\documentclass[aip,jcp,reprint,noshowkeys,superscriptaddress]{revtex4-1} \usepackage{graphicx,dcolumn,bm,xcolor,microtype,multirow,amsmath,amssymb,amsfonts,physics,mhchem,xspace,subfigure} \usepackage[T1]{fontenc} \usepackage[utf8]{inputenc} \usepackage{txfonts} \usepackage[ colorlinks=true, citecolor=blue, breaklinks=true ]{hyperref} \urlstyle{same} \newcommand{\ie}{\textit{i.e.}} \newcommand{\eg}{\textit{e.g.}} \newcommand{\alert}[1]{\textcolor{red}{#1}} \definecolor{darkgreen}{HTML}{009900} \usepackage[normalem]{ulem} \newcommand{\toto}[1]{\textcolor{blue}{#1}} \newcommand{\trashAS}[1]{\textcolor{blue}{\sout{#1}}} \newcommand{\titou}[1]{\textcolor{red}{#1}} \newcommand{\trashPFL}[1]{\textcolor{red}{\sout{#1}}} \newcommand{\PFL}[1]{\titou{(\underline{\bf PFL}: #1)}} \newcommand{\mc}{\multicolumn} \newcommand{\fnm}{\footnotemark} \newcommand{\fnt}{\footnotetext} \newcommand{\tabc}[1]{\multicolumn{1}{c}{#1}} \newcommand{\EPT}{E_{\text{PT2}}} \newcommand{\EDMC}{E_{\text{FN-DMC}}} \newcommand{\Ndet}{N_{\text{det}}} \newcommand{\hartree}{$E_h$} \newcommand{\LCT}{Laboratoire de Chimie Th\'eorique (UMR 7616), Sorbonne Universit\'e, CNRS, Paris, France} \newcommand{\ANL}{Argonne Leadership Computing Facility, Argonne National Laboratory, Argonne, IL 60439, United States} \newcommand{\LCPQ}{Laboratoire de Chimie et Physique Quantiques (UMR 5626), Universit\'e de Toulouse, CNRS, UPS, France} \begin{document} \title{Taming the fixed-node error in diffusion Monte Carlo via range separation} %\title{Enabling high accuracy diffusion Monte Carlo calculations with % range-separated density functional theory and selected configuration interaction} \author{Anthony Scemama} \email{scemama@irsamc.ups-tlse.fr} \affiliation{\LCPQ} \author{Emmanuel Giner} \email{emmanuel.giner@lct.jussieu.fr} \affiliation{\LCT} \author{Anouar Benali} \email{benali@anl.gov} \affiliation{\ANL} \author{Pierre-Fran\c{c}ois Loos} \email{loos@irsamc.ups-tlse.fr} \affiliation{\LCPQ} \begin{abstract} \end{abstract} \maketitle \section{Introduction} \label{sec:intro} Within a finite one-electron basis, full configuration interaction (FCI) delivers only an approximate solution of the Schr\"odinger equation. This solution is the eigenpair of an approximate Hamiltonian defined as the projection of the exact Hamiltonian onto the finite many-electron basis of all possible Slater determinants generated within this finite one-electron basis. The FCI wave function can be interpreted as a constrained solution of the true Hamiltonian forced to span the restricted space provided by the one-electron basis. In the complete basis set (CBS) limit, the constraint is lifted and the exact solution is recovered. Hence, the accuracy of a FCI calculation can be systematically improved by increasing the size of the one-electron basis set. Nevertheless, its exponential scaling with the number of electrons and with the size of the basis is prohibitive for most chemical systems. In recent years, the introduction of new algorithms \cite{Booth_2009} and the revival \cite{Abrams_2005,Bytautas_2009,Roth_2009,Giner_2013,Knowles_2015,Holmes_2016,Liu_2016,Garniron_2018} of selected configuration interaction (sCI) methods \cite{Bender_1969,Huron_1973,Buenker_1974} pushed the limits of the sizes of the systems that could be computed at the FCI level. \cite{Booth_2010,Cleland_2010,Daday_2012,Chien_2018,Loos_2018a,Loos_2019,Loos_2020b,Loos_2020c} However, the scaling remains exponential unless some bias is introduced leading to a loss of size consistency. \cite{Evangelisti_1983,Cleland_2010,Tenno_2017} Diffusion Monte Carlo (DMC) is a numerical scheme to obtain the exact solution of the Schr\"odinger equation with a different constraint. In DMC, the solution is imposed to have the same nodes (or zeroes) as a given trial (approximate) wave function. Within this so-called \emph{fixed-node} (FN) approximation, the FN-DMC energy associated with a given trial wave function is an upper bound to the exact energy, and the latter is recovered only when the nodes of the trial wave function coincide with the nodes of the exact wave function. The polynomial scaling with the number of electrons and with the size of the trial wave function makes the FN-DMC method particularly attractive. In addition, the total energies obtained are usually far below those obtained with the FCI method in computationally tractable basis sets because the constraints imposed by the FN approximation are less severe than the constraints imposed by the finite-basis approximation. However, it is usually harder to control the FN error in DMC, and this might affect energy differences such as atomization energies. Moreover, improving systematically the nodal surface of the trial wave function can be a tricky job as \trashAS{there is no variational principle for the nodes}\toto{the derivatives of the FN-DMC energy with respect to the variational parameters of the wave function can't be computed}. The qualitative picture of the electronic structure of weakly correlated systems, such as organic molecules near their equilibrium geometry, is usually well represented with a single Slater determinant. This feature is in part responsible for the success of density-functional theory (DFT) and coupled cluster theory. DMC with a single-determinant trial wave function can be used as a single-reference post-Hatree-Fock method, with an accuracy comparable to coupled cluster.\cite{Dubecky_2014,Grossman_2002} The favorable scaling of QMC, its very low memory requirements and its adequacy with massively parallel architectures make it a serious alternative for high-accuracy simulations of large systems. As it is not possible to minimize directly the FN-DMC energy with respect to the variational parameters of the trial wave function, the fixed-node approximation is much more difficult to control than the finite-basis approximation. The conventional approach consists in multiplying the trial wave function by a positive function, the \emph{Jastrow factor}, taking account of the electron-electron cusp and the short-range correlation effects. The wave function is then re-optimized within variational Monte Carlo (VMC) in the presence of the Jastrow factor and the nodal surface is expected to be improved. Using this technique, it has been shown that the chemical accuracy could be reached within FN-DMC.\cite{Petruzielo_2012} Another approach consists in considering the FN-DMC method as a \emph{post-FCI method}. The trial wave function is obtained by approaching the FCI with a selected configuration interaction (sCI) method such as CIPSI for instance.\cite{Giner_2013,Caffarel_2016_2} \titou{When the basis set is increased, the trial wave function gets closer to the exact wave function, so the nodal surface can be systematically improved.\cite{Caffarel_2016} WRONG} This technique has the advantage that using FCI nodes in a given basis set is well defined, so the calculations are reproducible in a black-box way without needing any expertise in QMC. But this technique cannot be applied to large systems because of the exponential scaling of the size of the trial wave function. Extrapolation techniques have been used to estimate the FN-DMC energies obtained with FCI wave functions,\cite{Scemama_2018} and other authors have used a combination of the two approaches where highly truncated CIPSI trial wave functions are re-optimized in VMC under the presence of a Jastrow factor to keep the number of determinants small,\cite{Giner_2016} and where the consistency between the different wave functions is kept by imposing a constant energy difference between the estimated FCI energy and the variational energy of the CI wave function.\cite{Dash_2018,Dash_2019} Nevertheless, finding a robust protocol to obtain high accuracy calculations which can be reproduced systematically, and which is applicable for large systems with a multi-configurational character is still an active field of research. The present paper falls within this context. \section{Combining CIPSI with range-separated DFT} \label{sec:rsdft-cipsi} In single-determinant DMC calculations, the degrees of freedom used to reduce the fixed-node error are the molecular orbitals on which the Slater determinant is built. Different molecular orbitals can be chosen: Hartree-Fock (HF), Kohn-Sham (KS), natural (NO) orbitals of a correlated wave function, or orbitals optimized under the presence of a Jastrow factor. The nodal surfaces obtained with the KS determinant are in general better than those obtained with the HF determinant,\cite{Per_2012} and of comparable quality to those obtained with a Slater determinant built with NOs.\cite{Wang_2019} Orbitals obtained in the presence of a Jastrow factor are generally superior to KS orbitals.\cite{Filippi_2000,Scemama_2006,HaghighiMood_2017,Ludovicy_2019} The description of electron correlation within DFT is very different from correlated methods. In DFT, one solves a mean field problem with a modified potential incorporating the effects of electron correlation, whereas in correlated methods the real Hamiltonian is used and the electron-electron interactions are considered. Nevertheless, as the orbitals are one-electron functions, the procedure of orbital optimization in the presence of the Jastrow factor can be interpreted as a self-consistent field procedure with an effective Hamiltonian,\cite{Filippi_2000} similarly to DFT. So KS-DFT can be viewed as a very cheap way of introducing the effect of correlation in the orbital parameters determining the nodal surface of a single Slater determinant. Nevertheless, even when using the exact exchange correlation potential at the CBS limit, a fixed-node error necessarily remains because the single-determinant ansätz does not have enough flexibility to describe the nodal surface of the exact correlated wave function of a generic $N$-electron system. If one wants to have to exact CBS limit, a multi-determinant parameterization of the wave functions is required. \subsection{CIPSI} Beyond the single-determinant representation, the best multi-determinant wave function one can obtain is the FCI. FCI is a \emph{post-Hartree-Fock} method, and there exists several systematic improvements between the Hartree-Fock and FCI wave functions: increasing the maximum degree of excitation of CI methods (CISD, CISDT, CISDTQ, \emph{etc}), or increasing the complete active space (CAS) wave functions until all the orbitals are in the active space. Selected CI methods take a shorter path between the Hartree-Fock determinant and the FCI wave function by increasing iteratively the number of determinants on which the wave function is expanded, selecting the determinants which are expected to contribute the most to the FCI eigenvector. At every iteration, the lowest eigenpair is extracted from the CI matrix expressed in the determinant subspace, and the FCI energy can be estimated by computing a second-order perturbative correction (PT2) to the variational energy, $\EPT$. The magnitude of $\EPT$ is a measure of the distance to the exact eigenvalue, and is an adjustable parameter controlling the quality of the wave function. Within the \emph{Configuration interaction using a perturbative selection made iteratively} (CIPSI)\cite{Huron_1973} method, the PT2 correction is computed along with the determinant selection. So the magnitude of $\EPT$ can be made the only parameter of the algorithm, and we choose this parameter as the convergence criterion of the CIPSI algorithm. Considering that the perturbatively corrected energy is a reliable estimate of the FCI energy, using a fixed value of the PT2 correction as a stopping criterion enforces a constant distance of all the calculations to the FCI energy. In this work, we target the chemical accuracy so all the CIPSI selections were made such that $|\EPT| < 1$~mE$_h$. \subsection{Range-separated DFT} \label{sec:rsdft} Following the seminal work of Savin,\cite{Savin_1996,Toulouse_2004} the Coulomb electron-electron interaction is split into a short-range (sr) and a long range (lr) interaction as \begin{equation} \frac{1}{r_{ij}} = w_{\text{ee}}^{\text{lr}, \mu}(r_{ij}) + \qty( \frac{1}{r_{ij}} - w_{\text{ee}}^{\text{lr}, \mu}(r_{ij}) ) \end{equation} where \begin{equation} w_{\text{ee}}^{\text{lr},\mu}(r_{ij}) = \frac{\erf \qty( \mu\, r_{ij})}{r_{ij}} \end{equation} The main idea is to treat the short-range electron-electron interaction with DFT, and the long range with wave function theory. The parameter $\mu$ controls the range of the separation, and allows to go continuously from the Kohn-Sham Hamiltonian ($\mu=0$) to the FCI Hamiltinoan ($\mu = \infty$). To rigorously connect wave function theory and DFT, the universal Levy-Lieb density functional\cite{Lev-PNAS-79,Lie-IJQC-83} is decomposed as \begin{equation} \mathcal{F}[n] = \mathcal{F}^{\mathrm{lr},\mu}[n] + \bar{E}_{\mathrm{Hxc}}^{\mathrm{sr,}\mu}[n], \label{Fdecomp} \end{equation} where $n$ is a one-particle density, $\mathcal{F}^{\mathrm{lr},\mu}$ is a long-range universal density functional and $\bar{E}_{\mathrm{Hxc}}^{\mathrm{sr,}\mu}$ is the complementary short-range Hartree-exchange-correlation (Hxc) density functional\cite{Savin_1996,Toulouse_2004}. One obtains the following expression for the ground-state electronic energy \begin{equation} \label{min_rsdft} E_0= \min_{\Psi} \left\{ \left \langle\Psi|\hat{T}+\hat{W}_\mathrm{{ee}}^{\mathrm{lr},\mu}+\hat{V}_{\mathrm{ne}}|\Psi\right \rangle + \bar{E}^{\mathrm{sr},\mu}_{\mathrm{Hxc}}[n_\Psi]\right\} \end{equation} with $\hat{T}$ the kinetic energy operator, $\hat{W}_\mathrm{ee}^{\mathrm{lr}}$ the long-range electron-electron interaction, $n_\Psi$ the one-particle density associated with $\Psi$, and $\hat{V}_{\mathrm{ne}}$ the electron-nucleus potential. The minimizing multi-determinant wave function $\Psi^\mu$ can be determined by the self-consistent eigenvalue equation \begin{equation} \label{rs-dft-eigen-equation} \hat{H}^\mu[n_{\Psi^{\mu}}] \ket{\Psi^{\mu}}= \mathcal{E}^{\mu} \ket{\Psi^{\mu}}, \end{equation} with the long-range interacting Hamiltonian \begin{equation} \label{H_mu} \hat{H}^\mu[n_{\Psi^{\mu}}] = \hat{T}+\hat{W}_{\mathrm{ee}}^{\mathrm{lr},\mu}+\hat{V}_{\mathrm{ne}}+ \hat{\bar{V}}_{\mathrm{Hxc}}^{\mathrm{sr},\mu}[n_{\Psi^{\mu}}], \end{equation} where $\hat{\bar{V}}_{\mathrm{Hxc}}^{\mathrm{sr},\mu}$ is the complementary short-range Hartree-exchange-correlation potential operator. Once $\Psi^{\mu}$ has been calculated, the electronic ground-state energy is obtained by \begin{equation} \label{E-rsdft} E_0= \mel{\Psi^{\mu}}{\hat{T}+\hat{W}_\mathrm{{ee}}^{\mathrm{lr},\mu}+\hat{V}_{\mathrm{ne}}}{\Psi^{\mu}}+\bar{E}^{\mathrm{sr},\mu}_{\mathrm{Hxc}}[n_{\Psi^\mu}]. \end{equation} Note that, for $\mu=0$, the long-range interaction vanishes, $w_{\mathrm{ee}}^{\mathrm{lr},\mu=0}(r_{12}) = 0$, and thus range-separated DFT (RS-DFT) reduces to standard KS-DFT and $\Psi^\mu$ is the KS determinant. For $\mu\to\infty$, the long-range interaction becomes the standard Coulomb interaction, $w_{\mathrm{ee}}^{\mathrm{lr},\mu\to\infty}(r_{12}) = 1/r_{12}$, and thus RS-DFT reduces to standard wave-function theory and $\Psi^\mu$ is the FCI wave function. \begin{figure*} \centering \includegraphics[width=0.7\linewidth]{algorithm.pdf} \caption{Algorithm showing the generation of the RS-DFT wave function.} \label{fig:algo} \end{figure*} Hence we have a continuous path connecting the KS determinant to the FCI wave function, and as the KS nodes are of higher quality than the HF nodes, we expect that using wave functions built along this path will always provide reduced fixed-node errors compared to the path connecting HF to FCI using an increasing number of selected determinants. We can follow this path by performing FCI calculations using the RS-DFT Hamiltonian with different values of $\mu$. In this work, we have used the CIPSI algorithm to peform approximate FCI calculations with the RS-DFT Hamiltonians,\cite{GinPraFerAssSavTou-JCP-18} $\hat{H}^\mu$ as shown in figure~\ref{fig:algo}. In the outer loop (red), a CIPSI selection is performed with a RS-Hamiltonian parameterized using the current density. An inner loop (blue) is introduced to accelerate the convergence of the self-consistent calculation, in which the set of determinants is kept fixed, and only the diagonalization of the RS-Hamiltonian is performed iteratively with the updated density. The convergence of the algorithm was further improved by introducing a direct inversion in the iterative subspace (DIIS) step to extrapolate the density both in the outer and inner loops. Note that any range-separated post-Hartree-Fock method can be implemented using this scheme by just replacing the CIPSI step by the post-HF method of interest. \section{Computational details} \label{sec:comp-details} All the calculations were made using BFD pseudopotentials\cite{Burkatzki_2008} with the associated double-, triple-, and quadruple-$\zeta$ basis sets (BFD-VXZ). CCSD(T) and KS-DFT calculations were made with \emph{Gaussian09},\cite{g16} using an unrestricted Hartree-Fock determinant as a reference for open-shell systems. All the CIPSI calculations were made with \emph{Quantum Package}.\cite{Garniron_2019,qp2_2020} We used the short-range version of the Perdew-Burke-Ernzerhof (PBE)~\cite{PerBurErn-PRL-96} exchange and correlation functionals of Ref.~\onlinecite{GolWerStoLeiGorSav-CP-06} (see also Refs.~\onlinecite{TouColSav-JCP-05,GolWerSto-PCCP-05}). The convergence criterion for stopping the CIPSI calculations was $\EPT < 1$~m\hartree{} $\vee \Ndet > 10^7$. All the wave functions are eigenfunctions of the $S^2$ operator, as described in ref~\onlinecite{Applencourt_2018}. Quantum Monte Carlo calculations were made with QMC=Chem,\cite{scemama_2013} in the determinant localization approximation (DLA),\cite{Zen_2019} where only the determinantal component of the trial wave function is present in the expression of the wave function on which the pseudopotential is localized. Hence, in the DLA the fixed-node energy is independent of the Jastrow factor, as in all-electron calculations. Simple Jastrow factors were used to reduce the fluctuations of the local energy. \section{Influence of the range-separation parameter on the fixed-node error} \label{sec:mu-dmc} \begin{table} \caption{Fixed-node energies (in hartree) and number of determinants in \ce{H2O} and \ce{F2} with various trial wave functions.} \label{tab:h2o-dmc} \centering \begin{ruledtabular} \begin{tabular}{ccrlrl} & & \multicolumn{2}{c}{BFD-VDZ} & \multicolumn{2}{c}{BFD-VTZ} \\ \cline{3-4} \cline{5-6} System & $\mu$ & $\Ndet$ & $\EDMC$ & $\Ndet$ & $\EDMC$ \\ \hline \ce{H2O} & $0.00$ & $11$ & $-17.253\,59(6)$ & $23$ & $-17.256\,74(7)$ \\ & $0.20$ & $23$ & $-17.253\,73(7)$ & $23$ & $-17.256\,73(8)$ \\ & $0.30$ & $53$ & $-17.253\,4(2)$ & $219$ & $-17.253\,7(5)$ \\ & $0.50$ & $1\,442$ & $-17.253\,9(2)$ & $16\,99$ & $-17.257\,7(2)$ \\ & $0.75$ & $3\,213$ & $-17.255\,1(2)$ & $13\,362$ & $-17.258\,4(3)$ \\ & $1.00$ & $6\,743$ & $-17.256\,6(2)$ & $256\,73$ & $-17.261\,0(2)$ \\ & $1.75$ & $54\,540$ & $-17.259\,5(3)$ & $207\,475$ & $-17.263\,5(2)$ \\ & $2.50$ & $51\,691$ & $-17.259\,4(3)$ & $858\,123$ & $-17.264\,3(3)$ \\ & $3.80$ & $103\,059$ & $-17.258\,7(3)$ & $1\,621\,513$ & $-17.263\,7(3)$ \\ & $5.70$ & $102\,599$ & $-17.257\,7(3)$ & $1\,629\,655$ & $-17.263\,2(3)$ \\ & $8.50$ & $101\,803$ & $-17.257\,3(3)$ & $1\,643\,301$ & $-17.263\,3(4)$ \\ & $\infty$ & $200\,521$ & $-17.256\,8(6)$ & $1\,631\,982$ & $-17.263\,9(3)$ \\ \\ \ce{F2} & $0.00$ & $23$ & $-48.419\,5(4)$ \\ & $0.25$ & $8$ & $-48.421\,9(4)$ \\ & $0.50$ & $1743$ & $-48.424\,8(8)$ \\ & $1.00$ & $11952$ & $-48.432\,4(3)$ \\ & $2.00$ & $829438$ & $-48.441\,0(7)$ \\ & $5.00$ & $5326459$ & $-48.445(2)$ \\ & $\infty$ & $8302442$ & $-48.437(3)$ \\ \end{tabular} \end{ruledtabular} \end{table} \begin{figure} \centering \includegraphics[width=\columnwidth]{h2o-dmc.pdf} \caption{Fixed-node energies of the water molecule for different values of $\mu$.} \label{fig:h2o-dmc} \end{figure} \begin{figure} \centering \includegraphics[width=\columnwidth]{f2-dmc.pdf} \caption{Fixed-node energies of difluorine for different values of $\mu$.} \label{fig:f2-dmc} \end{figure} The first question we would like to address is the quality of the nodes of the wave functions $\Psi^{\mu}$ obtained with an intermediate range separation parameter (\textit{i.e.} $0 < \mu < +\infty$). We generated trial wave functions $\Psi^\mu$ with multiple values of $\mu$, and computed the associated fixed node energy keeping all the parameters having an impact on the nodal surface fixed. We considered two weakly correlated molecular systems: the water molecule and the fluorine dimer, near their equilibrium geometry\cite{Caffarel_2016}. From table~\ref{tab:h2o-dmc} and figures~\ref{fig:h2o-dmc} and~\ref{fig:f2-dmc}, one can clearly observe that using FCI trial wave functions ($\mu = \infty$) gives FN-DMC energies which are lower than the energies obtained with a single Kohn-Sham determinant ($\mu=0$): a gain of $3.2 \pm 0.6$~m\hartree{} at the double-zeta level and $7.2 \pm 0.3$~m\hartree{} at the triple-zeta level are obtained for water, and a gain of $18 \pm 3$~m\hartree{} for F$_2$. Interestingly, using the RS-DFT-CIPSI trial wave function with a range-separation parameter $\mu=1.75$~bohr$^{-1}$ with the double-zeta basis one can obtain for water a FN-DMC energy $2.6 \pm 0.7$~m\hartree{} lower than the energy obtained with the FCI trial wave function. This can be explained by the inability of the basis set to properly describe the short-range correlation effects, shifting the nodes from their optimal position. Using DFT to take account of short-range correlation frees the determinant expansion from describing short-range effects, and enables a placement of the nodes closer to the optimum. In the case of F$_2$, a similar behavior with a gain of $8 \pm 4$ m\hartree{} is observed for $\mu\sim 5$~bohr$^{-1}$. The optimal value of $\mu$ is larger than in the case of water, and this is probably the signature of the fact that the average electron-electron distance in the valence is smaller in F$_2$ than in H$_2$O due to the larger nuclear charge shrinking the electron density. At the triple-zeta level, the short-range correlations can be better described by the determinant expansion, and the effect of sr-DFT on the trial wave function is insignificant on the fixed-node energy. However, it is important to note that the same FN-DMC energy can be obtained with a CI expansion which is eight times smaller when sr-DFT is introduced. One can also remark that the minimum has been slightly shifted towards the FCI, which is consistent with the fact that in the CBS limit we expect the minimum of the FN-DMC energy to be obtained for the FCI wave function, i.e. at $\mu=\infty$. \begin{figure} \centering \includegraphics[width=\columnwidth]{overlap.pdf} \caption{Overlap of the RS-DFT CI expansion with the CI expansion optimized in the presence of a Jastrow factor.} \label{fig:overlap} \end{figure} This data confirms that RS-DFT-CIPSI can give improved CI coefficients with small basis sets, similarly to the common practice of re-optimizing the trial wave function in the presence of the Jastrow factor. To confirm that the introduction of sr-DFT has an impact on the CI coefficients similar to the Jastrow factor, we have made the following numerical experiment. First, we extract the 200 determinants with the largest weights in the FCI wave function out of a large CIPSI calculation. Within this set of determinants, we diagonalize self-consistently the RS-DFT Hamiltonian with different values of $\mu$. This gives the CI expansions $\Psi^\mu$. Then, within the same set of determinants we optimize the CI coefficients in the presence of a simple one- and two-body Jastrow factor. This gives the CI expansion $\Psi^J$. In figure~\ref{fig:overlap}, we plot the overlaps $\braket{\Psi^J}{\Psi^\mu}$ obtained for water and the fluorine dimer. In the case of H$_2$O, there is a clear maximum of overlap at $\mu=1$~bohr$^{-1}$. This confirms that introducing short-range correlation with DFT has the an impact on the CI coefficients similar to the Jastrow factor. In the case of F$_2$, the Jastrow factor has very little effect on the CI coefficients, as the overlap $\braket{\Psi^J}{\Psi^{\mu=\infty}}$ is very close to $1$. Nevertheless, a slight maximum is obtained for $\mu=5$~bohr$^{-1}$. \section{Atomization energies} \label{sec:atomization} Atomization energies are challenging for post-Hartree-Fock methods because their calculation requires a perfect balance in the description of atoms and molecules. Basis sets used in molecular calculations are atom-centered, so they are always better adapted to atoms than molecules and atomization energies usually tend to be underestimated with variational methods. In the context of FN-DMC calculations, the nodal surface is imposed by the trial wavefunction which is expanded on an atom-centered basis set, so we expect the fixed-node error to be also tightly related to the basis set incompleteness error. Increasing the size of the basis set improves the description of the density and of electron correlation, but also reduces the imbalance in the quality of the description of the atoms and the molecule, leading to more accurate atomization energies. \subsection{Size consistency} An extremely important feature required to get accurate atomization energies is size-consistency (or strict separability), since the numbers of correlated electron pairs in the isolated atoms are different from those of the molecules. The energy computed within density functional theory is size-consistent, and as it is a mean-field method the convergence to the complete basis set (CBS) limit is relatively fast. Hence, DFT methods are very well adapted to the calculation of atomization energies, especially with small basis sets. But going to the CBS limit will converge to biased atomization energies because of the use of approximate density functionals. On the other hand, FCI is also size-consistent, but the convergence of the FCI energies to the CBS limit is much slower because of the description of short-range electron correlation using atom-centered functions. But ultimately the exact energy will be reached. In the context of selected CI calculations, when the variational energy is extrapolated to the FCI energy\cite{Holmes_2017} there is no size-consistency error. But when the truncated sCI wave function is used as a reference for post-Hartree-Fock methods such as sCI+PT2 or for QMC calculations, there is a residual size-consistency error originating from the truncation of the wave function. QMC energies can be made size-consistent by extrapolating the FN-DMC energy to estimate the energy obtained with the FCI as a trial wave function.\cite{Scemama_2018,Scemama_2018b} Alternatively, the size-consistency error can be reduced by choosing the number of selected determinants such that the sum of the PT2 corrections on the fragments is equal to the PT2 correction of the molecule, enforcing that the variational potential energy surface (PES) is parallel to the perturbatively corrected PES, which is a relatively accurate estimate of the FCI PES.\cite{Giner_2015} Another source of size-consistency error in QMC calculations originates from the Jastrow factor. Usually, the Jastrow factor contains one-electron, two-electron and one-nucleus-two-electron terms. The problematic part is the two-electron term, whose simplest form can be expressed as \begin{equation} J_\text{ee} = \sum_i \sum_{j exponentially fast convergence (as HF) with basis because of non divergence of effective KS potential (citer le papier de Gill) % \item With correlation consistent basis set, FCI nodes (which include correlation) are better than KS % \item With FCI, good limit at CBS ==> exact energy % \item But slow convergence with basis set because of divergence of the e-e interaction not well represented in atom centered basis set % \item Exponential increase of number of Slater determinants \item Cite papiers RS-DFT: there exists an hybrid scheme combining fast convergence wr to basis set (non divergent basis set) and short expansion in SCI (cite papier Ferté) \item Question: does such a scheme provide better nodal quality ? \item In RS-DFT we cannot optimize energy with $\mu$ , but in FNDMC % \item Factual stuffs: with optimal $\mu$, lower FNDMC energy than HF/KS/FCI % \begin{itemize} % \item less determinants $\Rightarrow$ large systems % \item only one parameter to optimize $\Rightarrow$ deterministic % \item $\Rightarrow$ reproducible % \end{itemize} % \item with the optimal $\mu$: % \begin{itemize} % \item Direct optimization of FNDMC with one parameter % \item Do we improve energy differences ? % \item system dependent % \item basis set dependent: $\mu \rightarrow \infty$ when $\mathcal{B}\rightarrow \text{CBS}$ % \item large wave functions % \end{itemize} \end{itemize} \end{enumerate} \end{document} % * Recouvrement avec Be : Optimization tous electrons % impossible. Abandon. On va prendre H2O. % * Manu doit faire des programmes pour des plots de ensite a 1 et 2 % corps le long des axes de liaison, et l'integrale de la densite a % 2 corps a coalescence. % 1 Manu calcule Be en cc-pvdz tous electrons: FCI -> NOs -> FCI -> % qp edit -n 200 % 2 Manu calcule qp_cipsi_rsh avec mu = [ 1.e-6 , 0.25, 0.5, 1.0, 2.0, 5.0, 1e6 ] % 3 Manu fait tourner ses petits programmes % 4 Manu envoie a toto un tar avec tous les ezfio % 5 Toto optimise les coefs en presence e jastrow % 6 Toto renvoie a manu psicoef % 7 Manu fait tourner ses petits programmes avec psi_J