1: %\documentclass[twocolumn,showpacs,preprintnumbers,aps,eqlabels]{revtex4}
2: \documentclass[twocolumn,prl]{revtex4}
3: %\documentclass[preprint,prl]{revtex4}
4:
5: \usepackage{graphicx}% Include figure files
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8:
9: %\nofiles
10:
11: \def\mb#1{{\boldsymbol #1}}
12: \def\bm{\boldsymbol}
13:
14: \def\bea{\begin{eqnarray}}
15: \def\eea{\end{eqnarray}}
16:
17: \def\ben{\begin{equation}}
18: \def\een{\end{equation}}
19:
20: \input rotate
21:
22:
23: \begin{document}
24:
25: \preprint{DIPC}
26:
27: \title{Hellman-Feynman operator sampling in Diffusion Monte Carlo
28: %On the correct sampling of operators in Diffusion Monte Carlo
29: calculations}
30: \author{R. Gaudoin$^1$ and J. M. Pitarke$^{2,3}$}
31: \affiliation{
32: $^1$Donostia International Physics Center (DIPC),
33: Manuel de Lardizabal Pasealekua, E-20018 Donostia, Basque Country, Spain\\
34: $^2$CIC nanoGUNE Consolider, Mikeletegi Pasealekua 56, E-2009 Donostia, Basque
35: Country, Spain\\
36: $^2$Materia Kondentsatuaren Fisika Saila, UPV/EHU, and Unidad F\'\i
37: sica Materiales
38: CSIC-UPV/EHU, 644 Posta kutxatila, E-48080
39: Bilbo, Basque Country, Spain}
40:
41:
42: \date{\today}
43:
44: \begin{abstract}
45: Diffusion Monte Carlo (DMC) calculations typically yield highly accurate
46: results in solid-state and quantum-chemical calculations. However, operators
47: that do not
48: commute with the Hamiltonian are at best sampled correctly up to second order
49: in the
50: error of the underlying trial wavefunction, once simple corrections have been
51: applied.
52: This error is of the same order as that for the energy in variational
53: calculations. Operators
54: that suffer from these problems include potential energies and the
55: density. This paper presents a new method, based on the Hellman-Feynman
56: theorem,
57: for the correct DMC sampling of
58: all operators diagonal in real space. Our method is easy to implement in
59: any standard DMC code.
60: \end{abstract}
61:
62: \pacs{71.10.Ca,71.15.-m}
63: %\keywords{Suggested keywords}%Use showkeys class option if keyword
64: %display desired
65: \maketitle
66:
67:
68: Diffusion Monte Carlo (DMC) is widely used for the computation of
69: properties of solids and molecules~\cite{qmc}.
70: Frequently, it is used as a check on other methods~\cite{check} or even as an
71: input~\cite{input}. It is
72: therefore very important that DMC be as accurate as possible.
73: However, other than for the total energy, standard DMC calculatioins are not
74: as definitive as one would hope, since operators that do not commute with the
75: Hamiltonian cannot be sampled exactly within standard DMC. Here we present a
76: simple yet effective addition
77: to standard DMC that plugs that gap and is easy to implement.
78:
79: DMC by construction yields the normalized
80: expectation value $\langle \hat{O}\rangle_{DMC}=
81: \langle \Psi_T| \hat{O}|\Psi^{fn}_0
82: \rangle/\langle \Psi_T|\Psi^{fn}_0\rangle$,
83: which is generally not the true ground-state expectation value
84: $\langle \hat{O}\rangle=
85: \langle \Psi_0| \hat{O}|\Psi_0
86: \rangle/\langle \Psi_0|\Psi_0\rangle$. In fact, it is not even
87: $\langle \hat{O}\rangle_{fn}=
88: \langle \Psi^{fn}_0| \hat{O}|\Psi^{fn}_0
89: \rangle/\langle \Psi^{fn}_0|\Psi^{fn}_0\rangle$, the ground-state expectation
90: value
91: constrained by a nodal structure of the fermionic many-body wavefunction that
92: is given by $\Psi_T$.
93: $\Psi_T$ is a trial wavefunction that approximates the generally unknown
94: ground-state
95: wavefunction $\Psi_0$ and is real. In its basic and most common form
96: $\Psi^{fn}_0$ is the ground state for a fixed nodal structure given by that
97: of $\Psi_T$. In addition to this fixed-node
98: approximation operators that do not commute with the Hamiltonian are
99: generally subject to a further error, the leading term of which is linear in
100: the
101: difference between $\Psi_T$ and $\Psi^{fn}_0$. In conjunction with Variational
102: Monte Carlo (VMC), this error can be reduced by one order~\cite{cep_err} by
103: using $\langle \hat{O}\rangle_{cDMC}=2\langle \hat{O}\rangle_{DMC}-\langle
104: \hat{O}\rangle_{VMC}$.
105: Correct sampling can be achieved, e.g. by
106: using forward walking~\cite{fwalking}, reptation Monte Carlo~\cite{rmc}, and
107: other methods~\cite{no_tagging}.
108: Many of these methods aim to sample $\Psi_0^{fn}\Psi_0^{fn}$, rather than the
109: usual DMC distribution
110: $\Psi_T\Psi_0^{fn}$. They are therefore not straight forward additions to the
111: DMC algorithm. Alternatively, the Virial theorem or the related
112: Hellman-Feynman (HF) theorem~\cite{hf} can be used to evaluate operator
113: expectation values~\cite{ob} which in the case
114: of DMC, however, involves numerical derivatives of noisy data.
115:
116: In this Letter, we present a method based on the application of the HF theorem
117: to the
118: DMC algorithm directly. Our method - Hellman-Feynman sampling (HFS) - can be
119: tagged onto the usual sampling of
120: operators with nearly no extra computational overhead. The aim is to maintain
121: the basic DMC
122: algorithm that samples $\Psi_T\Psi_0^{fn}$.
123: Now, the total energy is evaluated correctly within standard DMC, and
124: crucially operator expectation values can be cast as HF derivatives of the
125: total
126: energy.
127: Keeping in mind that ultimately the DMC algorithm is nothing but a large
128: sum that yields the total energy, we see the HF derivative can be applied
129: without problem
130: to the algorithm itself!
131: One advantage over numerical derivatives is that the resulting formula can
132: handle
133: several operators simultaneously in a single DMC run and maintaining
134: orbital occupancy for perturbed Hamiltonians
135: ceases to be a problem.
136: The DMC algorithm only involves numbers, so
137: non-commutability of operators - the source of the difficulties -
138: is no issue either. Writing down the DMC algorithm as a mathematical formula
139: and applying the
140: HF derivative to it yields an object that when sampled using standard DMC
141: produces
142: the exact operator expectation value. It has to by construction.
143:
144: In the following, we present a schematic overview of the DMC algorithm,
145: which however is sufficient to derive the relevant formulas.
146: The basic idea of DMC is to split the
147: imaginary-time propagator
148: $\exp(-\Delta t \hat{H}) \approx \exp(-\Delta t \hat{T})\exp(-\Delta t
149: \hat{V})$
150: for sufficiently small time intervals $\Delta t$
151: into a kinetic and potential term
152: and then to iterate it. This ultimately~\cite{note_dmc_alg} gives rise to a
153: real-space drift-diffusion process sampled using Monte Carlo (MC), augmented
154: by an exponential growth term whereby $N_w$ so-called walkers
155: are propagated in parallel. Courtesy of this growth term, at each propagation
156: or (imaginary) time
157: step $i$ the walker $j$ acquires a
158: multiplicative weight: $e^{-\Delta t \left(E^L_{i,j}-\tilde{E}^0_i\right)}$, where
159: $E^L_{i,j}=\hat{H}\Psi_T/\Psi_T$ evaluated at the real-space position of walker
160: $j$ at time step $i$ and $\tilde{E}^0_i$ is an estimate for the ground-state energy
161: also at time step $i$.
162: The total weight of walker $j$ at time step $i$ becomes
163: \ben
164: \label{def_omega}
165: \omega_{i,j}=\prod_{k=1}^{i}
166: e^{-\Delta t \left(E^L_{k,j}-E^0_{i}\right)},\, \mbox{where}\,
167: E^0_i=\frac{1}{i}\sum_{l=1}^{i} \tilde{E}^0_l
168: \een
169: and the presence of $E^0_i$ ensures normalization.
170: At time step $i$ the estimator for an operator that a DMC code yields is
171: \ben
172: \label{dmc_alg}
173: \overline{O^L_i} = \sum_j^{N_w} \omega_{i,j} O_{i,j}^L,
174: \een
175: where
176: $O_{i,j}^L=\hat{O}\Psi_T/\Psi_T$ and the wavefunction $\Psi_T$ is evaluated
177: for walker $j$ at time step $i$. For brevity, we use this bar-average
178: $\overline{O^L_i}$
179: where applicable and note that $\overline{O^L_i}$ has to be averaged over all
180: $i$
181: to yield the final DMC estimate $\langle \hat{O}\rangle_{DMC}$. Since the
182: ground-state energy is not
183: known, an estimate chosen such that Eq. (\ref{def_omega}) remains normalized
184: has to be used. This is the growth
185: estimator $E^0_i$~\cite{Etilde} and
186: is updated at each step, hence the index $i$.
187: Note that $E^0_i$ is independent of $j$, i.e. it is the same for every walker
188: and thus a property of the DMC process as a whole.
189: For reasons of numerical stability, DMC is implemented by allowing walkers to
190: die
191: or multiply such that the walker's survival probability optionally augmented by
192: residual weights corresponds to Eq. (\ref{def_omega}).
193:
194: Given a perturbed Hamiltonian $\hat{H}(\alpha)=\hat{H}+\alpha \hat{O}$ and
195: the associated fixed-node ground state energy
196: $ E_{fn}(\alpha)=\langle\hat{H}\rangle_{DMC}$,
197: first-order perturbation theory for $\Psi_0^{fn}$ yields a fixed node
198: equivalent of the HF theorem~\cite{HFproblem}
199: \ben
200: \label{hf_exp}
201: \langle O \rangle_{fn} =\left.
202: \frac{\partial E_{fn}(\alpha)}{\partial\alpha}
203: \right|_{\alpha=0},
204: \een
205: where $\langle O \rangle_{fn}$ converges to the correct
206: ground-state expectation value as the nodes of $\Psi_T$ become
207: exact though $\Psi_T$ itself need not. Note that while
208: $\langle\hat{H}\rangle_{DMC}=\langle\hat{H}\rangle_{fn}$ we have
209: $\langle\hat{O}\rangle_{DMC}\neq\langle\hat{O}\rangle_{fn}$,
210: unless $[\hat{O},\hat{H}]=0$, so Eq. (\ref{hf_exp}) is not trivial.
211: The energy $E_{fn}(\alpha)$ is
212: accessible exactly within standard DMC as the Hamiltonian
213: $\hat{H}(\alpha)$
214: commutes with itself.
215: Analytic operator estimators can then be derived by
216: applying the HF theorem to the formula expressing the DMC algorithm Eq.
217: (\ref{dmc_alg}).
218: Using Eqs. (\ref{def_omega}) and (\ref{dmc_alg}) the expectation value at time
219: step $i$ becomes
220: \ben
221: \label{def_eal}
222: E_i(\alpha) = \sum_j^{N_w} E^L_{i,j}(\alpha)\prod_{k=1}^{i} e^{-\Delta t
223: \left(E^L_{k,j}(\alpha)-E^0_i(\alpha)\right)}. \een
224: Here, $E^L_{i,j}(\alpha)=E^L_{i,j}+\alpha O^L_{i,j}$ and
225: $E^0_i(\alpha)= E^0_i+\Delta E^0_i(\alpha)$, so the weight of the wavefunction
226: is
227: \bea \Omega_i&=&\sum_j^{N_w} \underbrace{ \exp\left(-\Delta t
228: \sum_{k=1}^{i}\left(E^L_{k,j}-E^0_i\right)\right)}_{\omega_{i,j}}
229: \nonumber\\&& \times\exp\left({-\Delta t \sum_{k=1}^{i}\left(\alpha
230: O^L_{k,j}-\Delta E^0_i(\alpha)\right)}\right)
231: \\ &=&
232: \overline{
233: \exp\left(-t\alpha X_{i}\right)}
234: \exp\left(t\Delta E^0_i(\alpha)\right),
235: \eea
236: where $X_{i,j}=\frac{1}{i}\sum_{k=1}^{i} O^L_{k,j}$ and $t=i\Delta t$,
237: and we have made use of the fact that the growth estimator
238: $\Delta E^0_i(\alpha)$ is independent of the index $j$. $\Delta E^0_i(\alpha)$
239: ensures that $\Omega_i(\alpha)=1$ to all orders of $\alpha$, hence
240: $\Delta E^0_i(\alpha)=
241: -\frac{1}{t}\log\left[\overline{
242: \exp\left(-t\alpha X_{i}\right)}\right]$.
243: Eq. (\ref{def_eal}) then becomes
244: \ben
245: \label{def_eal2}
246: E_i(\alpha)=\frac{
247: \overline{ E^L_{i}(\alpha)e^{-t\alpha X_{i}}}
248: }{
249: \overline{ e^{-t\alpha X_{i}}}
250: }.
251: \een
252: Evaluating
253: $\Delta E^0_i$ to first order gives the growth estimator of an
254: operator:
255: \ben
256: \label{def_de0}
257: O^{GR}_i=\left.\frac{\partial E^0_i(\alpha)}{\partial\alpha}\right|_{\alpha=0}=
258: \left.\frac{\partial \Delta E^0_i(\alpha)}{\partial\alpha}\right|_{\alpha=0}=
259: \overline{ X_{i}}.
260: \een
261: In other words, the DMC sampling of $X_{i,j}$
262: by virtue of the HF theorem yields a growth estimator of the
263: true expectation value
264: of $\hat{O}$.
265: Interestingly, the growth estimator, if the residual weights are chosen to be
266: zero,
267: appears to be similar to Eq. (13) of Ref.~\cite{no_tagging}.
268: Applying the HF theorem to the energy estimator
269: Eq. (\ref{def_eal2}) yields
270: a second estimator
271: \ben
272: \label{def_de1}
273: O^{E}_i=\left.\frac{\partial E_i(\alpha)}{\partial\alpha}\right|_{\alpha=0}=
274: \overline{ O^L_{i} }
275: -t\left(
276: \overline{ E^L_{i} X_{i} }
277: -\overline{ E^L_{i}} \cdot \overline{ X_{i}}
278: \right).
279: \een
280: Equations (\ref{def_de0}) and (\ref{def_de1}) are of course
281: evaluated at $\alpha=0$ and are therefore
282: accessible in a regular DMC calculation.
283: We see that for $O^{E}_i$ the standard estimator $\overline{ O^L_{i}}$ is
284: augmented by a correction
285: term $\Delta O^{E}_i=-t\left( \overline{ E^L_{i} X_{i}}
286: -\overline{ E^L_{i}}\cdot
287: \overline{ X_{i}} \right)$. Several observations can be made. First,
288: in the case of the $\Psi_T$ being the ground state $\Psi^{fn}_0$
289: for a given nodal structure
290: the correction term is zero ($ E^L_{i,j}$ is a constant!)
291: and only $\overline{ O^L_{i}}$ contributes as it
292: should. Furthermore, the new estimator $O^{E}_{i}$ and the usual one
293: $\overline{ O^L_{i}}$
294: sample an observable and are both independent of the auxiliary
295: DMC parameter $t$. It follows that $
296: \overline{ E^L_{i} X_{i}} -\overline{ E^L_{i}}\cdot
297: \overline{ X_{i}}
298: \sim \frac{1}{t}$.
299: %Thirdly, in keeping with DMC traditions the growth estimator Eq.
300: %(\ref{def_de0}) seems
301: %noisier than the direct estimator Eq. (\ref{def_de1}); hence, only the latter
302: %is
303: %discussed in the rest of the paper.
304: Thirdly, since the growth estimator Eq.
305: (\ref{def_de0}) is derived from the ``averaged'' quantity $E^0_i$ rather than
306: $\tilde{E}^0_i$, Eq. (\ref{def_de0}) is itself already averaged over $i$ and therefore
307: the final estimate at $i$. This is in contrast to Eq. (\ref{def_de1}) which still has to be
308: averaged over all $i$ to yield the final DMC estimate. Using $\tilde{E}^0_i$ yields
309: an estimator $\tilde{O}^{GR}_i$ which when averaged over $i$ gives $O^{GR}_i$.
310: %
311: Finally, within the fixed-node approximation the correction term in Eq.
312: (\ref{def_de1}) can
313: be viewed as a direct measure of the error of the trial wavefunction with
314: respect to a certain operator.
315: In the remainder of the paper we will only discuss the direct
316: estimator Eq. (\ref{def_de1}).
317:
318: An important question is which operators are admissible and can be sampled
319: using the HF estimator Eq. (\ref{def_de1}) or for that matter the growth
320: estimator Eq. (\ref{def_de0})?
321: Looking at the
322: definition of the DMC algorithm one sees that it is based on splitting
323: the Hamiltonian into a kinetic energy kernel that gives rise to the diffusion
324: part of the algorithm and a potential energy term that has to be diagonal in
325: real
326: space. The diffusion part always being the same it follows that
327: $\Delta \hat{H}=\alpha \hat{O}$ has to be diagonal in real space too. Using
328: for example $O^L= T^L=\frac{\hat{T}\Psi_T}{\Psi_T}$ therefore
329: actually corresponds to sampling the real space many-body operator
330: given by the function $T^L$, rather than the kinetic energy.
331: The result using Eq. (\ref{def_de1}) is $\int
332: \frac{\hat{T}\Psi_T}{\Psi_T} \left[\Psi^{fn}_0\right]^2 dV$ which in general is
333: not the
334: desired expectation value
335: $\langle \hat{T}\rangle_{fn}=\int
336: \frac{\hat{T}\Psi^{fn}_0}{\Psi^{fn}_0} \left[\Psi^{fn}_0\right]^2 dV$.
337: Nevertheless, $\langle \hat{T}\rangle_{fn}$ is accessible within DMC by
338: using $\langle \hat{T}\rangle_{fn} =\langle \hat{H}\rangle_{fn} - \langle
339: \hat{V}\rangle_{fn}$ since the last two quantities can be sampled using
340: standard DMC and HFS, respectively.
341:
342: In the following, we give a few examples to demonstrate the applicability of
343: HFS. We apply the method to sample (i) the density of Helium
344: and (ii) the Ewald energy of a homogeneous electron gas with and without
345: interactions. All data are given in atomic units and we used the
346: CASINO~\cite{casino} package. The target for the number of walkers was between
347: 200 and 400 and the
348: residual weights were allowed to fluctuate between $0.5$ and $2$.
349: While we did not perform extensive studies it seems the
350: algorithm works with and without residual weights.
351: The only modification to the code consisted of adding
352: a variable $X$ to each walker, updating $X$ and applying Eq. (\ref{def_de1}).
353: Other than that we used the code as-is in a standard setup.
354:
355: Figure~\ref{heden} shows the electron density (arbitrary units) of He, as
356: obtained from standard DMC and from our HF method. When the
357: well-converged (i.e. $\Delta O^{E}_i \ll O^{E}_i$) correlated wavefunction
358: supplied with CASINO is used both calculations yield essentially the same
359: result (solid line); when an ``incorrect'' trial wavefunction (which we have
360: chosen to be the same as the ``correct'' one but with the radial term heavily
361: skewed) is used, only our new method (dotted line) recovers the correct
362: density, albeit the noise in the data is larger. Equally, the interaction
363: energy is also recovered (DMC correct
364: wavefunction: $0.947$, incorrect wavefunction $0.791$, incorrect wavefunction
365: HFS: $0.958$). We have also performed DMC calculations of the Hydrogen
366: density, where we systematically deformed the known exact wavefunction.
367: Suffice it to say, as for He we again see confirmation of our algorithm.
368: An interesting point to add here regards
369: the extent to which the wavefunction could be skewed. It turns out - rather
370: plausibly - that if the
371: wavfunction ceases to actually sample certain parts of phase space HFS cannot
372: recover
373: the true form. Nevertheless it seems capable of correcting relatively strong
374: errors
375: in the wavefunction (viz. the rarely sampled asymptoticically decaying part of the wavefunction
376: in Fig. (\ref{heden})),
377: but the details are clearly a topic for further investigation.
378: \begin{figure}
379: \centering
380: \includegraphics[width=0.48\textwidth]{he_den.eps}
381: \caption{
382: The ``exact'' Helium density (solid line) was derived using the well optimized
383: wavefunction provided by the CASINO package: The difference (not shown) between
384: standard
385: DMC sampling and HFS is essentially zero ($\Delta O^{E}_i \ll O^{E}_i$).
386: In addition, results using
387: a trial wavefunction with wrong radial function are
388: presented. Standard DMC yields a smooth but rather poor density. HFS,
389: while noisier (see inset), follows the correct density
390: even in the asymptotic region far from the nucleus where despite
391: little information HFS corrects for the wrong behaviour.}
392: \label {heden}
393: \end{figure}
394:
395: \begin{figure}
396: \centering
397: \includegraphics[width=0.48\textwidth]{fullint2.eps}
398: \caption{
399: Results for the Ewald energy of an unpolarized homogeneous electron gas
400: ($r_s=1$) with 54 electrons. Standard DMC, $\langle\hat{O}\rangle_{DMC}$,
401: yields the relatively smooth curves
402: at the top (see arrows). The noisier data below
403: use HFS (see arros) and the thin streight lines correspond to $\langle\hat{O}\rangle_{cDMC}$
404: at the end of the run. The partial Jastrow factor contains a correlation term
405: without cusp.
406: }
407: \label {cc=1}
408: \end{figure}
409:
410: As in standard DMC sampling, the worse the trial wavefucntion $\Psi_T$,
411: the larger the noise when using HFS.
412: However, when looking at the raw data before averaging over $i$ (not shown)
413: we observed that the noise in the HFS data rises
414: during the progression of the sampling, hence standard error estimation does not work. The source
415: can be traced to sampling over histories $X_{i}$. Limiting their depth results
416: in a constant noise term though also reintroduces a systematic bias.
417: Also, in a recent paper~\cite{wbias} Warren and Hinde
418: observe that using the forward-walking
419: method in DMC necessitates a rapidly growing number of walkers as the
420: dimensionality
421: of the quantum system is increased.
422: These two issues then lead us to the question as to whether HFS works
423: for larger systems.
424: We have therefore looked at an unpolarized homogeneous electron gas at
425: $r_s=1$. We used
426: a finite simulation cell (periodic boundary condition) with 54 electrons.
427: The data we plot shows the Ewald
428: interaction energy with no additional finite size corrections. We show in
429: Fig.~\ref{cc=1} results
430: for a fully interacting system that we have obtained by using trial
431: wavefunctions with either no Jastrow factor, a
432: partially optimized Jatrow factor, or a fully optimized one.
433: We show the mixed DMC estimate $\langle\hat{O}\rangle_{DMC}$,
434: the corrected estimate
435: $\langle \hat{O}\rangle_{cDMC},
436: =2\langle \hat{O}\rangle_{DMC}-\langle
437: \hat{O}\rangle_{VMC}$
438: which contains a second-order error,
439: %correct to first order in the wavefunction
440: and the results for HFS. The MC runs
441: start at 0 with a
442: short equilibration phase and we start sampling at time step $2000$.
443: The corrected estimate using the fully optimized Jastrow factor ought to give
444: the best result.
445: Clearly all three HFS estimates are very close but especially the non-optimized
446: wavefunctions yield quite noisy data. Nevertheless, even in that case the
447: results are a lot better
448: than using the pure DMC output for the best wavefunction. However, they are all
449: also better
450: than the corrected $\langle\hat{O}\rangle_{cDMC}$ results of
451: the partially/non-optimized wavefunction.
452: Regarding the noise one also has to keep
453: in mind the difficulty of the task: The interaction energy is dominated by the
454: region
455: where the electrons get close to each other but that is where the error of the
456: non-optimized
457: wavefunctions is largest. HFS essentially has to build a cusp from scratch.
458:
459:
460: \begin{figure}
461: \centering
462: \includegraphics[width=0.48\textwidth]{noint2.eps}
463: \caption{
464: As in Fig.~\ref{cc=1}, but for an interaction-free Hamiltonian.
465: The noisier data at the top have been sampled using the HFS estimator (see arrows). Below
466: follow the relatively smooth standard DMC values (see arrows)
467: sampling $\langle\hat{O}\rangle_{DMC}$, except for the no-Jastrow
468: case where the two estimators yield the same data as $\Delta O^{E}_i=0$.
469: The thin streight lines correspond to $\langle\hat{O}\rangle_{cDMC}$
470: at the end of the run, except in the case of
471: no Jastrow factor where the thin line gives the essentially exact VMC value.}
472: \label {cc=0}
473: \end{figure}
474:
475: Fig.~\ref{cc=0} repeats the same analysis
476: for a non-interaction Hamiltonian where the Slater determinant (no Jastrow
477: factor)
478: is the exact solution, whence the
479: HFS data and the standard DMC data in that case being identical.
480: This is of course consistent with
481: Eq. (\ref{def_de1}) and proves that given the correct nodes, HFS
482: yields the
483: correct answer. Apart from that Fig.~\ref{cc=0} is essentially a mirror image
484: of Fig.~\ref{cc=1}. In general, we see that unless the wavefunction is well
485: optimized
486: the HFS estimate is considerably better, despite the noise in the data. Such
487: situations might
488: occur when the system is dominated by the bulk while we are interested in
489: sampling data in the surface region. Optimization based on the total energy or
490: variance
491: would result in a sub-optimal wavefunction away from the bulk and hence
492: erroneous standard
493: sampling.
494:
495: In conclusion, by applying the HF theorem directly to the DMC algorithm
496: we have introduced a new method to sample a large class of operators exactly
497: within standard DMC.
498: Our method works for both small and large systems and is easy to add to
499: standard DMC, enabling the sampling of a large class of operators (densities,
500: interaction energies, etc.): Only one extra variable per operator ($X_{i,j}$)
501: need be added to the walkers, involving no more than an extra summation step
502: during sampling;
503: simple algebra (Eqs. (\ref{def_de0}) and (\ref{def_de1})) does the rest.
504: Future work
505: is needed to better understand, estimate, and deal with the noise and its slow increase
506: with simulation time.
507: This is currently under investigation. Similarly, the effect of residual
508: weights needs to be
509: looked at in more detail.
510: A promising line of research already under way
511: is to look at the second derivative. This might allow efficient DMC sampling of
512: the
513: fixed-node
514: density-response function and related quantities, the study of which is
515: currently not feasible due
516: to being numerically too demanding.
517:
518: R.G. would like to thank W.M.C. Foulkes, B. Wood, and N. Hine of Imperial
519: College
520: for helpful discussions.
521: The authors acknowledge partial support by the University of the
522: Basque Country, the Basque Unibertsitate eta Ikerketa Saila, the MCyT, and the
523: EC 6th framework Network of Excellence NANOQUANTA (Grant No.
524: NMP4-CT-2004-500198).
525:
526:
527:
528:
529:
530:
531:
532:
533:
534: \begin{thebibliography}{100}
535:
536: \bibitem{qmc} W. M. C. Foulkes, L. Mitas, R. J. Needs, and G. Rajagopal, Rev.
537: Mod. Phys. {\bf 73}, 33, 2001.
538: \bibitem{check} M. Nekovee and J. M. Pitarke, Comput. Physics Commun. {\bf 137}
539: , 123 (2001).
540: \bibitem{input} D. M. Ceperley and B. J. Alder, Phys. Rev. Lett.
541: {\bf 45}, 566 (1980).
542: See also R. M. Dreizler and E. K. U. Gross, {\it Density-Functional
543: Theory. An Approach to the Quantum Many-Body Problem}, Springer, 1990.
544: %\bibitem{cep_err} D. M. Ceperley and M. H. Kalos, in {\it Monte Carlo Methods
545: %in
546: %Statistical Physics}, edited by K. Binder (Springer, Berlin, 1979).
547: \bibitem{cep_err} P. A. Whitlock, D. M. Ceperley, G. V. Chester, and M. H.
548: Kalos, Phys.
549: Rev. B {\bf 19}, 5598 (1978).
550: \bibitem{fwalking} K. S. Liu, M. K. Kalos, and G. V. Chester,
551: Phys. Rev. A {\bf 10}, 303 (1974).
552: \bibitem{rmc} S. Baroni and S. Moroni, Phys. Rev. Lett. {\bf 82}, 4745 (1999).
553: \bibitem{no_tagging} J. Casulleras and J. Boronat, Phys. Rev. B {\bf 52}, 3654
554: (1995).
555: \bibitem{hf} R. P. Feynmann, Phys. Rev. {\bf 56}, 340 (1939).
556: \bibitem{ob} G. Ortiz and P. Ballone, Phys. Rev. B {\bf 50}, 1391 (1994).
557: \bibitem{note_dmc_alg} In practice one uses importance sampling
558: (R. C. Grimm and R. G. Storer, J. Comput. Phys. {\bf 7}, 134 (1971)), which
559: changes the kinetic diffusion kernel into drift-diffusion
560: and replaces the potential energy by the much smoother local energy
561: %$E_L=\frac{\hat{H}\Psi_T}{\Psi_T}$. The process then samples
562: %$\Psi_T\Psi^{fn}_0$
563: $E_L=[\hat{H}\Psi_T]/\Psi_T$. The process then samples $\Psi_T\Psi^{fn}_0$
564: rather than just $\Psi^{fn}_0$.
565: \bibitem{Etilde} Inverting $E^0_i=\frac{1}{i}\sum_{l=1}^{i} \tilde{E}^0_l$ then gives
566: $\tilde{E}^0_i=iE^0_i-(i-1)E^0_{i-1}$.
567: \bibitem{HFproblem} When using the HF theorem in conjunction with numerical
568: derivatives, problems appear if the underlying Hilbert space changes
569: with $\alpha$. This is not the case here.
570: \bibitem{casino} R. J. Needs, M. Towler, N. Drummond, and P. Kent, {\it CASINO
571: version 1.7 User manual} (University of Cambridge, 2004).
572: \bibitem{wbias} G. L. Warren and R. J. Hinde, Phys. Rev. E {\bf 73}, 056706
573: (2006).
574: %\bibitem{dft1} W. Kohn and L. J. Sham, Phys. Rev. {\bf 140}, A1133 (1965).
575:
576:
577:
578:
579: \end{thebibliography}
580:
581: \end{document}
582:
583: