\documentclass[aip,jcp,reprint,noshowkeys,superscriptaddress]{revtex4-1} \usepackage{graphicx,dcolumn,bm,xcolor,microtype,multirow,amscd,amsmath,amssymb,amsfonts,physics,longtable,wrapfig,txfonts,mleftright,bbold} \usepackage[version=4]{mhchem} \usepackage[utf8]{inputenc} \usepackage[T1]{fontenc} \usepackage{txfonts} \usepackage[ colorlinks=true, citecolor=blue, breaklinks=true ]{hyperref} \urlstyle{same} %============================================================% %%% NEWCOMMANDS %%% % ============================================================% \usepackage[normalem]{ulem} \newcommand{\titou}[1]{\textcolor{red}{#1}} \newcommand{\trashPFL}[1]{\textcolor{\red}{\sout{#1}}} \newcommand{\PFL}[1]{\titou{(\underline{\bf PFL}: #1)}} \newcommand{\ant}[1]{\textcolor{green}{#1}} % addresses \newcommand{\LCPQ}{Laboratoire de Chimie et Physique Quantiques (UMR 5626), Universit\'e de Toulouse, CNRS, UPS, France} \begin{document} \title{Notes on the project: Similarity Renormalization Group formalism applied to Green's function theory} \author{Antoine \surname{Marie}} \email{amarie@irsamc.ups-tlse.fr} \affiliation{\LCPQ} \author{Pierre-Fran\c{c}ois \surname{Loos}} \email{loos@irsamc.ups-tlse.fr} \affiliation{\LCPQ} %\begin{abstract} %Here comes the abstract. %\bigskip %\begin{center} % \boxed{\includegraphics[width=0.5\linewidth]{TOC}} %\end{center} %\bigskip %\end{abstract} \maketitle %=================================================================% \section{Introduction} \label{sec:intro} %=================================================================% This document is a compilation of various notes related to the similarity renormalisation group (SRG) and its application to many-body perturbation theory (MBPT). Before tackling its application to MBPT, we give the main SRG equations in Sec.~\ref{sec:intro}. Then, we derive in details the perturbative expressions obtained by Evangelista by applying the SRG to the electronic Hamiltonian (see Sec.~\ref{sec:srg}). Now turning to MBPT, its various flavors are presented in Sec.~\ref{sec:folded} and the corresponding unfolded equations are given in Sec.~\ref{sec:unfolded}. Then, a SRG perturbative analysis is performed on general matrices in Sec.~\ref{sec:matrix_srg}. Finally, in Sec.~\ref{sec:second_quant_mbpt} we investigate the possibility to find a second quantized effective Hamiltonian corresponding to the unfolded MBPT equations. %=================================================================% \section{The similarity renormalisation group} \label{sec:srg} %=================================================================% The similarity renormalization group aims at continuously transforming an Hamiltonian to a diagonal form, or more often to a block-diagonal form. Therefore, the transformed Hamiltonian \begin{equation} \bH(s) = \bU(s) \, \bH \, \bU^\dag(s) \end{equation} depends on a flow parameter $s$. The resulting Hamiltonian possess up to $N$-body operators with $N$ the number of particle. \begin{equation} \bH(s) = E_0(s) + \bF{}{}(s) + \bV{}{}(s) + \bW(s) + \dots \end{equation} In the following, we will truncate every contribution superior to two-body operators. We can easily derive an evolution equation for this Hamiltonian by taking the derivative of $\bH(s)$. This gives \begin{equation} \label{eq:flowEquation} \dv{\bH(s)}{s} = \comm{\boldsymbol{\eta}(s)}{\bH(s)} \end{equation} where $\boldsymbol{\eta}(s)$, the flow generator, is defined as \begin{equation} \boldsymbol{\eta}(s) = \dv{\bU(s)}{s} \bU^\dag(s) = - \boldsymbol{\eta}^\dag(s) . \end{equation} To solve this equation at a cost inferior to the one of diagonalizing the initial Hamiltonian, one needs to introduce approximation for $\boldsymbol{\eta}(s)$. Before doing so, we need to define what is the blocks to suppress in order to obtain a block-diagonal Hamiltonian. Therefore, the Hamiltonian is separated in two parts as \begin{equation} \bH(s) = \underbrace{\bH^\text{d}(s)}_{\text{diagonal}} + \underbrace{\bH^\text{od}(s)}_{\text{off-diagonal}}. \end{equation} By definition, we have the following condition on $\bH^\text{od}$ \begin{equation} \bH^\text{od}(\infty) = \boldsymbol{0}. \end{equation} In this work, we will use Wegner's canonical generator which is defined as \begin{equation} \boldsymbol{\eta}^\text{W}(s) = \comm{\bH^\text{d}(s)}{\bH(s)} = \comm{\bH^\text{d}(s)}{\bH^\text{od}(s)}. \end{equation} This generator has the advantage of defining a true renormalisation scheme, \ie the coupling coefficients with the highest energy determinants are removed first. One of the flaws of this generator is that it generates a \titou{stiff} set of ODE which is difficult to solve numerically. However, here we consider analytical perturbative expressions so we will not be affected by this problem. %=================================================================% \section{The SRG electronic Hamiltonian} \label{sec:electronic_ham} %=================================================================% In this part, we derive the perturbative expression for the SRG applied to the non-relativistic electronic Hamiltonian \begin{equation} \label{eq:hamiltonianSecondQuant} \hH = \sum_{pq} f_{pq} \cre{p}\ani{q} + \frac{1}{4} \sum_{pqrs} \aeri{pq}{rs} \cre{p}\cre{q}\ani{r}\ani{s} \end{equation} which can also be written in normal order wrt a reference determinant as \begin{equation} \label{eq:hamiltonianNormalOrder} \hH = E_0 + \sum_{pq} + f_p^q\no{q}{p} + \frac{1}{4} \sum_{pqrs}v_{pq}^{rs}\no{rs}{pq}. \end{equation} In this case, we want to decouple the reference determinant from every singly and doubly excited determinants. Hence, we define the off-diagonal Hamiltonian as \begin{equation} \label{eq:hamiltonianOffDiagonal} \hH^{\text{od}}(s) = \sum_{ia} f_i^a(s)\no{a}{i} + \frac{1}{4} \sum_{ijab}v_{ij}^{ab}(s)\no{ab}{ij}. \end{equation} Note that each coefficients depend on $s$. The perturbative parameter $\lambda$ is such that \begin{equation} \bH(0) = E_0(0) + \bF{}{}(0) + \lambda \bV{}{}(0) \end{equation} In addition, we know the following initial conditions. We use the HF basis set of the reference such that $\bF{}{\text{od}}(0) = 0$ and $\bF{}{\mathrm{d}}(0)_{pq}=\delta_{pq}\epsilon_p$. Therefore, we have \begin{align} \bH^\text{d}(0)&=E_0(0) + \bF{}{\mathrm{d}}(0) + \lambda \bV{}{\mathrm{d}}(0) & \bH^\text{od}(0)&= \lambda \bV{}{\mathrm{od}}(0) \end{align} Now, we want to compute the terms at each order of the following development \begin{equation} \label{eq:hamiltonianPTExpansion} \bH(s) = \bH^{(0)}(s) + \bH^{(1)}(s) + \bH^{(2)}(s) + \bH^{(3)}(s) + \dots \end{equation} by integrating Eq.~\eqref{eq:flowEquation}. %%%%%%%%%%%%%%%%%%%%%% \subsection{Zeroth order Hamiltonian} %%%%%%%%%%%%%%%%%%%%%% First, we start by showing that the zeroth order Hamiltonian is independant of $s$ and therefore equal to $\bH^{(0)}(0)$ \begin{align} \bH(\delta s) &= \bH(0) + \delta s \dv{\bH(s)}{s}\bigg|_{s=0} + O(\delta s^2) \\ \dv{\bH(s)}{s}\bigg|_{s=0} &= [ \boldsymbol{\eta}(0), \bH(0)] \\ \end{align} However, we have seen that $\bH^\text{od}(0)$ is of order 1. Hence, $\boldsymbol{\eta}(0)$ does not have a zero order contribution. Which gives us the following equality \begin{equation} \bH^{(0)}(\delta s) = \bH^{(0)}(0) \end{equation} meaning that the zeroth order Hamiltonian is independant of $s$. %%%%%%%%%%%%%%%%%%%%%% \subsection{First order Hamiltonian} %%%%%%%%%%%%%%%%%%%%%% The right-hand side of \eqref{eq:flowEquation} has no zeroth order, so we want to compute its first order term. To do so, we first need to compute the first order term of $\boldsymbol{\eta}(s)$. We have seen that $\bH^\text{od}(0)$ has no zeroth order contribution so we have \begin{equation} \boldsymbol{\eta}^{(1)}(s) = \comm{\bH^{\text{d},(0)}(0)}{\bH^{\text{od},(1)}(s)} \end{equation} According to the appendix the one-body part of $\boldsymbol{\eta}^{(1)}(s) $ has four contributions. However, two of them involve the two-body part of $\bH^{\text{d},(0)}(0)$ which is equal to zero. In addition, the term $\left[A_{1}, B_{2}\right]_{p}^{q}$ involves the coefficients $A_i^a = f_i^a(0) = 0$. So finally we only have one term \begin{align} \eta_a^{i,(1)}(s) &= \comm{\bH_1^{\text{d},(0)}(0)}{\bH_1^{\text{od},(1)}(s)}_a^i \\ &= \sum_r \left( f_a^r(0) f_r^{i,(1)}(s) - f_r^i(0) f_a^{r,(1)}(s) \right) \\ &= (\epsilon_a - \epsilon_i)f_a^{i,(1)}(s) \end{align} Now turning to the two-body part of $\boldsymbol{\eta}^{(1)}(s) $ and once again two terms are zero because the two-body part of $\bH^{\text{d},(0)}(0)$ is equal to zero. So we have \begin{align} \eta_{ab}^{ij,(1)}(s) &= \comm{\bH_1^{\text{d},(0)}(0)}{\bH_2^{\text{od},(1)}(s)}_{ab}^{ij} \\ &= \sum_t [P(ab) f_a^t(0) v_{tb}^{ij,(1)}(s) - P(ij) f_t^i(0) v_{ab}^{tj,(1)}(s) ] \notag \\ &= \sum_t [ P(ab) \epsilon_a \delta_{at}v_{tb}^{ij,(1)}(s) - P(ij) \epsilon_i \delta_{it} v_{ab}^{tj,(1)}(s) ] \notag \\ &= P(ab) \epsilon_a v_{ab}^{ij,(1)}(s) - P(ij) \epsilon_i v_{ab}^{ij,(1)}(s) \notag\\ &= \left( \epsilon_a + \epsilon_b - \epsilon_i - \epsilon_j \right) v_{ab}^{ij,(1)} \notag \end{align} We can now compute the first order contribution to Eq.~\eqref{eq:flowEquation}. We have seen that $\eta$ has no zeroth order contribution so \begin{equation} \dv{\bH^{(1)}(s)}{s} = \comm{\boldsymbol{\eta}^{(1)}(s)}{\bH^{(0)}(s)} \end{equation} We start with the scalar contribution, \ie the PT1 energy \begin{equation} \dv{E_0^{(1)}(s)}{s} = \mel{\phi}{\comm{\boldsymbol{\eta}_1^{(1)}(s)}{\bH_1^{(0)}(s)}}{\phi} + \mel{\phi}{\comm{\boldsymbol{\eta}_2^{(1)}(s)}{\bH_2^{(0)}(s)}}{\phi} \end{equation} where the second term is equal to zero. Thus we have \begin{align} \label{eq:diffEqScalPT1} \dv{E_0^{(1)}(s)}{s} &= \sum_{ip}\eta_i^{p,(1)}f_p^i(0) - \eta_p^{i,(1)}f_i^p(0) \\ &= \sum_i \epsilon_i(\eta_i^{i,(1)} - \eta_i^{i,(1)}) \notag \\ &= 0 \notag \end{align} Using the exact same reasoning as above we can show that there is only of the four terms of the one-body part of the commutator that is non-zero. \begin{align} \dv{f_a^{i,(1)}(s)}{s} &= \comm{\boldsymbol{\eta}_1^{(1)}(s)}{\bH_1^{(0)}(s)}_a^i \\ &= \sum_r \eta_a^{r,(1)}(s)f_r^i(0) - \eta_r^{i,(1)}f_a^r(0) \notag \\ &= (\epsilon_i - \epsilon_a) \eta_a^{i,(1)}(s) \notag \\ &= -(\epsilon_i - \epsilon_a) ^2f_a^{i,(1)}(s) \notag \end{align} The derivation for the two-body part to first order is once again similar \begin{align} \dv{v_{ab}^{ij,(1)}(s)}{s} &= \comm{\boldsymbol{\eta}_2^{(1)}(s)}{\bH_1^{(0)}(s)}_{ab}^{ij} \\ &= -(\epsilon_i + \epsilon_j - \epsilon_a - \epsilon_b) ^2v_{ab}^{ij,(1)}(s) \notag \end{align} These three differential equations can be integrated to obtain the analytical form of the Hamiltonian coefficients up to first order of perturbation theory. \begin{align} E_0^{(1)}(s) &= E_0^{(1)}(0) = 0 \\ f_a^{i,(1)}(s) &= f_a^{i,(1)}(0) e^{-s (\Delta_a^i )^2} = 0 \\ v_{ab}^{ij,(1)}(s) &= v_{ab}^{ij,(1)}(0) e^{-s (\Delta_{ab}^{ij})^2} = \aeri{ij}{ab} e^{-s (\Delta_{ab}^{ij})^2} \end{align} %%%%%%%%%%%%%%%%%%%%%% \subsection{Second order Hamiltonian} %%%%%%%%%%%%%%%%%%%%%% Now turning to the differential equations, we start by computing the scalar part of Eq.~\eqref{eq:flowEquation}, \ie the differential equation for the second order energy. \begin{align} \dv{E_0^{(2)}(s)}{s} & = \mel{\phi}{\comm{\boldsymbol{\eta}_1^{(2)}(s)}{\bH_1^{(0)}(s)}}{\phi} + \mel{\phi}{\comm{\boldsymbol{\eta}_2^{(2)}(s)}{\bH_2^{(0)}(s)}}{\phi} \\ &+ \mel{\phi}{\comm{\boldsymbol{\eta}_1^{(1)}(s)}{\bH_1^{(1)}(s)}}{\phi} + \mel{\phi}{\comm{\boldsymbol{\eta}_2^{(1)}(s)}{\bH_2^{(1)}(s)}}{\phi} \end{align} The two first terms are equal to zero for the same reason as the PT1 scalar differential equation (see Eq.~\eqref{eq:diffEqScalPT1}). In addition, the one-body hamiltonian has no first order contribution so \begin{align} \dv{E_0^{(2)}(s)}{s} &= \mel{\phi}{\comm{\boldsymbol{\eta}_2^{(1)}(s)}{\bH_2^{(1)}(s)}}{\phi} \\ &= \frac{1}{4} \sum_{i j} \sum_{a b}\left(\eta_{i j}^{ab,(1)} H_{ab}^{i j,(1)} - H_{i j}^{ab,(1)} \eta_{ab}^{i j,(1)}\right) \\ &=\frac{1}{4} \sum_{i j} \sum_{a b}\left(\eta_{i j}^{ab,(1)} - \eta_{ab}^{i j,(1)}\right) v_{ab}^{ij,(1)} \\ &= \frac{1}{4} \sum_{i j} \sum_{a b}\left(\Delta_{ab}^{ij}v_{ij}^{ab,(1)} - (-\Delta_{ab}^{ij} v_{ab}^{ij,(1)}) \right) \aeri{ij}{ab} e^{-s (\Delta_{ab}^{ij})^2} \\ &= \frac{1}{2} \sum_{i j} \sum_{a b} \Delta_{ab}^{ij} (v_{ab}^{ij,(1)})^2 \\ &= \frac{1}{2} \sum_{i j} \sum_{a b} \Delta_{ab}^{ij} \aeri{ij}{ab}^2 e^{-2s (\Delta_{ab}^{ij})^2} \end{align} After integration, using the initial condition $E_0^{(2)}(0)=0$, we obtain \begin{equation} \label{eq:SRG_MP2} E_0^{(2)}(s) = \frac{1}{4} \sum_{i j} \sum_{a b} \frac{\aeri{ij}{ab}^2}{\Delta_{ab}^{ij}}\left(1-e^{-2s (\Delta_{ab}^{ij})^2}\right) \end{equation} We can continue the derivation further than Evangelista's paper. To compute the second order contribution to the Hamiltonian coefficients, we first need to compute the second order contribution to $\boldsymbol{\eta}(s)$. \begin{equation} \boldsymbol{\eta}^{(2)}(s) = \comm{\bH^{\text{d},(0)}(0)}{\bH^{\text{od},(2)}(s)} + \comm{\bH^{\text{d},(1)}(s)}{\bH^{\text{od},(1)}(s)} \end{equation} The expressions for the first commutator are computed analogously to the one of the previous subsection. We focus on deriving expressions for the second term. The one-body part of $\bH^{\text{od},(1)}(s)$ is equal to zero so two of the four terms contributing to the one-body part of $\comm{\bH^{\text{d},(1)}(s)}{\bH^{\text{od},(1)}(s)}$ are zero. In addition, the term $\comm{A_1}{B_2}$ is equal to zero as well because the coefficients $(A_1)_{i}^a$ are zero (see expression in Appendix). So we have \begin{align} \eta_a^{i,(2)}(s) &= \comm{\bH_1^{\text{d},(0)}(0)}{\bH_1^{\text{od},(2)}(s)}_a^i + \comm{\bH_2^{\text{d},(1)}(s)}{\bH_2^{\text{od},(1)}(s)}_a^i \\ &= (\epsilon_a - \epsilon_i)f_a^{i,(2)}(s) + \dots \end{align} Need to continue this derivation but this not needed for EPT2. %=================================================================% \section{The various flavors of MBPT} \label{sec:mbpt} % =================================================================% The central equation of MBPT in practice is the following \begin{equation} \label{eq:quasipart_eq} \bF{}{} + \bSig(\omega) = \omega \mathbb{1}. \end{equation} \PFL{Not quite. You're missing the eigenvectors to make it a non-linear eigenvalue problem.} However, in order to use it we need to rely on approximations of the dynamical self-energy $\bSig(\omega)$. %%%%%%%%%%%%%%%%%%%%%% \subsection{Self-energies and quasiparticle equations} \label{sec:folded} %%%%%%%%%%%%%%%%%%%%%% In the following, we will focus on the GF(2), $GW$ and $GT$ approximations. The GF($n$) formalism is defined such that the self-energy includes every diagram up to $n$-th order of M\"oller-Plesset perturbation theory. \begin{equation} \label{eq:GF2_selfenergy} \Sigma_{pq}^{\text{GF(2)}}(\omega) = \sum_{ija} \frac{W_{pa,ij}^{\text{GF(2)}}W_{qa,ij}^{\text{GF(2)}}}{\omega + \epsilon _a -\epsilon_i -\epsilon_j - \ii \eta} + \sum_{iab} \frac{W_{pi,ab}^{\text{GF(2)}}W_{qi,ab}^{\text{GF(2)}}}{\omega + \epsilon _i -\epsilon_a -\epsilon_b + \ii \eta} \end{equation} with \begin{equation} \label{eq:GF2_sERI} W^{\GF}_{pq,rs}= \frac{1}{\sqrt{2}}\aeri{pq}{rs} \end{equation} On the other hand, the $GW$ self-energy is obtained by taking the RPA polarizability and removing the vertex correction in the exact definition of the self-energy. \begin{equation} \label{eq:GW_selfenergy} \Sigma_{pq}^{\GW}(\omega) = \sum_{iv} \frac{W_{pi,v}^{\GW} W_{qi,v}^{\GW}}{\omega - \epsilon_i + \Omega_{v}^{\dRPA} - \ii \eta} + \sum_{av} \frac{W_{pa,v}^{\GW}W_{qa,v}^{\GW}}{\omega - \epsilon_a - \Omega_{v}^{\dRPA} + \ii \eta} \notag \end{equation} with \begin{equation} \label{eq:GW_sERI} W_{pq,v}^{\GW} = \sum_{ia}\eri{pi}{qa}\qty( \bX_{v}^{\dRPA} + \bY_{v}^{\dRPA} )_{ia} \end{equation} Finally, the $GT$ approximation corresponds to another approximation to the polarizability than in $GW$, namely the one coming from pp-hh-RPA The corresponding self-energies read as \begin{equation} \label{eq:GT_selfenergy} \Sigma_{pq}^{\GT}(\omega) = \sum_{iv} \frac{W_{pi,v}^{N+2} W_{qi,v}^{N+2}}{\omega + \epsilon_i - \Omega_{v}^{N+2} + \ii \eta} + \sum_{av} \frac{W_{pa,v}^{N-2} W_{qa,v}^{N-2}}{\omega + \epsilon_a - \Omega_{v}^{N-2} - \ii \eta} \notag \end{equation} with \begin{align} \label{eq:GT_sERI} W_{pq,v}^{N+2} & = \sum_{c