cond-mat0702349/VdW.tex
1: \documentclass[aps,twocolumn,superscriptaddress,amsmath]{revtex4}
2: 
3: \usepackage[dvips]{graphicx}
4: \usepackage{dcolumn}
5: \usepackage{amsmath}
6: \usepackage{bm}
7: 
8: \addtolength{\voffset}{0cm}
9: \addtolength{\textheight}{0cm}
10: \addtolength{\footskip}{0cm}
11: 
12: \newcommand{\be}{\begin{equation}}
13: \newcommand{\ee}{\end{equation}}
14: \newcommand{\bea}{\begin{eqnarray}}
15: \newcommand{\eea}{\end{eqnarray}}
16: \newcommand{\bean}{\begin{eqnarray*}}
17: \newcommand{\eean}{\end{eqnarray*}}
18: \newcommand{\expect}[1]{\langle #1 \rangle}
19: \newcommand{\abs}[1]{\left| #1 \right|}
20: \newcommand{\grad}{\vec{\nabla}}
21: \newcommand{\ra}{\rightarrow}
22: \newcommand{\Det}[1]{\left| #1 \right|}
23: \newcommand{\half}{\frac{1}{2}}
24: \newcommand{\vc}[1]{{\bf #1}}
25: 
26: 
27: \begin{document}
28: 
29: \widetext
30: 
31: 
32: \title{Weak binding  between two aromatic rings: feeling the van der Waals
33:   attraction by quantum Monte Carlo methods} 
34: 
35: \author{Sandro Sorella}
36: \email[]{sorella@sissa.it}
37: \affiliation{ SISSA, International School for Advanced Studies, 
38:   34014, Trieste, Italy}
39: \affiliation{ DEMOCRITOS, National Simulation Center,
40:   34014, Trieste, Italy}
41: \author{Michele Casula}
42: \email[]{casula@uiuc.edu}
43: \affiliation{ Department of Physics, University of Illinois at Urbana-Champaign,
44: 1110 W. Green St, Urbana, IL 61801, USA} 
45: \author{Dario Rocca}
46: \email[]{roccad@sissa.it}
47: \affiliation{ SISSA, International School for Advanced Studies, 
48:   34014, Trieste, Italy}
49: \affiliation{ DEMOCRITOS, National Simulation Center,
50:   34014, Trieste, Italy}
51: 
52: 
53: \date{\today}
54: 
55: 
56: \begin{abstract}
57: We report a systematic study of the weak chemical bond between two benzene 
58: molecules. We first show that it is possible to obtain a very good description 
59: of the $C_2$ dimer and the benzene molecule, by using 
60: pseudopotentials for the chemically inert $1s$ electrons, and 
61: a resonating valence bond wave function as a variational ansatz,
62: expanded on a relatively small Gaussian basis set. 
63: We employ an improved version of the stochastic
64: reconfiguration technique to optimize the many-body wave function,
65: which is the starting point for highly accurate simulations
66: based on the lattice regularized diffusion Monte Carlo (LRDMC) method.
67: This projection technique provides
68: a rigorous variational upper bound for the total energy, 
69: even in the presence of pseudopotentials, 
70: and allows to improve systematically  
71: the accuracy of the trial wave function, 
72: which already yields a large fraction of the 
73: dynamical and non-dynamical electron correlation.
74: We show that the energy dispersion of two benzene molecules 
75: in the parallel displaced geometry is significantly 
76: deeper than the face-to-face configuration. 
77: However, contrary to previous studies based on post Hartree-Fock methods, 
78: the binding energy remains weak ($\simeq 2 kcal/mol$) also in this geometry, 
79: and its value is in agreement with the most accurate and recent  
80: experimental findings.\cite{bestexp}  
81: \end{abstract}
82: 
83: \maketitle
84: 
85: \section{Introduction}
86: \label{introduction}
87: The intermolecular interaction between  benzene rings has been 
88: a subject of intense theoretical and experimental studies in the last two
89: decades\cite{old,scoles,cc,cc2006,lind,guidoni}. 
90: Indeed the intermolecular bonds based on the corresponding 
91: $\pi$-$\pi$ interactions play an important
92: role in many interesting compounds. For instance, they stabilize 
93: the three-dimensional structures of biological systems such as proteins, DNA,
94: and RNA. Moreover, many drugs with a specific chemical target 
95: utilize these $\pi$-$\pi$ interactions and the long range 
96: forces for their stability. 
97: 
98: In order to understand the mechanism behind those attractions, 
99: we have considered here the benzene dimer as a 
100: prototype compound, because  both the $\pi$-$\pi$ interactions and  
101: the van der Waals (vdW) long range attraction
102: are already present and can be studied in a systematic way. 
103: Despite its simplicity, so far 
104: there is no general consensus about its equilibrium properties 
105: from both the theoretical and the experimental side.
106: Indeed, it is difficult to determine experimentally the 
107: complete energy dispersion, and only the total binding energy $D_0$ 
108: is known,\cite{grover,bestexp} with a relatively large experimental error
109: due to the weakness of the interactions. 
110: On the other hand, this compound represents 
111: a numerical challenge for theoretical methods,
112: because the local density approximation (LDA) and other standard 
113: treatments based on the density functional theory (DFT) 
114: are not supposed to work well, when 
115: dispersive forces  are the key ingredient in the chemical bond.
116: Despite some progress  
117: has been made recently,\cite{kohn,gross,scoles,lind,ursula,degironc}
118: a general and practical 
119: solution of this problem is still lacking in the DFT formalism.
120: Another family of methods, the accurate post Hartee-Fock methods
121: such as CCSD(T), have been extended only very recently to a 
122: larger basis set\cite{cc2006}, since their 
123: prohibitive computational cost has limited their application
124: to systems with few electrons and small basis set, and
125: the benzene dimer is already at the cutting edge of those approaches.
126: As a matter of fact, although the complete basis set (CBS) limit 
127: can now be estimated more precisely in the CCSD(T) framework, 
128: the most accurately determined  binding energy ($\simeq 2.8 kcal/mol$)
129: of the benzene dimer substantially disagrees from the most precise and
130: recent measurement\cite{bestexp}, as also 
131: honestly pointed out in Ref.~\onlinecite{cc}. Indeed, the CCSD(T) method seems to
132: overbind the dimer in the CBS limit. 
133: 
134: Quantum Monte Carlo (QMC) methods are 
135: a promising alternative to the aforementioned techniques.
136: They are able to deal with a highly correlated variational wave function, which
137: can explicitly contain all the key ingredients of the physical system.
138: Their computational cost scales favorably with the number of particles $N$,
139: usually as $N^3 - N^4$, depending on the method, which makes the QMC framework
140: generally faster than the most accurate post Hartree-Fock (HF) 
141: schemes for large enough $N$.
142: Moreover, recent important developments in the QMC field
143: allow now to optimize the
144: variational ansatz with much more parameters and higher accuracy.
145: In turn this   can be substantially improved 
146: by  projection QMC methods 
147: such as the diffusion Monte Carlo (DMC)\cite{dmc}
148: and its lattice regularized version (LRDMC).\cite{lrdmc}
149: These techniques  are able in principle to yield
150: the ground state energy of the system, since they are
151: based on a direct stochastic solution of the Schr\"odinger equation. 
152: However, the well known sign problem affects this kind of calculations, 
153: and the fixed node (FN) approximation is required to make those simulations feasible.
154: Within this approximation, it is possible to obtain 
155: the lowest variational state $\psi_{FN}(x)$ 
156: of the Hamiltonian with the constraint 
157: to have the same signs of a given variational wave function $\psi_G(x)$. 
158: The above condition is applied conveniently
159: in the space representation $\{ x\}$ of configurations with given electron positions 
160: and spins. It  turns out that good variational energies can be 
161: typically obtained with a projection QMC method even starting from a very poor
162: variational wave function,  
163: the method being clearly  exact in the case when $\psi_0(x) \psi_G(x) \ge 0,\forall x$, 
164: where $\psi_0(x)$ is the exact ground state of $H$.
165: 
166: Until few years ago, the FN approximation was applied\cite{gfmc} 
167: to simple variational wave functions obtained with 
168: basic methods, such as HF or 
169: LDA, because for large electron systems it was basically impossible to 
170: optimize several variational parameters within a statistical framework. 
171: On the other hand, on small dimer systems\cite{filippimol}, and even in the single 
172: benzene molecule\cite{casulamol}, it was clearly shown that
173: a highly correlated wave function $\psi_G(x)$ had to be carefully 
174: optimized  before applying the DMC method with the FN approximation.
175: Other examples of the importance of the 
176: optimization procedure have been recently discovered in significant 
177: chemical systems\cite{needs}, showing at the same time
178: that QMC is developing quite rapidly and may represent a promising tool 
179: for future calculations.
180:  
181: In the present work we report a systematic study of the benzene 
182: dimer, using the latest developments in the QMC framework: an improved 
183: optimization algorithm based on the stochastic reconfiguration (SR), and the 
184: LRDMC method,\cite{lrdmc} which allows to include non local potentials
185: (pseudopotentials) in the Hamiltonian with a rigorous variational approach. 
186: In principle, by means of the LRDMC method 
187: it is possible to estimate $E_{FN} = { \langle \psi_{FN} |H| \psi_{FN} \rangle \over 
188: \langle \psi_{FN} | \psi_{FN} \rangle }$  even in presence 
189: of pseudopotentials. Furthermore, a very stable and  efficient upper bound 
190: of $E_{FN}$ is obtained by the mixed estimator\cite{ceperley}:
191: \begin{equation}
192:  E_{LRDMC}= { \langle \psi_G | H |\psi_{FN} \rangle 
193: \over \langle \psi_G | \psi_{FN} \rangle }. 
194: \end{equation}
195: $E_{LRDMC}$ substantially improves the variational energy $E_G$ 
196: of the trial wave function $\psi_G$, and is  always very close to $E_{FN}$.
197: However, in the case of pseudopotentials, it has to be mentioned that  
198: $E_{FN}$ obtained with the effective Hamiltonian included in the LRDMC
199: is not necessarily the lowest variational energy compatible 
200: with the signs of $\psi_G(x)$.
201: 
202: The paper is organized as follows: in Sec.~\ref{wave_function} 
203: we describe the variational wave function and its corresponding basis set.
204: In Sec.~\ref{methods}, we introduce the QMC methods used.
205: We present some important improvements in the SR technique
206: to optimize the energy of a correlated wave functions 
207: containing several parameters. Moreover, 
208: we show how it is possible to reduce significantly
209: the lattice discretization error in the LRDMC method in order to 
210: improve its efficiency.
211: Finally, in Sec.~\ref{results} 
212: we discuss the results on the simple but strongly correlated 
213: carbon dimer, and the more demanding application to compute
214: the binding energy of the face-to-face and parallel
215: displaced configurations in the benzene dimer.
216: 
217: 
218: \section{Wave Function}
219: \label{wave_function}
220: We use the Jastrow correlated antisymmetrized geminal power (JAGP)
221: introduced in Refs.~\onlinecite{casulaagp} and \onlinecite{casulamol}, where
222: the determinantal 
223: part (AGP) is nothing but the particle number conserving version of the
224: Bardeen-Cooper-Schrieffer (BCS) wave function. The JAGP ansatz is the practical 
225: representation of the resonating valence bond idea, introduced by
226: L. Pauling for chemical systems~\cite{pauling}, and developed also 
227: by P. W. Anderson for strongly correlated spin systems~\cite{pwanderson}. 
228: Our variational wave function is defined by the product of two terms, namely
229: a Jastrow $J$ and an antisymmetric part ($\Psi=J \Psi_{AGP}$).
230: The Jastrow term is further split into one-body, two-body 
231: and a three-body factors ($J = J_1 J_2 J_3$) described 
232: in the following.
233: All the atomic and molecular cusp conditions are fulfilled through
234: the one-body $J_1$ and the two-body $J_2$ Jastrow factors.
235: The former treats the electron-ion cusp, while the latter cures 
236: the opposite-spin electron-electron cusp. They are both defined 
237: by means of a simple function $u(r)$ containing only 
238: one variational parameter $F$:
239: \be \label{simfahy}
240: u(r) = \frac{F}{2}\left ( 1 - e^{-{r}/F}\right),
241: \ee
242: where $u^\prime (r)=1/2$ in order to satisfy the
243: cusp condition for opposite spin electrons\cite{mitas}.
244: Then the two-body Jastrow factor reads:
245: \be
246: J_2(\textbf{r}_1,...,\textbf{r}_N) =
247: \exp{\left (\sum_{i<j} u(r_{ij}) \right )},
248: \label{two-body}
249: \ee
250: where  $r_{ij}= |\textbf{r}_i -\textbf{r}_j|$  
251: is the distance between two electrons.
252: On the other hand the electron-ion cusp condition can be satisfied 
253: by the one-body term:
254: \begin{equation}
255: J_1(\textbf{r}_1,...,\textbf{r}_N) =
256: \exp{\left (-\sum_{i,j} (2 Z_j)^{3/4} u( (2 Z_j)^{1/4}  |\textbf{r}_i -\textbf{R}_j| ) \right )},
257: \end{equation}
258: where $\textbf{R}_j$ are the atomic positions with corresponding atomic number 
259: $Z_j$.
260: The reason to take this form for the one-body Jastrow factor was inspired 
261: by the work of Holzmann \emph{et al.}\cite{pier} on dense Hydrogen: 
262: in the function $u$, the length scaling factor $(2 Z)^{1/4}$  
263: is used to satisfy the large distance RPA behavior,
264: whereas the
265: multiplicative factor $ (2 Z)^{3/4}$ is set by the electron-ion cusp 
266: condition:
267: \begin{equation}
268: < {  d J_1 \over d |\textbf{r}_i -\textbf{R}_j|} >= -Z_j J_1  ~~~{\rm for}~ |\textbf{r}_i-\textbf{R}_j|\to 0 
269: \end{equation}  
270: where $< >$ means the angular average. The above relation easily follows, 
271: since  $ u'(r)=1/2$. 
272:   
273: Once all the cusp conditions are satisfied, we can parametrize the 
274: remaining function $J_3$ and the AGP part of our resonating valence bond 
275: wave function $J \Psi_{AGP}$, and reach the CBS limit for both the full Jastrow
276: factor $J$ and the determinantal part, {\em with a Gaussian atomic basis set that 
277: does not contain any cusp}. This represents a clear 
278: advantage compared with the previous parametrization,\cite{casulamol} 
279: where it was not even possible to satisfy exactly all the electron-ions cusp 
280: conditions with a finite basis set. 
281: Furthermore, this parametrization is also particularly 
282: useful for interfacing a QMC code with standard packages for quantum 
283: chemistry calculations, which generally use a Gaussian basis set, and 
284: therefore are not supposed to satisfy {\em any} cusp conditions 
285: with a finite number of basis elements. 
286: Obviously this  approach applies in the same way  
287: also for all-electron calculations.
288: 
289: The AGP geminal function\cite{casulamol} is expanded over an atomic basis set:
290: \be
291: \Phi_{AGP}(\textbf{r}^\uparrow,\textbf{r}^\downarrow)
292: =\sum_{l,m,a,b}{\lambda^{l,m}_{a,b}\phi_{a,l}
293: (\textbf{r}^\uparrow)\phi_{b,m}(\textbf{r}^\downarrow)} ,
294: \label{expgem}
295: \ee
296: where the indices $l,m$ span different orbitals centered on corresponding 
297: atoms $a,b$.
298: In turn, the atomic orbitals $\phi_{a,l}$ are expanded with a set 
299: of primitive single zeta Gaussian functions.
300: All the coefficients and the exponents of the gaussians  are 
301: always consistently optimized. Notice that the largest number 
302: of variational parameters are contained in 
303:  the symmetric $\lambda$ matrix, the number of entries being proportional 
304: to the square of the atomic basis set size. For this reason, in order to
305: reduce the total number of parameters, it is useful to lower the dimension
306: of the atomic basis set, by introducing contracted orbitals.
307: 
308: The three-body $J_3$ Jastrow function takes care of what is 
309: missing in the one and the two-body Jastrow factors, namely the explicit 
310: dependence of the electron correlation on the ionic positions. Therefore, each
311: term in $J_3$ includes two electrons and one ion interacting each other
312: (this is the reason of the name ``three-body''):
313: \bea \label{3body}
314: J_3(\textbf{r}_1,...,\textbf{r}_N)
315: &=& \exp \left( \sum_{i<j} \Phi_J(\textbf{r}_i,\textbf{r}_j) \right)
316: \nonumber \\
317: \Phi_J(\textbf{r}_i,\textbf{r}_j) &=& \sum_{l,m,a,b} g_{l,m}^{a,b}\psi_{a,l}
318: (\textbf{r}_i)\psi_{b,m} (\textbf{r}_j),
319: \eea
320: where the indices $l$ and $m$ in the Jastrow geminal $\Phi_J$
321: indicate different orbitals  located around atoms $a$ and $b$,
322: respectively.
323: Again, since all cusp conditions are already satisfied by $J_1$
324: and $J_2$, in the pairing function $\Phi_J(\textbf{r}_i,\textbf{r}_j)$ 
325: we use single zeta gaussian orbitals, $\psi_{a,l} (r) = 
326: e^{ - z r^2} r^k \times ({\rm simple~polynomial~in~r_x,r_y,r_z})$, 
327: where $k\ge 0$  is an integer and  $z$ is the gaussian exponent. The
328: polynomials are related to the real space representation of the spherical
329: harmonics. For instance, to expand $J_3$ up to the angular momentum 
330: $l=1$, 
331: we have used two types of orbitals, with $k=0$ and $k=1$ respectively.
332: On simple dimer compounds we have tested that 
333: the inclusion of the latter Jastrow orbital is 
334: particularly useful for an accurate description of the weak vdW
335: interactions. Indeed, from a quantum mechanical
336: point of view this type of interactions is due to the correlated
337: transition (polarization) of a couple of electrons from s-wave states  
338: localized around two atoms to corresponding p-wave states.
339: Whenever these two atoms are at large distance, 
340: we can expand $J_3$ for small values of $g_{l,m}^{a,b}$, and
341: apply this term to a geminal product of two s-wave orbitals.
342: In this way, it is clearly possible to describe vdW interactions,
343: provided the gaussian basis set used for $J_3$ 
344: contains also suitable p-wave components. 
345: Moreover, we added in the $J_3$ pairing function also one body terms, 
346: which are the product of single zeta gaussian orbitals times a constant 
347: (i.e. like $g_{l,c}^{a,b}\psi_{a,l}$, 
348: where $c$ refers to the constant ``orbital'' $\psi_{b,c}=1$). 
349: Thus, our wave function can include a complete basis set expansion 
350: also for the one body Jastrow factor.
351: 
352: \section{Improved numerical methods}
353: \label{methods}
354: In this section we introduce some developments 
355: of two recently introduced QMC techniques, the SR\cite{sr}
356: and the LRDMC\cite{lrdmc} methods, 
357: reported in the first and second subsection, respectively.
358: The improvements described here are of fundamental importance 
359: in order to apply successfully those methods 
360: to realistic electronic systems with about
361: $100$ valence electrons. 
362: 
363: \subsection{Minimization method}
364: As described in the previous Section, the JAGP variational wave function can
365: contain a large number $p$ of non linear parameters $\{ \alpha_k \}$, 
366: which are usually difficult to optimize for three main reasons, 
367: listed below in order of difficulty:
368: \begin{itemize} 
369: \item[(i)] 
370: The occurrence of several local minima 
371: in the energy landscape, leading to the very complex numerical  
372: problem of finding the global minimum energy. 
373: \item[(ii)]
374: The strong dependence between  
375: several variational parameters. Sometimes,
376: the variation of some non linear parameters in the wave function 
377: can be almost exactly compensated 
378: by a corresponding change of other parameters. 
379: This may lead to instabilities and/or slow convergence 
380: to the minimum energy.
381: \item[(iii)] 
382: The slow convergence to the minimum energy can also be due to 
383: simple-minded and/or inefficient iterative methods.
384: \end{itemize}
385: 
386: In the QMC framework, the energy minimization 
387: is further complicated by the statistical uncertainty, which affects
388: all quantities computed, including the 
389: optimization target, namely the total energy. 
390: Despite these difficulties, a lot of progress has been made recently 
391: in the energy optimization of highly correlated wave function, 
392: especially for the alleviation of problems (ii) and (iii)
393: \cite{umrigarhess,srh,cyrc2}. As far as the problem (i) is concerned, the solution
394: remains only empirical and relies on the ability to find a good starting
395: point of the minimization procedure.   
396: 
397: In this work we have used a simple improvement of the SR method
398: introduced in Ref.~\onlinecite{sr} for lattice systems, and
399: applied later to small atoms\cite{casulaagp} 
400: and molecules\cite{casulamol}. The SR method
401: has shown to be an efficient and robust minimization scheme, although in
402: cases with many variational parameters the convergence to the minimum was
403: much slower and inefficient for a subset of parameters.
404: From this point of view, by using soft-pseudopotentials to remove the core
405: electrons, we have experienced a  speed up in 
406: the wave function optimization, because  
407: the too   short wave-length components, responsible of the 
408: slowing down, are no longer present.
409:  Moreover,
410: the recent methods based on the Hessian matrix\cite{umrigarhess,srh,cyrc2} 
411: provide  a further  improvement  in efficiency, since they  allow  
412: to converge to the minimum energy with fewer iterations.
413: 
414: Within the SR minimization, the variational parameters are changed at 
415: each iteration:
416: $$\alpha_k^\prime =\alpha_k + \delta \alpha_k$$
417: according to the simple rule:
418: \begin{equation} \label{itersr}
419: \delta \alpha_k = \Delta t \sum\limits_{k^\prime}  s^{-1}_{k,k^\prime} f_{k^\prime} 
420: \end{equation}
421: where $\Delta t>0$ is small enough to guarantee convergence to the minimum,
422: whereas  $f_k = - { \partial E   \over \partial \alpha_k }$ are the generalized 
423: forces. The SR matrix  $s$ can be any positive definite matrix 
424: (e.g. if $s$ is the identity matrix one recovers the standard steepest 
425: descent method), but to accelerate the convergence to the minimum 
426: and avoid the problem (ii) 
427: it is much more convenient, as explained in Ref.\onlinecite{casulamol}, to 
428: use the positive definite matrix defined by:
429: \begin{equation}
430: \label{overlap_matrix}
431: s_{k,k^\prime}= 
432:  \langle  O_k O_{k^\prime}  \rangle - \langle O_k \rangle
433:  \langle O_{k^\prime} \rangle,  
434: \end{equation}
435: where the brackets in $ \langle C \rangle $ denote the quantum expectation 
436: value of a generic  operator $C $ over the variational 
437: wave function $\psi_G$ with parameters $\{ \alpha_k  \}$. 
438: Moreover, $O_k$'s are operators diagonal in the Hilbert space spanned 
439: by configurations $\{ x \}$, where electrons 
440: have definite positions and spins:
441: \begin{equation}
442: O_k(x) = \partial_{\alpha_k} \ln | \langle x |\psi_G \rangle | . 
443: \end{equation}   
444: The symmetric matrix $s$ in Eq.~\ref{overlap_matrix} 
445: has certainly non negative eigenvalues  because it is 
446: just an overlap matrix. In the following, we will assume 
447: that the matrix $s$ is strictly positive definite, 
448: as this condition can be easily 
449: fulfilled by removing from the optimization   
450: those variational parameters which imply
451: strictly vanishing  eigenvalues for $s$.  
452: This possibility never occurs in practice, unless the wave function 
453: has not been efficiently parametrized and contains 
454: redundant variational parameters.
455: 
456: At each iteration the various quantities - the matrix elements $s_{k,k^\prime}$,
457: and the generalized forces $f_k$ - are evaluated stochastically 
458: over a set of $M$ configurations 
459: $x_i, i=1,\cdots M$, generated by the standard variational Monte Carlo method
460: according to the statistical weight 
461: $\pi_x=\frac{ |\langle x | \psi_G\rangle |^2 }{ \langle \psi_G | \psi_G \rangle }$. 
462: In order to avoid ergodicity problems, apparent
463: when the atoms are far apart, we have also 
464: included large hopping moves\cite{hopping} 
465: to  the standard Metropolis transition probability. 
466: In the limit $M\to \infty$, the statistical 
467: uncertainty vanishes like $1/\sqrt{M}$, and the above minimization 
468: strategy certainly converges to some local minimum for small enough $\Delta t$ 
469: and for large enough number of iterations.
470: 
471: In the QMC framework it is obviously important to work with a small number 
472: $M$ of configurations, because this number is proportional to the computer time 
473: required for the optimization. 
474: However, though the SR method is rather efficient, the statistical noise 
475: can deteriorate the stability of the method, especially because the 
476: matrix $s$ can be ill-conditioned, namely with very small eigenvalues, 
477: and its inverse can dramatically amplify the noise  present in the forces.
478: Indeed the SR matrix, even when computed with a finite number $M$ of 
479: samples, remains positive definite, but the lowest eigenvalues  
480: and corresponding eigenvectors can be very sensitive to the statistical noise.
481: In a previous work\cite{casulamol}, we described a simple strategy to work with a well conditioned matrix $s$, by disregarding some variational parameters 
482: at each iteration in the optimization procedure. 
483: This method has a problem, because sometimes it is necessary to disregard
484: a large fraction of the total number $p$ of parameters. 
485: Moreover, we have experienced 
486: that removing variational parameters from the optimization may be very
487: dangerous, as the probability to remain stuck in a local minimum 
488: or even in a saddle point grows dramatically, especially for large $p$. 
489: This occurs even when a relatively small number of parameters is not
490: considered in the optimization. 
491: 
492: In order to avoid the above problems, and improve the stability of the method, 
493: we have modified and simplified the conditioning of the matrix $s$.
494: At each step, we evaluate the SR matrix with a small bin length 
495: ($M \sim 1000-10000$), and we regularize it by 
496: the simple modification of its diagonal elements:
497: \begin{equation} \label{regularization}
498: s_{k,k} = s_{k,k} (1 + \epsilon),
499: \end{equation}
500: where $\epsilon$ can be considered a small Monte Carlo cut-off, 
501: which can be safely chosen smaller  than the average statistical  
502: accuracy of the diagonal matrix elements $s_{k,k}$. 
503: In this way the modified matrix appears well 
504: conditioned and without too small eigenvalues. 
505: Consequently, the improvement in stability can be substantial as  
506: shown in Fig.(\ref{graphsr}) for a simple lattice model test case\cite{srh}. 
507: At the same time, there is no need to disregard variational parameters as in
508: the previous scheme.  
509: It is important to emphasize that also the modified $s$ matrix is positive
510: definite, because the sum of two positive definite matrices,  
511: $s_{i,j}$ and $\epsilon \delta_{i,j} s_{i,i} $\cite{note3},
512: remains a positive definite matrix.  
513: As we have already mentioned, this is the only 
514: requirement for the iteration in Eq.~\ref{itersr} to converge to a minimum
515: ($f_k=0~ \forall k$). Therefore, since 
516: all force components $f_k$ are not {\em biased} by the $s$-matrix modification, 
517: by means of our approach  the exact minimum can be reached  
518: for arbitrary  values  of $\epsilon$ and $M\to \infty$.
519: 
520: Obviously other similar regularizations are possible and were also 
521: adopted elsewhere\cite{umrigarhess,rappe,night}. 
522: For instance, it is possible to add a simple rescaled identity
523: to $s$ ($ s_{k,k}  \to s_{k,k}  + \epsilon $), 
524: and obtain a well conditioned modified matrix with 
525: all eigenvalues greater than $\epsilon$.
526: However, we have preferred to use the less obvious modification 
527: in Eq.~\ref{regularization}, because in this way 
528: the relative change is the same for all diagonal elements, which 
529: are not deteriorated too much in the case they are very small.
530: This is particularly useful for the optimization of 
531: the present JAGP  wave function, as it contains 
532: some parameters  (e.g. the $\lambda_{i,j}$ in the determinant) 
533: ranging  in a very tiny interval (e.g. within   $~10^{-3}-10^{-6}$) and 
534: some others (e.g. the exponents $z_i$ in the gaussians) 
535: spanning a much wider range ( e.g. within $~1-100$).  
536: Without the appropriate scaling provided by the 
537: diagonal elements of the SR matrix in Eq.~\ref{itersr}, 
538: an exceedingly small $\Delta t$ should be used for a stable convergence, 
539: which would imply, on the other hand, a prohibitively slow convergence. 
540: 
541: \begin{figure}[!ht]
542: \includegraphics[width=\columnwidth]{graphsr.eps}
543: %\vspace{-1.2cm}
544: \caption{Optimization of the variational 
545: wave function in the simple 1D Heisenberg model $H=J \sum_i \vec S_i \cdot
546: \vec S_{i+1}$
547:  with  the 
548: standard SR  ($\epsilon=0$, open
549: circles), and with the present regularization ($\epsilon=0.001$, open
550: triangles). Further details on the wave function 
551: can be found in  Ref.~\onlinecite{srh}.
552: In the figure, the evolution
553: of the nearest neighbor spin-spin (n. n. Sz) Jastrow parameter is plotted.
554: For each iteration, the forces and the SR matrix in Eq.~\ref{itersr} were
555: evaluated over $M=2500$ samples, whereas $\Delta t J=0.125$.  
556: From this plot it is clear that the SR method with $\epsilon=0.001$ is several 
557: order of magnitudes more efficient than the standard SR for determining the 
558: variational parameter with given statistical accuracy.
559: The inset shows the first few iterations. 
560: \label{graphsr}}
561: \end{figure}
562:  
563: The present optimization scheme is in practice very efficient. 
564: For a given bin length $M$, the SR method  becomes optimal for
565: $\epsilon$ equal to a {\em finite} value, which may be even much smaller 
566: than  the statistical accuracy of the matrix elements $s_{k,k^\prime}$.  
567: In the optimal limit, the statistical fluctuations of  the variational 
568: parameters are substantially suppressed without slowing down too much the 
569: convergence to the energy minimum (see e.g. Fig.~\ref{graphsr}). 
570: Probably, the value $\epsilon=0$ is 
571: optimal only for a noiseless infinite precision arithmetic. 
572: 
573: \subsection{LRDMC method with a better $a\to 0$ limit}
574: After the energy minimization of a given variational wave function 
575: $\psi_G$, a substantial improvement in the correlation energy is obtained by using 
576: the DMC method, with the so called FN approximation.
577: This method allows in principle to determine statistically the lowest energy
578: wave function $\psi_{FN}(x)$ with the same nodal surface as $\psi_G(x)$, 
579: namely $ \psi_{FN}( x) \psi_G(x) \ge 0$ (FN constraint). In other words
580: the  corresponding energy $E_{FN} = { \langle \psi_{FN}| H| \psi_{FN} \rangle 
581:   \over \langle \psi_{FN} | \psi_{FN} \rangle }  $ is the minimum possible
582: within the FN constraint. Only recently
583: this idea has been generalized\cite{lrdmc,casuladmc} to include non local
584: potentials in a rigorous variational formulation.
585: The LRDMC method is based on 
586: a lattice discretization of the exact Hamiltonian included 
587: in the standard DMC framework.   
588: In short, the exact Hamiltonian $H$ is replaced by a lattice regularized one 
589: $H^{a}$, such that $H^a \to H$ for 
590: $a \to 0$, where $a$ is some lattice space which allows to 
591: discretize the kinetic energy using finite difference schemes, 
592: e.g. $\partial_y^2 \psi(y)= { \psi (y+a) + \psi(y-a) - 2 \psi (y) \over 
593: a^2}$, where $\psi(y)$ is an arbitrary function.
594: Indeed, our approximate laplacian is:
595: \begin{eqnarray} \label{laplace}
596: &\Delta^{a,p} &  f(x,y,z)= \nonumber \\
597: & \eta/a^2 & \left[ p(x + a/2,y,z) (f(x+a,y,z) - f(x,y,z)) \right. 
598: \nonumber \\
599: & +   &      \left. p(x-a/2,y,z)  (f(x-a,y,z)-f(x,y,z))   \right] 
600: \nonumber \\ 
601: & + & x\leftrightarrow y \leftrightarrow z.
602: \end{eqnarray}
603: where $(x,y,z)\equiv \mathbf{r}$ are Cartesian coordinates, 
604: and the function $p$ is given by  
605: \begin{equation}
606: p(\mathbf{r}) = 1/(1+ Z^2 |\mathbf{r} - \mathbf{R}|^2 / 4)\,, \label{p}
607: \end{equation}
608: where $\mathbf{R}$ is  the   atom position closest  
609: to the electron in $\mathbf{r}$, and $Z$ is the largest atomic 
610: number considered in the system.
611: In particular for the carbon atom, we used $Z=4$ throughout this study,     
612: as  the 1s electrons  are removed by the pseudopotential.
613: The constant $\eta$ behaves as $1 +O(a^2)$ and is introduced to
614: further reduce the error coming from the discretization of the kinetic term.
615: 
616: As pointed out in Ref.\onlinecite{lrdmc}, an appropriate use of two lattice
617: spaces $a$ and $a^\prime$  with fixed irrational ratio $a^\prime/a=
618: \sqrt{Z^2/4+1}$ allows to define $H^{a}$ in the continuous space even for finite $a$.
619: In the same work, the constant $\eta$ was determined by 
620: requiring that the discretized
621: kinetic energy is equal to the continuous one calculated on the state
622: $\psi_{G}$. Here we have found that this requirement is not particularly
623: useful for obtaining a very small lattice discretization error in the 
624: total energy. Indeed as shown in Fig.~\ref{evsa}, it is much more convenient  
625: to define   $\eta = 1 + K a^2$, with $K$  determined empirically in order 
626: to reduce the systematic finite $a$ error. 
627: The optimal value of $K=3.2 a.u$ ($10.8 a.u.$),  with $a'/a=\sqrt{5} 
628: (\sqrt{10})$, has been determined for the 
629: carbon (oxygen) pseudoatom, 
630: and can be then used also for larger systems containing 
631: the same atom, as we have done in the forthcoming studies.
632:  
633: In principle, the LRDMC method allows to calculate the expectation value of the 
634: Hamiltonian $H$ on the more accurate FN wave function $\psi_{FN}$. 
635: However, this approach is rather time consuming because several runs have 
636: to be performed and some extrapolation is  required, which increases  
637: the statistical error by at least a factor of $3$. Since the LRDMC method -
638: like any other projection method- is quite expensive, in the following 
639: we have preferred to evaluate
640: the simplest upper bound, indicated here by $E_{LRDMC}$, valid for the directly 
641: computable mixed average  $E_{LRDMC} = 
642: { \langle \psi_G | H | \psi_{FN} \rangle \over \langle \psi_G |
643:  \psi_{FN} \rangle  }  > E_{FN}$, 
644:  an inequality that follows by applying the variational theorem on
645:  the Hamiltonian $H^{a}$ .\cite{ceperley,lrdmc,notelrdmc}  
646: 
647: 
648: 
649: \begin{figure}[!ht]
650: \includegraphics[width=\columnwidth]{evsa.eps}
651: %\vspace{-1.2cm}
652: \caption{Energy (Hartree) vs lattice space $a$ for various ways to approach the $a\to 0$ 
653: limit. The symbol $KE_a=KE$ refers to the choice made in Ref.
654: \protect\onlinecite{lrdmc}, where $\eta$ was obtained  by 
655: setting the lattice regularized kinetic energy  equal to the 
656: continuous one.  
657:  \label{evsa}}
658: \end{figure}
659: 
660: 
661: \section{Results}
662: \label{results}
663: Before tackling the calculation of the weak interaction between two benzene
664: molecules, we studied the effect of the basis set 
665: on our results, and the size consistency of our variational JAGP wave
666: function. The basis set dependence has been analyzed extremely carefully on
667: the carbon and oxygen pseudoatoms, as reported in Subsection~\ref{basis_set}, while the
668: size consistency problem of the JAGP ansatz applied to carbon-based compounds 
669: is described in Subsection~\ref{size_consistency}. We studied the relation  between 
670: the size consistency and the binding energy, computed for the carbon
671: dimer and the benzene ring. Finally, in Subsection~\ref{benzene_dimer} we
672: report the main results on the benzene dimer.
673: 
674: In all those calculations, we used soft
675: pseudopotentials to replace the $1s$ electron pair in the carbon and oxygen
676: atoms. The former contains a norm-conserving HF pseudopotential, generated  
677: using the Vanderbilt construction~\cite{vanderbilt}, while
678: an ab-initio energy-adjusted HF pseudopotential is included in the oxygen.
679: In the latter case, the effective core potential has been fitted\cite{filippipseudo} to reproduce 
680: a wide range of HF excitations from the neutral, the cation, and the anion
681: atom. The transferability and the accuracy of the energy-adjusted pseudopotentials have
682: shown to be excellent in a recent systematic study of the carbon dimer binding
683: energy\cite{cyrc2}. However in this work we have not adopted this recent 
684: pseudopotentials for the Carbon based compound.
685: 
686: 
687: \subsection{VMC/LRDMC basis set dependence on carbon and oxygen pseudoatoms}
688: \label{basis_set}
689: As shown in Tabs.~(\ref{table_C}) and (\ref{table_O}), the convergence with the basis set appears 
690: quite rapid in the carbon and oxygen atoms. 
691: Within  the present QMC framework, based on the JAGP,  
692: there is no need of a large basis set,
693: probably 
694: because all the cusp conditions can be fulfilled exactly by the variational 
695: wave function,
696: even if it is expanded over a finite basis set. 
697: In this way the polarization orbitals (e.g. with angular momentum d) have not 
698: to be included in the QMC ansatz for an 
699: accuracy of $\simeq 1 mH$.  This is quite remarkable  
700: if we consider  the sensitivity to the  basis  set 
701: commonly observed  in  conventional  quantum chemistry methods.
702: Indeed, as shown by Dunning\cite{dunning}, 
703: the contribution of the first polarization d-orbital 
704: to the correlation energy of the oxygen atom is  
705: $\simeq 60mH$, an effect about {\em two} order of magnitudes larger 
706: than the one reported in Tab.~\ref{table_O}, both for the VMC and the 
707: LRDMC oxygen atom calculations, where the gain in energy (if any) is within
708: the statistical accuracy of the simulations ($\approx 0.5mH$).
709: In these tables it is interesting to observe that, while the oxygen 
710: is well described by a Jastrow-Slater wave function, in the carbon 
711: atom the  AGP plays a crucial role for characterizing the non-dynamical 
712: correlations, providing an energy gain of about $10mH$ even within  the
713: LRDMC method. This shows that    
714: our approach  based on the JAGP is particularly useful 
715: for generic (saturated and unsaturated) carbon based compounds.
716:   
717: In order  to extend the calculation to large 
718: electronic systems,  an appropriate   contraction  
719: of a large primitive basis set (up to 6s6p) is important to reduce the
720: dimension of the $\lambda$ matrix in the AGP part (Eq.~\ref{expgem}). 
721: Notice that there is a substantial gain in the LRDMC 
722: correlation energy, by slightly increasing the HF 1s1p contracted basis
723: with another contracted s shell.
724: Indeed, already the $[2s1p]$ contraction provides a much better LRDMC energy, implying that 
725: within our JAGP wave function it is possible  to improve substantially  
726: the nodes of the  HF Slater determinant, with a little extension 
727: of the variational freedom. 
728: It is important to emphasize that we have also 
729: optimized the HF determinant in the presence of the  
730: Jastrow factors described in Sec.~\ref{wave_function}. 
731: On the other hand, as shown in Tab.~\ref{table_C}, 
732: we have obtained the HF energy  
733: within our general Monte Carlo optimization scheme, even though, in this 
734: case, it is obviously not necessary to use a statistical method.
735: The LRDMC calculation  in the HF  case was done after 
736: optimizing  the two-body and three-body Jastrow factors, without 
737: changing the HF determinant.
738: Although the variational energy of this Slater-Jastrow wave function  
739: is higher than the corresponding fully optimized HF+Jastrow one, their LRDMC  
740: energies are the same. This implies that,  within a single determinant 
741: wave function, it is difficult in this case to improve the nodes even when 
742: the Jastrow and the determinantal parts are optimized together.
743: 
744: Nevertheless, our optimization scheme 
745: is very stable and reliable and allows to optimize a rather large number of 
746: variational parameters in a systematic way.
747: Within the JAGP ansatz   and in particular for the benzene molecule, 
748: it is extremely important to optimize the wave function in order 
749: to improve the nodal structure, and obtain a good LRDMC total energy. 
750: This was previously pointed out by two of us, 
751: in an all-electron calculation within the standard DMC
752: framework.\cite{casulamol}  
753: In that case, the DMC method
754: provides the same total energies as the LRDMC method, used here, since for  
755: both methods  
756: the FN approximation is  exactly the same in absence of pseudopotentials.
757:  
758: \begin{table*}
759: \caption {\label{table_C} LRDMC ground state energies (Hartree units) 
760:  for carbon pseudoatom  
761: using various basis sets. The LRDMC value is a rigorous upper bound 
762: for the ground state energy. The limit $a\to 0$ was obtained by using
763: $\eta=1+3.2 a^2$ in Eq.~\ref{laplace}.
764: The 2-body Jastrow factor has the form reported in Eq.~\ref{two-body}. 
765: The Jastrow and the AGP (or HF) geminals are 
766: expanded on a primitive gaussian basis denoted by $(ns~mp)$, where $n$ 
767: ($m$) is  the number of s-wave (p-wave) gaussian orbitals.
768: Analogously, the number and type of contracted orbitals follow 
769: the slash symbol.
770: In particular $(6s6p)/[1s1p]$  denotes the standard HF Slater 
771: determinant. For comparison the HF energy  obtained 
772: in the CBS limit is  $-5.319505$ Hartree. 
773: }
774: \begin{ruledtabular}
775: \begin{tabular}{|c|c|c|c|c|}
776:  Wave function   & 3-body J basis & AGP basis   &   VMC  &   
777:   LRDMC      \\
778: \hline
779: \hline
780: AGP+2-body & -   &  (2s2p) &   -5.266 (1)  & -5.397(1)  \\
781: \hline
782: AGP+2-body & -  & (3s3p) &     -5.392 (1)  & -5.416(1)  \\
783: \hline
784: AGP+2-body & - & (4s4p) & -5.4066(4)  &  -5.4178(3)  \\
785: \hline
786: AGP+2-body & - & (5s5p)  & -5.4095(3)  &  -5.4180(1) \\
787: \hline
788: AGP+2-body & - & (6s6p)  & -5.4096(2)  &  -5.4181(1) \\
789: \hline
790: AGP+2-body & - & (5s5p1d)   & -5.4096(2) &  -5.4182(1) \\
791: \hline
792: AGP+2\&3-body & (1s1p) & (5s5p) & -5.4103(2) & -5.4181(1) \\ 
793: \hline
794: HF & -  & (6s6p)/[1s1p]  & -5.3193(3) &  -5.4107(3)  \\
795: \hline
796: HF+2\&3-body & (1s1p)  & (6s6p)/[1s1p]  & -5.3991(3) &  -5.4107(2)  \\
797: \hline
798: AGP+2\&3-body & (2s)  & (6s6p)/[2s1p]  &  -5.4075(2) &   -5.4160(1)  \\
799: \hline 
800: AGP+2\&3-body & (3s2p) & (4s5p)/[2s2p] &  -5.4115(1) &   -5.4182(1) \\ 
801: \hline 
802: AGP+2\&3-body & (3s2p) & (6s6p) &  -5.4113(1) &   -5.4183(1) \\ 
803: \end{tabular}
804: \end{ruledtabular}
805: \end{table*}
806: 
807: \begin{table*}
808: \caption {\label{table_O}   
809: Same as in Tab.(\ref{table_C}) for the oxygen pseudoatom. For a comparison 
810: with the reported values, the unrestricted HF, the MP2 and the CCSD(T) 
811: on the VTZ basis set have total energies of $-15.7149$, $-15.8636$, and $-15.8822$ 
812: respectively, calculated with Gaussian 03, Revision C.02~\cite{gaussian}. 
813: The limit $a\to 0$ was obtained by using
814: $\eta=1+10.8 a^2$ in Eq.~\ref{laplace}.}
815: \begin{ruledtabular}
816: \begin{tabular}{|c|c|c|c|c|}
817:  Wave function   & 3-body J basis & AGP basis   &   VMC  &   
818:   LRDMC      \\
819: \hline
820: \hline
821: AGP+2-body & -   &  (2s2p) &   -15.410 (3)  & -15.834(2)  \\
822: \hline
823: AGP+2-body & -  & (3s3p) &     -15.813(1)   &  -15.884(1)  \\
824: \hline
825: AGP+2-body & - & (4s4p) & -15.8611(6)  & -15.8901(3)  \\
826: \hline
827: AGP+2-body & - & (5s5p)  & -15.8687(4)  &  -15.8916(2) \\
828: \hline
829: AGP+2-body & - & (6s6p)  & -15.8685(6)   &  -15.8918(3) \\
830: \hline
831: AGP+2-body & - & (5s5p1d)   & -15.8679(5)  &  -15.8920(3) \\
832: \hline
833: HF+2-body & - & (5s5p)/[1s1p]   & -15.8674(5)  &  -15.8920(3) \\
834: \end{tabular}
835: \end{ruledtabular}
836: \end{table*}
837: 
838: 
839: \subsection{Binding energy of $C_2$ and benzene molecule, 
840: a size consistency study}
841: \label{size_consistency}
842: 
843: Though the $[2s1p]$ contraction is a rather small basis and does not provide the 
844: converged result in the total energy of the carbon atom, it represents 
845: a good compromise between accuracy and efficiency, because it can 
846: describe satisfactorily the chemical bond in all carbon-based 
847: compounds studied, as it is shown in Tab.~\ref{binding}.
848: 
849: To this purpose, in this Table we have reported two methods 
850: to calculate the binding energy. 
851: In the standard method (method I), we compute the difference between the 
852: total energy at the equilibrium distance and the 
853: sum of the energies of the independent fragments for a chosen atomic basis set.
854: The second method (method II) is based on the evaluation of 
855: the difference between the total energy at the equilibrium distance and the 
856: energy directly obtained when the constituents of the compound are still together, but pulled apart at large distance. 
857: 
858: In the following we show that method II is more appropriate
859: to compute the binding energy within the JAGP ansatz.
860: Indeed, the AGP part is the particle conserving BCS version 
861: only for the total number, not for the number in a local sector of the wave function.  
862: Therefore, if more than one fragment is included in the same AGP wavefunction,
863: the number of electrons  on each fragment is 
864: not conserved, and this leads to unphysical charge fluctuations which are energetically expensive. The Jastrow factor can significantly lower the energy, by imposing the right occupation number, but the local conservation of charge is fully restored only in the CBS limit of the Jastrow expansion. Thus, 
865: with a finite basis set in the Jastrow factor, the JAGP wavefunction
866: is clearly more accurate for a single fragment 
867: than for the whole system,  and method I usually underestimates the binding energy of the compound.
868: On the other hand, method II is much more accurate, as it includes the cancellation of 
869: the finite basis set errors in the Jastrow term.
870: 
871: Moreover, in order to exploit a better cancellation of errors, 
872: it is important that the energy of the  
873: fragments at large distance is obtained by 
874: iteratively optimizing the wavefunction of the compound  for 
875: larger and larger 
876: separations of the fragments. In this way, one follows adiabatically the 
877: fragmentation process, and avoids possible 
878: spurious energy minima, that may  occur in the optimization of a non linear 
879: function such as the JAGP.
880: 
881: For a perfectly size consistent wave function, methods I and II 
882: should coincide in the CBS limit. 
883: The JAGP wavefunction is perfectly size consistent  
884: for fragments which are singlet, and with the Jastrow factor
885: in the CBS limit. In the case of two singlets at large distance, it is 
886: enough to define the matrix $\lambda$ of the compound as the
887: sum of the two fragments $A$ and $B$  ($ \lambda= \lambda^{A,A}+ \lambda^{B,B}$),
888: with an appropriate Jastrow factor freezing the charge in $A$ 
889: and and $B$ when these two are far apart. 
890: In presence of unpaired orbitals, e.g. for the triplet Carbon atom, 
891: size consistency is very difficult to fulfill in general.
892: For instance,  a singlet $S=0$ $C_2$ wavefunction 
893: corresponding to two entangled Carbon atoms at large distance
894: can be obtained only with six independent Slater determinants, by appropriately combining the  two unpaired $p-$orbitals of each Carbon HF wavefunction, 
895: i.e.:
896: \begin{eqnarray}
897: && | S=0,  {\rm A~far~from~B} \rangle =
898: \nonumber \\
899: && \frac{1}{\sqrt{3}} ~
900: | p_x \uparrow A, p_y \uparrow A,  p_x \downarrow B, p_y \downarrow B \rangle  \nonumber \\ 
901: &+&  
902: \frac{1}{\sqrt{3}} ~
903: | p_x \downarrow A, p_y \downarrow A , p_x \uparrow B, 
904: p_y \uparrow  B \rangle  \nonumber \\
905:  &-& \frac{1}{2\sqrt{3}} ~
906:  | p_x \uparrow A, p_y \downarrow A , p_x \uparrow B, p_y \downarrow B \rangle \nonumber \\
907: &-& \frac{1}{2\sqrt{3}} ~ 
908: | p_x \uparrow A, p_y \downarrow A , p_x \downarrow B, p_y \uparrow B \rangle   \nonumber \\
909: &-& \frac{1}{2\sqrt{3}} ~
910: | p_x \downarrow A, p_y \uparrow A , p_x \uparrow B, p_y \downarrow B \rangle \nonumber \\
911: &- & \frac{1}{2\sqrt{3}} ~
912: | p_x \downarrow A, p_y \uparrow A , p_x \downarrow B, p_y \uparrow B \rangle  \label{expc2}
913: \end{eqnarray}  
914: where each term in the above expression is a single determinant,
915: with  the orbitals indicated inside the brackets, toghether with 
916: the four $2s$ orbitals ($2s  \uparrow  A $, $2s  \downarrow  A $, 
917:  $2s \uparrow  B$, $2s  \downarrow  B $).
918: 
919: The JAGP wavefunction can be perfectly 
920: size consistent even for triplet fragments in the ideal but important limit 
921: of strong repulsion between electrons in the same orbital (strong Hubbard $U$). 
922: In this limit the occupation of  the same unpaired $p-$orbital by electrons of opposite 
923: spins is forbidden as in the singlet expansion for $C_2$ (Eq.~\ref{expc2}), and the two Carbon Slater 
924: determinants, each with two unpaired orbitals $p_x$ and $p_y$,  can be joined in a single determinant 
925: (AGP singlet), by turning on matrix elements such as: 
926: $$\lambda^{p_x,p_x}_{A,B} = \lambda^{p_y,p_y}_{A,B}= 
927: \lambda_{A,B}^{p_x,p_y} =- \lambda^{p_y,p_x}_{A,B},  $$
928: where $A$ and $B$ indicate the two Carbon atoms at large distance, and 
929: these matrix elements are assumed to be small compared with  the occupied 
930: $2s$ orbitals (e.g. $\lambda^{2s,2s}_{A,A}=\lambda^{2s,2s}_{B,B}=1$).
931: Then, it is simple to show that the $6$ Slater determinants defining the $C_2$ 
932: singlet can be obtained with the correct coefficients 
933: by { \em a single determinant}  AGP wavefunction,
934: provided the double occupations of the $p_x$ and $p_y$ orbitals can be 
935:  projected out by an appropriate  Jastrow factor. 
936: However,  the Jastrow factor used here, 
937: within the present JAGP expansion, can only partially project out double 
938: occupation of the same orbital because it  depends only on  the 
939: total electron density and not  
940: explicitly on the corresponding 
941:   angular momentum orbital components. 
942: By consequence,  the present JAGP wavefunction 
943: can be only approximately size consistent in this special case. 
944: However, since this  loss of size consistency is clearly due to a local 
945: effect of the correlation on the same atomic orbital,
946: one expects that this contribution should 
947: be almost the same both at the equilibrium and at large $A-B$ distance, 
948: and therefore it should affect weakly the chemical bond.
949:  
950: Within this hypothesis, that looks very well confirmed in 
951: Fig(\ref{evsrc2}), it is possible to obtain a good chemical accuracy 
952: ($\simeq 0.1eV$), by using only a single geminal JAGP ansatz. 
953: Notice that in this picture the LRDMC energy appears very smooth and 
954: reasonable at large distance, even without approaching the energy of 
955: two isolated Carbon atoms. This calculation suggests that in the
956: exact size consistent framework, which includes  
957: many AGP or determinants, the total 
958: energy should acquire only an irrelevant rigid shift, at least 
959: within the LRDMC method.
960: Unfortunately we are not aware of very 
961: accurate calculation of the full energy 
962: dispersion in $C_2$, but the zero point energy (ZPE), computed 
963: from the data in Fig.(\ref{evsrc2}), is in very good agreement 
964: with the experimental value ($4.2mH$)\cite{nonso}, clearly supporting the 
965: accuracy of our calculation apart for an irrelevant energy shift.  
966: 
967: \begin{figure}[!ht]
968: \includegraphics[width=\columnwidth]{evsrc2.eps}
969: %\vspace{-1.2cm} 
970: \caption{\label{evsrc2} Energy for two Carbon atoms 
971: as a function of their distance. We used a (4s5p)/[2s2p] GTO basis set for the AGP part, and a (3s2p) uncontracted GTO basis set for the Jastrow factor. The atomic basis set convergence has been reached within $1mH$ at the VMC level for the molecule at the 
972: equilibrium distance.
973: The LRDMC and VMC energies are not fully size consistent 
974: (for large $R$ they should approach the energies  indicated by the full lines standing below). 
975: The VMC curve shows a maximum at $R=6$ and a shallow minimum at 
976: $R=10$,  which are almost completely 
977: removed by the LRDMC energies (within an accuracy of $0.1eV$). 
978: The inset shows an expansion of the picture around the equilibrium distance. 
979: Here a cubic polynomial has been used for  fitting the data 
980: in the range $ 2.1 \le R \le 3$. From this interpolation 
981: the resulting VMC [LRDMC] equilibrium distance is 
982: $2.357(4)~a.u.$ [$2.358(5)~a.u.$], and the ZPE is 
983: $4.20(4)~mH$ [$4.20(5)~mH$].
984: }
985: \end{figure}
986: 
987: The very remarkable outcome of this careful analysis is that 
988: it is possible to describe well the chemical bond in most of the
989: interesting carbon-based compounds, as shown in Tab.~\ref{binding}. 
990: As a further confirmation that this hypothesis is plausible,
991: we have computed the  equilibrium distance of the carbon dimer (Tab.~\ref{Rminc2}) using the simultaneous optimization of the bond length and the variational parameters as 
992: described in Ref.~\onlinecite{casulamol}. 
993: By means of this technique, based on energy derivatives, we can compute the 
994: bond length much more accurately than by fitting only the total energy around the minimum (e.g. in the calculation of the bond length in Fig.~\ref{evsrc2}, 
995: a much larger statistical error is obtained for this quantity). We found a perfect agreement with the experimental bond length in the large basis set limit. 
996: \begin{table*}
997: \caption {\label{binding} 
998: Binding energy (eV) for carbon-based compounds, obtained with the JAGP 
999: wave function described in the text for a given atomic basis set, reported in
1000: the table. The most accurate binding energy is 
1001: obtained by evaluating the difference between the total energy at the equilibrium 
1002: and at large distance (method  II). For the benzene molecule 
1003: we exploited the size consistency of the JAGP ansatz valid for singlet  
1004: fragments and complete Jastrow factor. Therefore, 
1005: we considered first the fragmentation process $ C_6 H_6 \to 3 H_2 + 3 C_2$, 
1006: and then we used the already determined $C_2$ binding energy with method 
1007: $II$ (for $H_2$ the JAGP is clearly size consistent, since it is exact for 
1008: two electrons).
1009: The less accurate method is the standard one (method I), obtained   
1010: by computing the large distance energy 
1011: by summing  the energy  of each individual fragment with the same basis.  
1012: In the contracted  $[2s2p1d^*]$, $[2s2p^*]$  or   $[2s1p^*]$ cases,
1013:  used generally here  in the large system cases, 
1014: the coefficients of the contracted orbitals 
1015: are   assumed to be independent of the angular momentum projection $l_z$.  
1016: Notice also that the inclusion of the polarization d-orbital 
1017: does not affect the binding of $C_2$  within $1mH$.
1018: The DMC HF binding energy (I) for $C_2$ is 
1019: $5.66eV$\protect\cite{cyrc2}. The last column refers to the non relativistic 
1020: value estimated either by experiments or by a very accurate calculation for 
1021: $C_2$.}
1022: \begin{ruledtabular}
1023: %{\scriptsize
1024: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
1025:  Compound    &  3-body J basis & AGP basis &   \#  par &
1026:   VMC (I) &  LRDMC (I) & VMC(II)  & LRDMC (II)   & Estimated  \\
1027: \hline
1028: $C_2$ &  (3s2p) & (6s6p)/[2s1p$^*$] &  69 & 5.806 (16)  & 5.946(4) & 6.766(25)  &  6.267(4)  &  6.36(1) \footnotemark[1] \\ 
1029: \hline
1030: $C_2$ &  (3s2p) & (6s6p)/[2s1p]  &  74 & 5.884(16) & 5.959(4) & 6.862(6) &  6.283(5)  &  6.36(1) \footnotemark[1] \\ 
1031: \hline 
1032: $C_2$ &  (3s2p) & (4s5p)/[2s2p$^*$] & 95 & 5.688(8) & 5.883(4) & 6.910(7) & 6.318(9) & 6.36(1) \footnotemark[1] \\ 
1033: \hline 
1034: $C_2$ &  (3s2p1d) & (4s5p1d)/[2s2p1d$^*$] & 136  & 5.724(8) & 5.887(4) & 6.893(7) & 6.314(7) & 6.36(1) \footnotemark[1] \\ 
1035: \hline
1036: $C_2$ &  (3s2p) & (6s6p) &  255 & 5.763(12) & 5.812(4) &  6.737(5) &  6.289(4) &   6.36(1) \footnotemark[1] \\ 
1037: \hline
1038: \hline
1039: $C_6 H_6$ & (3s2p) & (6s6p)/[2s1p$^*$] & 505 & 57.06(3) &  58.105(9) &  59.942(60)  &  59.067(8) &   59.24(11) \footnotemark[2] \\ 
1040: \end{tabular}
1041: \footnotetext[1]{Ref.~\onlinecite{cyrc2}}
1042: \footnotetext[2]{Ref.~\onlinecite{ermler}}
1043: %}
1044: \end{ruledtabular}
1045: \end{table*}
1046: 
1047: \begin{table}
1048: \caption {\label{Rminc2} 
1049: Equilibrium distance of the $C_2$ molecule obtained by minimizing 
1050: the energy of the JAGP with the given basis  set.  The symbols used refer 
1051: to the ones defined in previous tables.}
1052: \begin{ruledtabular}
1053: \begin{tabular}{|c|c|c|c|c|}
1054: 3-body J basis & AGP basis & \#  par &  R (VMC)  & R (exp)    \\
1055: \hline
1056: \hline
1057:  (3s2p) & (6s6p)/[2s1p$^*$] &  69 & 2.3555(8)    &  2.3481  \\ 
1058: \hline
1059: (3s2p) & (6s6p)/[2s1p]  &  74 & 2.3559(9)  & 2.3481  \\ 
1060: \hline
1061: (3s2p) & (6s6p) &  255 & 2.3480(6)  & 2.3481  \\ 
1062: \end{tabular}
1063: \end{ruledtabular}
1064: \end{table}
1065: 
1066: \subsection{The benzene dimer}
1067: \label{benzene_dimer}
1068: As discussed in the previous subsection for two singlet molecules A and B  with 
1069: electron number $N_A$ and $N_B$ respectively, the JAGP is size consistent 
1070: whenever the three-body Jastrow factor is optimized in the CBS limit.
1071: In this way, this term can fully project out the charge fluctuations present in the AGP 
1072: part of the wave function, which would erroneously allow  a number of 
1073: electrons different from $N_A$ and $N_B$ even when the molecules A and B are 
1074: at very large distance.  
1075: In our variational wave function the Jastrow geminal (Eq.~\ref{3body}) 
1076: is defined only on a (3s2p) single zeta 
1077: gaussian  basis set for the carbon atom, and a (1s) single zeta for the
1078: hydrogen atom.
1079: Nevertheless, the wave function is very close to be 
1080: size consistent, because the total 
1081: energy evaluated at a fairly large distance, i.e. $12~a.u.$, 
1082: is given by $E_{A+B}=-75.0825 \pm 0.0003 H$ after a full 
1083: energy optimization,  
1084: whereas the energy of a single benzene molecule within the same basis set
1085: is given by  $E_{A}=-37.5422 \pm 0.0002 H$, i.e. exactly $E_{A+B}/2$ 
1086: within error bars. Therefore, the JAGP ansatz with the chosen basis set 
1087: is supposed to be accurate enough to describe the weak interactions in the benzene dimer, 
1088: as both the basis set convergence and the size consistent behavior are taken
1089: into account.
1090: 
1091: The full dispersion curve of the benzene dimer is reported in
1092: Fig.(\ref{dimpar}) for a face-to-face 
1093: geometry, together with the more accurate LRDMC results. 
1094: As it is apparent from this picture, the 
1095: LRDMC result does not change qualitatively  the variational outcome, showing 
1096: a very weak dispersion, much less deep if compared to the 
1097: most accurate CCSD(T) results. Our best value of the binding energy is
1098: $0.5(3)$ kcal/mol. 
1099: It is possible that the LRDMC method 
1100: reduces the VMC bond length  ($9-10 a.u.$) 
1101: by  $1-2 a.u.$, though an accurate determination of this quantity is 
1102: rather difficult due to the very shallow minimum.
1103: 
1104: We have extended the calculation to the parallel displaced geometry 
1105: (see Fig.\ref{c12h12pd}), which has been proposed 
1106:  to be the most stable configuration.
1107: However, since in this case the number of variational parameters is larger
1108: ($\simeq 10000$), we have used partial information of the Hessian matrix, 
1109: following the scheme introduced in Ref.~(\onlinecite{srh}) to accelerate
1110: the convergence of the minimization.
1111: In the first iteration we move the parameters along
1112: the direction $\vec g_1=s^{-1}\vec  f$, with $s$ the regularized
1113: SR matrix in Eq.~\ref{regularization}. At this step, 
1114: the Hessian matrix gives the optimal amplitude $\gamma_1$ of  the
1115: parameter change $\delta \vec \alpha = \gamma_1 \vec  g_1 $.
1116: Analogously, after $n$ iterations 
1117: the variation of  the  parameters is given by 
1118: $\delta \vec \alpha = \sum_{i=1}^n \gamma_i \vec g_i $, where 
1119: $\{\gamma_i\}_{i=1,n}$ are  determined by using the Hessian matrix 
1120: of the last iteration $n$. After 
1121: changing the variational parameters $ \vec \alpha_k \to \vec \alpha_k+ 
1122: \delta\vec  \alpha_k$, a new vector $\vec g_{n+1}=s^{-1}\vec  f $ 
1123: is computed with the new wave function, and then the procedure is repeated
1124: iteratively. Notice that a single optimal direction is ``collective'' in the
1125: parameter space, as it involves many degrees of  freedom. 
1126: In this way the minimization proceeds in a very stable and fast way,
1127: as shown in Fig.~\ref{minwr}. The main advantage of this method  is that
1128: the Hessian matrix can be calculated in a small basis 
1129: set, and it is not necessary to introduce 
1130: any further regularization parameter other than $\epsilon=10^{-4}$ (Eq.~\ref{regularization}).
1131: \begin{figure}[!ht]
1132: \includegraphics[width=\columnwidth]{minwr.eps}
1133: %\vspace{-1.2cm} 
1134: \caption{\label{minwr} 
1135: Total energy (Hartree) of the variational wave function during the 
1136: optimization of all the $10405$ variational parameters 
1137: consistent with the chosen basis
1138: in the parallel displaced geometry shown in Fig.(\ref{c12h12pd}).
1139: The case $R_1=7~a.u.$ and $R_2=3.4~a.u.$ is considered here. In the inset the 
1140: evolution of the variational parameter $F$ (Eq.~\ref{two-body}) is shown.
1141: }
1142: \end{figure}
1143: 
1144: The optimization of the parallel displaced benzene dimer is rather heavy
1145: (about two  days on a 64 processor SP5 parallel machine) 
1146: because in every iteration shown in Fig.(\ref{minwr}) a 
1147: very high statistical accuracy is required, 
1148: due to the so many variational parameters, 
1149: otherwise all the matrices involved in the iteration process (especially the 
1150: large overlap matrix $s$) are too much noisy. 
1151: For this reason we have performed the wave function optimization 
1152: only for two  particular geometries reported in Tab.~\ref{bindingbe}.
1153: From  the force components in the two inequivalent directions, 
1154: it is clear that the minimum energy  occurs at a value of
1155: $R_2=7.5  \pm 0.2 a.u.$, while $R_1$ is about unchanged. The binding energy is 
1156: $2.2(3)$ kcal/mol.
1157: 
1158: \begin{figure}[!ht]
1159: \includegraphics[width=\columnwidth]{evsr.eps}
1160: %\vspace{-1.2cm} 
1161: \caption{\label{dimpar} Energy for two face-to-face benzene molecules 
1162: as a function of their distance for different methods.  
1163: The reference was taken at $R=12$. The LRDMC kinetic parameters are
1164: $\eta=1.8$, $a=0.5 a.u.$, and $a^\prime/a=\sqrt{5}$.
1165: The nearest neighbor $C-C$ ($C-H$) distance was set 
1166: to $2.636$ ($2.038$) a.u. in the two molecules.}
1167: \end{figure}
1168: 
1169: \section{Conclusion}
1170: In  this work, we have devised a QMC framework which is able to provide
1171: reliable estimates of weak chemical bonds, mainly driven by vdW dispersive
1172: forces. We used a  
1173: highly correlated variational wave function, the JAGP ansatz, which contains
1174: all the necessary ingredients to describe intermolecular interactions: (i) a very
1175: high ``on site'' accuracy, through the inclusion of near degeneracy
1176: correlation effects in the AGP part, 
1177: (ii)  the possibility to control the molecular charge
1178: distribution  through a local 3-body Jastrow factor, (iii)  the capability to
1179: take into account the intermolecular correlation, responsible for the weak
1180: dispersive forces, by means of a ``long-range'' Jastrow  term,  which connects
1181: the molecules (or the fragments) involved in the interaction.
1182: Although the JAGP ansatz is not size consistent in general,
1183: we have shown that in the carbon-based compounds analyzed here it is possible
1184: to obtain accurate and reliable  results, by taking 
1185: the calculation of the system in the large distance geometry as reference point.
1186: 
1187: We have described an improved optimization method, based on a proper
1188: regularization of the overlap matrix in the SR scheme,
1189: which can be further boosted by the information of the Hessian matrix.
1190: With this method it is possible to optimize a number of parameters of the
1191: order of $10000$. Our fully optimized variational wave function has been used 
1192: as initial guess in projection LRDMC calculations. We also found the optimal
1193: setting of the kinetic parameters in the LRDMC method, in order to speed up
1194: the diffusion MC simulations with pseudopotentials.
1195: After the optimization step and the LRDMC projection, our results are very weakly
1196: dependent on  the basis set used, at variance with the post-HF quantum chemistry
1197: methods.
1198: 
1199: We studied the face-to-face and displaced parallel geometry and energetics 
1200: of the benzene dimer, which is a prototype compound to understand
1201: intermolecular dispersive 
1202: forces. After a full optimization of both the Jastrow and the 
1203: AGP part, the VMC binding energy remains in qualitative agreement 
1204: with the LRDMC result, which is supposed to be the most accurate
1205: QMC calculation. All these findings strongly support the reliability of our
1206: numerical study. 
1207: 
1208: The binding of the benzene dimer appears small 
1209: and almost negligible ($\simeq 0.5$ kcal/mol) in the face-to-face geometry.  
1210: On the other hand, in the  parallel displaced configuration
1211: where the two molecules are shifted by a distance  $R_1=3.4 a.u.$, 
1212: there is a sizable gain in energy, 
1213: which reaches its optimal value of $~2.2(3)$ kcal/mol at $R_2 \simeq 7 a.u.$.
1214: Apparently, this is smaller than the most recent post HF value ($~2.8$
1215: kcal/mol \cite{cc}) obtained with the $CCSD(T)$ method, after
1216: a careful extrapolation to the CBS limit. However, 
1217: by considering the reduction of the binding energy due to the 
1218: zero point vibrational energy $ZPE$ ($\Delta ZPE=0.37$kcal/mol), 
1219: our result goes clearly in the direction of the best experimental estimate 
1220: of the binding energy, which is $1.6 \pm 0.2$ kcal/mol.\cite{bestexp}. The
1221: agreement between the experiment and our theoretical prediction is another
1222: striking sign of the capability of the QMC techniques to describe accurately
1223: not only a strong intramolecular bond, but also 
1224: the very weak intermolecular attractions based on vdW dispersive forces.
1225: 
1226: 
1227: \begin{table}
1228: \caption {\label{bindingbe} 
1229: Binding energies $\Delta E$ (kcal/mol) 
1230: and forces (kcal/(mol a.u.)) acting on the two independent directions
1231: $\vec R_1$ and $\vec R_2$ shown in Fig.~\ref{c12h12pd}. Energies differences
1232: are evaluated  with respect to the large separation geometry
1233: ($R_1=0$,$R_2=12  a.u.$), used also in Fig.~\ref{dimpar}. The 
1234: forces are computed in a VMC calculation with the optimized variational
1235: wave function, and include both Feynman and Pulay contributions.
1236:  $R_1$ and $R_2$ are given  in a.u. }
1237: \begin{ruledtabular}
1238: \begin{tabular}{|c|c|c|c|c|c|}
1239:  $R_1$ & $R_2$  &  $F_1$ &  $F_2$ & $\Delta E_{VMC}$  &$\Delta  E_{LRDMC}$ \\
1240:  \hline  
1241:   0  &   7     &    0        &   2.1(2)    &   -1.4(4)  &  0.2(3) \\   
1242: \hline
1243:   0  &   8     &    0        &   0.1(2)   &     0.7(3)  &   0.5(3)  \\
1244:   \hline 
1245:  3.4 &   7     &   0.20(8) & 0.6(1)  &     1.4(3)  &   2.2(3)  \\ 
1246:  \hline  
1247:  3.4 &   8     &   -0.22(6) & -0.7(1) &    2.0 (3)  &  1.8(3)   \\  
1248: \end{tabular}
1249: \end{ruledtabular}
1250: \end{table}
1251: 
1252: \begin{figure}[!ht]
1253: \includegraphics[width=\columnwidth]{c12h12pd.eps}
1254: %\vspace{-1.2cm}
1255: \caption{ 
1256:  Geometry of the benzene dimer with the $R_1$ and $R_2$ distances studied in
1257:  this work.\label{c12h12pd}}
1258: \end{figure}
1259: 
1260: 
1261: \acknowledgments 
1262: We thank Leonardo Guidoni and Giacinto Scoles for helpful discussions, and Claudia Filippi for sending 
1263: us accurate pseudopotentials for oxygen and carbon atoms.
1264: We are also indebted to J. Toulouse and C. Umrigar for their careful reading of
1265: this manuscript. 
1266: This work was partially supported by COFIN2005, and CNR. 
1267: One of us (M.C.) acknowledges support in the form of the NSF grant DMR-0404853.
1268: 
1269: \begin{thebibliography}{99}
1270: \bibitem{bestexp} H. Krause, B. Ernstberger, H. J. Neusser, 
1271: Chem. Phys. Lett. {\bf 184}, 411 (1991).
1272: 
1273: \bibitem{old} K. C. Janda, J. C. Hemminger, J. S. Winn, S. E. Novick,
1274:   S. J. Harris, and W. Klemperer, J. Chem. Phys. {\bf 63}, 1419 (1975).  
1275: 
1276: \bibitem{scoles} R. Schmied, P. Carcabal, A. M. Dokter, V. P. A. Lonij,
1277:   K. K. Lehmann, and G. Scoles, J. Chem. Phys. {\bf 121}, 2701 (2004).  
1278: 
1279: \bibitem{cc} S. Tsuzuki, T. Uchimaru, K. Matsumura, M. Mikami, and K. Tanabe,
1280: Chem. Phys. Lett. {\bf 319}, 547 (2000);
1281: S. Tsuzuki and H. P. L\"uthi, J. Chem. Phys. {\bf 114}, 3949 (2001); 
1282: S. Tsuzuki, T. Uchimaru, K. Sugawara, and M. Mikami,
1283: J. Chem. Phys. {\bf 117}, 11216 (2002);   
1284: S. Tsuzuki, K. Honda,
1285: T. Uchimaru, M. Mikami, and K. Tanabe, J. Am. Chem. Soc. {\bf 124}, 104 (2002). 
1286: 
1287: 
1288: \bibitem{cc2006} Y. C. Park and J. S. Lee, J. Phys. Chem. A {\bf 110}, 5091 (2006).
1289: 
1290: \bibitem{lind} M. Dion, H. Rydberg, E. Schr\"oder, D. C. Langreth, and
1291:  B. I. Lundqvist,  \prl {\bf 92}, 246401 (2004). 
1292: 
1293: \bibitem{guidoni} M. O. Sinnokrot and C. D. Sherrill, J. Phys. Chem A, 
1294: {\bf 110}, 10656 (2006).  
1295: 
1296: \bibitem{grover} J. R. Grover, E. A. Walters, and E. T. Hui, J. Phys. Chem.,
1297:   {\bf 91} 3233 (1987).  
1298: 
1299: \bibitem{kohn} W. Kohn, Y. Meier, and D. E. Makarov, \prl {\bf 80}, 4153 (1998).
1300: 
1301: \bibitem{gross} M. Lein, J. F. Dobson, and E. K. U. Gross, J. Comp. Chem. {\bf
1302:   20}, 12 (1999).
1303: 
1304: \bibitem{ursula} O. Anatole von Lilienfeld, I. Tavernelli, 
1305:   U. Rothlisberger, and D. Sebastiani, \prl {\bf 93}, 153004 (2004).  
1306: 
1307: \bibitem{degironc} V. H. Nguyen, and S. de Gironcoli, to be published.
1308: 
1309: \bibitem{dmc} P. J. Reynolds, D. M. Ceperley, B. J. Alder, and W. A. Lester
1310:   Jr., J. Chem. Phys. {\bf 77}, 5593 (1982); C. J. Umrigar, M. P. Nightingale,
1311:   and K. J. Runge, J. Chem. Phys. {\bf 99}, 2865 (1993). 
1312: 
1313: \bibitem{lrdmc}  M. Casula, C. Filippi, and S. Sorella, Phys. Rev. Lett.,
1314:   {\bf 95}, 100201 (2005). 
1315: 
1316: \bibitem{gfmc} C. Diedrich, A. L\"uchov, and S. Grimme, J. Chem. Phys. {\bf
1317:   123}, 184106 (2005). 
1318: 
1319: \bibitem{filippimol} C. Filippi and C. J. Umrigar, J. Chem. Phys. {\bf 105},
1320:   213 (1996).  
1321: 
1322: \bibitem{casulamol}  M. Casula, C. Attaccalite, and S. Sorella, 
1323: J. Chem. Phys. {\bf 121} 7110 (2004).
1324: 
1325: \bibitem{needs} I. G. Gurtubay, N. D. Drummond, M. D. Towler, and R. J. Needs,
1326:   J. Chem. Phys. {\bf 124}, 024318 (2006); 
1327: N. D. Drummond, P. Lopez Rios, A. Ma, J. R. Trail, G. G. Spink, 
1328: M. D. Towler, and R.J. Needs, J. Chem. Phys. {\bf 124}, 224104 (2006).
1329: 
1330: \bibitem{ceperley} D. F. B. ten Haaf, H. J. M. van Bemmel, 
1331: J. M. J.  van Leeuwen, W. van Saarloos, and D. M. Ceperley,
1332: \prb   {\bf 51}, 13039 (1995).
1333: 
1334: \bibitem{casulaagp}  M. Casula and S. Sorella, J. Chem.  Phys. {\bf 119}, 6500
1335:   (2003). 
1336: 
1337: \bibitem{pauling} L. Pauling, in \emph{The nature of the chemical bond}, Third
1338:   edition, Cornell University Press, Ithaca, New York, page 204.
1339: 
1340: \bibitem{pwanderson} see e.g. 
1341: P. Fazekas and P.W. Anderson, Philos. Mag. {\bf 30}, 423 (1974);
1342: P.W. Anderson, Science {\bf 235}, 1196 (1987).
1343: 
1344: \bibitem{mitas} see e.g. 
1345:  W. M. C. Foulkes, L. Mitas, R. J. Needs, and G. Rajagopal, 
1346: Rev. Mod. Phys. {\bf 73}, 33 (2001). 
1347: 
1348: \bibitem{pier} M. Holzmann, D. M. Ceperley, C. Pierleoni, and K. Esler, \pre {\bf 68}, 046707 (2003).   
1349: 
1350: \bibitem{sr}  S. Sorella, Phys. Rev. B \textbf{64}, 024512 (2001).
1351: 
1352: \bibitem{umrigarhess} C. J. Umrigar and C. Filippi, 
1353:  \prl {\bf 94}, 150201 (2005). 
1354: 
1355: \bibitem{srh} S. Sorella,  Phys. Rev. B {\bf 71}, 241103 (2005).
1356: 
1357: \bibitem{cyrc2} C. J. Umrigar, J. Toulouse, C. Filippi, S. Sorella, and 
1358: R. G. Hennig, cond-mat/0611094 (2006), to appear in \prl.
1359: 
1360: \bibitem{hopping} With probability $1/2$ we perform the standard Metropolis 
1361: algorithm: in this case each electron coordinate 
1362: is moved  only in a neighborhood of the initial position $\vec r_i$.
1363: With the same probability  we make a large move in the direction of an
1364: arbitrary atom $R_j$ ( chosen  
1365: randomly), by a vector equal to $\vec R_j-\vec R_{min}$ 
1366:  where $\vec R_{min}$ represents  the  
1367: atomic  position closest to $\vec r_i$.   
1368: 
1369: \bibitem{note3} This diagonal matrix is positive definite because 
1370: every diagonal element  $s_{i,i}> 0$, as  $s$ is positive definite. 
1371: 
1372: \bibitem{rappe}  M. Casalegno, M. Mella, and  A. M. Rappe,
1373:   J. Chem. Phys. \textbf{118}, 7193 (2003).
1374: 
1375: \bibitem{night} M. P. Nightingale and V. Melik-Alaverdian, \prl {\bf 87}, 043401   
1376: (2001). 
1377: 
1378: \bibitem{casuladmc} M. Casula, Phys. Rev. B  {\bf 74}, 161102(R) (2006).
1379: 
1380: \bibitem{notelrdmc} In practice we set 
1381: the locality parameter $\alpha=0$ in the LRDMC scheme\cite{lrdmc}. 
1382: 
1383: \bibitem{vanderbilt} D. Vanderbilt, Phys. Rev. B {\bf 32}, 8412 (1985).
1384: 
1385: \bibitem{filippipseudo} M. Burkatzki, C. Filippi, and M. Dolg, to be published. 
1386: 
1387: \bibitem{dunning}  T. H. Dunning Jr., J. Chem. Phys.  {\bf 90}, 1007 (1989). 
1388: 
1389: \bibitem{gaussian} Gaussian 03, Revision C.02, M. J. Frisch, G. W. Trucks,
1390:   H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman,
1391:   J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam,
1392:   S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani,
1393:   N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota,
1394:   R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao,
1395:   H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross,
1396:   V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev,
1397:   A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala,
1398:   K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski,
1399:   S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick,
1400:   A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui,
1401:   A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu,
1402:   A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith,
1403:   M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill,
1404:   B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A. Pople, Gaussian,
1405:   Inc., Wallingford CT, 2004. 
1406: 
1407: 
1408: \bibitem{nonso} see e.g. L. Bytautas and K. Ruedenberg J. Chem. Phys. {\bf 122},and references therein. 
1409: 
1410: \bibitem{ermler} W. C. Ermler and C. W. Kern, J. Chem. Phys. {\bf 58},
1411: 3458 (1973).
1412: 
1413: 
1414: 
1415: \end{thebibliography}
1416: 
1417: \end{document}
1418: