moving Hubbard MP section

This commit is contained in:
Pierre-Francois Loos 2020-11-18 22:01:21 +01:00
parent ff78e32218
commit b9dccf97cd
2 changed files with 73 additions and 67 deletions

View File

@ -6,12 +6,13 @@
\BOOKMARK [1][-]{section*.7}{Perturbation theory}{section*.2}% 6
\BOOKMARK [2][-]{section*.8}{Rayleigh-Schr\366dinger perturbation theory}{section*.7}% 7
\BOOKMARK [2][-]{section*.9}{The Hartree-Fock Hamiltonian}{section*.7}% 8
\BOOKMARK [2][-]{section*.11}{M\370ller-Plesset perturbation theory}{section*.7}% 9
\BOOKMARK [1][-]{section*.12}{Historical overview}{section*.2}% 10
\BOOKMARK [2][-]{section*.13}{Behavior of the M\370ller-Plesset series}{section*.12}% 11
\BOOKMARK [2][-]{section*.16}{Insights from a two-state model}{section*.12}% 12
\BOOKMARK [2][-]{section*.17}{The singularity structure}{section*.12}% 13
\BOOKMARK [2][-]{section*.18}{The physics of quantum phase transitions}{section*.12}% 14
\BOOKMARK [1][-]{section*.19}{Conclusion}{section*.2}% 15
\BOOKMARK [1][-]{section*.20}{Acknowledgments}{section*.2}% 16
\BOOKMARK [1][-]{section*.21}{References}{section*.2}% 17
\BOOKMARK [2][-]{section*.10}{Complex adiabatic connection}{section*.7}% 9
\BOOKMARK [2][-]{section*.12}{M\370ller-Plesset perturbation theory}{section*.7}% 10
\BOOKMARK [1][-]{section*.14}{Historical overview}{section*.2}% 11
\BOOKMARK [2][-]{section*.15}{Behavior of the M\370ller-Plesset series}{section*.14}% 12
\BOOKMARK [2][-]{section*.17}{Insights from a two-state model}{section*.14}% 13
\BOOKMARK [2][-]{section*.18}{The singularity structure}{section*.14}% 14
\BOOKMARK [2][-]{section*.19}{The physics of quantum phase transitions}{section*.14}% 15
\BOOKMARK [1][-]{section*.20}{Conclusion}{section*.2}% 16
\BOOKMARK [1][-]{section*.21}{Acknowledgments}{section*.2}% 17
\BOOKMARK [1][-]{section*.22}{References}{section*.2}% 18

View File

@ -512,6 +512,10 @@ by swapping $\ta^{\text{UHF}}$ and $\tb^{\text{UHF}}$ in Eqs.~\eqref{eq:ta_uhf}
Note that the RHF wave function remains a genuine solution of the HF equations for $U \ge 2t$, but corresponds to a saddle point
of the HF energy rather than a minimum.
%============================================================%
\subsection{Complex adiabatic connection}
%============================================================%
\titou{In a recent paper, \cite{Burton_2019} using holomorphic Hartree-Fock (h-HF) \cite{Hiscock_2014,Burton_2018,Burton_2016} as an analytic continuation of conventional HF theory, we have demonstrated, on a simple model, that one can interconvert states of different symmetries and natures by following well-defined contours in the complex $\lambda$-plane, where $\lambda$ is the strength of the electron-electron interaction (see Fig.~\ref{fig:iAC}).
In particular, by slowly varying $\lambda$ in a similar (yet different) manner to an adiabatic connection in density-functional theory, \cite{Langreth_1975,Gunnarsson_1976,Zhang_2004} one can ``morph'' a ground-state wave function into an excited-state wave function via a stationary path of HF solutions. \cite{Seidl_2018}
In such a way, we could obtain a doubly-excited state wave function starting from the ground state wave function, a process which is not as easy as one might think. \cite{Gilbert_2008,Thom_2008,Shea_2018}
@ -573,12 +577,71 @@ In order to better understand the behavior of the MP series and how it is connec
For small systems, one can access the whole terms of the series using full configuration interaction (FCI).
If the Hamiltonian $H(\lambda)$ is diagonalized in the FCI space, one gets the exact energies (in this finite Hilbert space) and the Taylor expansion with respect to $\lambda$ allows to access the MP perturbation series at any order.
Let us illustrate the behaviour of the RMP and UMP series on the Hubbard dimer (see Fig.~\ref{fig:RMP}).
Within the RMP partition technique, we have
\begin{equation}
\label{eq:H_RMP}
\bH_\text{RMP} =
\begin{pmatrix}
-2t + U - \lambda U/2 & 0 & 0 & \lambda U/2 \\
0 & U - \lambda U/2 & \lambda U/2 & 0 \\
0 & \lambda U/2 & U - \lambda U/2 & 0 \\
\lambda U/2 & 0 & 0 & 2t + U - \lambda U/2 \\
\end{pmatrix},
\end{equation}
which yields the ground-state energy
\begin{equation}
\label{eq:E0MP}
E_{-}(\lambda) = U - \frac{\lambda U}{2} - \frac{1}{2} \sqrt{(4t)^2 + \lambda ^2 U^2}.
\end{equation}
From this expression, it is easy to identify that the radius of convergence is defined by the EPs at $\lambda = \pm i 4t / U$ (which are actually located at the same position than the exact EPs discussed in Sec.~\ref{sec:example}).
Equation \eqref{eq:E0MP} can be Taylor expanded in terms of $\lambda$ to obtain the $n$th-order MP correction
\begin{equation}
E_\text{MP}^{(k)} = U \delta_{0,k} - \frac{1}{2} \frac{U^k}{(4t)^{k-1}} \mqty( 1/2 \\ k/2),
\end{equation}
with
\begin{equation}
E_{-}(\lambda) = \sum_{k=0}^\infty E_\text{MP}^{(k)} \lambda^k.
\end{equation}
We illustrate this by plotting the energy expansions at orders of $\lambda$ just below and above the radius of convergence in Fig.~\ref{fig:RMP}.
At the UMP level now, we have
\begin{widetext}
\begin{equation}
\label{eq:H_UMP}
\bH_\text{UMP} =
\begin{pmatrix}
-2t^2 \lambda/U & 0 & 0 & +2t^2 \lambda/U \\
0 & U - 2t^2 \lambda/U & +2t^2\lambda/U & +2t \sqrt{U^2 - (2t)^2} \lambda/U \\
0 & +2t^2\lambda/U & U - 2t^2 \lambda/U & -2t \sqrt{U^2 - (2t)^2} \lambda/U \\
+2t^2 \lambda/U & +2t \sqrt{U^2 - (2t)^2} \lambda/U & -2t \sqrt{U^2 - (2t)^2} \lambda/U & 2U(1-\lambda) + 6t^2\lambda/U \\
\end{pmatrix},
\end{equation}
\end{widetext}
A closed-form expression for the ground-state energy can be obtained but it is cumbersome, so we eschew reporting its expression.
The convergence of the UMP as a function of the ratio $U/t$ is depicted in Fig.~\ref{fig:UMP}.
%%% FIG 3 %%%
\begin{figure*}
\centering
\includegraphics[height=0.23\textwidth]{fig3a}
\hspace{0.05\textwidth}
\includegraphics[height=0.23\textwidth]{fig3b}
\hspace{0.05\textwidth}
\includegraphics[height=0.23\textwidth]{fig3c}
\caption{
Convergence of the URMP series as a function of the perturbation order $n$ energies for the Hubbard dimer at $U/t = 3$ and $7$.
The Riemann surfaces of the FCI energies are also represented for these two values of $U/t$.
\label{fig:UMP}}
\end{figure*}
Obviously, although practically convenient for electronic structure calculations, the MP partitioning is not the only possibility, and alternative partitioning have been proposed in the literature:
i) the Epstein-Nesbet (EN) partitioning which consists in taking the diagonal elements of $\hH$ as the zeroth-order Hamiltonian. \cite{Nesbet_1955,Epstein_1926}
Hence, the off-diagonal elements of $\hH$ are the perturbation operator,
ii) the weak correlation partitioning in which the one-electron part is consider as the unperturbed Hamiltonian $\hH^{(0)}$ and the two-electron part is the perturbation operator $\hV$, and
iii) the strong coupling partitioning where the two operators are inverted compared to the weak correlation partitioning. \cite{Seidl_2018}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Historical overview}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
@ -640,64 +703,6 @@ The authors illustrate this protocol on the dissociation curve of \ce{LiH} and t
\label{fig:RMP}}
\end{figure*}
Let us illustrate the behaviour of the RMP and UMP series on the Hubbard dimer (see Fig.~\ref{fig:RMP}).
Within the RMP partition technique, we have
\begin{equation}
\label{eq:H_RMP}
\bH_\text{RMP} =
\begin{pmatrix}
-2t + U - \lambda U/2 & 0 & 0 & \lambda U/2 \\
0 & U - \lambda U/2 & \lambda U/2 & 0 \\
0 & \lambda U/2 & U - \lambda U/2 & 0 \\
\lambda U/2 & 0 & 0 & 2t + U - \lambda U/2 \\
\end{pmatrix},
\end{equation}
which yields the ground-state energy
\begin{equation}
\label{eq:E0MP}
E_{-}(\lambda) = U - \frac{\lambda U}{2} - \frac{1}{2} \sqrt{(4t)^2 + \lambda ^2 U^2}.
\end{equation}
From this expression, it is easy to identify that the radius of convergence is defined by the EPs at $\lambda = \pm i 4t / U$ (which are actually located at the same position than the exact EPs discussed in Sec.~\ref{sec:example}).
Equation \eqref{eq:E0MP} can be Taylor expanded in terms of $\lambda$ to obtain the $n$th-order MP correction
\begin{equation}
E_\text{MP}^{(k)} = U \delta_{0,k} - \frac{1}{2} \frac{U^k}{(4t)^{k-1}} \mqty( 1/2 \\ k/2),
\end{equation}
with
\begin{equation}
E_{-}(\lambda) = \sum_{k=0}^\infty E_\text{MP}^{(k)} \lambda^k.
\end{equation}
We illustrate this by plotting the energy expansions at orders of $\lambda$ just below and above the radius of convergence in Fig.~\ref{fig:RMP}.
At the UMP level now, we have
\begin{widetext}
\begin{equation}
\label{eq:H_UMP}
\bH_\text{UMP} =
\begin{pmatrix}
-2t^2 \lambda/U & 0 & 0 & +2t^2 \lambda/U \\
0 & U - 2t^2 \lambda/U & +2t^2\lambda/U & +2t \sqrt{U^2 - (2t)^2} \lambda/U \\
0 & +2t^2\lambda/U & U - 2t^2 \lambda/U & -2t \sqrt{U^2 - (2t)^2} \lambda/U \\
+2t^2 \lambda/U & +2t \sqrt{U^2 - (2t)^2} \lambda/U & -2t \sqrt{U^2 - (2t)^2} \lambda/U & 2U(1-\lambda) + 6t^2\lambda/U \\
\end{pmatrix},
\end{equation}
\end{widetext}
A closed-form expression for the ground-state energy can be obtained but it is cumbersome, so we eschew reporting its expression.
The convergence of the UMP as a function of the ratio $U/t$ is depicted in Fig.~\ref{fig:UMP}.
%%% FIG 3 %%%
\begin{figure*}
\centering
\includegraphics[height=0.23\textwidth]{fig3a}
\hspace{0.05\textwidth}
\includegraphics[height=0.23\textwidth]{fig3b}
\hspace{0.05\textwidth}
\includegraphics[height=0.23\textwidth]{fig3c}
\caption{
Convergence of the URMP series as a function of the perturbation order $n$ energies for the Hubbard dimer at $U/t = 3$ and $7$.
The Riemann surfaces of the FCI energies are also represented for these two values of $U/t$.
\label{fig:UMP}}
\end{figure*}
%==========================================%
\subsection{Insights from a two-state model}
%==========================================%