saving work on Notes
This commit is contained in:
parent
c0a901619e
commit
17eaacefc9
28
BSEdyn.bib
28
BSEdyn.bib
@ -1,7 +1,7 @@
|
||||
%% This BibTeX bibliography file was created using BibDesk.
|
||||
%% http://bibdesk.sourceforge.net/
|
||||
|
||||
%% Created for Pierre-Francois Loos at 2020-06-19 15:03:36 +0200
|
||||
%% Created for Pierre-Francois Loos at 2020-06-19 21:54:18 +0200
|
||||
|
||||
|
||||
%% Saved with string encoding Unicode (UTF-8)
|
||||
@ -1242,15 +1242,10 @@
|
||||
@article{Huix-Rotllant_2011,
|
||||
Author = {{Huix-Rotllant}, Miquel and Ipatov, Andrei and Rubio, Angel and Casida, Mark E.},
|
||||
Date-Added = {2020-05-18 21:40:28 +0200},
|
||||
Date-Modified = {2020-05-18 21:40:28 +0200},
|
||||
Date-Modified = {2020-06-19 21:34:49 +0200},
|
||||
Doi = {10.1016/j.chemphys.2011.03.019},
|
||||
File = {/Users/loos/Zotero/storage/A4JUV4M4/Huix-Rotllant et al. - 2011 - Assessment of dressed time-dependent density-funct.pdf},
|
||||
Issn = {03010104},
|
||||
Journal = {Chem. Phys.},
|
||||
Language = {en},
|
||||
Month = nov,
|
||||
Number = {1},
|
||||
Pages = {120-129},
|
||||
Pages = {120--129},
|
||||
Title = {Assessment of Dressed Time-Dependent Density-Functional Theory for the Low-Lying Valence States of 28 Organic Chromophores},
|
||||
Volume = {391},
|
||||
Year = {2011},
|
||||
@ -3539,14 +3534,10 @@
|
||||
@article{Boggio-Pasqua_2007,
|
||||
Author = {{Boggio-Pasqua}, Martial and Bearpark, Michael J. and Robb, Michael A.},
|
||||
Date-Added = {2020-01-01 21:36:51 +0100},
|
||||
Date-Modified = {2020-01-01 21:36:51 +0100},
|
||||
Date-Modified = {2020-06-19 21:54:14 +0200},
|
||||
Doi = {10.1021/jo070452v},
|
||||
Issn = {0022-3263, 1520-6904},
|
||||
Journal = {J. Org. Chem.},
|
||||
Language = {en},
|
||||
Month = jun,
|
||||
Number = {12},
|
||||
Pages = {4497-4503},
|
||||
Pages = {4497--4503},
|
||||
Title = {Toward a {{Mechanistic Understanding}} of the {{Photochromism}} of {{Dimethyldihydropyrenes}}},
|
||||
Volume = {72},
|
||||
Year = {2007},
|
||||
@ -4149,15 +4140,10 @@
|
||||
@article{Cave_2004,
|
||||
Author = {Cave, Robert J. and Zhang, Fan and Maitra, Neepa T. and Burke, Kieron},
|
||||
Date-Added = {2020-01-01 21:36:51 +0100},
|
||||
Date-Modified = {2020-01-01 21:36:51 +0100},
|
||||
Date-Modified = {2020-06-19 21:18:58 +0200},
|
||||
Doi = {10.1016/j.cplett.2004.03.051},
|
||||
File = {/Users/loos/Zotero/storage/6L9X6HT4/Cave et al. - 2004 - A dressed TDDFT treatment of the 21Ag states of bu.pdf},
|
||||
Issn = {00092614},
|
||||
Journal = {Chem. Phys. Lett.},
|
||||
Language = {en},
|
||||
Month = may,
|
||||
Number = {1-3},
|
||||
Pages = {39-42},
|
||||
Pages = {39--42},
|
||||
Title = {A Dressed {{TDDFT}} Treatment of the {{21Ag}} States of Butadiene and Hexatriene},
|
||||
Volume = {389},
|
||||
Year = {2004},
|
||||
|
@ -22,11 +22,11 @@
|
||||
\affiliation{\LCPQ}
|
||||
|
||||
\begin{abstract}
|
||||
We discuss the physical properties and accuracy of three distinct dynamical (\ie, frequency-dependent) kernels for the computation of excitation energies within linear response theory:
|
||||
i) an \textit{a priori} built kernel inspired by the dressed time-dependent density-functional theory (TD-DFT) kernel proposed by Maitra and coworkers [\href{https://doi.org/10.1063/1.1651060}{J.~Chem.~Phys.~120, 5932 (2004)}],
|
||||
We discuss the physical properties and accuracy of three distinct dynamical (\ie, frequency-dependent) kernels for the computation of optical excitations within linear response theory:
|
||||
i) an \textit{a priori} built kernel inspired by the dressed time-dependent density-functional theory (TDDFT) kernel proposed by Maitra and coworkers [\href{https://doi.org/10.1063/1.1651060}{J.~Chem.~Phys.~120, 5932 (2004)}],
|
||||
ii) the dynamical kernel stemming from the Bethe-Salpeter equation (BSE) formalism derived originally by Strinati [\href{https://doi.org/10.1007/BF02725962}{Riv.~Nuovo Cimento 11, 1--86 (1988)}], and
|
||||
iii) the second-order BSE kernel derived by Yang and coworkers [\href{https://doi.org/10.1063/1.4824907}{J.~Chem.~Phys.~139, 154109 (2013)}].
|
||||
In particular, using a simple two-level model, we analyze the appearance of spurious excitations, as first evidenced by Romaniello and collaborators [\href{https://doi.org/10.1063/1.3065669}{J.~Chem.~Phys.~130, 044108 (2009)}], due to the approximate nature of the kernels.
|
||||
In particular, using a simple two-level model, we analyze, for each kernel, the appearance of spurious excitations, as first evidenced by Romaniello and collaborators [\href{https://doi.org/10.1063/1.3065669}{J.~Chem.~Phys.~130, 044108 (2009)}], due to the approximate nature of the kernels.
|
||||
%\\
|
||||
%\bigskip
|
||||
%\begin{center}
|
||||
@ -39,8 +39,9 @@ In particular, using a simple two-level model, we analyze the appearance of spur
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\section{Linear response theory}
|
||||
\label{sec:LR}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
Linear response 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{Casida_1995}
|
||||
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{Casida_1995}
|
||||
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}
|
||||
@ -62,22 +63,40 @@ From a practical point of view, these quantities are obtained by solving non-lin
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
where the explicit expressions of the resonant and coupling blocks, $\bR(\omega)$ and $\bC(\omega)$, depend on the level of approximation that one employs.
|
||||
Neglecting the coupling block between the resonant and anti-resonants parts, $\bR(\omega)$ and $-\bR(-\omega_s)$, is known as the Tamm-Dancoff approximation (TDA).
|
||||
Neglecting the coupling block between the resonant and anti-resonants parts, $\bR(\omega)$ and $-\bR(-\omega)$, is known as the Tamm-Dancoff approximation (TDA).
|
||||
The non-linear eigenvalue problem defined in Eq.~\eqref{eq:LR} has particle-hole symmetry which means that it is invariant via the transformation $\omega \to -\omega$, and, thanks to its non-linear nature stemming from its frequency dependence, it potentially generates more than just single excitations.
|
||||
|
||||
In a wave function context, introducing a spatial orbital basis $\lbrace \MO{p} \rbrace$, we assume here that the elements of the matrices defined in Eq.~\eqref{eq:LR} read
|
||||
In a wave function context, introducing a spatial orbital basis $\lbrace \MO{p} \rbrace$, we assume here that the elements of the matrices defined in Eq.~\eqref{eq:LR} have the following generic form:
|
||||
\begin{subequations}
|
||||
\begin{gather}
|
||||
R_{ia,jb}(\omega) = (\e{a} - \e{i}) \delta_{ij} \delta_{ab} + 2 \sigma \ERI{ia}{jb} - \ERI{ib}{ja} + f_{ia,jb}^\sigma(\omega)
|
||||
\begin{split}
|
||||
R_{ia,jb}(\omega)
|
||||
& = \iint \MO{i}(\br) \MO{a}(\br) \bR(\omega) \MO{j}(\br') \MO{b}(\br') d\br d\br'
|
||||
\\
|
||||
& = (\e{a} - \e{i}) \delta_{ij} \delta_{ab} + 2 \sigma \ERI{ia}{jb} - \ERI{ib}{ja} + f_{ia,jb}^\sigma(\omega)
|
||||
\end{split}
|
||||
\\
|
||||
C_{ia,jb}(\omega) = 2 \sigma \ERI{ia}{bj} - \ERI{ij}{ba} + f_{ia,bj}^\sigma(\omega)
|
||||
\begin{split}
|
||||
C_{ia,jb}(\omega)
|
||||
& = \iint \MO{i}(\br) \MO{a}(\br) \bC(\omega) \MO{j}(\br') \MO{b}(\br') d\br d\br'
|
||||
\\
|
||||
& = 2 \sigma \ERI{ia}{bj} - \ERI{ij}{ba} + f_{ia,bj}^\sigma(\omega)
|
||||
\end{split}
|
||||
\end{gather}
|
||||
where $\sigma = 1 $ or $0$ for singlet ($\updw$) and triplet ($\upup$) excited states (respectively), $i$ and $j$ are occupied orbitals, $a$ and $b$ are unoccupied orbitals, and $f(\omega)^{\sigma}$ is the correlation part of the spin-resolved kernel.
|
||||
\end{subequations}
|
||||
where $\sigma = 1 $ or $0$ for singlet ($\updw$) and triplet ($\upup$) excited states (respectively), and
|
||||
\begin{equation}
|
||||
\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 (bare) two-electron integrals.
|
||||
Here, $i$ and $j$ are occupied orbitals, $a$ and $b$ are unoccupied orbitals, and $f^{\sigma}(\omega)$ is the correlation part of the spin-resolved kernel.
|
||||
(Note that, usually, only the correlation part of the kernel is frequency dependent.)
|
||||
In the case of a spin-independent kernel, we will drop the superscrit $\sigma$.
|
||||
Unless otherwise stated, atomic units are used and we assume real quantities throughout this manuscript.
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\section{The concept of dynamical quantities}
|
||||
\label{sec:dyn}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%s
|
||||
As a chemist, it is maybe difficult to understand the concept of dynamical properties, the motivation behind their introduction, and their actual usefulness.
|
||||
Here, we will try to give a pedagogical example showing the importance of dynamical quantities and their main purposes. \cite{Romaniello_2009b,Sangalli_2011,ReiningBook}
|
||||
@ -89,7 +108,7 @@ In most cases, this can be done by solving a set of linear equations of the form
|
||||
\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.
|
||||
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
|
||||
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}
|
||||
\begin{pmatrix}
|
||||
@ -129,28 +148,34 @@ with
|
||||
\Tilde{\bA}_1(\omega) = \bA_1 + \T{\bb} (\omega \bI - \bA_2)^{-1} \bb
|
||||
\end{equation}
|
||||
which has, by construction, exactly the same solutions than the linear system \eqref{eq:lin_sys} but a smaller dimension.
|
||||
For example, an operator $\Tilde{\bA}_1(\omega)$ built in the basis of single excitations 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}
|
||||
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}
|
||||
|
||||
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.
|
||||
This procedure converting degrees of freedom into frequency or energy dependence is very general and can be applied in various contexts. \cite{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 usually no free lunch, this non-linear system is obviously harder to solve than its corresponding linear analog given by Eq.~\eqref{eq:lin_sys}.
|
||||
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}.
|
||||
Nonetheless, approximations can be now applied to Eq.~\eqref{eq:non_lin_sys} in order to solve it efficiently.
|
||||
For example, assuming that $\bA_2$ is a diagonal matrix is of common practice (see, for example, Ref.~\onlinecite{Garniron_2018} and references therein).
|
||||
|
||||
Another of these approximations is the so-called \textit{static} approximation, which corresponds to fixing the frequency to a particular value.
|
||||
For example, as commonly done within the Bethe-Salpeter equation (BSE) formalism, \cite{Strinati_1988} $\Tilde{\bA}_1(\omega) = \Tilde{\bA}_1 \equiv \Tilde{\bA}_1(\omega = 0)$.
|
||||
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 (TD-DFT) \cite{Runge_1984} is provided by the ubiquitous adiabatic approximation, \cite{Tozer_2000} which neglects all memory effects by making the exchange-correlation (xc) kernel static (\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 basis of single excitations cannot provide double excitations anymore, and the only $N_1$ excitation energies are associated with single excitations.
|
||||
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.
|
||||
In the next section, we illustrate these concepts and the various levels of approximation that can be used to recover some of these dynamical effects.
|
||||
In the next section, we illustrate these concepts and the various tricks that can be used to recover some of these dynamical effects starting from the static eigenproblem.
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\section{A two-level model}
|
||||
\section{Dynamical kernels}
|
||||
\label{sec:kernel}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\subsection{Exact Hamiltonian}
|
||||
\label{sec:exact}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
|
||||
Let us consider a two-level quantum system made of two orbitals \cite{Romaniello_2009b} in its singlet ground state (\ie, the lowest orbital is doubly occupied).
|
||||
@ -161,7 +186,7 @@ There is then only one single excitation possible which corresponds to the trans
|
||||
As usual, this can produce a singlet singly-excited state $\ket{S} = (\ket{v\bar{c}} + \ket{c\bar{v}})/\sqrt{2}$, and a triplet singly-excited state $\ket{T} = (\ket{v\bar{c}} - \ket{c\bar{v}})/\sqrt{2}$. \cite{SzaboBook}
|
||||
|
||||
For the singlet manifold, the exact Hamiltonian in the basis of the (spin-adapted) configuration state functions reads
|
||||
\begin{equation}
|
||||
\begin{equation} \label{eq:H-exact}
|
||||
\bH^{\updw} =
|
||||
\begin{pmatrix}
|
||||
\mel{0}{\hH}{0} & \mel{0}{\hH}{S} & \mel{0}{\hH}{D} \\
|
||||
@ -170,32 +195,110 @@ For the singlet manifold, the exact Hamiltonian in the basis of the (spin-adapte
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
with
|
||||
\begin{gather}
|
||||
\mel{0}{\hH}{0} \equiv \EHF = 2\e{v} - \ERI{vv}{vv}
|
||||
\begin{subequations}
|
||||
\begin{align}
|
||||
\mel{0}{\hH}{0} & = 2\e{v} - \ERI{vv}{vv} = \EHF
|
||||
\\
|
||||
\mel{1}{\hH - \EHF}{1} = \Delta\e{} + \ERI{vc}{cv} - \ERI{vv}{cc}
|
||||
\mel{S}{\hH - \EHF}{S} & = \Delta\e{} + \ERI{vc}{cv} - \ERI{vv}{cc}
|
||||
\\
|
||||
\mel{1}{\hH - \EHF}{1} = 2\Delta\e{} + \ERI{vv}{vv} + \ERI{cc}{cc} + 2\ERI{vc}{cv} - 4\ERI{vv}{cc}
|
||||
\begin{split}
|
||||
\mel{D}{\hH - \EHF}{D}
|
||||
& = 2\Delta\e{} + \ERI{vv}{vv} + \ERI{cc}{cc}
|
||||
\\
|
||||
& + 2\ERI{vc}{cv} - 4\ERI{vv}{cc}
|
||||
\end{split}
|
||||
\\
|
||||
\mel{0}{\hH - \EHF}{1} = 0
|
||||
\mel{0}{\hH}{S} & = 0
|
||||
\\
|
||||
\mel{1}{\hH - \EHF}{2} = \sqrt{2}[\ERI{vc}{cc} - \ERI{cv}{vv}]
|
||||
\mel{S}{\hH}{D} & = \sqrt{2}[\ERI{vc}{cc} - \ERI{cv}{vv}]
|
||||
\\
|
||||
\mel{0}{\hH - \EHF}{2} = \ERI{vc}{cv}
|
||||
\end{gather}
|
||||
\mel{0}{\hH}{D} & = \ERI{vc}{cv}
|
||||
\end{align}
|
||||
\end{subequations}
|
||||
and $\Delta\e{} = \e{c} - \e{v}$.
|
||||
The energy of the only triplet state is simply $\mel{T}{\hH}{T} = \EHF + \Delta\e{} - \ERI{vv}{cc}$.
|
||||
|
||||
For the sake of illustration, we will use the same numerical example throughout this study, and consider the singlet ground state of the \ce{He} atom in Pople's 6-31G basis set.
|
||||
This system contains two orbitals and the numerical values of the various quantities defined above are
|
||||
\begin{subequations}
|
||||
\begin{align}
|
||||
\e{v} & = -0.914\,127
|
||||
&
|
||||
\e{c} & = + 1.399\,859
|
||||
\\
|
||||
\ERI{vv}{vv} & = 1.026\,907
|
||||
&
|
||||
\ERI{cc}{cc} & = 0.766\,363
|
||||
\\
|
||||
\ERI{vv}{cc} & = 0.858\,133
|
||||
&
|
||||
\ERI{vc}{cv} & = 0.227\,670
|
||||
\\
|
||||
\ERI{vv}{vc} & = 0.316\,490
|
||||
&
|
||||
\ERI{vc}{cc} & = 0.255\,554
|
||||
\end{align}
|
||||
\end{subequations}
|
||||
This yields the following exact singlet and triplet excitation energies
|
||||
\begin{align} \label{sec:exact}
|
||||
\omega_{1}^{\updw} & = 1.92145
|
||||
&
|
||||
\omega_{3}^{\updw} & = 3.47880
|
||||
&
|
||||
\omega_{1}^{\upup} & = 1.47085
|
||||
\end{align}
|
||||
that we are going to use a reference for the remaining of this study.
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\subsection{Maitra's kernel}
|
||||
\subsection{Maitra's dynamical kernel}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
The kernel proposed by Maitra in the context of dressed TD-DFT corresponds to a static kernel to which
|
||||
where a frequency-dependent kernel is build \textit{a priori} and manually for a particular excitation
|
||||
The kernel proposed by Maitra and coworkers \cite{Maitra_2004,Cave_2004} in the context of dressed TDDFT (D-TDDFT) corresponds to an \textit{ad hoc} many-body theory correction to TDDFT.
|
||||
More specifically, D-TDDFT adds manually to the static kernel a frequency-dependent part by reverse-engineering the exact Hamiltonian \eqref{eq:H-exact}.
|
||||
The very same idea was taking further by Huix-Rotllant, Casida and coworkers. \cite{Huix-Rotllant_2011}
|
||||
For the singlet states, we have
|
||||
\begin{equation} \label{eq:f-Maitra}
|
||||
f_M^{\updw}(\omega) = \frac{\abs*{\mel{S}{\hH}{D}}^2}{\omega - (\mel{D}{\hH}{D} - \mel{0}{\hH}{0}) }
|
||||
\end{equation}
|
||||
while $f_M^{\upup}(\omega) = 0$.
|
||||
The expression \eqref{eq:f-Maitra} can be easily obtained by folding the double excitation onto the single excitation starting from the Hamiltonian \eqref{eq:H-exact}, as explained in Sec.~\ref{sec:dyn}.
|
||||
It is clear that one must know \textit{a priori} the structure of the Hamiltonian to construct such dynamical kernel, and this obviously hampers its applicability to realistic photochemical systems where it is sometimes hard to get a clear picture of the interplay between excited states. \cite{Boggio-Pasqua_2007}
|
||||
|
||||
For the two-level model, the non-linear equations defined in Eq.~\eqref{eq:LR} provides the following effective Hamiltonian
|
||||
\begin{equation} \label{eq:H-M}
|
||||
\bH_{M}(\omega) =
|
||||
\begin{pmatrix}
|
||||
R_M(\omega) & C_M(\omega)
|
||||
\\
|
||||
-C_M(-\omega) & -R_M(-\omega)
|
||||
\end{pmatrix}
|
||||
\end{equation}
|
||||
with
|
||||
\begin{subequations}
|
||||
\begin{gather}
|
||||
R_M(\omega) = \Delta\e{} + 2 \sigma \ERI{vc}{vc} - \ERI{vc}{vc} + f_M^{\sigma}(\omega)
|
||||
\\
|
||||
C_M(\omega) = 2 \sigma \ERI{vc}{cv} - \ERI{vv}{cc} + f_M^{\sigma}(\omega)
|
||||
\end{gather}
|
||||
\end{subequations}
|
||||
which provides the following excitation energies when diagonalized:
|
||||
\begin{align} \label{sec:M}
|
||||
\omega_{1}^{\updw} & = 1.89314
|
||||
&
|
||||
\omega_{3}^{\updw} & = 3.44865
|
||||
&
|
||||
\omega_{1}^{\upup} & = 1.43794
|
||||
\end{align}
|
||||
Although not particularly accurate, this kernel provides exactly the right number of solutions (2 singlets and 1 triplet).
|
||||
Its accuracy could be certainly improved in a DFT context.
|
||||
However, this is not the point of the present investigation.
|
||||
Because $f_M^{\upup}(\omega) = 0$, the triplet excitation energy is equivalent to the TDHF excitation energy.
|
||||
In the static approximation where $f_M^{\updw}(\omega) = 0$, the singlet excitations are also TDHF excitation energies.
|
||||
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
\subsection{Dynamical BSE kernel}
|
||||
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
|
||||
|
||||
Within many-body perturbation theory (MBPT), one can easily compute the quasiparticle energies associated with the valence and conduction orbitals.
|
||||
Within MBPT, one can easily compute the quasiparticle energies associated with the valence and conduction orbitals.
|
||||
Assuming that the dynamically-screened Coulomb potential has been calculated at the random-phase approximation (RPA) level and within the Tamm-Dancoff approximation (TDA), the expression of the $\GW$ quasiparticle energy is
|
||||
\begin{equation}
|
||||
\e{p}^{\GW} = \e{p} + Z_{p}^{\GW} \SigGW{p}(\e{p})
|
||||
@ -209,11 +312,7 @@ are the correlation parts of the self-energy associated with wither the valence
|
||||
\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 \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.
|
||||
is the renormalization factor.
|
||||
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
|
||||
@ -298,25 +397,7 @@ are the eigenvectors of $\bH^{\BSE}$, and
|
||||
\end{equation}
|
||||
This corresponds to a dynamical correction to the static excitations, and the TDA can be applied to the dynamical correction, a scheme we label as dTDA in the following.
|
||||
|
||||
We now take a numerical example by considering the singlet ground state of the \ce{He} atom in the 6-31G basis set.
|
||||
This system contains two orbitals and the numerical values of the various quantities defined above are
|
||||
\begin{align}
|
||||
\e{v} & = -0.914\,127
|
||||
&
|
||||
\e{c} & = + 1.399\,859
|
||||
\\
|
||||
\ERI{vv}{vv} & = 1.026\,907
|
||||
&
|
||||
\ERI{cc}{cc} & = 0.766\,363
|
||||
\\
|
||||
\ERI{vv}{cc} & = 0.858\,133
|
||||
&
|
||||
\ERI{vc}{cv} & = 0.227\,670
|
||||
\\
|
||||
\ERI{vv}{vc} & = 0.316\,490
|
||||
&
|
||||
\ERI{vc}{cc} & = 0.255\,554
|
||||
\end{align}
|
||||
|
||||
which yields
|
||||
\begin{align}
|
||||
\Omega & = 2.769\,327
|
||||
@ -395,17 +476,9 @@ while the static values are
|
||||
\omega_{1,\upup}^{\TDABSE} & = 1.49603
|
||||
\end{align}
|
||||
|
||||
It is now instructive to provide the exact results, \ie, the excitation energies obtained by diagonalizing the exact Hamiltonian in the same basis set.
|
||||
A quick configuration interaction with singles and doubles (CISD) calculation provide the following excitation energies:
|
||||
\begin{align}
|
||||
\omega_{1}^{\updw} & = 1.92145
|
||||
&
|
||||
\omega_{1}^{\upup} & = 1.47085
|
||||
&
|
||||
\omega_{3}^{\updw} & = 3.47880
|
||||
\end{align}
|
||||
This evidences that BSE reproduces qualitatively well the singlet and triplet single excitations, but quite badly the double excitation which is off by more than 1 hartree.
|
||||
All these numerical results are gathered in Table \ref{tab:BSE}.
|
||||
This evidences that BSE reproduces qualitatively well the singlet and triplet single excitations, but quite badly the double excitation which is off by more than 1 hartree.
|
||||
A
|
||||
|
||||
The perturbatively-corrected values are also reported, which shows that this scheme is very efficient at reproducing the dynamical value.
|
||||
Note that, although the BSE1(dTDA) value is further from the dBSE value than BSE1, it is quite close to the exact excitation energy.
|
||||
|
BIN
Notes/Maitra.pdf
Normal file
BIN
Notes/Maitra.pdf
Normal file
Binary file not shown.
Loading…
Reference in New Issue
Block a user