Sangalli
This commit is contained in:
parent
dc37db6258
commit
bb02d36819
10
BSEdyn.bib
10
BSEdyn.bib
@ -1,7 +1,7 @@
|
||||
%% This BibTeX bibliography file was created using BibDesk.
|
||||
%% http://bibdesk.sourceforge.net/
|
||||
|
||||
%% Created for Pierre-Francois Loos at 2020-06-10 22:40:48 +0200
|
||||
%% Created for Pierre-Francois Loos at 2020-06-17 15:22:32 +0200
|
||||
|
||||
|
||||
%% Saved with string encoding Unicode (UTF-8)
|
||||
@ -2068,16 +2068,10 @@
|
||||
@article{Sangalli_2011,
|
||||
Author = {Sangalli, Davide and Romaniello, Pina and Onida, Giovanni and Marini, Andrea},
|
||||
Date-Added = {2020-05-18 21:40:28 +0200},
|
||||
Date-Modified = {2020-05-18 21:40:28 +0200},
|
||||
Date-Modified = {2020-06-17 15:22:30 +0200},
|
||||
Doi = {10.1063/1.3518705},
|
||||
File = {/Users/loos/Zotero/storage/9S3XW2FJ/Sangalli et al. - 2011 - Double excitations in correlated systems A many--b.pdf},
|
||||
Issn = {0021-9606, 1089-7690},
|
||||
Journal = {J. Chem. Phys.},
|
||||
Language = {en},
|
||||
Month = jan,
|
||||
Number = {3},
|
||||
Pages = {034115},
|
||||
Shorttitle = {Double Excitations in Correlated Systems},
|
||||
Title = {Double Excitations in Correlated Systems: {{A}} Many\textendash{}Body Approach},
|
||||
Volume = {134},
|
||||
Year = {2011},
|
||||
|
@ -1,10 +1,10 @@
|
||||
\documentclass{article}
|
||||
\usepackage{graphicx,dcolumn,bm,xcolor,microtype,multirow,amscd,amsmath,amssymb,amsfonts,physics,longtable,wrapfig}
|
||||
\documentclass[aps,prb,reprint,noshowkeys,superscriptaddress]{revtex4-2}
|
||||
\usepackage{graphicx,dcolumn,bm,xcolor,microtype,multirow,amscd,amsmath,amssymb,amsfonts,physics,longtable,wrapfig,txfonts}
|
||||
\usepackage[version=4]{mhchem}
|
||||
|
||||
%\usepackage[utf8]{inputenc}
|
||||
%\usepackage[T1]{fontenc}
|
||||
%\usepackage{txfonts}
|
||||
\usepackage[utf8]{inputenc}
|
||||
\usepackage[T1]{fontenc}
|
||||
\usepackage{txfonts}
|
||||
|
||||
\usepackage[
|
||||
colorlinks=true,
|
||||
@ -111,6 +111,7 @@
|
||||
\newcommand{\bb}{\mathbf{b}}
|
||||
\newcommand{\bA}{\mathbf{A}}
|
||||
\newcommand{\bB}{\mathbf{B}}
|
||||
\newcommand{\bc}{\mathbf{c}}
|
||||
\newcommand{\bx}{\mathbf{x}}
|
||||
|
||||
% units
|
||||
@ -137,14 +138,32 @@
|
||||
|
||||
\newcommand{\LCPQ}{Laboratoire de Chimie et Physique Quantiques (UMR 5626), Universit\'e de Toulouse, CNRS, UPS, France}
|
||||
|
||||
\title{Notes on the Dynamical Bethe-Salpeter Equation}
|
||||
|
||||
\author{Pierre-Fran\c{c}ois Loos}
|
||||
|
||||
\begin{document}
|
||||
|
||||
\title{Dynamical Kernels}
|
||||
|
||||
\author{Pierre-Fran\c{c}ois \surname{Loos}}
|
||||
\email{loos@irsamc.ups-tlse.fr}
|
||||
\affiliation{\LCPQ}
|
||||
|
||||
\begin{abstract}
|
||||
Similar to the ubiquitous adiabatic approximation in time-dependent density-functional theory, the static approximation, which substitutes a dynamical (\ie, frequency-dependent) kernel by its static limit, is usually enforced in most implementations of the Bethe-Salpeter equation (BSE) formalism.
|
||||
Here, going beyond the static approximation, we compute the dynamical correction in the electron-hole screening for molecular excitation energies thanks to a renormalized first-order perturbative correction to the static BSE excitation energies.
|
||||
The present dynamical correction goes beyond the plasmon-pole approximation as the dynamical screening of the Coulomb interaction is computed exactly.
|
||||
Moreover, we investigate quantitatively the effect of the Tamm-Dancoff approximation by computing both the resonant and anti-resonant dynamical corrections to the BSE excitation energies.
|
||||
%\\
|
||||
%\bigskip
|
||||
%\begin{center}
|
||||
% \boxed{\includegraphics[width=0.5\linewidth]{TOC}}
|
||||
%\end{center}
|
||||
%\bigskip
|
||||
\end{abstract}
|
||||
|
||||
\maketitle
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\section{Introduction}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\section{The concept of dynamical quantities}
|
||||
@ -155,11 +174,11 @@ To do so, let us consider the usual chemical scenario where one wants to get the
|
||||
In most cases, this can be done by solving a set of linear equations of the form
|
||||
\begin{equation}
|
||||
\label{eq:lin_sys}
|
||||
\bA \bx = \omega \bx
|
||||
\bA \bc = \omega \bc
|
||||
\end{equation}
|
||||
where $\omega$ is one of the neutral excitation energies of the system associated with the transition vector $\bx$.
|
||||
If we assume that the operator $\bA$ has a matrix representation of size $K \times K$, this \textit{linear} set of equations yields $K$ excitation energies.
|
||||
However, in practice, $K$ might be very large, and it might therefore be practically useful to recast this system as two smaller coupled systems, such that
|
||||
where $\omega$ is one of the neutral excitation energies of the system associated with the transition vector $\bc$.
|
||||
If we assume that the operator $\bA$ has a matrix representation of size $N \times N$, this \textit{linear} set of equations yields $N$ excitation energies.
|
||||
However, in practice, $N$ might be very large, and it might therefore be practically useful to recast this system as two smaller coupled systems, such that
|
||||
\begin{equation}
|
||||
\label{eq:lin_sys_split}
|
||||
\begin{pmatrix}
|
||||
@ -167,32 +186,32 @@ However, in practice, $K$ might be very large, and it might therefore be practic
|
||||
\bb & \bA_2 \\
|
||||
\end{pmatrix}
|
||||
\begin{pmatrix}
|
||||
\bx_1 \\
|
||||
\bx_2 \\
|
||||
\bc_1 \\
|
||||
\bc_2 \\
|
||||
\end{pmatrix}
|
||||
= \omega
|
||||
\begin{pmatrix}
|
||||
\bx_1 \\
|
||||
\bx_2 \\
|
||||
\bc_1 \\
|
||||
\bc_2 \\
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
where the blocks $\bA_1$ and $\bA_2$, of sizes $K_1 \times K_1$ and $K_2 \times K_2$ (with $K_1 + K_2 = K$), can be associated with, for example, the single and double excitations of the system.
|
||||
where the blocks $\bA_1$ and $\bA_2$, of sizes $N_1 \times N_1$ and $N_2 \times N_2$ (with $N_1 + N_2 = N$), can be associated with, for example, the single and double excitations of the system.
|
||||
Note that this \textit{exact} decomposition does not alter, in any case, the values of the excitation energies, not their eigenvectors.
|
||||
|
||||
Solving separately each row of the system \eqref{eq:lin_sys_split} yields
|
||||
\begin{subequations}
|
||||
\begin{gather}
|
||||
\label{eq:row1}
|
||||
\bA_1 \bx_1 + \T{\bb} \bx_2 = \omega \bx_1
|
||||
\bA_1 \bc_1 + \T{\bb} \bc_2 = \omega \bc_1
|
||||
\\
|
||||
\label{eq:row2}
|
||||
\bx_2 = (\omega \bI - \bA_2)^{-1} \bb \bx_1
|
||||
\bc_2 = (\omega \bI - \bA_2)^{-1} \bb \bc_1
|
||||
\end{gather}
|
||||
\end{subequations}
|
||||
Substituting Eq.~\eqref{eq:row2} into Eq.~\eqref{eq:row1} yields the following effective \textit{non-linear}, frequency-dependent operator
|
||||
\begin{equation}
|
||||
\label{eq:non_lin_sys}
|
||||
\Tilde{\bA}_1(\omega) \bx_1 = \omega \bx_1
|
||||
\Tilde{\bA}_1(\omega) \bc_1 = \omega \bc_1
|
||||
\end{equation}
|
||||
with
|
||||
\begin{equation}
|
||||
@ -240,16 +259,16 @@ where $p = v$ or $c$,
|
||||
\label{eq:SigC}
|
||||
\SigGW{p}(\omega) = \frac{2 \ERI{pv}{vc}^2}{\omega - \e{v} + \Omega} + \frac{2 \ERI{pc}{cv}^2}{\omega - \e{c} - \Omega}
|
||||
\end{equation}
|
||||
are the correlation parts of the self-energy associated with the valence of conduction orbitals,
|
||||
are the correlation parts of the self-energy associated with wither the valence of conduction orbitals,
|
||||
\begin{equation}
|
||||
Z_{p}^{\GW} = \qty( 1 - \left. \pdv{\SigGW{p}(\omega)}{\omega} \right|_{\omega = \e{p}} )^{-1}
|
||||
\end{equation}
|
||||
is the renormalization factor, and
|
||||
\begin{equation}
|
||||
\ERI{pq}{rs} = \iint p(\br) q(\br) \frac{1}{\abs{\br - \br'}} r(\br') s(\br') d\br d\br'
|
||||
\ERI{pq}{rs} = \iint \MO{p}(\br) \MO{q}(\br) \frac{1}{\abs{\br - \br'}} \MO{r}(\br') \MO{s}(\br') d\br d\br'
|
||||
\end{equation}
|
||||
are the usual (bare) two-electron integrals.
|
||||
In Eq.~\eqref{eq:SigC}, $\Omega = \Delta\e{} + 2 \ERI{vc}{vc}$ is the sole (singlet) RPA excitation energy of the system, with $\Delta\eGW{} = \eGW{c} - \eGW{v}$.
|
||||
In Eq.~\eqref{eq:SigC}, $\Omega = \Delta\eGW{} + 2 \ERI{vc}{vc}$ is the sole (singlet) RPA excitation energy of the system, with $\Delta\eGW{} = \eGW{c} - \eGW{v}$.
|
||||
|
||||
One can now build the dynamical Bethe-Salpeter equation (dBSE) Hamiltonian, which reads
|
||||
\begin{equation} \label{eq:HBSE}
|
||||
@ -444,14 +463,12 @@ The perturbatively-corrected values are also reported, which shows that this sch
|
||||
Note that, although the BSE1(dTDA) value is further from the dBSE value than BSE1, it is quite close to the exact excitation energy.
|
||||
|
||||
%%% TABLE I %%%
|
||||
\begin{table}
|
||||
\begin{table*}
|
||||
\caption{BSE singlet and triplet excitation energies at various levels of theory.
|
||||
\label{tab:BSE}
|
||||
}
|
||||
\begin{center}
|
||||
\footnotesize
|
||||
\begin{ruledtabular}
|
||||
\begin{tabular}{|c|ccccccc|c|}
|
||||
\hline
|
||||
Singlets & BSE & BSE1 & BSE1(dTDA) & dBSE & BSE(TDA) & BSE1(TDA) & dBSE(TDA) & Exact \\
|
||||
\hline
|
||||
$\omega_1$ & 1.92778 & 1.90022 & 1.91554 & 1.90527 & 1.95137 & 1.94004 & 1.94005 & 1.92145 \\
|
||||
@ -463,10 +480,9 @@ Note that, although the BSE1(dTDA) value is further from the dBSE value than BSE
|
||||
$\omega_1$ & 1.48821 & 1.46860 & 1.46260 & 1.46636 & 1.49603 & 1.47070 & 1.47070 & 1.47085 \\
|
||||
$\omega_2$ & & & & 2.76178 & & & & \\
|
||||
$\omega_3$ & & & & 4.91545 & & & 4.91517 & \\
|
||||
\hline
|
||||
\end{tabular}
|
||||
\end{center}
|
||||
\end{table}
|
||||
\end{ruledtabular}
|
||||
\end{table*}
|
||||
%%% %%% %%% %%%
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
@ -529,48 +545,102 @@ and
|
||||
Note that the coupling blocks $B$ are frequency independent, as they should.
|
||||
This has an important consequence as this lack of frequency dependence removes one of the spurious pole.
|
||||
The singlet manifold has then the right number of excitations.
|
||||
However, one spurious triplet excitation remains.
|
||||
Numerical results for the two-level model are reported in Table \ref{tab:RPAx} with the usual approximations and perturbative treatmenets.
|
||||
However, one spurious triplet excitation remains (see Fig.~\ref{fig:dBSE2}).
|
||||
Numerical results for the two-level model are reported in Table \ref{tab:RPAx} with the usual approximations and perturbative treatments.
|
||||
In the case of dRPAx, the perturbative partitioning is simply
|
||||
\begin{equation}
|
||||
\bH^{\dRPAx}(\omega) = \underbrace{\bH^{\RPAx}}_{\bH^{(0)}} + \underbrace{\qty[ \bH^{\dRPAx}(\omega) - \bH^{\RPAx} ]}_{\bH^{(1)}}
|
||||
\end{equation}
|
||||
This might not be the smartest way of decomposing the Hamiltonian though but it seems to work fine.
|
||||
|
||||
%%% TABLE II %%%
|
||||
\begin{table}
|
||||
\begin{table*}
|
||||
\caption{RPAx singlet and triplet excitation energies at various levels of theory.
|
||||
\label{tab:RPAx}
|
||||
}
|
||||
\begin{center}
|
||||
\footnotesize
|
||||
\begin{ruledtabular}
|
||||
\begin{tabular}{|c|ccccccc|c|}
|
||||
\hline
|
||||
Singlets & RPAx & RPAx1 & RPAx1(dTDA) & dRPAx & RPAx(TDA) & RPAx1(TDA) & dRPAx(TDA) & Exact \\
|
||||
\hline
|
||||
$\omega_1$ & 1.84903 & 1.90927 & 1.90950 & 1.90362 & 1.86299 & 1.92356 & 1.92359 & 1.92145 \\
|
||||
$\omega_1$ & 1.84903 & 1.90941 & 1.90950 & 1.90362 & 1.86299 & 1.92356 & 1.92359 & 1.92145 \\
|
||||
$\omega_2$ & & & & & & & & \\
|
||||
$\omega_3$ & & & & 4.47124 & & & 4.47097 & 3.47880 \\
|
||||
\hline
|
||||
Triplets & RPAx & RPAx1 & RPAx1(dTDA) & dRPAx & RPAx(TDA) & RPAx1(TDA) & dRPAx(TDA) & Exact \\
|
||||
\hline
|
||||
$\omega_1$ & 1.38912 & 1.44267 & 1.44304 & 1.42564 & 1.40765 & 1.46154 & 1.46155 & 1.47085 \\
|
||||
$\omega_1$ & 1.38912 & 1.44285 & 1.44304 & 1.42564 & 1.40765 & 1.46154 & 1.46155 & 1.47085 \\
|
||||
$\omega_2$ & & & & & & & & \\
|
||||
$\omega_3$ & & & & 4.47797 & & & 4.47767 & \\
|
||||
\hline
|
||||
\end{tabular}
|
||||
\end{center}
|
||||
\end{table}
|
||||
\end{ruledtabular}
|
||||
\end{table*}
|
||||
%%% %%% %%% %%%
|
||||
|
||||
%%% FIGURE 3 %%%
|
||||
\begin{figure}
|
||||
\includegraphics[width=\linewidth]{dBSE2}
|
||||
\caption{
|
||||
$\det[\bH^{\dRPAx}(\omega) - \omega \bI]$ as a function of $\omega$ for both the singlet (red) and triplet (blue) manifolds.
|
||||
\label{fig:dBSE2}
|
||||
}
|
||||
\end{figure}
|
||||
%%% %%% %%% %%%
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\subsection{Sangalli's kernel}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
In Ref.~\cite{Sangalli_2011}, Sangalli proposed a norm-conserving kernel without (he claims) spurious excitations.
|
||||
For the two-level model, this kernel (based on the second RPA) reads
|
||||
In Ref.~\cite{Sangalli_2011}, Sangalli proposed a dynamical kernel (based on the second RPA) without (he claims) spurious excitations thanks to the design of a number-conserving approach which correctly describes particle indistinguishability and Pauli exclusion principle.
|
||||
We will first start by writing down explicitly this kernel as it is given in obscure physicist notations.
|
||||
|
||||
The dynamical BSE Hamiltonian with Sangalli's kernel is
|
||||
\begin{equation}
|
||||
\bH^\text{NC}(\omega) =
|
||||
\begin{pmatrix}
|
||||
H(\omega) & K(\omega)
|
||||
\\
|
||||
-K(-\omega) & -H(-\omega)
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
with
|
||||
\begin{align}
|
||||
H_{ia,jb}(\omega) & = \delta_{ij} \delta_{ab} (\eGW{a} - \eGW{i}) + \Xi_{ia,jb} (\omega)
|
||||
\\
|
||||
K_{ia,jb}(\omega) & = \Xi_{ia,bj} (\omega)
|
||||
\end{align}
|
||||
and
|
||||
\begin{gather}
|
||||
\Xi_{ia,jb} (\omega) = \sum_{m \neq n} \frac{ C_{ia,mn} C_{jb,mn} }{\omega - ( \omega_{m} + \omega_{n} + 2i\eta)}
|
||||
\\
|
||||
C_{ia,mn} = \frac{1}{2} \sum_{jb,kc} \qty{ \qty[ \ERI{ij}{kc} \delta_{ab} + \ERI{kc}{ab} \delta_{ij} ] \qty[ R_{m,jc} R_{n,kb}
|
||||
+ R_{m,kb} R_{n,jc} ] }
|
||||
\end{gather}
|
||||
where $R_{m,ia}$ are the elements of the RPA eigenvectors.
|
||||
Here $i$, $j$, and $k$ are occupied orbitals, $a$, $b$, and $c$ are unoccupied orbitals, and $m$ and $n$ label single excitations.
|
||||
|
||||
For the two-level model, Sangalli's kernel reads
|
||||
\begin{align}
|
||||
H(\omega) & = \Delta\eGW{} + \Xi_H (\omega)
|
||||
\\
|
||||
K(\omega) & = \Xi_K (\omega)
|
||||
\end{align}
|
||||
|
||||
\begin{align}
|
||||
\Xi_H (\omega) = \sum_{m \neq n} \frac{ C_{ia,mn} C_{jb,mn} }{\omega - ( \omega_{m} + \omega_{n} + 2i\eta)}
|
||||
\end{align}
|
||||
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\section{Take-home messages}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
What have we learnt here?
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\acknowledgements{
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
The author thanks the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (Grant agreement No.~863481) for financial support.
|
||||
|
||||
|
||||
% BIBLIOGRAPHY
|
||||
\bibliographystyle{unsrt}
|
||||
\bibliography{../BSEdyn}
|
||||
|
||||
\end{document}
|
||||
|
Binary file not shown.
BIN
Notes/dBSE.pdf
BIN
Notes/dBSE.pdf
Binary file not shown.
Loading…
Reference in New Issue
Block a user