
374: \end{equation}
375: By choosing various approximate expressions for 
376: the exact energy in Eq.(\ref{eq:exact}), it is possible to derive various
377: approximate estimates for the average. In the next sections we present
378: two choices which turn out to be particularly efficient 
379: in practical applications.
380: 
381: \subsection{Improved estimator built from the variational approximation of 
382: the energy}
383: 
384: A most natural choice consists in replacing the exact energy 
385: of the HF theorem by a high-quality variational approximation. 
386: To do that, we introduce some $\lambda$-dependent approximate trial 
387: wave function, $\psi_T(\lambda)$ to describe the ground-state of 
388: $H(\lambda)$ 
389: [Note that, for the sake of clarity and simplicity, we shall denote 
390: in what follows ${\psi_T}(0)$, $H(0)$, and $E_0(0)$ as ${\psi_T}$, $H$, 
391: and $E_0$, respectively]. 
392: 
393: The exact average of the observable can be decomposed as
394: \begin{equation}
395: \frac{ \langle \psi_0|O|\psi_0 \rangle } 
396: {\langle \psi_0|\psi_0 \rangle }
397: =\frac{d {E_v}(\lambda)} {d\lambda}  \big|_{\lambda=0} 
398: + \epsilon(\delta \psi,\delta \psi^\prime)
399: \label{approx}
400: \end{equation}
401: where ${E_v}(\lambda)$ is the variational energy associated with 
402: $\psi_T(\lambda)$
403: \begin{equation}
404: {E_v}(\lambda) \equiv \langle {E_L}(\lambda) \rangle_{{\psi_T}^2(\lambda)} = 
405: \langle \frac{H(\lambda) {\psi_T}(\lambda)}{{\psi_T}(\lambda)} \rangle_{{\psi_T}^2(\lambda)}
406: \end{equation}
407: and $\epsilon$ some correction depending on $\delta\psi=\psi_0-\psi_T$ and 
408: its derivative, and vanishing when the exact wave function is used as 
409: trial wave function.
410: 
411: Now, the important point is that the derivative of the variational energy, 
412: $\frac{d{E_v}(\lambda)} {d\lambda} \big|_{\lambda=0}$,
413: is expected to be a better estimate of the exact average than the ordinary 
414: average of the bare estimator, Eq.(\ref{averageobs}), when properly chosen 
415: $\lambda$-dependent trial wave functions are used. This is true since
416: the standard estimator, Eq.(\ref{averageobs}), can be re-expressed as 
417: a particular case of the derivative of the variational energy for 
418: a $\lambda$-{\it independent} trial wave function, 
419: a choice which is clearly not optimal.
420: Before justifying more quantitatively this statement, let 
421: us rewrite the derivative as an ordinary 
422: average over the density ${\psi_T}^2$. This can be easily done, it gives
423: \begin{equation}
424: \frac{d{E_v}(\lambda)}{d\lambda} \big|_{\lambda=0} 
425: = \langle \tilde{O} \rangle_{{\psi_T}^2}
426: \label{evprime}
427: \end{equation}
428: where $\tilde{O}$ is a new modified local operator written as
429: \begin{equation}
430: \tilde{O} \equiv  O + \frac{(H-{E_L}){\psi_T}^\prime}{{\psi_T}}
431: +2 ({E_L}-{E_v}) \frac{{{\psi_T}}^\prime}
432: {{\psi_T}}
433: \label{tildeobias}
434: \end{equation}
435: In this latter formula, and in the formulae to follow, we shall use 
436: the following simplified notation  
437: \begin{equation}
438: f^\prime \equiv \frac{d f(\lambda)}{d\lambda} 
439: \big|_{\lambda =0}
440: \end{equation}
441: where $f(\lambda)$ is some arbitrary function of $\lambda$.
442: 
443: Now, we have to justify the first important result,
444: that the new estimator $\tilde{O}$ is a 
445: better estimator for the exact average than the bare observable $O$. 
446: For that purpose, we compute the systematic error in the corresponding 
447: average and the variance of the new operator.
448: Regarding the systematic error we can write
449: \begin{equation}
450: \Delta_{\tilde O} \equiv
451: \langle \tilde O \rangle_{{{\psi_T}}^2} - \langle O \rangle_{\psi_0^2}
452: = \frac{d[{E_v}(\lambda)-E_0(\lambda)]}{d\lambda} \big|_{\lambda=0}.
453: \end{equation}
454: 
455: Let us denote ${\psi_0}(\lambda)$ the exact groundstate of $H(\lambda)$ 
456: [with $\psi_0(0)=\psi_0$]. Using the equality
457: $$
458: E_v(\lambda)-E_0(\lambda)=
459: $$
460: \begin{equation}
461: \frac {\langle {\psi_T}(\lambda) - {\psi_0}(\lambda)| H(\lambda) - E_0(\lambda) 
462:  | {\psi_T}(\lambda) - {\psi_0}(\lambda)  \rangle}
463:                   {\langle {\psi_T}(\lambda)|{\psi_T}(\lambda) \rangle }
464: \end{equation}
465: and choosing the following convention of normalization 
466: \begin{equation}
467: \langle {\psi_T}(\lambda)|{\psi_T}(\lambda) \rangle= 1,
468: \nonumber
469: \end{equation}
470: the derivative can be easily computed, we get
471: \begin{eqnarray}
472:  \Delta_{\tilde O} & = & 
473:  \langle {\psi_T}
474: -\psi_0 | O - \langle O \rangle_{\psi_0^2} |  {\psi_T} -\psi_0 \rangle  
475: \nonumber \\ 
476: & + &
477:  2 \langle {\psi_T}-\psi_0 | H -E_0 | {{\psi_T}}^\prime-{\psi_0}^\prime \rangle 
478: \label{biasderiv}
479: \end{eqnarray}
480: As it can be seen, the systematic error is now of order two in the errors ${\psi_T}-\psi_0$ and
481: ${{\psi_T}}^\prime-{\psi_0}^\prime$
482: \begin{equation}
483: \Delta_{\tilde O} \sim O[({\psi_T}-\psi_0)
484: ({\psi_T}^\prime-{\psi_0}^\prime)].
485: \label{eqint3}
486: \end{equation}
487: 
488: Now, let us compute the variance defined as
489: \begin{equation}
490: \sigma^2(\tilde{O}) = \langle 
491: (\tilde{O} - \langle \tilde{O} \rangle_{{\psi_T}^2})^2 \rangle_{{\psi_T}^2}.
492: \label{varotilde}
493: \end{equation}
494: Using Eqs. (\ref{evprime}),(\ref{tildeobias}) and the fact that 
495: $E_L^\prime= O +  \frac{(H-{E_L}){\psi_T}^\prime}{{\psi_T}}$
496: we can express the difference $\tilde{O} - 
497: \langle \tilde{O}\rangle_{{\psi_T}^2}$ as follows
498: \begin{equation}
499: \tilde{O}- \langle{\tilde{O}} 
500: \rangle_{{{\psi_T}}^2}  =  {E_L}^\prime -{E_v}^\prime
501: +2 ({E_L} -{E_v}) \frac{{\psi_T}^\prime}
502: {{\psi_T}}. 
503: \label{eqint2}
504: \end{equation} 
505: For the sake of clarity, let us distinguish two different contributions 
506: in the difference. The first contribution is given by
507: \begin{equation}
508: {E_L}^\prime -{E_v}^\prime = \frac{d[{E_L}(\lambda) -{E_v}(\lambda)]}
509: {d\lambda}\mid_{\lambda=0}.
510: \end{equation}
511: Using expression (\ref{eqint1}) for $[{E_L}(\lambda)-{E_v}(\lambda)]$ 
512: and performing the derivative one obtains
513: $$
514: {E_L}^\prime -{E_v}^\prime = \frac{(O-\langle O \rangle_{\psi_0^2})
515:  ({\psi_T}-\psi_0)}{{\psi_T}}  
516: $$
517: $$
518:  +
519: \frac{(H-E_0)({\psi_T}^\prime-\psi_0^\prime)}{{\psi_T}} 
520:  - 
521: \frac{(H-E_0)({\psi_T}-\psi_0)}{{\psi_T}} \frac{ {\psi_T}^\prime }{\psi_T}
522: $$
523: 
524: \begin{equation}
525:  + 
526: \langle O \rangle_{\psi_0^2}- \langle \tilde{O} \rangle_{\psi_T^2}
527: \end{equation}
528: This latter expression is clearly of order one 
529: in ${{\psi_T}}-{\psi_0}$ and its derivative,${\psi_T}^\prime-\psi_0^\prime $.
530: The second contribution in the R.H.S. of Eq.(\ref{eqint2}) is 
531: proportional to ${E_L} -{E_v}$. We have already seen that 
532: it is of order one in ${\psi_T}-{\psi_0}$, Eqs.(\ref{deltaorder}),
533: (\ref{eqint1}). Finally, $\tilde{O}-\langle\tilde{O}\rangle_{{\psi_T}^2}$
534: is found to be of order one in 
535: ${\psi_T}-\psi_0$ and ${\psi_T}^\prime-\psi_0^\prime$.
536: The variance, Eq.(\ref{varotilde}), is therefore of order two 
537: \begin{equation}
538: \sigma^2(\tilde{O}) \sim O[({\psi_T}-\psi_0)({\psi_T}^\prime-{\psi_0}^\prime)].
539: \label{eqint4}
540: \end{equation}
541: 
542: To summarize, using the HF theorem we are able to construct 
543: an improved observable $\tilde{O}$, Eq.(\ref{tildeobias}), 
544: having a quadratic zero-variance-zero-bias property, 
545: Eqs.(\ref{eqint3},\ref{eqint4}), similar to what is known for the energy case, 
546: Eqs.(\ref{deltaorder},\ref{varorder}). The improved estimator
547: $\tilde{O}$ depends only on one single quantity, namely ${\psi_T}(\lambda)$. 
548: Accordingly, to get accurate results we need to choose in the 
549: neighborhood of $\lambda=0$ a trial function accurate enough to get
550: not only a small difference in wave functions but also 
551: in the derivative of the wave functions. In practice, this latter point 
552: is particularly difficult to fulfill. Indeed, at fixed values of $\lambda$,
553: it is known that the minimization of the fluctuations of the local energy can allow 
554: an important reduction of the error in the trial wave function. However, 
555: there is no reason why it should also lead to a satisfactory representation of
556: the derivative of the trial wave function.
557: 
558: In order to escape from this difficulty we propose here to 
559: work directly at $\lambda=0$ and to optimize {\it independently} the 
560: trial wave function ${\psi_T}$ and its derivative ${\psi_T}^\prime$.
561: Such procedure is justified since it corresponds to choose as 
562: $\lambda$-dependent trial wave function the following expression
563: \begin{equation} 
564: {\psi_T}(\lambda)= {\psi_T} + \lambda \tilde{\psi},
565: \end{equation} 
566: where $\tilde{\psi}$ is some new independent function playing the role 
567: of a trial function for the derivative of the ground-state at $\lambda=0$.
568: In this case, the renormalized observable can be rewritten under the
569: final form
570: \begin{equation}
571: \tilde{O} \equiv  O + \frac{(H-E_L)\tilde{\psi}}{\psi_T}
572: +2 (E_L-E_v) \frac{\tilde{\psi}} {\psi_T}.
573: \label{tildeobias2}
574: \end{equation}
575: where the pair of functions $(\psi_T,\tilde{\psi})$ is the current 
576: guess for the exact solution $(\psi_0,{\psi_0}^\prime)$.
577: 
578: Let us now turn our attention on the problem of optimizing the two 
579: trial functions $(\psi_T,\tilde{\psi})$.
580: Regarding  ${\psi_T}$ we know that the standard procedure consists
581: in minimizing the variance of the local energy with respect to the 
582: parameters of the trial function.
583: Quite remarkably, we have here a similar result for $\tilde{\psi}$:
584: the best choice is obtained by minimizing the 
585: variance of the renormalized operator $\tilde{O}$ with respect to 
586: the parameters of $\tilde{\psi}$.
587:  
588: To prove this property it is sufficient to show that 
589: the zero-variance (or zero-fluctuations) equations for ${E_L}$  
590: and $\tilde{O}$:
591: \begin{eqnarray}
592: {E_L} & = &  \langle {E_L} \rangle_{\psi_T^2} \nonumber \\
593:  \tilde{O} & = & \langle
594: \tilde{O} \rangle_{{{\psi_T}}^2}
595: \label{2var0}
596: \end{eqnarray}
597: are equivalent to the equations defining $\psi_0$ and $\psi_0^\prime$,
598: namely
599: \begin{eqnarray}
600: (H-{E_0})\psi_0 & = & 0 \nonumber  \\
601: (H-{E_0})\psi_0^\prime & + & (O- \langle O \rangle_{\psi_0^2}  )\psi_0
602: =  0
603: \label{eqfip}
604: \end{eqnarray}
605: In these formulae, the first equation is just the ordinary 
606: Schroedinger equation. The second one is obtained by deriving 
607: the Schroedinger equation:
608: \begin{equation}
609: H(\lambda)\psi(\lambda)=E_0(\lambda) \psi(\lambda)
610: \end{equation}
611: with respect to $\lambda$ at $\lambda=0$.
612: Note that equations (\ref{eqfip}) determine an unique solution,
613: ($\psi_0,\psi_0^\prime,E_0,\langle O \rangle_{\psi_0^2}$),
614: as soon as $H$ has a non-degenerate ground-state.
615: Now, using Eqs.(\ref{tildeobias}) and (\ref{localenergy}) for the 
616: definition of $\tilde{O}$ and 
617: $E_L$, respectively, the system of equations (\ref{2var0}) 
618: can be rewritten under the form
619: \begin{eqnarray}
620: (H-{E_v}) {\psi_T} & =  & 0 \\
621: (H-{E_v}) {\psi_T}^\prime & + & (O-\langle O \rangle_{{{\psi_T}}^2})
622:  {\psi_T} = 0
623: \label{eqfippsi}
624: \end{eqnarray}
625: which are nothing but Eqs.(\ref{eqfip}) with $({\psi_T},{\psi_T}^\prime)
626: =(\psi_0,\psi_0^\prime)$. Accordingly, the zero-variance equations (\ref{2var0}) 
627: admits this latter pair of functions as unique solution.
628: 
629: In practical calculations, different strategies of optimization 
630: can be employed. A first approach consists in minimizing {\it separately}
631: the variance of the local energy with respect to the wave function 
632: ${\psi_T}$ and the variance of $\tilde{O}$ with respect
633: to $\tilde{\psi}$. In this way, we get an optimal trial wave function 
634: $\psi_T$ for the energy and the best derivative at fixed $\psi_T$. However,
635: let us emphasize that this approach is not the most general: we can also
636: minimize both variances simultaneously with respect to the two independent 
637: functions. Another remark is that the second equation of system (\ref{eqfip}) 
638: can be viewed as an ordinary first-order perturbation equation.
639: This is expected since, when $\lambda O$ is considered as a 
640: perturbation of the Hamiltonian $H$, $\psi_0^\prime$ is nothing but 
641: the first-order correction to the ground-state 
642: and $\langle O \rangle_{\psi_0^2}$ the first-order correction 
643: to the energy.
644: 
645: Finally, let us end this section by commenting in more detail 
646: the various terms entering expression (\ref{tildeobias2}) of 
647: the improved operator. Three different contributions can be distinguished:
648: 
649: (i) The ordinary bare estimator $O$ corresponding to $\tilde{\psi}=0$.
650: 
651: (ii) A second contribution given by $(H-{E_L})\tilde{\psi}/{\psi_T}$. 
652: It is easy to verify that this contribution has a zero average
653: over the density ${\psi_T}^2$
654: \begin{equation}
655: \langle (H-{E_L})\tilde{\psi}/{\psi_T} \rangle_{{{\psi_T}}^2}=0.
656: \end{equation}
657: Accordingly, its role is to lower the variance 
658: of the improved estimator without changing the average of the observable
659: (no influence on the systematic error). 
660: Note that for applications where the stationary density is known and can 
661: be exactly sampled (that is, there is no systematic error in the average) the use of 
662: contributions (i) and (ii) is sufficient. Important examples 
663: include all ``classical'' Monte Carlo simulations based on the Metropolis 
664: algorithm or one of its variants. Such a possibility was the subject of a 
665: previous work.\cite{zvp1}
666: (iii) A third term given by $2({E_L}-{E_v})\frac{\tilde{\psi_T}}{{\psi_T}}$.
667: This contribution has a very small impact on the statistical
668: fluctuations since the variance of $({E_L}-{E_v})$ is of order 
669: two in the trial wave function error for any choice of $\tilde{\psi}$. 
670: Its main effect 
671: is to take into account the change of stationary density under the external 
672: perturbation defined by the observable 
673: and, therefore, to lower the systematic error in the expectation value of the observable.
674: Note that in the limit ${\psi_T}=\psi_0$, 
675: this contribution reduces to zero and, therefore, the average of this term can be 
676: understood as a correction to the Hellmann-Feynman formula
677: when ${\psi_T}$ is not the exact ground-state (note that similar corrections to the HF formula 
678: exist also in more traditional {\it ab initio} calculations, e.g. the
679: ``Pulay force''\cite{pulay} resulting from approximate Hartree-Fock (or LDA) orbitals 
680: in self-consistent schemes).
681: 
682: \subsection{More improved estimators: use of coordinate transformations}
683: 
684: In this section it is shown how to generalize further our 
685: renormalized operators.
686: The basic idea of the generalization is based on an original idea 
687: recently proposed by Filippi and Umrigar in their work 
688: on the computation of forces \cite{fil3}. Working in a finite difference formalism 
689: they have proposed to compute the forces as a small but finite difference
690: of energies for two close enough geometries. In order to minimize the 
691: fluctuations they have proposed to use a correlated sampling method in 
692: which a common Monte Carlo density (the so-called primary one) 
693: is used for the two close geometries. 
694: Written within our notations and taking the limit of the two geometries 
695: infinitely close ($\delta R \rightarrow$ 0 is equivalent to $\lambda \rightarrow$ 0) it means that 
696: the variational energy is written under the form
697: \begin{equation}
698: E_v(\lambda)=
699: \frac{\langle {E_L}(\lambda) \frac{{{\psi_T}^2(\lambda)}}{\psi_T^2} 
700: \rangle_{\psi_T^2}}
701: {\langle \frac{{{\psi_T}^2(\lambda)}}{\psi_T^2} \rangle_{\psi_T^2}}
702: \end{equation}
703: where ${\psi_T}(\lambda)$ is the trial wave function chosen for a parameter
704: $\lambda$ and $\psi_T$ is the reference (primary) trial wave function. 
705: 
706: The price to pay when doing that is the introduction of 
707: some additional fluctuations associated with the weight
708: $\frac{{{\psi_T}^2(\lambda})}{\psi_T^2}$.
709: The remedy they propose to deal with this problem is to use  
710: a specific coordinate transformation (space-warp transformation)
711: based on physical motivations: The transformation is built 
712: so that the electrons close to a given nucleus move almost rigidly with 
713: that nucleus when the geometry is changed. 
714: Here, we generalize this idea: coordinate 
715: transformation can help to minimize the relative fluctuations when 
716: varying the external parameter $\lambda$. As a physical consequence, estimators 
717: built from the derivative are expected to have smaller fluctuations 
718: and smaller systematic errors. 
719: 
720: Let us write a general coordinate transformation as follows
721: \begin{equation}    
722: \vec{y}= \vec{T}(\lambda,\vec{x})
723: \end{equation}  
724: where the vector $\vec{x}$ (or $\vec{y}$) denotes the set of 
725: the 3$n_{elec}$ electronic coordinates. Using this transformation the variational energy 
726: at a given $\lambda$ can be written as
727: \begin{equation}
728: E_v(\lambda)=
729: \frac{\langle 
730: E_L[\lambda,\vec{T}(\lambda,\vec{x})] 
731: J(\lambda,\vec{x})
732: \frac{ {\psi_T}^2[\lambda,\vec{T}(\lambda,\vec{x})] }
733:      { {\psi_T}^2(\vec{x}) }
734: \rangle_{{\psi_T}^2} }
735: {
736: \langle 
737: J(\lambda,\vec{x}) 
738: \frac{{\psi_T}^2[\lambda,\vec{T}(\lambda,\vec{x})]}
739:      {{\psi_T}^2(\vec{x})} \rangle_{{\psi_T}^2}
740: }
741: \label{eJacprime}
742: \end{equation}
743: where $J(\lambda,\vec{x})$ is the Jacobian of the transformation. 
744: Introducing the vector field ${\vec{v}}$ such that at first order
745: in $\lambda$ we have:
746: \begin{equation}
747: \vec{T}(\lambda,\vec{x})=\vec{x}+\lambda \vec{v}(\vec{x})+O(\lambda^2)
748: \end{equation}
749: we can compute the derivative of the variational energy with respect to 
750: $\lambda$ at $\lambda=0$. After some simple but tedious algebra we get 
751: the following equality
752: \begin{equation}
753: \frac{d{E_v}(\lambda)}{d\lambda} \big|_{\lambda=0}
754: = \langle \tilde{O} \rangle_{{\psi_T}^2}
755: \label{evprime2}
756: \end{equation}
757: where $\tilde{O}$ is a new renormalized operator given by
758: \begin{equation}
759: \tilde{O} \equiv  O + \frac{(H-{E_L}){\psi_T}^\prime}{{\psi_T}}
760: +2 ({E_L}-{E_v}) \frac{{{\psi_T}}^\prime}
761: {{\psi_T}}
762: + 
763: \frac{\vec{\nabla} [(E_L -E_v) \psi_T^2 \vec{v}]}
764: {\psi_T^2}.
765: \label{tildeobias3}
766: \end{equation}
767: To derive this expression we have used the fact that the Jacobian
768: defined as
769: \begin{equation}
770: J(\lambda,\vec{x})=det[ \frac{\partial T_i(\lambda,\vec{x})  }{\partial x_j}]
771: \label{jacobian}
772: \end{equation}
773: has the following small-$\lambda$ expression
774: \begin{equation}
775: J(\lambda,\vec{x})=det[ \delta_{ij} + \lambda \frac{\partial v_i}{\partial x_j}]
776: + O(\lambda^2)
777: \label{jacobian2}
778: \end{equation}
779: and, therefore,
780: \begin{equation}
781: J(0,\vec{x})=1
782: \end{equation}
783: \begin{equation}
784: \frac{\partial J}{\partial \lambda}(0,\vec{x})= \vec{\nabla}.\vec{v}
785: \end{equation}
786: This more general operator is identical to the operator derived in the previous 
787: section plus a new contribution resulting from the derivative of the 
788: coordinate transformation. This new term has a zero average over
789: the VMC distribution $\psi_T^2$. Accordingly, its main role is 
790: to reduce further the statistical error. However, it is important to emphasize that, 
791: when the trial function 
792: $\tilde{\psi}$ and the vector field $\vec{v}$ are optimized simultaneously 
793: it has also an influence on the magnitude of the systematic error.
794: 
795: \section{Beyond Variational Monte Carlo}
796: 
797: In the preceding section we have shown how to construct improved observables,
798: $\tilde{O}$, associated with accurate expectation values:
799: \begin{equation}
800: \frac {\langle \psi_T |\tilde{O}| \psi_T \rangle}
801:                  { \langle \psi_T | \psi_T \rangle}
802: =\frac {\langle \psi_0 |O| \psi_0 \rangle}
803:                  { \langle \psi_0 | \psi_0 \rangle} 
804:  + O[({\psi_T}-\psi_0)
805: (\tilde{\psi}-{\psi_0}^\prime)].
806: \label{bvmc1}
807: \end{equation}
808: When the error $\delta \psi^\prime=\tilde{\psi}-{\psi_0}^\prime$ in the trial function for the derivative
809: is comparable to the error in the trial function for the ground-state, 
810: $\delta \psi={\psi}_T -{\psi_0}$, the accuracy reached with the preceding variational 
811: estimate, Eq.(\ref{bvmc1}), can be comparable to the very good accuracy usually obtained for 
812: total energies. However, despite this remarkable improvment,
813: we are still left with some small residual systematic error 
814: associated with approximate $\psi_T$ and $\tilde{\psi}$. In the energy case
815: it is known that this error can be entirely suppressed (at least for systems with no nodes or 
816: known nodes) by averaging the local 
817: energy over the mixed Diffusion Monte Carlo (DMC) probability distribution, 
818: $\pi_{DMC} \sim \psi_T \psi_0$ instead of the VMC distribution, $\pi_{VMC} \sim {\psi_T}^2$. 
819: Unfortunately, we have no such result for the improved 
820: observables defined here. However, as we shall see now, we can still define some 
821: approximate way for recovering most of the error.
822: 
823: A natural way of defining an exact extimator for the observable is to consider 
824: the derivative of the exact DMC energy estimator instead of the VMC one
825: \begin{equation}
826: {E_0}(\lambda) \equiv \langle {E_L}(\lambda) \rangle_{{\psi_T}(\lambda){\psi_0}
827: (\lambda) } =
828: \langle \frac{H(\lambda) {\psi_T}(\lambda)}{{\psi_T}(\lambda)} 
829: \rangle_{{\psi_T}(\lambda) {\psi_0}(\lambda)}
830: \end{equation}
831: Making the derivative and rewritting the result as an ordinary average we get:
832: \begin{equation}
833: \frac{d{E_0}(\lambda)}{d\lambda} \big|_{\lambda=0}
834: = \langle \tilde{O}\rangle_{{\psi_T}{\psi_0}}
835: \label{e0vprime}
836: \end{equation}
837: where $\tilde{O}$ is written as
838: \begin{equation}
839: \tilde{O} \equiv  O + \frac{(H-{E_L}){\psi_T}^\prime}{{\psi_T}}
840: +({E_L}-{E_0}) (\frac{{{\psi_T}}^\prime} {{\psi_T}}
841: + \frac{{{\psi_0}}^\prime} {{\psi_0}}).
842: \label{tildeobias4}
843: \end{equation}
844: Of course, written under the above form, this exact estimator is useless since the exact 
845: wave function is not known. Here, we propose to make the following natural approximation
846: \begin{equation}
847: \frac{{{\psi_0}}^\prime} {{\psi_0}} = \frac{{{\psi_T}}^\prime} {{\psi_T}}.
848: \label{approxdmc}
849: \end{equation}
850: Therefore, our final approximate DMC estimator is written as
851: \begin{equation}
852: \tilde{O}_{DMC} \equiv  O + \frac{(H-{E_L})\tilde{\psi}}{{\psi_T}}
853: +2 ({E_L}-{E_0}) \frac{{\tilde{\psi}}} {{\psi_T}}
854: \label{tildeobias5}
855: \end{equation}
856: where $\tilde{\psi}$ is as usual our trial function for the derivative of 
857: the exact wave function. Note that this estimator is very similar to the VMC one, Eq.(\ref{tildeobias2}).
858: The only difference lies in the value of the average energy, $E_0 =\langle E_L \rangle$, 
859: entering the definition of $\tilde{O}_{DMC}$. More precisely, we have
860: \begin{equation}
861: \tilde{O}_{DMC}- \tilde{O}_{VMC} = 2 (E_v -E_0) \frac{{\tilde{\psi}}} {{\psi_T}}.
862: \label{f10}
863: \end{equation}
864: 
865: Now, in order to reduce further the error let us show that we can generalize the 
866: usual ``hybrid formula'' ${\langle O \rangle}_{hybrid} \equiv 2{\langle O \rangle }_{DMC}
867: - {\langle O \rangle }_{VMC}$ defined for bare observables to the case of improved observables. 
868: To do that, let us develop the quantity $\langle \delta \psi | \tilde{O}_{DMC}
869:  |\delta \psi \rangle$
870: where $\delta \psi= \psi_T - \psi_0$
871: \begin{equation}
872: \langle \delta \psi | \tilde{O}_{DMC}|\delta \psi \rangle = 
873: \langle \psi_T | \tilde{O}_{DMC}|\psi_T \rangle 
874: -2 \langle \psi_T | \tilde{O}_{DMC}|\psi_0 \rangle
875: + \langle \psi_0 | \tilde{O}_{DMC}|\psi_0 \rangle
876: \end{equation}
877: which leads to
878: $$
879: 2 \langle \psi_T | \tilde{O}_{DMC}|\psi_0 \rangle
880: -\langle \psi_T | \tilde{O}_{DMC}|\psi_T \rangle 
881: = \langle \psi_0 | O |\psi_0 \rangle
882:  + A + O( {\delta \psi}^2)
883: $$
884: where the intermediate quantity $A$ is defined as
885: $$
886: A \equiv 
887: \langle \psi_0 | \frac{(H-{E_L})\tilde{\psi}}{{\psi_T}}
888: +2 ({E_L}-{E_0}) \frac{{\tilde{\psi}}} {{\psi_T}} | \psi_0 \rangle
889: $$
890: Expanding $A$ in terms of $\delta \psi= \psi_T-\psi_0$ we get 
891: $$
892: A = 2 \langle \psi_T |E_L - E_0|\tilde{\psi} \rangle 
893:    -2 \langle  \delta \psi| (H-E_L) |\tilde{\psi}\rangle 
894: -4 \langle \delta \psi |E_L - E_0|\tilde{\psi} \rangle 
895:  + O( {\delta \psi}^2)
896: $$
897: Using now the equality
898: $$
899: E_L-E_0= \frac{ (H -E_0)\delta \psi} {\psi_T}
900: $$
901: we obtain
902: $$
903: A = O( {\delta \psi}^2)
904: $$
905: This latter result shows that the error in the hybrid estimator is 
906: of order two in $\delta \psi$
907: \begin{equation}
908: 2 \langle \tilde{O}_{DMC} \rangle_{\psi_T  \psi_0} - 
909:   \langle \tilde{O}_{DMC} \rangle_{\psi_T^2} = \langle \psi_0 | O |\psi_0 \rangle + O( {\delta \psi}^2),
910: \label{hybridfinal}
911: \end{equation}
912: thus generalizing the standard result for the bare observable.
913: Note that we can use either $\tilde{O}_{DMC}$ or $\tilde{O}_{VMC}$ 
914: [Eq.(\ref{tildeobias2})] in this latter 
915: formula since the difference between the two renormalized operators is proportional to $E_v-E_0$, Eq.(\ref{f10}), 
916: and, therefore, is also of order two in $\delta \psi$ [Eq.(\ref{deltaorder})].
917: 
918: When using coordinate transformation we have similar results. The 
919: exact DMC estimator is found to be
920: \begin{equation}
921: \tilde{O} \equiv  O + \frac{(H-{E_L}){\psi_T}^\prime}{{\psi_T}}
922: +({E_L}-{E_0}) (\frac{{{\psi_T}}^\prime} {{\psi_T}}
923: + \frac{{{\psi_0}}^\prime} {{\psi_0}})
924: +
925: \frac{\vec{\nabla} [(E_L -E_0) \psi_T \psi_0 \vec{v}]}
926: {\psi_T \psi_0}
927: \label{tildeobias6}
928: \end{equation}
929: and we propose to use the following approximate form
930: \begin{equation}
931: \tilde{O}_{DMC} \equiv  O + \frac{(H-E_L)\tilde{\psi}}{{\psi_T}}
932: +2 ({E_L}-\langle E_L \rangle ) \frac{{\tilde{\psi}}} {{\psi_T}}
933: +
934: \frac{\vec{\nabla} [(E_L - \langle E_L \rangle ) {\psi_T}^2 \vec{v}]}
935: {{\psi_T}^2}
936: \label{tildeobias7}
937: \end{equation}
938: Because the difference $(E_L -\langle E_L \rangle )$ is of order $\delta \psi$ it is easy to 
939: verify that the error in the hydrid estimator given by Eq.(\ref{hybridfinal}) 
940: remains here also of order two.
941: 
942: Before ending this section let us emphasize that it is possible to write a closed 
943: computable expression for the exact estimator of the observable,  
944: Eq.(\ref{tildeobias4}), by expressing the unknown quantity 
945: $ \frac{{{\psi_0}}^\prime} {{\psi_0}}$ as a computable 
946: stochastic average. Choosing a $\lambda$-independent trial 
947: wave function $\psi_T$ we can write\cite{ceperley2,cafstat,caffaclav}
948: \begin{equation}
949: \psi_0(\lambda,x) = \psi_T(x)  \lim_{T \rightarrow +\infty}
950: << e^{-\int_0^{T} ds E_L[\lambda,x(s)]} >>_{x(0)=x}
951: \label{psiPDMC}
952: \end{equation}
953: where $x$ denotes an arbitrary point in configuration space and 
954: $<<...>>_{x(0)=x}$ denotes the sum over all drifted random walks of length $T$ starting 
955: at $x$ as obtained in a Pure Diffusion Monte Carlo (PDMC) scheme 
956: (DMC without branching).\cite{caffaclav} 
957: Of course, a similar formula can also be obtained in a DMC scheme.\cite{liu1} 
958: Now, using formula (\ref{psiPDMC}) we get
959: \begin{equation}
960: \frac{{{\psi_0}}^\prime} {{\psi_0}}= \lim_{T \rightarrow +\infty}
961: \int_0^{T} dt \frac{
962: << O[x(t)] e^{-\int_0^{T} ds E_L[\lambda,x(s)]}>>_{x(0)=x} }
963: {<< e^{-\int_0^{T} ds E_L[\lambda,x(s)]}>>_{x(0)=x} }
964: \end{equation}
965: and, therefore, the exact estimator can be written in terms of a standard part plus 
966: a time-integral of the two-point correlation function between the local energy 
967: and the observable
968: $$
969: \tilde{O} = O + \frac{(H-{E_L})\tilde{\psi}}{\psi_T}
970: +(E_L-\langle E_L \rangle) \frac{{\tilde{\psi}}} {\psi_T}
971: $$
972: \begin{equation}
973: + \lim_{T \rightarrow +\infty}
974: \int_0^{T} dt  \frac{
975: << (E_L- \langle E_L \rangle )[x(0)] O[x(t)] e^{-\int_0^{T} ds E_L}>>_{x(0)=x} }
976: {<< e^{-\int_0^{T} ds E_L}>>_{x(0)=x} }
977: \end{equation}
978: It is important to emphasize that this latter estimator is exact: averaged 
979: over the mixed DMC distribution it leads to an unbiased estimate of the exact 
980: average. However, the correlator can only obtained within a Forward Walking 
981: scheme and, therefore, the stability in time is not guaranteed. In this work, we shall 
982: not use this expression, its implementation will be presented in a forthcoming work.
983: 
984: \section{Application to forces}
985: 
986: The average force between atoms in a molecular system is defined as
987: \begin{equation}
988: \bar{F}_{q_i} \equiv -\frac{\partial E_0({\bf q})}{ \partial q_i},
989: \label{force}
990: \end{equation}
991: where $E_0({\bf q})$ is the total electronic ground-state energy for a given nuclear configuration;
992: ${\bf q}$ represents the $3N_{nucl}$ nuclear coordinates ($N_{nucl}$, number of nuclei) and
993: $q_i$ the particular force component in which we are interested.
994: 
995: Defining the local force as follows
996: \begin{equation}
997: F_{q_i}({\bf x}, {\bf q}) \equiv - \frac{\partial V({\bf x},{\bf q})}{ \partial q_i}.
998: \label{localforces}
999: \end{equation}
1000: where ${\bf x}$ represents the $3n_{elec}$ electronic coordinates ($n_{elec}$, number of electrons)
1001: and $V$ the total potential energy operator, and making use of the
1002: Hellmann-Feynman (HF) theorem the average force can be rewritten as
1003: the statistical average of the local force over the exact distribution $\psi_0^2({\bf x})$:
1004: \begin{equation}
1005: \bar{F}_{q_i} = \langle F_{q_i}(\bf x, q) \rangle_{\psi_0^2({\bf x})}.
1006: \end{equation}
1007: Written under this form the various proposals presented in the preceding sections can be applied 
1008: to the calculation of the average force.
1009: It is important to emphasize that for approximate probability densities (VMC or DMC) the HF theorem is no 
1010: longer valid and a systematic error in the statistical average $\langle F_{q_i}(\bf x, q)\rangle$ is introduced.
1011: However, it is not a problem here since it is the 
1012: purpose of this work to show that, by using suitable improved estimators, this error 
1013: can be reduced and even suppressed in the zero-bias limit. 
1014: 
1015: In order to discuss the various aspects of the method we shall restrict ourselves to the 
1016: case of diatomic molecules. Let us consider a diatomic molecule $AB$ with atom $A$ located 
1017: at $(R,0,0)$ and atom $B$ located at the origin. The only non-zero component of the local force 
1018: acting on the nucleus $A$ is the $x$-component given by
1019: \begin{equation}
1020: F = - \frac{\partial V}{\partial R}
1021: = \frac{Z_A Z_B}{R^2} - Z_A \sum_{i=1}^{n_{elec}} \frac{(x_i-R)}{|{\bf r_i -R}|^3}.
1022: \label{localforcediatomics}
1023: \end{equation}
1024: 
1025: In this work we present a number of VMC and fixed-node DMC calculations for the diatomic molecules 
1026: H$_2$,LiH, and Li$_2$. Implementation of the quantum Monte Carlo methods is well-known and will not be
1027: discussed here. For the H$_2$ molecule the trial wave function used has the following simple form
1028: \begin{equation}
1029: \psi_T= (1s_A 1s_B + 1s_B 1s_A) + c (1s_A 1s_A + 1s_B 1s_B)
1030: \end{equation}
1031: where $1s_M$ is a 1s-Slater function centered at nucleus $M=A,B$ with 
1032: exponent $\mu$ and $c$ a parameter describing the amount of ionic contribution 
1033: into the wave function. Of course, much more accurate trial wave functions can be constructed for H$_2$. However, 
1034: our purpose here is to show that such a simple form for $\psi_T$ is already sufficient to 
1035: get accurate values of the force.
1036: 
1037: For LiH and Li$_2$ we have employed two types of trial wave function. Our main choice
1038: is standard in QMC calculations for molecules. The trial 
1039: wave function is made of a determinant of single-particle orbitals multiplied by a Jastrow 
1040: factor. The determinantal part is obtained from a RHF calculation 
1041: and only the Jastrow factor is optimized.
1042: As we shall see below, we have also used Valence-Bond (VB)-type wave functions
1043: consisting of a number of determinants multiplied by a Jastrow factor. We have used such 
1044: a multideterminantal description to reproduce correctly 
1045: the large interatomic distance regime (dissociation limit). In the case of LiH 
1046: the determinantal part consists of three determinants corresponding to 
1047: the covalent VB resonating structure: $(1s_{Li})^2[ 2s_{Li} 1s_H + 1s_H 2s_{Li}]$ ($\{1s_{Li},2s_{Li},1s_H\}$ 
1048: optimized {\it atomic} orbitals for the Li and H atoms, orbitals occupied 
1049: by electrons $\alpha$ and $\beta$ antisymmetrized separately) 
1050: and one ionic VB structure: $(1\tilde{s}_{Li})^2 1\tilde{s}_H 1\tilde{s}_H$ 
1051: ($1\tilde{s}_{Li},1\tilde{s}_H$ optimized atomic orbitals for the Li$^+$ and H$^-$ ions). 
1052: In the case of Li$_2$ we have considered a six-determinant 
1053: representation consisting of 
1054: the three covalent VB structures describing the resonance between atomic orbitals $(2s_A,2s_B)$, 
1055: $(2p_{y_A},2p_{y_B})$, and $(2p_{z_A},2p_{z_B})$. 
1056: This latter trial wave function reproduces not only the dissociation limit 
1057: but also a major part of the 2s-2p near-degeneracy.
1058: 
1059: In Figures \ref{figureEH2},\ref{figureELiH}, and \ref{figureELi2}
1060: the energy curves obtained for H$_2$,LiH, and Li$_2$ are presented.
1061: Upper curves are the VMC curves (open squares joined by a dotted line). 
1062: For H$_2$ the two parameters $c$ and $\mu$ have been optimized for each 
1063: interatomic distance. For LiH and Li$_2$ the Jastrow-RHF wave function (one determinant) has been used.
1064: All the parameters entering the Jastrow factor have been optimized for all distances.
1065: Optimizations have been performed by minimizing the variance of the local energy 
1066: using the correlated sampling method of Umrigar {\it et al.}\cite{umrigar}. 
1067: The first important observation is that, except for H$_2$, VMC curves are not 
1068: smooth as a function of $R$. Such a result is not surprising: 
1069: It is typical of a situation where an approximate trial wave function 
1070: is optimized {\it independently} for different 
1071: values of an external parameter (here, $R$) with respect 
1072: to a large number of variables (for LiH and Li$_2$ we have used about 30 independent 
1073: variational parameters). Depending on the initial conditions (which are themselves very dependent on $R$)
1074: the algorithm used for minimizing the variance can be trapped within one of the various {\it local}
1075: minima. As a consequence, the actual value obtained for the variance
1076: (and the corresponding energy) can vary abruptly even when the 
1077: external parameter is changed smoothly. 
1078: Of course, this problem can be solved in principle by making very careful optimizations on very large samples.
1079: Indeed, the functional form of the trial wave function being identical 
1080: at all distances a smooth curve must be got when the correct lowest minimum 
1081: of the variance is obtained at each distance. Here, this is the case for
1082: H$_2$ whose trial wave function contains only two variational parameters.
1083: However, for large systems including a much larger number 
1084: of variational parameters and nuclear degrees of freedom, the possibility of fully optimizing 
1085: the trial wave function is just irrealistic.
1086: As an important consequence, let us emphasize
1087: that, in practice, there is no hope of 
1088: obtaining meaningful forces by making straight finite differences of optimized variational energies
1089: without using some sort of correlated sampling scheme. This is a good illustration of how difficult the 
1090: calculation of forces is within a QMC framework. 
1091: 
1092: Intermediate points (filled squares) are the DMC results obtained from
1093: fixed-node calculations using the optimized VMC trial wave functions. 
1094: In sharp contrast with VMC, the DMC curves are now regular. 
1095: This is so because, unlike VMC, fixed-node DMC averages 
1096: do not depend on the particular form of the trial function used, except for 
1097: the nodal structure. Here, the nodal hypersurfaces vary smoothly as a function 
1098: of the distance and, therefore, the corresponding DMC energy curves are smooth 
1099: within error bars.
1100: The second important observation is related to the global shape of the curves.
1101: Ideally, we are interested in having energy curves which differ from the exact curve 
1102: only by a constant independent on the distance. This is indeed the condition 
1103: to obtain accurate derivatives. Here, it is not the case for the 
1104: VMC curves of LiH and Li$_2$: the difference between the VMC and  
1105: exact energy curves is an increasing function of the distance. 
1106: It is not a surprise since in both cases the trial wavefunction is built 
1107: from a single RHF determinant based on delocalized molecular orbitals which leads to a 
1108: wrong description of the dissociation 
1109: limit. However and, very interestingly, fixed-node DMC results have a much better 
1110: behavior at large distances. As a consequence, one may expect at this stage to obtain 
1111: accurate forces from the derivative of the fixed-node energy curve even when 
1112: relatively crude wave functions are used.
1113: Finally, let us note that the quality of the fixed-node calculations for the 
1114: molecules considered here is quite good. To give an example, at the equilibrium 
1115: distance of the Li$_2$ molecule, the total energy obtained is -14.9901(6) 
1116: to be compared with the exact non-relativistic value of $E_0$=-14.9954. 
1117: The amount of correlation energy recovered within the fixed-node approximation 
1118: is about $95.7\%$. A similar quality is obtained for other 
1119: distances and also for the LiH molecule. 
1120: In the case of the nodeless H$_2$ molecule (no fixed-node approximation), the DMC 
1121: energies agree perfectly well with the exact ones.
1122: For the H$_2$ molecule the variance of the local energy varies between 0.3 at $R=0.8$ and 
1123: 0.02 at $R=3.5$; for LiH the variance is about 0.07, and for Li$_2$ it varies between 0.09 and 0.2.
1124: 
1125: The crucial point when implementing the various formulae presented in the preceding 
1126: section is the choice of the trial function $\tilde{\psi}$ for the derivative.
1127: In our previous study on forces\cite{zvp2} 
1128: where we have focused our attention on
1129: the reduction of statistical fluctuations only, we have proposed to employ
1130: the minimal form leading to a finite variance of the renormalized local force.
1131: As can be viewed from Eq.(\ref{localforcediatomics}), at short electron-nucleus 
1132: distance $r$ the local force behaves as $F\sim 1/r^2$ and, therefore, the 
1133: variance $ \langle F^2 \rangle -{\langle F \rangle }^2$ is infinite. This  
1134: well-known problem has been discussed in different places 
1135: (See, {\it e.g.} \cite{lester2} Chap. 8.2 or \cite{vrbik2}). Here, 
1136: the ``minimal'' form removing the singular part responsible for the infinite 
1137: variance is written as
1138: \begin{equation}
1139: \tilde{\psi}_{min}({\bf x})= Q {\psi_T}
1140: \label{psi}
1141: \end{equation}
1142: where $Q$ is given by
1143: \be
1144: Q=  Z_A \sum_{i=1}^{n_{elec}} \frac{(x_i -R)} { {|{\bf r}_i
1145:  - {\bf R}|}  }.
1146: \label{q}
1147: \ee
1148: To see this, we just need to compute the following quantity
1149: \be
1150: \frac{ (H-E_L) \tilde{\psi}_{min} } {\psi_T}
1151: = Z_A \sum_{i=1}^{n_{elec}} \frac{(x_i-R)}{|{\bf r_i -R}|^3}
1152:  - {\bf \nabla} Q \cdot {\bf \nabla}\psi_T /\psi_T.
1153: \ee
1154: By adding this latter quantity to the bare local force, Eq.(\ref{localforcediatomics}),
1155: the singular part is exactly removed, the remaining contribution 
1156: having a finite variance. In what follows, $\tilde{\psi}_{min}$ will be referred to as
1157: the minimal form for $\tilde{\psi}$.
1158: 
1159: In Figure \ref{forcevar} we present various VMC calculations of the average 
1160: force for the Li$_2$ molecule as a function of the interatomic distance.
1161: A first set of points (filled squares points with very large error bars at $R$=5.,$R$=6.5, and $R$=7.5) are results 
1162: obtained from the ordinary bare estimator, Eq.(\ref{localforcediatomics}).
1163: Open squares (with small error bars) joined by the dashed curve  correspond to results obtained by using 
1164: ${\tilde{\psi}_{min}}$ as trial function for the derivative
1165: \begin{equation}
1166: \tilde{F}_{VMC-ZV}[\psi_T,\tilde{\psi}_{min}] 
1167: = F + \frac{(H-{E_L}) {\tilde{\psi}_{min}}}{{\psi_T}}.
1168: \label{f1}
1169: \end{equation}
1170: This estimator can be viewed as the simplest improved estimator we can think of 
1171: having a finite variance; it corresponds to the form employed in our previous study.\cite{zvp2}
1172: The subscript $ZV$ (Zero-Variance) is used here to 
1173: emphasize that the improved estimator is built to decrease the statistical 
1174: error only. Circles joined by a dotted line are results obtained from
1175: the ZVZB improved estimator derived in the preceding section, Eq.(\ref{tildeobias2}):
1176: \begin{equation}
1177: \tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{min}]
1178: = F + \frac{(H-{E_L}) {\tilde{\psi}_{min}}}{{\psi_T}} + 2(E_L- \langle E_L \rangle )
1179: \frac{\tilde{\psi}_{min}}
1180: {\psi_T}.
1181: \label{f2}
1182: \end{equation}
1183: Note the use of the subscript $ZVZB$ to emphasize on the two aspects: reduction of statistical and 
1184: systematic errors. Finally, the solid line represents the ``exact''
1185: non-relativistic force curve for Li$_2$.
1186: 
1187: A first important observation is that using improved estimators 
1188: is extremely efficient in reducing the statistical error. This can be seen by comparing 
1189: the magnitude of the error bars on data obtained from the ordinary bare estimator 
1190: (filled squares at $R=5.,6.5$, and 7.5) with those corresponding to other calculations based 
1191: on improved estimators. A reduction of at least two orders of magnitude is observed.
1192: As already discussed this remarkable result is a direct consequence of the fact that 
1193: the infinite variance of the bare estimator has been reduced to a finite value.
1194: At the scale of the figure error bars associated with improved estimators are almost not visible. 
1195: A more quantitative analysis will be given later (see, Table \ref{tableLi2}).
1196: 
1197: Now, regarding systematic errors, results are much more disappointing. Using 
1198: the pure Zero-Variance (ZV) renormalized estimator, Eq.(\ref{f1}), the behavior of the average 
1199: force (open squares joined by the dashed line in Fig.\ref{forcevar}) as a function of $R$ appears erratic. 
1200: This can be easily understood since the term added 
1201: to the bare force in Eq.(\ref{f1}) has a zero-average and, therefore, the erratic behavior is a direct 
1202: consequence of the irregular VMC energy curve presented in Figure \ref{figureELi2}.
1203: When adding the term correcting the average, the 
1204: results are improved. As seen on the figure the behavior of 
1205: $\langle \tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{min}]\rangle$, Eq.(\ref{f2}), as a function of $R$, 
1206: is much less irregular, thus illustrating the important role played by the zero-bias
1207: additional contribution [third term of the R.H.S. of Eq.(\ref{f2})] to correct the error due to the approximate trial wave function.
1208: Despite of that, the resulting curve is far from being satisfactory.
1209: To weaken the role played by $\psi_T$ we can think of going 
1210: beyond VMC calculations. In Figure \ref{forcedmc}
1211: we present such calculations for Li$_2$ using the DMC-ZVZB improved estimator 
1212: $\tilde{F}_{DMC-ZVZB}[\psi_T,\tilde{\psi}_{min}]$ written as
1213: \begin{equation}
1214: \tilde{F}_{DMC-ZVZB}[\psi_T,\tilde{\psi}_{min}]
1215: = F + \frac{(H-E_L) {\tilde{\psi}_{min}}}{{\psi_T}} + 2(E_L- \langle E_L \rangle)
1216: \frac{\tilde{\psi}_{min}}
1217: {\psi_T}
1218: \label{f3}
1219: \end{equation}
1220: where the energy average is a fixed-node DMC average,
1221: and, also, results obtained by using the generalized hybrid formula, 
1222: Eq.(\ref{hybridfinal})
1223: \begin{equation}
1224: \bar{F} \simeq 2 \langle \tilde{F}_{DMC-ZVZB} \rangle_{\psi_T  \psi_0} -
1225: \langle \tilde{F}_{VMC-ZVZB} \rangle_{\psi_T^2} 
1226: \label{hybridfinal2}
1227: \end{equation}
1228: A clear improvment is observed when going from VMC (open circles) to DMC (filled squares) and, then, 
1229: to hybrid calculations (open squares): The systematic error present in VMC calculations is reduced. 
1230: However, the resulting curves are still not satisfactory. Extracting from them a meaningful equilibrium 
1231: distance or first derivative of the force curve (calculation of $\omega_e$) is impossible.
1232: Very similar behaviors have been obtained for H$_2$ and LiH. They do not need to be reproduced here.
1233: 
1234: The main reason for the poor results just presented is the 
1235: low quality of the trial function $\tilde{\psi}_{min}$ used. According to our 
1236: general presentation of Sec. II we know that a good trial function $\tilde{\psi}$ must 
1237: be close to the derivative of the exact ground-state wave function with respect to $R$.
1238: Here, this is only true when an electron approaches the nucleus A (note that nucleus B has been fixed  
1239: at the origin and, thus, has no pathological contribution). In that case the non-vanishing part of the 
1240: exact wave function is expected to behave as
1241: \begin{equation}
1242: \psi_0 \sim_{ {\bf r}_i \rightarrow {\bf R}} \exp{(-Z_A | {\bf r}_i- {\bf R}|)}
1243: \end{equation}
1244: which leads to 
1245: \begin{equation} 
1246: \frac{\partial \psi_0}{\partial R} \sim_{ {\bf r}_i \rightarrow {\bf R} }
1247: -Z_A \frac{ (x_i -R)} { {|{\bf r}_i - {\bf R}|} } \psi_0
1248: \end{equation}
1249: which is nothing but (up to a minus sign) the minimal form for $\tilde{\psi}$ given above, Eqs.(\ref{psi},\ref{q}).
1250: 
1251: In order to improve our trial wave function $\tilde{\psi}$ we propose to use 
1252: the following finite-difference form 
1253: 
1254: \begin{equation}   
1255: \tilde{\psi}_{Deriv}= \frac{\psi_T[R+\Delta R, p(R+\Delta R)]- 
1256: \psi_T[R, p(R)] } { \Delta R }.
1257: \label{psideriv}
1258: \end{equation}
1259: In this expression $p(R)$ denotes the complete set of variational parameters entering 
1260: the trial wave function (coefficients of molecular orbitals, basis set exponents, 
1261: Jastrow parameters, etc...). The main advantage of using a finite-difference form 
1262: instead of the exact derivative is practical:  To estimate 
1263: the derivative we only need to compute two additional local energies and, thus, we 
1264: avoid deriving and programming the lengthy expressions resulting 
1265: from the explicit derivative. Note also that  using an approximate
1266: finite-difference representation is not a problem here:
1267: In any case $\tilde{\psi}_{Deriv}$ must be considered 
1268: as an approximate trial function for the exact derivative and $\Delta R$ 
1269: can always be interpreted as a new additional variational parameter for $\tilde{\psi}$.
1270: In practical calculations, the complete set of parameters we use for 
1271: minimizing the fluctuations of the various improved estimators
1272: consists of $ \{ p(R)$, $p(R+\Delta R)$, and $\Delta R \}$.
1273: 
1274: At the VMC level we consider the following form for the improved estimator
1275: $$
1276: \tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{Deriv},\vec{v}]= 
1277: $$
1278: \begin{equation}
1279: F+ \frac{(H-E_L) \tilde{\psi}_{Deriv}}{\psi_T} 
1280: + 2(E_L - \langle E_L \rangle ) \frac{\tilde{\psi}_{Deriv}}{\psi_T}
1281: +
1282: \frac{\vec{\nabla} [(E_L - \langle E_L \rangle) \psi_T^2 \vec{v}]}
1283: {\psi_T^2}.
1284: \label{fpsir}
1285: \end{equation}
1286: At the DMC level the expression used is very similar, see Eq.(\ref{tildeobias7}).
1287: In this expression the vector field $\vec{v}$ associated with the coordinate
1288: transformation is chosen as follows
1289: \begin{equation}
1290: \vec{v}=\sum_{i=1}^{n_{elec}} e^{-\alpha r_{iA}-\beta r_{iA}^2} \vec{u_x}
1291: \end{equation}
1292: where $\vec{u_x}$ is the unit vector along the x-axis.
1293: The vector field depends on two parameters $\alpha$ and $\beta$, which are
1294: optimized to lower the variance of $\tilde{F}_{VMC-ZVZB}$.
1295: The vector field is built so that electrons close to the nucleus A translate with the 
1296: nucleus, while electrons far away do not move.
1297: In practice, we compute the additional term associated with the coordinate 
1298: transformation using a finite-difference scheme along the direction defined by the 
1299: vector $\vec{v}$
1300: $$
1301: \frac{\vec{\nabla} [(E_L - \langle E_L \rangle ) \psi_T^2 \vec{v}]}{\psi_T^2} = 
1302: $$
1303: \begin{equation}
1304: (E_L - \langle E_L \rangle) \nabla . {\vec{v}} + 
1305: [(E_L- \langle E_L \rangle)(\vec{x}+\epsilon \vec{v}) \frac{\psi_T^2( \vec{x}+\epsilon \vec{v})}
1306: {\psi_T^2 (\vec{x})} - (E_L- \langle E_L \rangle)(\vec{x})]/ \epsilon
1307: \label{bestestimator}
1308: \end{equation}
1309: where $\vec{x}$ represents the electronic coordinates and  $\epsilon$ a small 
1310: positive quantity whose magnitude can also be optimized.
1311: 
1312: In Figure \ref{fLi2psiR} we present VMC calculations 
1313: for Li$_2$ using 
1314: $\tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{min}]$ and 
1315: $\tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{Deriv},\vec{v}]$ as 
1316: improved estimators.  The two estimators have been evaluated on the same Monte Carlo samples.
1317: There are two striking differences when using the second estimator 
1318: $\tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{Deriv},\vec{v}]$.
1319: First, the gain in statistical error is spectacular (about one order of magnitude, for a quantitative analysis see 
1320: discussion below, Table \ref{tableLi2}). Second, the curve is much more regular and closer to the exact 
1321: result (solid line). The VMC results are very satisfactory at small distances 
1322: (between R=3. and R=4.). However, at larger interatomic distances, the VMC curve 
1323: begins to separate from the exact one. This is due to the fact that the wave function 
1324: is built from a RHF calculation and, therefore, the dissociation limit 
1325: is not correctly described. To address this problem we have considered a more 
1326: sophisticated trial wave function consisting of a product of a Jastrow factor 
1327: and a six-determinant one-particle part (for a more detailed description, see above).
1328: This VB-type trial wave function has been 
1329: used only for the largest distances: R=7.,R=7.5,R=8, and R=8.5
1330: In figures \ref{fLi2finalvmc} and \ref{fLi2finaldmc} the comparison between results 
1331: obtained with the Jastrow-RHF (one determinant) and the Jastrow-VB (six determinants) wave functions 
1332: is presented. At the VMC level (Fig.\ref{fLi2finalvmc}), the improvment resulting from the multideterminant 
1333: wave function is clearly seen, the forces computed are much closer to the exact curve than in the one-determinant 
1334: case. At the DMC level (Fig.\ref{fLi2finaldmc}) we could expect that this error disappears even with 
1335: the Jastrow-RHF (one determinant) wave function since the DMC results depend only on the nodal structure of 
1336: the wave function. 
1337: However, it is not true. The difference between the 
1338: DMC curve and the exact one is still important at large distances like in the VMC case. 
1339: This result takes its origin in the approximation made for the exact derivative of 
1340: the wave function in the DMC estimator, Eq.(\ref{approxdmc})
1341: ($\frac{{{\psi_0}}^\prime} {{\psi_0}}$ is replaced by $\frac{{{\psi_T}}^\prime} {{\psi_T}}$).
1342: When using the Jastrow-VB wave function the DMC results obtained are much better.
1343: 
1344: We are now in a position to present our final curves for the three molecules obtained 
1345: with our best fully-optimized estimator $\tilde{F}_{ZVZB}[\psi_T,\tilde{\psi}_{Deriv},
1346: \vec{v}]$ and the hybrid formula. Results for the molecules H$_2$,LiH, and Li$_2$
1347: are presented in Figures \ref{fH2final},\ref{fLiHfinal},\ref{fLi2finalhyb}, respectively.
1348: As seen from these figures 
1349: the overall agreement between the exact curves (solid lines) and QMC results (open squares) is very good. 
1350: To be more quantitative we have extracted from these curves an estimate of the spectroscopic 
1351: constants $R_e$ (equilibrium distance) and $\omega_e$ (harmonic frequency). To do that, the data
1352: have been fitted with a functional form given by the derivative of a Morse potential curve 
1353: $E(R) = D [\exp{-2 \beta (R-R_e)} - 2 \exp{-2 \beta (R-R_e)}]$ over some interval of distances around 
1354: the equilibrium geometry (R between 1.1 and 2. for H$_2$, between 2.6 and 4. for LiH, and between 
1355: 4. and 6. for Li$_2$).
1356: Parameters $D,\beta$, and $R_e$ have been determined via a generalized least-squares fit.
1357: Our results at the VMC, DMC, and Hybrid levels are presented in Table \ref{fitdata} and 
1358: compared to experimental values.\cite{herz} 
1359: As seen from the Table results for the equilibrium distances are excellent. The largest
1360: systematic errors are obtained at the VMC level (relative errors of 4.3$\%$,3.3$\%$, and 5.7$\%$ for H$_2$,LiH, and 
1361: Li$_2$, respectively). A reduction of a factor of about two is gained when DMC calculatons are 
1362: performed. Finally, using the hybrid formula, the exact equilibrium distances are recovered 
1363: within statistical errors (the relative statistical errors being 1.1$\%$,0.3$\%$, and 0.5$\%$ for H$_2$,LiH, and
1364: Li$_2$, respectively). In contrast, results for the harmonic frequencies are less accurate but still satisfactory. 
1365: For the H$_2$ molecule, the exact experimental result is almost recovered within statistical error at the VMC, 
1366: DMC, and Hybrid levels, the relative statistical error being between 3 and 4$\%$. For LiH and Li$_2$ the 
1367: relative statistical errors are of the same order of magnitude. However, a non-negligible systematic 
1368: error of about $10\%$ is found for these molecules. This result illustrates that obtaining accurate 
1369: harmonic frequencies is more difficult than obtaining
1370: accurate equilibrium geometries.
1371: 
1372: Now, we would like to present a more quantitative discussion 
1373: of the performance of the various force estimators introduced in this work. 
1374: This will allow us to summarize the various aspects of the method,
1375: to present some comparisons with the recent results obtained from the improved estimator 
1376: implicitly used by Casalegno and collaborators\cite{cmr}, and, also, to emphasize 
1377: on some important {\it quantitative} issues. In Table \ref{tableLi2} the systematic and statistical errors 
1378: associated with the various force estimators at the VMC, DMC and Hybrid levels of calculation are presented.
1379: The results shown are for the Li$_2$ molecule at the equilibrium 
1380: geometry (R=5.051 a.u.) where the exact force average, denoted as
1381: $\langle F \rangle_{ex}$, is equal to zero. To allow direct comparisons between force estimators all averages have been 
1382: computed in a common Monte Carlo calculation (identical MC samples).
1383: To compare with, we also give the systematic and statistical errors on the 
1384: total energy. To give a measure of the fluctuations of each estimator the
1385: corresponding variances at the VMC level, $\sigma^2$(VMC), are reported. To facilitate comparisons between 
1386: data all averages (except for the VMC variances) are given with five significant figures after the 
1387: decimal point and all statistical errors are given on the fifth decimal place (magnitude $10^{-5}$). 
1388: The first estimator presented is the bare estimator, $F$ 
1389: [Eq.(\ref{localforcediatomics})]. As already pointed out, this estimator, which 
1390: has an infinite variance, displays very large statistical fluctuations (between two and three orders 
1391: of magnitude with respect to the improved estimators to follow)
1392: and is, therefore, not at all suitable for practical calculations. The second estimator, 
1393: $\tilde{F}_{ZV} [\psi_T, \tilde{\psi}_{min}]$ [Eq.(\ref{f1})],
1394: introduced in our previous work on forces\cite{zvp2} is the simplest estimator having 
1395: a finite variance. However, as explained above, when using such an estimator no control on the systematic error 
1396: exists. The third estimator presented,
1397: $\tilde{F}_{ZVZB} [\psi_T, \tilde{\psi}_{min}]$, is the simplest estimator 
1398: having the ZVZB property. We can see that the introduction of the contribution associated 
1399: with the ZB property, $2(E_L - \langle E_L\rangle) \frac{\tilde{\psi}_{min}}{\psi_T}$ is efficient in 
1400: reducing the bias (the DMC and hybrid errors are roughly divided by a factor two). 
1401: However, as already discussed, the derivative of the 
1402: trial wave function is not correctly reproduced as a function of the 
1403: interatomic distance and the corresponding force curve is not smooth 
1404: (see, Figures \ref{forcevar},\ref{forcedmc}). To get accurate and well-behaved (as a function of $R$) 
1405: values of the force it is important to introduce an auxiliary function close to the 
1406: exact derivative of the wave function. The most simple estimator based on this idea and having a finite 
1407: variance can be constructed by using the minimal form $\tilde{\psi}_{min}$ [Eqs.(\ref{psi}),(\ref{q})]
1408: for the Zero-Variance part and $\tilde{\psi}_{Deriv}$,[Eq.(\ref{psideriv})], for the Zero-Bias part. Such an 
1409: estimator is written as
1410: \begin{equation}
1411: \tilde{F}[\psi_T, \tilde{\psi}_{min} | \tilde{\psi}_{Deriv}]  \equiv F+ \frac{(H-E_L) \tilde{\psi}_{min}}{\psi_T}
1412: + 2(E_L - \langle E_L\rangle) \frac{\tilde{\psi}_{Deriv}}{\psi_T}
1413: \label{minder}
1414: \end{equation}
1415: Written with our notations this is in fact the estimator implicitly used by Casalegno and collaborators in their 
1416: very recent work\cite{cmr}. In contrast with the other estimators presented here this ``mixed'' estimator 
1417: (different trial functions $\tilde{\psi}$ are used for the ZV and ZB parts) 
1418: has no ZVZB property. Accordingly, our general optimization procedure based on the minimization 
1419: of the improved-estimator variance is no longer meaningful here. Nevertheless, as pointed out 
1420: by Casalegno {\it et al.}, to optimize estimator (\ref{minder}) 
1421: we still have the possibility of optimizing the parameters of the trial wave 
1422: function $\psi_T$ via energy minimization. Such a procedure is justified
1423: because fully-optimized trial wave functions are known to verify the Hellmann-Feynman theorem. 
1424: Statistical errors associated with this estimator are reasonable and roughly similar to those obtained 
1425: with the two previous 
1426: estimators. Systematic errors are also comparable. However, in sharp contrast with all ZVZB-improved-estimators 
1427: introduced in this work, the DMC (and hybrid) calculations do not improve the results and, as seen in the Table 
1428: the hybrid results can be even bad, despite the fact that VMC results are reasonable. 
1429: Regarding the dependence of the results on the interatomic distance 
1430: we have been able to recover a relatively smooth force curve for the smallest molecules H$_2$ and LiH but not for Li$_2$. 
1431: In this latter case, the systematic error is found to be too much sensitive on 
1432: the quality of the optimization of the trial wave function to lead to reliable results. 
1433: The last improved estimator presented in Table \ref{tableLi2} is our best proposal for the 
1434: force estimator, Eq.(\ref{fpsir}). 
1435: We report results with ($\vec{v} \neq 0$) and without ($\vec{v} = 0$) to enlighten the role 
1436: of the coordinate-transformation term. 
1437: As seen, the introduction of the $\vec{v}$-term is extremely efficient in reducing the statistical error. 
1438: For example, at the VMC level the statistical error without this term is 218.10$^{-5}$, while the optimized 
1439: improved estimator using the $\vec{v}$-term is decreased down to 9.10$^{-5}$. 
1440: The reduction gained in statistical error is more than one order of magnitude. 
1441: This remarkable result is general : It is valid for all molecules and all distances treated here. 
1442: Another most important point is that our best improved estimator (\ref{fpsir}) is the only estimator presented in this work 
1443: whose statistical error (here, 9.10$^{-5}$) is (much) smaller than the energy one (here, 32.10$^{-5}$). 
1444: Note that it is also true for the systematic error (whatever the level of calculation). 
1445: Such a result is particularly important since it has been found that a precise control of the magnitude of 
1446: the systematic error through variance minimization 
1447: of the improved force estimators is possible only when such a condition is verified. 
1448: In contrast, when the statistical error on 
1449: the force is larger than the energy error, the variance minimization can lead to various results and to get 
1450: a smooth force curve is very difficult.
1451: Actually, we would like to emphasize that obtaining results of the quality presented in Figure (\ref{fLi2finalhyb})
1452: for the Li$_2$ molecule has only been possible with the improved estimator (\ref{fpsir}). Using other estimators 
1453: we have not been capable of constructing a reasonable force curve (smooth and accurate) for this molecule.
1454: 
1455: Finally, let us say a word about
1456: the dependence of our results on the optimization process (determination of the optimal parameters entering $\tilde{\psi}$, 
1457: $\psi_T$, and $\vec{v}$ by minimization of the variance of the improved estimator). Clearly, the method presented in this work 
1458: is useful only if the results obtained do not depend too much on the way the optimization is performed and on which particular 
1459: minimum has been found for the variance (as already emphasized, when a large number of parameters are 
1460: considered the location of such a minimum can depend very crucially on the initial conditions 
1461: and/or on the random numbers series used).
1462: To quantify this aspect we have made 9 independent optimizations over 9 independent sets of 2000 walkers for the Li$_2$ molecule 
1463: at the equilibrium geometry with our best estimator. Results show that the VMC average force results may vary in a 
1464: significant way for the different sets of optimized parameters found. In this case, the domain of variation is about twenty 
1465: times the magnitude of the statistical error and about 30$\%$ of the average itself.
1466: However, it has been observed that the DMC and Hybrid averages are much less sensitive on the optimized parameters.
1467: The error on the DMC and hybrid average forces due to an incomplete optimization has been found to be of the order 
1468: of magnitude of the statistical error which is rather small. 
1469: 
1470: \section{Summary and conclusions}
1471: 
1472: In this work we have shown how to construct improved VMC or DMC estimators 
1473: for observables. By improved it is meant that, compared to the standard 
1474: bare estimator $O$, the new estimators $\tilde{O}$ have a lower variance 
1475: and a reduced systematic error when averaged over the approximate VMC or 
1476: (fixed-node) DMC probability densities.
1477: 
1478: At the Variational Monte Carlo level the most general form we propose
1479: for $\tilde{O}$ is given by
1480: \begin{equation}
1481: \tilde{O}_{VMC}[\psi_T,\tilde{\psi},\vec{v}] \equiv  O 
1482: + \frac{(H-{E_L})\tilde{\psi}}{{\psi_T}}
1483: +2 ({E_L}- \langle E_L\rangle) \frac{ \tilde{\psi}}{\psi_T} +
1484: \frac{\vec{\nabla} [(E_L - \langle E_L\rangle) \psi_T^2 \vec{v}]}
1485: {\psi_T^2}
1486: \label{tildeOconcVMC}
1487: \end{equation}
1488: where averages are defined over the VMC distribution. At the Diffusion Monte Carlo level
1489: the expression proposed is essentially similar, except that 
1490: the average of the local energy entering the definition of $\tilde{O}_{DMC}$ is
1491: defined over the DMC distribution, Eq.(\ref{tildeobias7}).
1492: The various terms defining the improved observables have a well-defined 
1493: physical origin: The three first contributions result from the change 
1494: of the energy average when the magnitude of the observable considered 
1495: as an external field is varied, while the 
1496: last contribution comes from the use of a coordinate transformation 
1497: correlating electron displacements and change of the external field.
1498: The functions $\psi_T$ and $\tilde{\psi}$ appearing in 
1499: the improved observables play the role of trial functions:
1500: $\psi_T$ is the ordinary trial wave function for the ground-state of $H$ and $\tilde{\psi}$ is a 
1501: guess for the derivative of the exact ground-state wave function, 
1502: $\psi_0(\lambda)$, of the perturbed Hamiltonian 
1503: $$
1504: H(\lambda) \equiv H + \lambda O
1505: $$
1506: with respect to $\lambda$ at $\lambda=0$.
1507: When the trial functions are exact, $(\psi_T,\tilde{\psi})=
1508: (\psi_0,\frac{d\psi_0(\lambda)}{d\lambda} 
1509: \big|_{\lambda =0})$, the improved estimator reduces to a constant, namely 
1510: the exact average for the observable. In that case both statistical 
1511: and systematic errors vanish. We have called this 
1512: remarkable property ``Zero-Variance Zero-Bias property''(ZVZB).
1513: In the neighborhood of the exact solution, 
1514: a local expansion of the various quantities obtained from the 
1515: approximate guess $(\psi_T,\tilde{\psi})$ can be 
1516: done. It is found that there is a {\it quadratic} 
1517: behavior in the errors $\delta \psi= \psi_T-\psi_0$ and 
1518: $\delta \psi^\prime= \tilde{\psi}-\psi_0^\prime$. At the VMC level it 
1519: reads
1520: \begin{equation}
1521: \sigma^2(\tilde{O}) \equiv \langle (\tilde{O} - \langle \tilde{O} \rangle_{{\psi_T}^2})^2 
1522: \rangle_{{\psi_T}^2} \sim O[\delta \psi \delta \psi^\prime]
1523: \label{conc1}
1524: \end{equation}
1525: and
1526: \begin{equation}
1527: \Delta_{\tilde O} \equiv
1528: \langle \tilde O \rangle_{{{\psi_T}}^2} - \langle O \rangle_{\psi_0^2}
1529: \sim O[\delta \psi \delta \psi^\prime],
1530: \label{conc2}
1531: \end{equation}
1532: with a similar result in the DMC case.
1533: This important result generalizes the well-known quadratic Zero-Variance 
1534: Zero-Bias property of the energy where the local energy, 
1535: $E_L = H\psi_T/\psi_T$ plays the role of the improved estimator:
1536: \begin{equation}
1537: \sigma^2(E_L) \sim O[{\delta \psi}^2]
1538: \label{conc3}
1539: \end{equation}
1540: and
1541: \begin{equation}
1542: \Delta_{E_L} \sim O[{\delta \psi}^2].
1543: \label{conc4}
1544: \end{equation}
1545: In the case of the energy we can write a Zero-Variance Zero-Bias equation 
1546: defining the optimal trial wave function by imposing that the local energy 
1547: reduces to a constant, namely the
1548: exact energy
1549: \begin{equation}
1550: \frac{H\psi_T}{\psi_T} = E_0.
1551: \label{zvzbeqe0}
1552: \end{equation}
1553: Of course, this equation is nothing but the Schroedinger equation. 
1554: Here, the Zero-Variance Zero-Bias equation for the observable is obtained 
1555: by imposing that the improved observable $\tilde{O}$ reduces to the 
1556: exact average
1557: \begin{equation}
1558: \tilde{O}= \langle O \rangle_{\psi_0^2}.
1559: \end{equation}
1560: By optimizing the three quantities $(\psi_T,\tilde{\psi},\vec{v})$ so that
1561: fluctuations of $\tilde{O}$ are minimal we can obtain the optimal improved estimator
1562: for the observable. In practice, it is done in two steps. First, functional forms 
1563: for the trial functions $\psi_T$ and $\tilde{\psi}$ are chosen in order to reproduce 
1564: the best as possible the exact solution of the zero-variance equations. The choice 
1565: of the vector field $\vec{v}$ is done on physical grounds: It corresponds to the 
1566: electron-coordinate transformation, $\vec{y}=\vec{x}+\lambda \vec{v}(\vec{x})
1567: +O(\lambda^2)$, correlating as much as possible the electron displacements and 
1568: the change of density associated with the external field defined by the bare 
1569: observable. Second, the various parameters entering the three quantities 
1570: are optimized by minimizing the fluctuations of $\tilde{O}$ over a large but 
1571: finite number of configurations (typically, several thousands) drawn according to 
1572: the VMC or DMC distributions.
1573: 
1574: It is important to emphasize that by using the improved estimators presented here 
1575: it is possible to get an accuracy on expectation values of observables which is
1576: comparable to the very good one obtained for total energies. As it can be seen from 
1577: Eqs.(\ref{conc1},\ref{conc2},\ref{conc3},\ref{conc4}) this is true when we are able 
1578: to reduce the error on the derivative of the wave function at the level 
1579: of the error on the wave function itself, that is $\delta \psi \sim {\delta \psi}^\prime$.
1580: 
1581: Another fact worth pointing out is that there is not an unique way of constructing 
1582: improved estimators. Here, we have built our estimators by considering the derivative 
1583: of the variational, Eq.(\ref{evprime}), or the exact DMC energy average, Eq.(\ref{e0vprime}), 
1584: We have also considered the possibility of making a coordinate 
1585: transformation before making the derivative, Eqs.(\ref{eJacprime},\ref{evprime2}). 
1586: Of course, we can think of many other choices and/or transformations. Ultimately, 
1587: the better strategy will depend very much on the specific problem considered.
1588: 
1589: Finally, in order to go beyond VMC or DMC calculations, 
1590: we have shown that the reduction of error of one order in $\delta \psi$ associated 
1591: with the popular ``hybrid''(or ``second-order'') formula mixing DMC 
1592: and VMC averages can be generalized to the case of our improved estimators:
1593: \begin{equation}
1594: 2 \langle \tilde{O}_{DMC} \rangle_{\psi_T  \psi_0} -
1595:   \langle \tilde{O}_{VMC} \rangle_{\psi_T^2} \sim \langle \psi_0 | O |\psi_0 \rangle.
1596: \label{hybridconc}
1597: \end{equation}
1598: 
1599: As an important application we have applied our formalism to the case of the 
1600: computation of forces for some simple diatomic molecules. In our preceding work 
1601: on forces\cite{zvp2} we have focused our attention only on the zero-variance part of 
1602: the problem. More precisely, we have employed i) a simplified version of the renormalized force, Eq.(\ref{f1}) 
1603: ii.) the minimal expression for $\tilde{\psi}$ leading to a finite variance, 
1604: Eqs.(\ref{psi},\ref{q}), and iii.) the hybrid formula mixing 
1605: VMC and DMC calculations, Eq.(\ref{hybridconc}). Results obtained for 
1606: the vanishing force at the equilibrium distance for a number of 
1607: small diatomic molecules were reasonably good.
1608: Here, we have illustrated that such a strategy is in fact not valid for 
1609: describing the global shape of the force curve. It has been shown that results 
1610: depend very much on the trial wave function used and, particularly, on the quality 
1611: of the optimization process of the numerous parameters of the trial wave function. 
1612: As a result, the force curves obtained are not regular as a function of the 
1613: interatomic distance and important spectroscopic quantities such as the equilibrium 
1614: distance $R_e$ and the harmonic frequency $\omega_e$
1615: cannot be obtained reliably. To get accurate curves
1616: we need not only to have a small amount of statistical fluctuations but also 
1617: a control of the systematic error.
1618: By exploiting the general ZVZB principle presented in this work it has been shown that 
1619: obtaining accurate curves is now possible. The basic ingredients are: i.) the use of 
1620: a trial wave function for the derivative, $\tilde{\psi}$, built as a 
1621: finite-difference of the trial wave function with respect to the nuclear coordinate ii.) the use of a
1622: coordinate transformation in the spirit of the work of Filippi and Umrigar\cite{fil3} and, finally, iii.) 
1623: the systematic minimization of the variance of the improved estimator 
1624: with respect to all the parameters entering the two trial functions ($\psi_T,\tilde{\psi}$), and the vector field, $\vec{v}$,
1625: associated with the coordinate transformation. Let us emphasize that to get a well-balanced optimization of the 
1626: two trial functions (leading to smooth curves for the forces) , it is essential to reduce the variance of the improved 
1627: estimator for the force at the level of the variance of the local energy. Although such a condition may appear as 
1628: very difficult to fulfill (local energies have usually very small variances), we have shown that
1629: it is in fact possible thanks to the coordinate-transformation term.
1630: Such a result is remarkable and is certainly one of most important 
1631: practical aspect of the approach proposed in this work.
1632: 
1633: Finally, let us remark that the price to pay with respect to the minimal scheme presented in our previous work\cite{zvp2}
1634: lies on the need of computing about $3N_{nucl}$ local energies to calculate the various
1635: components of the force. However, we do not think it represents a major difficulty for 
1636: the realistic applications to come. Indeed, the few-percent accuracy needed on average forces
1637: will be obtained with relatively small statistics and, therefore, 
1638: it will not be necessary to compute the force vector at each Monte Carlo step
1639: (the expensive $3N_{nucl}$-local energy-calculation step will be done rarely). 
1640: In addition to this, in applications where the nuclear geometry is varied 
1641: during the simulation (Molecular Dynamics-type applications) it should also be possible to use 
1642: suitable re-actualization schemes to avoid re-computing entirely the $3N_{nucl}$ local energies 
1643: for close nuclear configurations. Of course, the validity of these various strategies as well as
1644: the quality of the improved estimators presented in this work need now to be 
1645: checked for realistic applications involving many nuclear degrees of freedom.
1646: 
1647: {\bf Acknowledgments}
1648: This work was supported by the ``Centre National de la Recherche 
1649: Scientifique'' (CNRS).\\
1650: 
1651: \begin{thebibliography}{99}
1652: 
1653: \bibitem{wil1}
1654: A.J. Williamson, J.C. Grossman, R.Q. Hood, A. Puzder, and G. Galli, Phys. Rev. Lett. {\bf 89}, 
1655: 196803 (2002).
1656: \bibitem{wil2}
1657: A. Puzder, A.J. Williamson, J.C. Grossman, and G. Galli, J. Chem. Phys. {\bf 117}, 6721 (2002).
1658: \bibitem{wil3}
1659: A. Puzder, A.J. Williamson, J.C. Grossman, and G. Galli, Phys. Rev. Lett. {\bf 88}, 097401 (2002).
1660: \bibitem{mitas1}
1661: G. Belomoin, E. Rogozhina, J. Therrien, P.V. Braun, L. Abuhassan, M.H. Nayfeh, L. Wagner, and L. Mitas,
1662: Phys. Rev. B {\bf 65}, 193406 (2002).
1663: \bibitem{fil1}
1664: C. Filippi, S.B. Healy, P. Kratzer, E. Pehlke, and M. Scheffler {\bf 89}, 166102 (2002).
1665: \bibitem{needs1}
1666: A.R. Porter, M.D. Towler, and R.J. Needs, Phys. Rev. B {\bf 64}, 035320 (2001).
1667: \bibitem{mitas2} 
1668: L. Mitas, J. Therrien, R. Twesten, G. Belomoin, and M.H. Nayfeh, Appl. Phys. Lett. {\bf 78}, 1918 (2001).
1669: \bibitem{lester1}
1670: I.V. Ovcharenko, W.A. Lester,Jr.,C. Xiao, and F. Hagelberg, J. Chem. Phys. {\bf 114}, 9028 (2001).
1671: \bibitem{wil4}
1672: A.J. Williamson, R.Q. Hood, and J.C. Grossman, Phys. Rev. Lett. {\bf 87}, 246406 (2001)
1673: \bibitem{fil2}
1674: S.B. Healy, C. Filippi, P. Kratzer, E. Penev, and M. Scheffler, Phys. Rev. Lett. {\bf 87}, 016105 (2001).
1675: \bibitem{needs2}
1676: P.R.C. Kent, M.D. Towler, R.J. Needs, and G. Rajagopal, Phys. Rev. B {\bf 62}, 15394 (2000).
1677: \bibitem{mitas3}
1678: T. Torelli and L. Mitas, Phys. Rev. Lett. {\bf 85}, 1702 (2000).
1679: \bibitem{mitas4}
1680: L. Mitas, J.C. Grossman, I. Stich, and J. Tobik, Phys. Rev. Lett. {\bf 84}, 1479 (2000).
1681: \bibitem{anderson0}
1682: Y. Shlyakhter, S. Sokolova, A. L\"{u}chow, 
1683: and J.B. Anderson, J. Chem. Phys. {\bf 110}, 10725 (1999).
1684: \bibitem{lester2}
1685: See e.g. B.L. Hammond, W.A. Lester,Jr., and P.J. Reynolds in {\it Monte Carlo Methods in Ab Initio
1686: Quantum Chemistry}, World Scientific Lecture and course notes in chemistry Vol.1 (1994).
1687: \bibitem{ceperley1} 
1688: D.M. Ceperley and M.H. Kalos, in {\it Monte Carlo Methods in Statistical Physics}, edited 
1689: by K. Binder (Springer, Berlin, 1979), pp.145-194.
1690: \bibitem{liu1}
1691: K.S. Liu, M.H. Kalos, and G.V. Chester, Phys. Rev. A {\bf 10}, 303 (1974).
1692: \bibitem{caffaclav}
1693: M. Caffarel and P. Claverie, J. Chem. Phys. {\bf 88}, 1088 (1988).
1694: \bibitem{casul1}
1695: J. Casulleras and J. Boronat, Phys. Rev. B {\bf 52}, 3654 (1995).
1696: \bibitem{zvp1}
1697: R. Assaraf and M. Caffarel, Phys. Rev. Lett. {\bf 83 }, 4682 (1999).
1698: \bibitem{zvp2}
1699: R. Assaraf and M. Caffarel, J. Chem. Phys. {\bf 113},4028 (2000).
1700: \bibitem{wells1}
1701: B.H. Wells, Chem. Phys. Lett. {\bf 115}, 89 (1985).
1702: \bibitem{traynor1}
1703: C.A. Traynor and J.B. Anderson, Chem. Phys. Lett. {\bf 147}, 389 (1988).
1704: \bibitem{umrigar1} 
1705: C.J. Umrigar, Int. J. Quantum Chem. {\bf 23}, 217 (1989).
1706: \bibitem{sun1}
1707: Z. Sun, W.A. Lester,Jr.,and B.L. Hammond, J. Chem. Phys. {\bf 97}, 7585 (1992).
1708: \bibitem{vrbik1}
1709: J. Vrbik, D.A. Lagare and S.M. Rothstein, J. Chem. Phys. {\bf 92}, 1221 (1990).
1710: \bibitem{vrbik2}
1711: J. Vrbik and S.M. Rothstein, J. Chem. Phys. {\bf 96}, 2071 (1991).
1712: \bibitem{belohorec1}
1713: P. Belohorec, S.M. Rothstein, and J. Vrbik, J. Chem. Phys. {\bf 98}, 6401 (1993)
1714: \bibitem{fil3}
1715: C. Filippi and C.J. Umrigar, Phys. Rev. B {\bf 61}, R16291 (2000).
1716: \bibitem{baer}
1717: R. Baer, J. Chem. Phys. {\bf 113}, 473 (2000).
1718: \bibitem{umrigarwrapped}
1719: C. J. Umrigar, Int J. Quantum Chem., Symp. {\bf 23}, 217 (1989).
1720: \bibitem{cmr}
1721: M. Casalegno, M. Mella, and A.M. Rappe, J. Chem. Phys. {\bf 118}, 7193 (2003).
1722: \bibitem{pulay} 
1723: P. Pulay, Mol. Phys. {\bf 17} 197 (1969).
1724: \bibitem{ceperley2}
1725: E.L. Pollock and D.M. Ceperley, Phys.Rev. B {\bf 30}, 2555 (1984).
1726: \bibitem{cafstat}
1727: M. Caffarel and P. Claverie, J. Stat. Phys. {\bf 43}, 797 (1986).
1728: \bibitem{umrigar}
1729: C. J. Umrigar, K. G. Wilson, and J. W. Wilkins, Phys. Rev. Lett. {\bf 60},
1730: 1719 (1988).
1731: \bibitem{herz}
1732: K. P. Huber and G. Herzberg, {\it Molecular Spectra and Molecular Structure IV, Constants
1733: of Diatomic Molecules} (van Nostrand Reinhold, New York, 1979).
1734: \bibitem{amu} A. H. Wapstra and G. Audi, Nucl. Phys. A {\bf 432}, 1 (1985).
1735: \end{thebibliography}
1736: 
1737: \textwidth=18truecm
1738: \begin{table}[htp]
1739: \caption{VMC, DMC, and Hybrid estimates of the equilibrium geometry $R_e$ (a.u.)
1740: and harmonic frequency $\omega_e$ (cm$^{-1}$). The atomic isotopic masses taken$^a$ are 1.007825035 
1741: amu for $^1$H and 7.0160030 amu for $^7$Li.}
1742: \begin{tabular}{llll}
1743: \multicolumn{1}{c}{               } &
1744: \multicolumn{1}{c}{H$_2$          } &
1745: \multicolumn{1}{c}{LiH            } &
1746: \multicolumn{1}{c}{Li$_2$         } \\
1747: \hline
1748: R$_e$ (VMC) & 1.463(12)    & 3.111(17)   &  5.346(27) \\
1749: R$_e$ (DMC) & 1.426(13)    & 3.056(6)   &   5.200(16) \\
1750: R$_e$ (Hybrid) & 1.395(15)    & 3.001(15)   &  5.068(27)  \\
1751: R$_e$ (Exp.)$^b$ & 1.401    & 3.015   &  5.051             \\
1752: \hline
1753: $\omega_e$ (VMC) & 4194(130) & 1559(40)   &    366(9)           \\
1754: $\omega_e$ (DMC) &  4432(165) & 1549(22)   &   373(5)            \\
1755: $\omega_e$ (Hybrid) & 4662(205) & 1519(31)   &  387(8)         \\
1756: $\omega_e$ (Exp.)$^b$ & 4395.2     & 1405.65   &  351.4            \\
1757: \end{tabular}
1758: \raggedright
1759: \small{}
1760: $^a$ Ref.\cite{amu}\\
1761: $^b$ Ref.\cite{herz}\\
1762: \label{fitdata}
1763: \end{table}
1764: 
1765: \begin{table}[htp]
1766: \caption{VMC, DMC, and Hybrid systematic (bias) and statistical errors for the total
1767: energy and various force estimators for Li$_2$ at R=5.051 a.u. The VMC variances,
1768: $\sigma^2$ (VMC), are also given. $\langle E_L\rangle_{ex}$ and $\langle F\rangle_{ex}$ denote the exact
1769: total energy, $\langle E_L\rangle_{ex}$=-14.9954 a.u. and exact force $\langle F\rangle_{ex}=0.$
1770: (equilibrium geometry), respectively. To facilitate comparisons between
1771: energy and force results all averages are given with five significant figures after the
1772: decimal point and statistical errors are given on the fifth decimal place (magnitude $10^{-5}$).
1773: Statistical errors on VMC variances are on the last digit.}
1774: \begin{tabular}{ccccc}
1775: \multicolumn{1}{c}{Estimator      } &
1776: \multicolumn{1}{c}{VMC Average    } &
1777: \multicolumn{1}{c}{DMC Average    } &
1778: \multicolumn{1}{c}{Hybrid         } &
1779: \multicolumn{1}{c}{$\sigma^2$(VMC)} \\
1780: \hline
1781: $E_L - \langle E_L\rangle_{ex}$
1782: & 0.03871(32)    & 0.00531(50)   & -               & 0.113(5)    \\
1783: $^a F - \langle F \rangle_{ex} $
1784: &0.18217(23216) & 0.15462(12293)& 0.12707(33185) & $+\infty$    \\
1785: $^b \tilde{F}_{ZV} [\psi_T, \tilde{\psi}_{min}] - \langle F \rangle_{ex}$
1786: & -0.06352(84)& -0.04003(151)&-0.01654(313)& 1.27(5) \\
1787: $^c \tilde{F}_{ZVZB} [\psi_T, \tilde{\psi}_{min}]- \langle F\rangle_{ex}$
1788: &  -0.05802(104) &-0.02484(184)&0.00834(382) & 1.3(2)      \\
1789: $^d \tilde{F}[\psi_T,\tilde{\psi}_{min} | \tilde{\psi}_{Deriv}]- \langle F\rangle_{ex}$
1790: & 0.00619(109) &0.02993(187) &0.05367(390) & 2.8(1)      \\
1791: $^e \tilde{F}_{ZVZB} [\psi_T, \tilde{\psi}_{Deriv},\vec{v}=0]- \langle F \rangle_{ex}$
1792: & 0.00871(218)&0.00474(147)&0.00077(366) & 14(3)       \\
1793: $^e \tilde{F}_{ZVZB} [\psi_T,\tilde{\psi}_{Deriv},\vec{v}]- \langle F\rangle_{ex}$
1794: &0.00692(9) &0.00358(19)&0.00024(39)& 0.016(1)    \\
1795: \end{tabular}
1796: \raggedright
1797: %\small{}
1798: a. Eq.(\ref{localforcediatomics})\\
1799: b. Eq.(\ref{f1})\\
1800: c. Eq.(\ref{f2})\\
1801: d. Eq.(\ref{minder})\\
1802: e. Eq.(\ref{fpsir})\\
1803: \label{tableLi2}
1804: \end{table}
1805: \textwidth=16truecm
1806: 
1807: \newpage
1808: \begin{center}
1809: FIGURE CAPTIONS
1810: \end{center}
1811: 
1812: \begin{itemize}
1813: \item Fig.\ref{figureEH2} H$_2$ molecule. Variational Monte Carlo (VMC) energies (open squares), Diffusion Monte Carlo (DMC)
1814: energies (filled squares) and exact non-relativistic curve (solid line). The dotted line between
1815: VMC results is a simple linear interpolation to guide the eye.
1816: 
1817: \item Fig.\ref{figureELiH}
1818: LiH molecule. Variational Monte Carlo (VMC) energies (open squares), fixed-node
1819: Diffusion Monte Carlo (DMC) energies (filled squares),
1820: and exact non-relativistic curve (solid line). The dotted line between
1821: VMC results is a simple linear interpolation to guide the eye.
1822: 
1823: \item Fig.\ref{figureELi2} Li$_2$ molecule. Variational Monte Carlo (VMC) energies (open squares),
1824: fixed-Node Diffusion Monte Carlo (DMC) energies (filled squares),
1825: and exact non-relativistic curve (solid line). The dotted line between
1826: VMC results is a simple linear interpolation to guide the eye.
1827: 
1828: \item Fig.\ref{forcevar} Various VMC average forces for Li$_2$. Filled squares 
1829: with large error bars: $\langle F \rangle$, Eq.(\ref{localforcediatomics}).
1830: Open squares joined by the dashed line: $\langle \tilde{F}_{VMC-ZV}[\psi_T,\tilde{\psi}_{min}]\rangle$, Eq.(\ref{f1});
1831: Circles joined with the dotted line: $\langle \tilde{F}_{VMC-ZVZB}[\psi_T,\tilde{\psi}_{min}]\rangle$,
1832: Eq.(\ref{f2}). Solid line: exact non-relativistic force curve.
1833: 
1834: \item Fig.\ref{forcedmc} Li$_2$ molecule. Average forces using $\tilde{F}_{ZVZB} (\tilde{\psi_T},
1835: \tilde{\psi_{min}})$,Eq.(\ref{f2},\ref{f3},\ref{hybridfinal2}).
1836: VMC average: lowest curve with open circles. DMC average: intermediate curve with filled squares.
1837: Hybrid average: highest curve with open squares. Solid line: exact non-relativistic force curve.
1838: Dashed lines between QMC results are a simple linear interpolation to guide the eye.
1839: 
1840: \item Fig.\ref{fLi2psiR} VMC force for Li$_2$.
1841: Lowest irregular curve with filled squares: $\langle \tilde{F}_{VMC-ZV}[\psi_T,\tilde{\psi}_{min}]\rangle$, Eq.(\ref{f1}).
1842: Upper curve with open squares: $\langle \tilde{F}_{VMC-ZV}[\psi_T,\tilde{\psi}_{Deriv},\vec{v}]\rangle$, Eq.(\ref{fpsir}).
1843: Solid line: exact non-relativistic force curve.
1844: 
1845: \item Fig.\ref{fLi2finalvmc} VMC force for Li$_2$.
1846: Open squares: Average VMC forces from estimator (\ref{fpsir}) using the Jastrow-RHF one-determinant wave function.
1847: Open circles: Average VMC forces from estimator (\ref{fpsir}) using the Jastrow-VB six-determinant wave function.
1848: Solid line: exact non-relativistic force curve.
1849: 
1850: \item Fig.\ref{fLi2finaldmc} DMC force for Li$_2$.
1851: Open squares: Average fixed-node DMC forces using the Jastrow-RHF one-determinant wave function.
1852: Open circles: Average fixed-node DMC forces using the Jastrow-VB six-determinant wave function.
1853: Solid line: exact non-relativistic force curve.
1854: 
1855: \item Fig.\ref{fH2final} Hybrid force for H$_2$.
1856: Solid line: exact non-relativistic force curve.
1857: 
1858: \item Fig.\ref{fLiHfinal} Hybrid force for LiH.
1859: Solid line: exact non-relativistic force curve. 
1860: 
1861: \item Fig.\ref{fLi2finalhyb} Hybrid force for Li$_2$.
1862: Solid line: exact non-relativistic force curve.
1863: 
1864: \end{itemize}
1865: 
1866: %Fig1
1867: \begin{center}
1868: \begin{figure}[htp]
1869: \includegraphics[height=18cm,width=18cm,angle=0]{eH2.ps}
1870: \caption{}
1871: \label{figureEH2}
1872: \end{figure}
1873: \end{center}
1874: 
1875: %Fig2
1876: \begin{center}
1877: \begin{figure}[htp]
1878: \includegraphics[height=18cm,width=18cm,angle=0]{eLiH.ps}
1879: \caption{}
1880: \label{figureELiH}
1881: \end{figure}
1882: \end{center}
1883: 
1884: %Fig3
1885: \begin{center}
1886: \begin{figure}[htp]
1887: \includegraphics[height=18cm,width=18cm,angle=0]{eLi2.ps}
1888: \caption{}
1889: \label{figureELi2}
1890: \end{figure}
1891: \end{center}
1892: 
1893: %Fig4
1894: \begin{center}
1895: \begin{figure}[htp]
1896: \includegraphics[height=18cm,width=18cm,angle=0]{fLi2min.ps}
1897: \caption{}
1898: \label{forcevar}
1899: \end{figure}
1900: \end{center}
1901: 
1902: %Fig5
1903: \begin{center}
1904: \begin{figure}[htp]
1905: \includegraphics[height=18cm,width=13cm,angle=-90]{fLi2minDMC.ps}
1906: \caption{}
1907: \label{forcedmc}
1908: \end{figure}
1909: \end{center}
1910: 
1911: %Fig6
1912: \begin{center}
1913: \begin{figure}[htp]
1914: \includegraphics[height=18cm,width=13cm,angle=-90]{fLi2psiR.ps}
1915: \caption{}
1916: \label{fLi2psiR}
1917: \end{figure}
1918: \end{center}
1919: 
1920: %Fig7
1921: \begin{center}
1922: \begin{figure}[htp]
1923: \includegraphics[height=18cm,width=13cm,angle=-90]{fLi2finalvar.ps}
1924: \caption{}
1925: \label{fLi2finalvmc}
1926: \end{figure}
1927: \end{center}
1928: 
1929: %Fig8
1930: \begin{center}
1931: \begin{figure}[htp]
1932: \includegraphics[height=18cm,width=13cm,angle=-90]{fLi2finaldmc.ps}
1933: \caption{}
1934: \label{fLi2finaldmc}
1935: \end{figure}
1936: \end{center}
1937: 
1938: %Fig9
1939: \begin{center}
1940: \begin{figure}[htp]
1941: \includegraphics[height=18cm,width=13cm,angle=-90]{fH2final.ps}
1942: \caption{}
1943: \label{fH2final}
1944: \end{figure}
1945: \end{center}
1946: 
1947: %Fig10
1948: \begin{center}
1949: \begin{figure}[htp]
1950: \includegraphics[height=18cm,width=13cm,angle=-90]{fLiHfinal.ps}
1951: \caption{}
1952: \label{fLiHfinal}
1953: \end{figure}
1954: \end{center}
1955: 
1956: %Fig11
1957: \begin{center}
1958: \begin{figure}[htp]
1959: \includegraphics[height=18cm,width=13cm,angle=-90]{fLi2finalhyb.ps}
1960: \caption{}
1961: \label{fLi2finalhyb}
1962: \end{figure}
1963: \end{center}
1964: 
1965: \end{document}
1966: