working on the manuscript

This commit is contained in:
Emmanuel Giner 2019-04-08 12:13:44 +02:00
parent 45aaef12e4
commit a4f89808bb

View File

@ -139,8 +139,8 @@ For example, CCSD(T)+LDA/cc-pVTZ and CCSD(T)+PBE/cc-pVTZ return mean absolute de
Contemporary quantum chemistry has developed in two directions --- wave function theory (WFT) \cite{Pop-RMP-99} and density-functional theory (DFT). \cite{Koh-RMP-99}
Although both spring from the same Schr\"odinger equation, each of these philosophies has its own advantages and shortcomings.
WFT is attractive as it exists a well-defined path for systematic improvement.
For example, the coupled cluster (CC) family of methods offers a powerful WFT approach for the description of weakly correlated systems and is well regarded as the gold standard of quantum chemistry.
WFT is attractive as it exists a well-defined path for systematic improvement \manu{and powerful tools, such as perturbation theory, to guide the development of new attractive WFT models}.
\trashMG{For example, t} The coupled cluster (CC) family of methods \trashMG{offers a powerful WFT approach} \manu{are a typical example of the WFT philosophy} for the description of weakly correlated systems and is well regarded as the gold standard of quantum chemistry.
By increasing the excitation degree of the CC expansion, one can systematically converge, for a given basis set, to the exact, full configuration interaction (FCI) limit, although the computational cost associated with such improvement is usually pricey.
One of the most fundamental drawback of conventional WFT methods is the slow convergence of energies and properties with respect to the size of the one-electron basis set.
This undesirable feature was put into light by Kutzelnigg more than thirty years ago. \cite{Kut-TCA-85}
@ -203,9 +203,9 @@ Importantly, in the limit of a complete basis set $\Bas$ (which we refer to as $
\label{eq:limitfunc}
\lim_{\Bas \to \infty} \qty( \E{\modX}{\Bas} + \bE{}{\Bas}[\n{\modY}{\Bas}] ) = \E{\modX}{} \approx E,
\end{equation}
where $\E{\modX}{}$ is the energy associated with the method $\modX$ in the complete basis set.
where \trashMG{$\E{\modX}{}$} \manu{$\E{\modX}{\infty}$} is the energy associated with the method $\modX$ in the complete basis set.
In the case $\modX = \FCI$, we have as strict equality as $E_{\FCI}^\infty = E$.
Provided that the functional $\bE{}{\Bas}[\n{}{}]$ is known exactly, \manu{the only source of error at this stage lies in the potential approximate nature of the methods $\modX$ and $\modY$}.
Provided that the functional $\bE{}{\Bas}[\n{}{}]$ is known exactly, \manu{the only source of error at this stage lies in the potential approximate nature of the methods $\modX$ and $\modY$} \manu{for the FCI energy and density within $\Bas$}.
%Here we propose to generalize such approach to a general WFT model, referred here as $\model$, projected in a basis set $\Bas$ which must provides a density $\denmodel$ and an energy $\emodel$.
%As any wave function model is necessary an approximation to the FCI model, one can write
@ -297,10 +297,11 @@ Provided that the functional $\bE{}{\Bas}[\n{}{}]$ is known exactly, \manu{the o
% \EexFCIinfty \approx \EexFCIbasis + \efuncden{\dencipsi}
%\end{equation}
However, in addition of being unknown, the functional $\bE{}{\Bas}[\n{}{}]$ is obviously \textit{not} universal as it depends on $\Bas$.
One of the consequences of the incompleteness of $\Bas$ is that $\wf{}{\Bas}$ does not have a cusp (i.e.~a discontinuous derivative) at the electron-electron (e-e) coalescence points.
\trashMG{However, in addition of being unknown,} \manu{Rigorously speaking, } the functional $\bE{}{\Bas}[\n{}{}]$ is obviously \textit{not} universal as it depends on $\Bas$.
\manu{Nevertheless, from the physical point of view $\bE{}{\Bas}[\n{}{}]$ plays a quite universal role as it aims at fixing the main }
\trashMG{One of the} consequences of the incompleteness of $\Bas$\manu{, which} is that $\wf{}{\Bas}$ does not have a cusp (i.e.~a discontinuous derivative) at the electron-electron (e-e) coalescence points.
As the e-e cusp originates from the divergence of the Coulomb operator at $r_{12} = 0$, a cuspless wave function could equivalently originate from a Hamiltonian with a non-divergent Coulomb interaction at $r_{12} = 0$.
Therefore, as we shall do later on, it feels natural to evaluate $\bE{}{\Bas}[\n{}{}]$ with short-range density functionals.
Therefore, as we shall do later on, it feels natural to \trashMG{evaluate} \manu{approximate} $\bE{}{\Bas}[\n{}{}]$ with short-range density functionals.
Contrary to the conventional RS-DFT scheme which requires a range-separated \textit{parameter} $\rsmu{}{}$, the spatial inhomogeneity of $\Bas$ forces us to define a range-separated \textit{function} $\rsmu{}{}(\br{})$ as the value of $\rsmu{}{}$ must be known at any point in space.
The first step of our basis set correction consists in obtaining an effective operator $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ which i) is finite at the e-e coalescence point as long as an incomplete basis set is used, and ii) tends to the genuine, unbounded $r_{12}^{-1}$ Coulomb operator in the limit of a complete basis set.
@ -323,61 +324,48 @@ In the final step, we employ $\rsmu{}{}(\br{})$ within short-range density funct
%=================================================================
%\subsection{Effective Coulomb operator}
%=================================================================
To compute the effective operator $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ defined as
\trashMG{To compute} \manu{We define} the effective operator $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ \trashMG{defined} as \manu{(see Appendix A of Ref.~\onlinecite{GinPraFerAssSavTou-JCP-18})}
\begin{equation}
\label{eq:int_eq_wee}
\mel*{\wf{}{\Bas}}{\hWee{\Bas}}{\wf{}{\Bas}} = \iint \W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) \n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2}) \dbx{1} \dbx{2},
\label{eq:def_weebasis}
\W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) = \f{\wf{}{\Bas}}{}(\bx{1},\bx{2})/\n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2})
\end{equation}
where
where $\n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2})$ is the two-body density associated with $\wf{}{\Bas}$
\begin{equation}
\n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2})
= \sum_{pqrs \in \Bas} \SO{p}{1} \SO{q}{2} \Gam{pq}{rs}[\wf{}{\Bas}] \SO{r}{1} \SO{s}{2}
= \sum_{pqrs \in \Bas} \SO{p}{1} \SO{q}{2} \Gam{pq}{rs}[\wf{}{\Bas}] \SO{r}{1} \SO{s}{2},
\end{equation}
is the two-body density associated with $\wf{}{\Bas}$, $\Gam{pq}{rs}[\wf{}{\Bas}] = \mel*{\wf{}{\Bas}}{ \aic{r}\aic{s}\ai{p}\ai{q} }{\wf{}{\Bas}}$ is the two-body density tensor of $\wf{}{\Bas}$, $\SO{i}{}$ are spinorbitals, $\bx{} = \qty(\br{},\sigma)$ collects space and spin variables, and $\int \dbx{} = \sum_{\sigma}\,\int_{\mathbb{R}^3} \dbr{}$), one must realize that
\begin{equation}
\mel*{\wf{}{\Bas}}{\hWee{}}{\wf{}{\Bas}} = \mel*{\wf{}{\Bas}}{\hWee{\Bas}}{\wf{}{\Bas}},
\end{equation}
which states that the expectation value of $\hWee{}$ over $\wf{}{\Bas}$ is equal to the expectation value of its projected version in $\Bas$
\begin{equation}
\label{eq:WeeB}
\hWee{\Bas} = \frac{1}{2} \sum_{pqrs \in \Bas}\V{pq}{rs} \aic{r} \aic{s} \ai{q} \ai{p}
\end{equation}
over the same wave function $\wf{}{\Bas}$, where the indices run over all spinorbitals in $\Bas$ and $\V{pq}{rs}$ are the usual two-electron Coulomb integrals.
Because one can show (see Appendix A of Ref.~\onlinecite{GinPraFerAssSavTou-JCP-18}) that
\begin{subequations}
\begin{align}
\label{eq:expectweeb}
\mel*{\wf{}{\Bas}}{\hWee{\Bas}}{\wf{}{\Bas}} & = \frac{1}{2} \iint \f{\wf{}{\Bas}}{}(\bx{1},\bx{2}) \dbx{1} \dbx{2},
\\
\label{eq:expectwee}
\mel*{\wf{}{\Bas}}{\hWee{}}{\wf{}{\Bas}} & = \frac{1}{2} \iint r_{12}^{-1} \n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2}) \dbx{1} \dbx{2},
\end{align}
\end{subequations}
where
$\Gam{pq}{rs}[\wf{}{\Bas}] = \mel*{\wf{}{\Bas}}{ \aic{r}\aic{s}\ai{p}\ai{q} }{\wf{}{\Bas}}$ is the two-body density tensor of $\wf{}{\Bas}$, $\SO{i}{}$ are spinorbitals, $\bx{} = \qty(\br{},\sigma)$ collects space and spin variables, $\int \dbx{} = \sum_{\sigma}\,\int_{\mathbb{R}^3} \dbr{}$), $\f{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ is defined as
\begin{multline}
\label{eq:fbasis}
\f{\wf{}{\Bas}}{}(\bx{1},\bx{2})
\\
= \sum_{pqrstu \in \Bas} \SO{p}{1} \SO{q}{2} \V{pq}{rs} \Gam{rs}{tu}[\wf{}{\Bas}] \SO{t}{1} \SO{u}{2},
\end{multline}
it comes naturally that
and $\V{pq}{rs}$ are the usual Coulomb two-electron integrals.
With such a definition, $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ verifies
\begin{equation}
\label{eq:def_weebasis}
\W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) = \f{\wf{}{\Bas}}{}(\bx{1},\bx{2})/\n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2}).
\label{eq:int_eq_wee}
\mel*{\wf{}{\Bas}}{\hWee{}}{\wf{}{\Bas}} = \iint \W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) \n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2}) \dbx{1} \dbx{2},
\end{equation}
which therefore can be rewritten as
\begin{equation}
\iint r_{12}^{-1} \n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2}) \dbx{1} \dbx{2} = \iint \W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) \n{\wf{}{\Bas}}{(2)}(\bx{1},\bx{2}) \dbx{1} \dbx{2},
\end{equation}
intuitively motivating $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ as a potential candidate for an effective interaction.
As already discussed in Ref.~\onlinecite{GinPraFerAssSavTou-JCP-18}, $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$ is symmetric, \textit{a priori} non translational nor rotational invariant if $\Bas$ does not have such symmetries and is necessarily \textit{finite} at $r_{12} = 0$ for an incomplete basis set.
Also, as demonstrated in Appendix B of Ref.~\onlinecite{GinPraFerAssSavTou-JCP-18}
Of course, there exists \textit{a priori} an infinite set of functions satisfying \eqref{eq:int_eq_wee}, but thanks to its very definition one can show (see Appendix B of Ref.~\onlinecite{GinPraFerAssSavTou-JCP-18}) that
\begin{equation}
\label{eq:lim_W}
\lim_{\Bas \to \infty}\W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) = r_{12}^{-1}.
\lim_{\Bas \to \infty}\W{\wf{}{\Bas}}{}(\bx{1},\bx{2}) = r_{12}^{-1}\quad \forall \,\,(\bx{1},\bx{2}),
\end{equation}
which therefore guarantees a physically satisfying limit. An important point here is that, with the present definition of $\W{\wf{}{\Bas}}{}(\bx{1},\bx{2})$, one can see the effect of the incompleteness of $\Bas$ on the Coulomb operator itself as a removal of the divergence of the two-electron interaction near the electron coalescence.
%=================================================================
%\subsection{Range-separation function}
%=================================================================
To be able to approximate the complementary functional $\bE{}{\Bas}[\n{}{}]$ thanks to functionals developed in the field of RS-DFT, we associate the effective interaction to a long-range interaction characterized by a range-separation function $\rsmu{}{}(\br{})$.
\manu{Working on that paragraph}
As we can map the Coulomb operator within a basis set $\Bas$ with a non divergent two-electron interaction, we can link the present theory with the RS-DFT which uses a smooth bounded two-electron interactions.
In pracice, to be able to approximate the complementary functional $\bE{}{\Bas}[\n{}{}]$, we use the functionals developed in the field of RS-DFT thanks to we associate the effective interaction to a long-range interaction characterized by a range-separation function $\rsmu{}{}(\br{})$.
Although this choice is not unique, the long-range interaction we have chosen is
\begin{equation}
\w{}{\lr,\rsmu{}{}}(\br{1},\br{2}) = \frac{1}{2} \qty{ \frac{\erf[ \rsmu{}{}(\br{1}) r_{12}]}{r_{12}} + \frac{\erf[ \rsmu{}{}(\br{2}) r_{12}]}{ r_{12}} }.