saving work before going to the market
This commit is contained in:
parent
addcb7ee11
commit
6873af824a
@ -114,16 +114,16 @@ For example, modeling core electron spectroscopy requires core ionization energi
|
||||
Many-body perturbation theory can also be used to access optical excitation energies through the Bethe-Salpeter equation. \cite{Salpeter_1951,Strinati_1988,Blase_2018,Blase_2020} However, the accuracy is not yet satisfying for triplet excited states, where instabilities often occur. \cite{Bruneval_2015,Jacquemin_2017a,Jacquemin_2017b,Holzer_2018a}
|
||||
Therefore, even if $GW$ offers a good trade-off between accuracy and computational cost, some situations might require higher precision.
|
||||
Unfortunately, defining a systematic way to go beyond $GW$ via the inclusion of vertex corrections has been demonstrated to be a tricky task. \cite{Baym_1961,Baym_1962,DeDominicis_1964a,DeDominicis_1964b,Bickers_1989a,Bickers_1989b,Bickers_1991,Hedin_1999,Bickers_2004,Shirley_1996,DelSol_1994,Schindlmayr_1998,Morris_2007,Shishkin_2007b,Romaniello_2009a,Romaniello_2012,Gruneis_2014,Hung_2017,Maggio_2017b,Mejuto-Zaera_2022}
|
||||
For example, Lewis and Berkelbach have shown that naive vertex corrections can even worsen the quasi-particle energies with respect to $GW$. \cite{Lewis_2019}
|
||||
For example, Lewis and Berkelbach have shown that naive vertex corrections can even worsen the quasiparticle energies with respect to $GW$. \cite{Lewis_2019}
|
||||
We refer the reader to the recent review by Golze and co-workers (see Ref.~\onlinecite{Golze_2019}) for an extensive list of current challenges in many-body perturbation theory.
|
||||
|
||||
Many-body perturbation theory also suffers from the infamous intruder-state problem,\cite{Andersson_1994,Andersson_1995a,Roos_1995,Forsberg_1997,Olsen_2000,Choe_2001} where they manifest themselves as solutions of the quasi-particle equation with non-negligible spectral weights.
|
||||
Many-body perturbation theory also suffers from the infamous intruder-state problem,\cite{Andersson_1994,Andersson_1995a,Roos_1995,Forsberg_1997,Olsen_2000,Choe_2001} where they manifest themselves as solutions of the quasiparticle equation with non-negligible spectral weights.
|
||||
In some cases, this transfer of spectral weight makes it difficult to distinguish between a quasiparticle and a satellite.
|
||||
These multiple solutions hinder the convergence of partially self-consistent schemes such as quasiparticle self-consistent $GW$ (qs$GW$) and eigenvalue-only self-consistent $GW$ (ev$GW$). \cite{Veril_2018,Forster_2021,Monino_2022}
|
||||
The simpler one-shot $G_0W_0$ scheme is also impacted by these intruder states, leading to discontinuities in a variety of physical quantities including charged and neutral excitation energies, correlation and total energies.\cite{Loos_2018b,Veril_2018,Loos_2020e,Berger_2021,DiSabatino_2021,Monino_2022,Scott_2023}
|
||||
These convergence problems and discontinuities can even happen in the weakly correlated regime where the $GW$ approximation is supposed to be valid.
|
||||
|
||||
In a recent study, Monino and Loos showed that the discontinuities could be removed by the introduction of a regularizer inspired by the similarity renormalization group (SRG) in the quasi-particle equation. \cite{Monino_2022}
|
||||
In a recent study, Monino and Loos showed that the discontinuities could be removed by the introduction of a regularizer inspired by the similarity renormalization group (SRG) in the quasiparticle equation. \cite{Monino_2022}
|
||||
Encouraged by the recent successes of regularization schemes in many-body quantum chemistry methods, such as in single- and multi-reference perturbation theory, \cite{Lee_2018a,Shee_2021,Evangelista_2014b,ChenyangLi_2019a,Battaglia_2022} the present work investigates the application of the SRG formalism to many-body perturbation theory in its $GW$.
|
||||
In particular, we focus here on the possibility of curing the qs$GW$ convergence problems using the SRG.
|
||||
|
||||
@ -160,7 +160,7 @@ This section starts by
|
||||
\label{sec:gw}
|
||||
%%%%%%%%%%%%%%%%%%%%%%
|
||||
|
||||
The central equation of many-body perturbation theory based on Hedin's equations is the so-called quasi-particle equation which, within the $GW$ approximation, reads
|
||||
The central equation of many-body perturbation theory based on Hedin's equations is the so-called quasiparticle equation which, within the $GW$ approximation, reads
|
||||
\begin{equation}
|
||||
\label{eq:quasipart_eq}
|
||||
\qty[ \bF + \bSig(\omega = \epsilon_p) ] \psi_p(\bx) = \epsilon_p \psi_p(\bx),
|
||||
@ -169,7 +169,8 @@ where $\bF$ is the Fock matrix \cite{SzaboBook} and $\bSig(\omega)$ is (the corr
|
||||
Both are $K \times K$ matrices with $K$ the number of one-electron orbitals.
|
||||
The self-energy can be physically understood as a correction to the Hartree-Fock (HF) problem (represented by $\bF$) accounting for dynamical screening effects.
|
||||
Similarly to the HF case, Eq.~\eqref{eq:quasipart_eq} has to be solved self-consistently.
|
||||
\ant{Note that $\bSig(\omega)$ is dynamical which implies that it depends on both the one-electron orbitals $\psi_p(\bx)$ and their corresponding energies $\epsilon_p$, while $\bF$ depends only on the orbitals.}
|
||||
\titou{Note that $\bSig(\omega)$ is dynamical which implies that it depends on both the one-electron orbitals $\psi_p(\bx)$ and their corresponding energies $\epsilon_p$, while $\bF$ depends only on the orbitals.}
|
||||
\PFL{I still don't like it.}
|
||||
|
||||
The matrix elements of $\bSig(\omega)$ have the following analytic expression \cite{Hedin_1999,Tiago_2006,Bruneval_2012,vanSetten_2013,Bruneval_2016}
|
||||
\begin{equation}
|
||||
@ -181,56 +182,62 @@ The matrix elements of $\bSig(\omega)$ have the following analytic expression \c
|
||||
where $\eta$ is a positive infinitesimal and the screened two-electron integrals are
|
||||
\begin{equation}
|
||||
\label{eq:GW_sERI}
|
||||
W_{p,q\nu} = \sum_{ia}\eri{pi}{qa}\qty( \bX_{\nu}+\bY_{\nu})_{ia},
|
||||
W_{p,q\nu} = \sum_{ia}\eri{pi}{qa}\qty(\bX+\bY)_{ia,\nu},
|
||||
\end{equation}
|
||||
where $\bX$ and $\bY$ are the components of the eigenvectors of the particle-hole direct RPA problem defined as
|
||||
where $\bX$ and $\bY$ are the components of the eigenvectors of the particle-hole direct (\ie without exchange) RPA problem defined as
|
||||
\begin{equation}
|
||||
\label{eq:full_dRPA}
|
||||
\mqty( \bA & \bB \\ -\bA & -\bB ) \mqty( \bX \\ \bY ) = \mqty( \bX \\ \bY ) \mqty( \boldsymbol{\Omega} & \bO \\ \bO & -\boldsymbol{\Omega} ),
|
||||
\mqty( \bA & \bB \\ -\bB & -\bA ) \mqty( \bX \\ \bY ) = \mqty( \bX \\ \bY ) \mqty( \boldsymbol{\Omega} & \bO \\ \bO & -\boldsymbol{\Omega} ),
|
||||
\end{equation}
|
||||
with
|
||||
\begin{subequations}
|
||||
\begin{align}
|
||||
A_{ia,jb} & = (\epsilon_i - \epsilon_a) \delta_{ij}\delta_{ab} + \eri{ib}{aj}.
|
||||
A_{ia,jb} & = (\epsilon_i - \epsilon_a) \delta_{ij}\delta_{ab} + \eri{ib}{aj},
|
||||
\\
|
||||
B_{ia,jb} & = \eri{ij}{ab}.
|
||||
B_{ia,jb} & = \eri{ij}{ab},
|
||||
\end{align}
|
||||
\end{subequations}
|
||||
The diagonal matrix $\boldsymbol{\Omega}$ contains the positive eigenvalues and its elements $\Omega_\nu$ appear in Eq.~\eqref{eq:GW_selfenergy}.
|
||||
The TDA case is discussed in Appendix \ref{sec:nonTDA}.
|
||||
Throughout the manuscript, the indices $p,q,r,s$ are general orbitals while $i,j,k,l$ and $a,b,c,d$ refers to occupied and virtual orbitals, respectively.
|
||||
The indices $\mu$ and $\nu$ are composite indices, \eg $\nu=(ia)$, referring to neutral excitations.
|
||||
and
|
||||
\begin{equation}
|
||||
\braket{pq}{rs} = \iint \frac{\SO{p}(\bx_1) \SO{q}(\bx_2)\SO{r}(\bx_1) \SO{s}(\bx_2) }{\abs{\br_1 - \br_2}} d\bx_1 d\bx_2
|
||||
\end{equation}
|
||||
are two-electron integrals in the spin-orbital basis.
|
||||
|
||||
Because of the frequency dependence, fully solving the quasi-particle equation is a rather complicated task.
|
||||
The diagonal matrix $\boldsymbol{\Omega}$ contains the positive eigenvalues and its elements $\Omega_\nu$ appear in Eq.~\eqref{eq:GW_selfenergy}.
|
||||
\titou{In the Tamm-Dancoff approximation (TDA), which is discussed in Appendix \ref{sec:nonTDA}, one sets $\bB = \bO$ in Eq.~\eqref{eq:full_dRPA} which reduces to a Hermitian eigenvalue problem of the form $\bA \bX = \bX \bOm$.}
|
||||
Throughout the manuscript, the indices $p,q,r,s$ are general orbitals while $i,j,k,l$ and $a,b,c,d$ refers to occupied and virtual orbitals, respectively.
|
||||
The indices $\mu$ and $\nu$ are composite indices, \eg $\nu=(ia)$, referring to neutral (single) excitations.
|
||||
|
||||
Because of the frequency dependence of the self-energy, fully solving the quasiparticle equation \eqref{eq:quasipart_eq} is a rather complicated task.
|
||||
Hence, several approximate schemes have been developed to bypass self-consistency.
|
||||
The most popular one is the one-shot (perturbative) scheme, known as $G_0W_0$, where the self-consistency is completely abandoned, and the off-diagonal elements of Eq.~\eqref{eq:quasipart_eq} are neglected.
|
||||
Assuming a HF starting point, this results in $K$ quasi-particle equations that read
|
||||
The most popular strategy is the one-shot (perturbative) scheme, known as $G_0W_0$, where the self-consistency is completely abandoned, and the off-diagonal elements of Eq.~\eqref{eq:quasipart_eq} are neglected.
|
||||
Assuming a HF starting point, this results in $K$ quasiparticle equations that read
|
||||
\begin{equation}
|
||||
\label{eq:G0W0}
|
||||
\epsilon_p^{\HF} + \Sigma_{pp}(\omega) - \omega = 0,
|
||||
\end{equation}
|
||||
where $\Sigma_{pp}(\omega)$ are the diagonal elements of $\bSig$ and $\epsilon_p^{\HF}$ are the HF orbital energies.
|
||||
The previous equations are non-linear with respect to $\omega$ and therefore can have multiple solutions $\epsilon_{p,s}$ for a given $p$ (where the index $s$ is numbering solutions).
|
||||
The previous equations are non-linear with respect to $\omega$ and therefore have multiple solutions $\epsilon_{p,s}$ for a given $p$ (where the index $s$ is numbering solutions).
|
||||
These solutions can be characterized by their spectral weight given by the renormalisation factor $Z_{p,s}$
|
||||
\begin{equation}
|
||||
\label{eq:renorm_factor}
|
||||
0 \leq Z_{p,s} = \qty[ 1 - \eval{\pdv{\Sigma_{pp}(\omega)}{\omega}}_{\omega=\epsilon_{p,s}} ]^{-1} \leq 1.
|
||||
\end{equation}
|
||||
The solution with the largest weight is referred to as the quasi-particle while the others are known as satellites (or shake-up transitions).
|
||||
However, in some cases, Eq.~\eqref{eq:G0W0} can have two (or more) solutions with similar weights, hence the quasi-particle is not well-defined.
|
||||
The solution with the largest weight is referred to as the quasiparticle while the others are known as satellites (or shake-up transitions).
|
||||
However, in some cases, Eq.~\eqref{eq:G0W0} can have two (or more) solutions with similar weights, hence the quasiparticle is not well-defined.
|
||||
These additional solutions with large weights are the previously mentioned intruder states.
|
||||
|
||||
One obvious flaw of the one-shot scheme mentioned above is its starting point dependence.
|
||||
Indeed, in Eq.~\eqref{eq:G0W0} we chose to use the HF orbital energies but this is arbitrary and one could have chosen Kohn-Sham orbitals for example.
|
||||
Therefore, one can \ant{tune} the starting point to obtain the best one-shot energies possible, which is commonly done. \cite{Korzdorfer_2012,Marom_2012,Bruneval_2013,Gallandi_2015,Caruso_2016, Gallandi_2016}
|
||||
Alternatively, one could solve this set of quasi-particle equations self-consistently leading to the eigenvalue-only self-consistent scheme (ev$GW$). \cite{Shishkin_2007a,Blase_2011b,Marom_2012,Kaplan_2016,Wilhelm_2016}
|
||||
The solutions $\epsilon_p$ are used to build Eq.~\eqref{eq:G0W0} instead of the HF ones and then these equation are solved for $\omega$ again.
|
||||
Therefore, one can ``tune'' the starting point to obtain the best one-shot energies possible, which is commonly done. \cite{Korzdorfer_2012,Marom_2012,Bruneval_2013,Gallandi_2015,Caruso_2016, Gallandi_2016}
|
||||
Alternatively, one could solve this set of quasiparticle equations self-consistently leading to the eigenvalue-only self-consistent scheme (ev$GW$). \cite{Shishkin_2007a,Blase_2011b,Marom_2012,Kaplan_2016,Wilhelm_2016}
|
||||
The solutions $\epsilon_p$ are used to build Eq.~\eqref{eq:G0W0} instead of the HF ones and then these equations are solved for $\omega$ again.
|
||||
This procedure is iterated until convergence for $\epsilon_p$ is reached.
|
||||
However, if one of the quasi-particle equations does not have a well-defined quasi-particle solution, reaching self-consistency can be quite difficult, if not impossible.
|
||||
However, if one of the quasiparticle equations does not have a well-defined quasiparticle solution, reaching self-consistency can be quite difficult, if not impossible.
|
||||
Even at convergence, the starting point dependence is not totally removed as the results still depend on the initial molecular orbitals. \cite{Marom_2012}
|
||||
|
||||
In order to update both the orbital energies and coefficients, one must consider the off-diagonal elements in $\bSig(\omega)$.
|
||||
To take into account the off-diagonal elements without solving the dynamic quasi-particle equation [Eq.~\eqref{eq:quasipart_eq}], one can resort to the quasi-particle self-consistent (qs) $GW$ scheme in which $\bSig(\omega)$ is replaced by a static approximation $\bSig^{\qs}$.
|
||||
To take into account the off-diagonal elements without solving the dynamic quasiparticle equation [Eq.~\eqref{eq:quasipart_eq}], one can resort to the quasiparticle self-consistent (qs) $GW$ scheme in which $\bSig(\omega)$ is replaced by a static approximation $\bSig^{\qs}$.
|
||||
Then the qs$GW$ problem is solved using the usual HF algorithm with $\bF$ replaced by $\bF + \bSig^{\qs}$.
|
||||
Various choices for $\bSig^\qs$ are possible but the most popular one is the following Hermitian approximation
|
||||
\begin{equation}
|
||||
@ -243,7 +250,7 @@ This will be done in the next sections.
|
||||
|
||||
Once again, in cases where multiple solutions have large spectral weights, qs$GW$ self-consistency can be difficult to reach.
|
||||
Multiple solutions of Eq.~\eqref{eq:G0W0} arise due to the $\omega$ dependence of the self-energy.
|
||||
Therefore, by suppressing this dependence the static approximation relies on the fact that there is one well-defined quasi-particle solution.
|
||||
Therefore, by suppressing this dependence the static approximation relies on the fact that there is one well-defined quasiparticle solution.
|
||||
If it is not the case, the qs$GW$ self-consistent scheme will oscillate between the solutions with large weights. \cite{Forster_2021}
|
||||
|
||||
The satellites causing convergence problems are the so-called intruder states.
|
||||
@ -253,10 +260,10 @@ Various other regularisers are possible and in particular one of us has shown th
|
||||
But it would be more rigorous, and more instructive, to obtain this regulariser from first principles by applying the SRG formalism to many-body perturbation theory.
|
||||
This is the aim of the rest of this work.
|
||||
|
||||
Applying the SRG to $GW$ could gradually remove the coupling between the quasi-particle and the satellites resulting in a renormalized quasi-particle.
|
||||
Applying the SRG to $GW$ could gradually remove the coupling between the quasiparticle and the satellites resulting in a renormalized quasiparticle.
|
||||
However, to do so one needs to identify the coupling terms in Eq.~\eqref{eq:quasipart_eq}, which is not straightforward.
|
||||
The way around this problem is to transform Eq.~\eqref{eq:quasipart_eq} to its upfolded version and the coupling terms will elegantly appear in the process.
|
||||
The upfolded $GW$ quasi-particle equation is \cite{Bintrim_2021,Tolle_2022}
|
||||
The upfolded $GW$ quasiparticle equation is \cite{Bintrim_2021,Tolle_2022}
|
||||
\begin{equation}
|
||||
\label{eq:GWlin}
|
||||
\begin{pmatrix}
|
||||
@ -277,7 +284,7 @@ The upfolded $GW$ quasi-particle equation is \cite{Bintrim_2021,Tolle_2022}
|
||||
\end{pmatrix}
|
||||
\boldsymbol{\epsilon},
|
||||
\end{equation}
|
||||
where $\boldsymbol{\epsilon}$ is a diagonal matrix collecting the quasi-particle and satellite energies, the 2h1p and 2p1h matrix elements are
|
||||
where $\boldsymbol{\epsilon}$ is a diagonal matrix collecting the quasiparticle and satellite energies, the 2h1p and 2p1h matrix elements are
|
||||
\begin{subequations}
|
||||
\begin{align}
|
||||
C^\text{2h1p}_{i\nu,j\mu} & = \left(\epsilon_i - \Omega_\nu\right)\delta_{ij}\delta_{\nu\mu},
|
||||
@ -400,8 +407,8 @@ Then, the aim is to solve order by order the flow equation [see Eq.~\eqref{eq:fl
|
||||
\bHod{1}(0) &= \begin{pmatrix} \bO & \bW \\ \bW^{\dagger} & \bO \end{pmatrix},
|
||||
\end{align}
|
||||
and $ \bHod{0}(0) = \bHd{1}(0) = \bO$, where we have defined the matrices $\bC$ and $\bV$ that collects the 2h1p and 2p1h channels for the sake of conciseness.
|
||||
Once the analytical low-order perturbative expansions are known they can be inserted in Eq.~\eqref{eq:GWlin} before downfolding to obtain a renormalized quasi-particle equation.
|
||||
In particular, in this manuscript, the focus will be on the second-order renormalized quasi-particle equation.
|
||||
Once the analytical low-order perturbative expansions are known they can be inserted in Eq.~\eqref{eq:GWlin} before downfolding to obtain a renormalized quasiparticle equation.
|
||||
In particular, in this manuscript, the focus will be on the second-order renormalized quasiparticle equation.
|
||||
|
||||
%///////////////////////////%
|
||||
\subsection{Zeroth-order matrix elements}
|
||||
@ -450,7 +457,7 @@ Note the close similarity of the first-order element expressions with the ones o
|
||||
\subsection{Second-order matrix elements}
|
||||
% ///////////////////////////%
|
||||
|
||||
The second-order renormalized quasi-particle equation is given by
|
||||
The second-order renormalized quasiparticle equation is given by
|
||||
\begin{equation}
|
||||
\label{eq:GW_renorm}
|
||||
\left( \widetilde{\bF}(s) + \widetilde{\bSig}(\omega; s) \right) \bX = \omega \bX,
|
||||
@ -461,7 +468,7 @@ with
|
||||
\label{eq:srg_sigma}
|
||||
\widetilde{\bSig}(\omega; s) &= \bV^{(1)}(s) \left(\omega \bI - \bC^{(0)}\right)^{-1} (\bV^{(1)}(s))^{\dagger}.
|
||||
\end{align}
|
||||
As can be readily seen above, $\bF^{(2)}$ is the only second-order block of the effective Hamiltonian contributing to the second-order SRG quasi-particle equation.
|
||||
As can be readily seen above, $\bF^{(2)}$ is the only second-order block of the effective Hamiltonian contributing to the second-order SRG quasiparticle equation.
|
||||
Collecting every second-order terms in the flow equation and performing the block matrix products results in the following differential equation for $\bF^{(2)}$
|
||||
\begin{multline}
|
||||
\label{eq:diffeqF2}
|
||||
@ -732,7 +739,7 @@ Within the TDA the renormalized matrix elements have the same $s$ dependence as
|
||||
% The fact that the integrals are not screened in GF(2) manifests itself in the fact that the $\bC$ matrices are already diagonal.
|
||||
|
||||
% Applying the SRG formalism to this matrix is completely analog to the derivation exposed in the main text.
|
||||
% We only give the analytical expressions of the matrix elements needed for the second-order SRG-GF(2) quasi-particle equations.
|
||||
% We only give the analytical expressions of the matrix elements needed for the second-order SRG-GF(2) quasiparticle equations.
|
||||
|
||||
% \begin{equation}
|
||||
% (V^\text{2h1p}_{p,ija})^{(1)}(s) = \frac{1}{\sqrt{2}}\aeri{pa}{ij} e^{- (\epsilon_p + \epsilon_a - \epsilon_i - \epsilon_j)^2 s}
|
||||
|
Loading…
Reference in New Issue
Block a user