minor corrections up to Sec III
This commit is contained in:
parent
8830045aaf
commit
ef4e33e878
36
dynker.bib
36
dynker.bib
@ -1,13 +1,24 @@
|
||||
%% This BibTeX bibliography file was created using BibDesk.
|
||||
%% http://bibdesk.sourceforge.net/
|
||||
|
||||
%% Created for Pierre-Francois Loos at 2020-06-26 09:45:06 +0200
|
||||
%% Created for Pierre-Francois Loos at 2020-07-20 08:52:20 +0200
|
||||
|
||||
|
||||
%% Saved with string encoding Unicode (UTF-8)
|
||||
|
||||
|
||||
|
||||
@article{Lowdin_1963,
|
||||
Author = {P. L{\"o}wdin},
|
||||
Date-Added = {2020-07-20 08:49:22 +0200},
|
||||
Date-Modified = {2020-07-20 08:52:15 +0200},
|
||||
Doi = {10.1016/0022-2852(63)90151-6},
|
||||
Journal = {J. Mol. Spectrosc.},
|
||||
Pages = {12--33},
|
||||
Title = {Studies in perturbation theory: Part I. An elementary iteration-variation procedure for solving the Schr{\"o}dinger equation by partitioning technique},
|
||||
Volume = {10},
|
||||
Year = {1963}}
|
||||
|
||||
@article{Petersilka_1996,
|
||||
Author = {M. Petersilka and U. J. Gossmann and and E. K. U. Gross},
|
||||
Date-Added = {2020-06-26 09:43:33 +0200},
|
||||
@ -17,7 +28,8 @@
|
||||
Pages = {1212},
|
||||
Title = {Excitation Energies From Time-Dependent Density-Functional Theory},
|
||||
Volume = {76},
|
||||
Year = {1996}}
|
||||
Year = {1996},
|
||||
Bdsk-Url-1 = {https://doi.org/10.1103/PhysRevLett.76.1212}}
|
||||
|
||||
@article{Nielsen_1980,
|
||||
Author = {Egon S. Nielsen and Poul Jorgensen},
|
||||
@ -14426,14 +14438,12 @@
|
||||
Bdsk-Url-2 = {https://doi.org/10.1103/PhysRevB.93.235113}}
|
||||
|
||||
@article{Boulanger_2014,
|
||||
author = {Boulanger, Paul and Jacquemin, Denis and Duchemin, Ivan and Blase, Xavier},
|
||||
title = {Fast and Accurate Electronic Excitations in Cyanines with the Many-Body Bethe-Salpeter Approach},
|
||||
journal = {J. Chem. Theory Comput.},
|
||||
volume = {10},
|
||||
number = {3},
|
||||
pages = {1212--1218},
|
||||
year = {2014},
|
||||
doi = {10.1021/ct401101u},
|
||||
}
|
||||
|
||||
|
||||
Author = {Boulanger, Paul and Jacquemin, Denis and Duchemin, Ivan and Blase, Xavier},
|
||||
Doi = {10.1021/ct401101u},
|
||||
Journal = {J. Chem. Theory Comput.},
|
||||
Number = {3},
|
||||
Pages = {1212--1218},
|
||||
Title = {Fast and Accurate Electronic Excitations in Cyanines with the Many-Body Bethe-Salpeter Approach},
|
||||
Volume = {10},
|
||||
Year = {2014},
|
||||
Bdsk-Url-1 = {https://doi.org/10.1021/ct401101u}}
|
||||
|
35
dynker.tex
35
dynker.tex
@ -41,26 +41,26 @@ In particular, using a simple two-level model, we analyze, for each kernel, the
|
||||
\section{Linear response theory}
|
||||
\label{sec:LR}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
Linear response theory is a powerful approach that allows to directly access the optical excitations $\omega_s$ of a given electronic system (such as a molecule) and their corresponding oscillator strengths $f_s$ [extracted from their eigenvectors $\T{(\bX_s \bY_s)}$] via the response of the system to a weak electromagnetic field. \cite{Oddershede_1977,Casida_1995,Petersilka_1996}
|
||||
Linear response theory is a powerful approach that allows to directly access the optical excitations $\omega_S$ of a given electronic system (such as a molecule) and their corresponding oscillator strengths $f_s$ [extracted from their eigenvectors $\T{(\bX_S \bY_S)}$] via the response of the system to a weak electromagnetic field. \cite{Oddershede_1977,Casida_1995,Petersilka_1996}
|
||||
From a practical point of view, these quantities are obtained by solving non-linear, frequency-dependent Casida-like equations in the space of single excitations and de-excitations \cite{Casida_1995}
|
||||
\begin{equation} \label{eq:LR}
|
||||
\begin{pmatrix}
|
||||
\bR^{\sigma}(\omega_s) & \bC^{\sigma}(\omega_s)
|
||||
\bR^{\sigma}(\omega_S) & \bC^{\sigma}(\omega_S)
|
||||
\\
|
||||
-\bC^{\sigma}(-\omega_s)^* & -\bR^{\sigma}(-\omega_s)^*
|
||||
-\bC^{\sigma}(-\omega_S)^* & -\bR^{\sigma}(-\omega_S)^*
|
||||
\end{pmatrix}
|
||||
\cdot
|
||||
\begin{pmatrix}
|
||||
\bX_s^{\sigma}
|
||||
\bX_S^{\sigma}
|
||||
\\
|
||||
\bY_s^{\sigma}
|
||||
\bY_S^{\sigma}
|
||||
\end{pmatrix}
|
||||
=
|
||||
\omega_s
|
||||
\omega_S
|
||||
\begin{pmatrix}
|
||||
\bX_s^{\sigma}
|
||||
\bX_S^{\sigma}
|
||||
\\
|
||||
\bY_s^{\sigma}
|
||||
\bY_S^{\sigma}
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
where the explicit expressions of the resonant and coupling blocks, $\bR^{\sigma}(\omega)$ and $\bC^{\sigma}(\omega)$, depend on the spin manifold ($\sigma =$ $\updw$ for singlets and $\sigma =$ $\upup$ for triplets) and the level of approximation that one employs.
|
||||
@ -68,7 +68,7 @@ Neglecting the coupling block [\ie, $\bC^{\sigma}(\omega) = 0$] between the reso
|
||||
In the absence of symmetry breaking, \cite{Dreuw_2005} the non-linear eigenvalue problem defined in Eq.~\eqref{eq:LR} has particle-hole symmetry which means that it is invariant via the transformation $\omega \to -\omega$.
|
||||
Therefore, without loss of generality, we will restrict our analysis to positive frequencies.
|
||||
|
||||
In the one-electron basis of (real) spatial orbitals $\lbrace \MO{p} \rbrace$, we will assume that the elements of the matrices defined in Eq.~\eqref{eq:LR} have the following generic forms: \cite{Dreuw_2005}
|
||||
In the one-electron basis of (real) spatial orbitals $\lbrace \MO{p}(\br) \rbrace$, we will assume that the elements of the matrices defined in Eq.~\eqref{eq:LR} have the following generic forms: \cite{Dreuw_2005}
|
||||
\begin{subequations}
|
||||
\begin{gather}
|
||||
R_{ia,jb}^{\sigma}(\omega) = (\e{a} - \e{i}) \delta_{ij} \delta_{ab} + f_{ia,jb}^{\Hxc,\sigma}(\omega)
|
||||
@ -76,7 +76,7 @@ In the one-electron basis of (real) spatial orbitals $\lbrace \MO{p} \rbrace$, w
|
||||
C_{ia,jb}^{\sigma}(\omega) = f_{ia,bj}^{\Hxc,\sigma}(\omega)
|
||||
\end{gather}
|
||||
\end{subequations}
|
||||
where $\delta_{pq}$ is the Kronecker delta, $\e{p}$ is the one-electron energy associated with $\MO{p}$, and
|
||||
where $\delta_{pq}$ is the Kronecker delta, $\e{p}$ is the one-electron (or quasiparticle) energy associated with $\MO{p}(\br)$, and
|
||||
\begin{equation} \label{eq:kernel}
|
||||
f_{ia,jb}^{\Hxc,\sigma}(\omega)
|
||||
= \iint \MO{i}(\br) \MO{a}(\br) f^{\Hxc,\sigma}(\omega) \MO{j}(\br') \MO{b}(\br') d\br d\br'
|
||||
@ -97,7 +97,7 @@ where $\sigma = 1 $ or $0$ for singlet and triplet excited states (respectively)
|
||||
\ERI{ia}{jb} = \iint \MO{i}(\br) \MO{a}(\br) \frac{1}{\abs{\br - \br'}} \MO{j}(\br') \MO{b}(\br') d\br d\br'
|
||||
\end{equation}
|
||||
are the usual two-electron integrals.
|
||||
The central point here is that, thanks to its non-linear nature stemming from their frequency dependence, a dynamical kernel potentially generates more than just single excitations.
|
||||
The launchpad of the present study is that, thanks to its non-linear nature stemming from its frequency dependence, a dynamical kernel potentially generates more than just single excitations.
|
||||
Unless otherwise stated, atomic units are used and we assume real quantities throughout this manuscript.
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
@ -110,10 +110,10 @@ 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 \cdot \bc = \omega \, \bc
|
||||
\bA \cdot \bc = \omega_S \, \bc
|
||||
\end{equation}
|
||||
where $\omega$ is one of the optical excitation energies of interest and $\bc$ its transition vector .
|
||||
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.
|
||||
If we assume that the matrix $\bA$ is diagonalizable and of size $N \times N$, the \textit{linear} set of equations \eqref{eq:lin_sys} yields $N$ excitation energies.
|
||||
However, in practice, $N$ might be (very) large (\eg, equal to the total number of single and double excitations generated from a reference Slater determinant), 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}
|
||||
@ -133,9 +133,9 @@ However, in practice, $N$ might be (very) large (\eg, equal to the total number
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
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.
|
||||
This decomposition technique is often called L\"owdin partitioning in the literature. \cite{Lowdin_1963}
|
||||
|
||||
Solving separately each row of the system \eqref{eq:lin_sys_split} and assuming that $\omega \bI - \bA_2$ is invertible, it follows that
|
||||
Solving separately each row of the system \eqref{eq:lin_sys_split} and assuming that $\omega \bI - \bA_2$ is invertible, we get
|
||||
\begin{subequations}
|
||||
\begin{gather}
|
||||
\label{eq:row1}
|
||||
@ -156,10 +156,11 @@ with
|
||||
\end{equation}
|
||||
which has, by construction, exactly the same solutions as the linear system \eqref{eq:lin_sys} but a smaller dimension.
|
||||
For example, an operator $\Tilde{\bA}_1(\omega)$ built in the single-excitation basis can potentially provide excitation energies for double excitations thanks to its frequency-dependent nature, the information from the double excitations being ``folded'' into $\Tilde{\bA}_1(\omega)$ via Eq.~\eqref{eq:row2}. \cite{ReiningBook}
|
||||
Note that this \textit{exact} decomposition does not alter, in any case, the values of the excitation energies.
|
||||
|
||||
How have we been able to reduce the dimension of the problem while keeping the same number of solutions?
|
||||
To do so, we have transformed a linear operator $\bA$ into a non-linear operator $\Tilde{\bA}_1(\omega)$ by making it frequency dependent.
|
||||
In other words, we have sacrificed the linearity of the system in order to obtain a new, non-linear systems of equations of smaller dimension.
|
||||
In other words, we have sacrificed the linearity of the system in order to obtain a new, non-linear systems of equations of smaller dimension [see Eq.~\eqref{eq:non_lin_sys}].
|
||||
This procedure converting degrees of freedom into frequency or energy dependence is very general and can be applied in various contexts. \cite{Sottile_2003,Garniron_2018,QP2}
|
||||
Thanks to its non-linearity, Eq.~\eqref{eq:non_lin_sys} can produce more solutions than its actual dimension.
|
||||
However, because there is no free lunch, this non-linear system is obviously harder to solve than its corresponding linear analog given by Eq.~\eqref{eq:lin_sys}.
|
||||
@ -169,7 +170,7 @@ For example, assuming that $\bA_2$ is a diagonal matrix is of common practice (s
|
||||
Another of these approximations is the so-called \textit{static} approximation, where one sets the frequency to a particular value.
|
||||
For example, as commonly done within the Bethe-Salpeter equation (BSE) formalism of many-body perturbation theory (MBPT), \cite{Strinati_1988} $\Tilde{\bA}_1(\omega) = \Tilde{\bA}_1 \equiv \Tilde{\bA}_1(\omega = 0)$.
|
||||
In such a way, the operator $\Tilde{\bA}_1$ is made linear again by removing its frequency-dependent nature.
|
||||
A similar example in the context of time-dependent density-functional theory (TDDFT) \cite{Runge_1984} is provided by the ubiquitous adiabatic approximation, \cite{Tozer_2000} which neglects all memory effects by making static the exchange-correlation (xc) kernel (\ie, frequency-independent). \cite{Maitra_2016}
|
||||
A similar example in the context of time-dependent density-functional theory (TDDFT) \cite{Runge_1984} is provided by the ubiquitous adiabatic approximation, \cite{Tozer_2000} which neglects all memory effects by making static the exchange-correlation (xc) kernel (\ie, frequency independent). \cite{Maitra_2016}
|
||||
These approximations come with a heavy price as the number of solutions provided by the system of equations \eqref{eq:non_lin_sys} has now been reduced from $N$ to $N_1$.
|
||||
Coming back to our example, in the static (or adiabatic) approximation, the operator $\Tilde{\bA}_1$ built in the single-excitation basis cannot provide double excitations anymore, and the $N_1$ excitation energies are associated with single excitations.
|
||||
All additional solutions associated with higher excitations have been forever lost.
|
||||
|
Loading…
Reference in New Issue
Block a user