From 53fbd8aa2eb891cc9c540e9ce0937a80c1062693 Mon Sep 17 00:00:00 2001 From: Pierre-Francois Loos Date: Tue, 28 Jul 2020 09:52:19 +0200 Subject: [PATCH] modifications in Sec. 1 --- RapportStage/Rapport.tex | 81 ++++++++++++++++++++++------------------ 1 file changed, 45 insertions(+), 36 deletions(-) diff --git a/RapportStage/Rapport.tex b/RapportStage/Rapport.tex index fb1d26b..a831cb7 100644 --- a/RapportStage/Rapport.tex +++ b/RapportStage/Rapport.tex @@ -25,6 +25,7 @@ \usepackage{tabularx} % gestion avancée des tableaux \usepackage{physics} +\usepackage{wrapfig} \usepackage{amsmath} % collection de symboles mathématiques \usepackage{amssymb} % collection de symboles mathématiques @@ -160,12 +161,17 @@ In other words, our view of the quantized nature of conventional Hermitian quant The realization that ground and excited states both emerge from one single mathematical structure with equal importance suggests that excited-state energies can be computed from first principles in their own right. One could then exploit the structure of these Riemann surfaces to develop methods that directly target excited-state energies without needing ground-state information. +\begin{wrapfigure}{r}{0.5\textwidth} + \centering + \includegraphics[width=\linewidth]{TopologyEP.pdf} + \caption{A generic EP with the square root branch point topology. A loop around the EP interconvert the states.} + \label{fig:TopologyEP} +\end{wrapfigure} By analytically continuing the electronic energy $E(\lambda)$ in the complex domain (where $\lambda$ is a coupling parameter), the ground and excited states of a molecule can be smoothly connected. This connection is possible because by extending real numbers to the complex domain, the ordering property of real numbers is lost. Hence, electronic states can be interchanged away from the real axis since the concept of ground and excited states has been lost. Amazingly, this smooth and continuous transition from one state to another has recently been experimentally realized in physical settings such as electronics, microwaves, mechanics, acoustics, atomic systems and optics \cite{Bittner_2012, Chong_2011, Chtchelkatchev_2012, Doppler_2016, Guo_2009, Hang_2013, Liertzer_2012, Longhi_2010, Peng_2014, Peng_2014a, Regensburger_2012, Ruter_2010, Schindler_2011, Szameit_2011, Zhao_2010, Zheng_2013, Choi_2018, El-Ganainy_2018}. - Exceptional points (EPs) are branch point singularities where two (or more) states become exactly degenerate \cite{Heiss_1990, Heiss_1999, Heiss_2012, Heiss_2016}. They are the non-Hermitian analogs of conical intersections (CIs) \cite{Yarkony_1996}. CIs are ubiquitous in non-adiabatic processes and play a key role in photo-chemical mechanisms. @@ -177,13 +183,6 @@ In contrast, encircling Hermitian degeneracies at CIs only introduces a geometri More dramatically, whilst eigenvectors remain orthogonal at CIs, at non-Hermitian EPs the eigenvectors themselves become equivalent, resulting in a \textit{self-orthogonal} state \cite{MoiseyevBook}. More importantly here, although EPs usually lie off the real axis, these singular points are intimately related to the convergence properties of perturbative methods and avoided crossing on the real axis are indicative of singularities in the complex plane \cite{Olsen_1996, Olsen_2000}. -\begin{figure}[h!] - \centering - \includegraphics[width=0.7\textwidth]{TopologyEP.pdf} - \caption{\centering A generic EP with the square root branch point topology. A loop around the EP interconvert the states.} - \label{fig:TopologyEP} -\end{figure} - \subsection{An illustrative example} In order to highlight the general properties of EPs mentioned above, we propose to consider the following $2 \times 2$ Hamiltonian commonly used in quantum chemistry @@ -200,9 +199,10 @@ This Hamiltonian could represent, for example, a minimal-basis configuration int \begin{figure}[h!] \centering - \includegraphics[width=0.45\textwidth]{2x2.pdf} - \includegraphics[width=0.45\textwidth]{i2x2.pdf} - \caption{\centering Energies, as given by Eq.~\ref{eq:E_2x2}, of the Hamiltonian \eqref{eq:H_2x2} as a function of $\lambda$ with $\epsilon_1 = -1/2$ and $\epsilon_2 = +1/2$.} + \includegraphics[width=0.3\textwidth]{2x2.pdf} + \hspace{0.2\textwidth} + \includegraphics[width=0.3\textwidth]{i2x2.pdf} + \caption{Energies, as given by Eq.~\ref{eq:E_2x2}, of the Hamiltonian \eqref{eq:H_2x2} as a function of $\lambda$ with $\epsilon_1 = -1/2$ and $\epsilon_2 = +1/2$.} \label{fig:2x2} \end{figure} @@ -232,7 +232,7 @@ Around $\lambda = \lambda_\text{EP}$, Eq.~\eqref{eq:E_2x2} behaves as \cite{Mois \begin{equation} \label{eq:E_EP} E_{\pm} = E_\text{EP} \pm \sqrt{2\lambda_\text{EP}} \sqrt{\lambda - \lambda_\text{EP}}, \end{equation} -and following a complex contour around the EP, i.e.~$\lambda = \lambda_\text{EP} + R \exp(i\theta)$, yields +and following a complex contour around the EP, i.e., $\lambda = \lambda_\text{EP} + R \exp(i\theta)$, yields \begin{equation} E_{\pm}(\theta) = E_\text{EP} \pm \sqrt{2\lambda_\text{EP} R} \exp(i\theta/2), \end{equation} @@ -246,38 +246,47 @@ This evidences that encircling non-Hermitian degeneracies at EPs leads to an int The eigenvectors associated to the eigenenergies \eqref{eq:E_2x2} are \begin{equation}\label{eq:phi_2x2} -\phi_{\pm}=\begin{pmatrix} -(\epsilon_1 - \epsilon_2 \pm \sqrt{(\epsilon_1 - \epsilon_2)^2 + 4\lambda^2})/2\lambda \\ 1 -\end{pmatrix}, -\end{equation} -and, for $\lambda=\lambda_\text{EP}$, they become -\begin{align} - \phi_{\pm}(i\,\frac{\epsilon_1 - \epsilon_2}{2}) & = \begin{pmatrix} -i \\ 1\end{pmatrix}, - & - \phi_{\pm}(-i\,\frac{\epsilon_1 - \epsilon_2}{2}) & = \begin{pmatrix} i \\ 1\end{pmatrix}. -\end{align} -which are self-orthogonal i.e. their norm is equal to zero. -Using Eq.~\eqref{eq:E_EP}, Eq.~\eqref{eq:phi_2x2} can be recast as -\begin{equation}\label{eq:phi_EP} -\phi_{\pm}= +\begin{split} + \phi_{\pm}(\lambda) + & = + \begin{pmatrix} + (\epsilon_1 - \epsilon_2 \pm \sqrt{(\epsilon_1 - \epsilon_2)^2 + 4\lambda^2})/2\lambda + \\ + 1 + \end{pmatrix} + \\ + & = \begin{pmatrix} (\epsilon_1-\epsilon_2)/2\lambda \pm \sqrt{2\lambda_\text{EP}} \sqrt{\lambda - \lambda_\text{EP}}/\lambda \\ 1 - \end{pmatrix}. + \end{pmatrix}, +\end{split} \end{equation} -We have seen that the EP inherits its topology from the double valued function $\sqrt{\lambda - \lambda_\text{EP}}$. Then if the eigenvectors are normalised they behave as $(\lambda - \lambda_\text{EP})^{-1/4}$ which gives the following pattern when looping around one EP: +and, for $\lambda=\lambda_\text{EP}$, they become +\begin{align} + \phi_{\pm}\qty(i\,\frac{\epsilon_1 - \epsilon_2}{2}) & = \begin{pmatrix} -i \\ 1\end{pmatrix}, + & + \phi_{\pm}\qty(-i\,\frac{\epsilon_1 - \epsilon_2}{2}) & = \begin{pmatrix} i \\ 1\end{pmatrix}. +\end{align} +which are clearly self-orthogonal, i.e., their norm is equal to zero. +%Using Eq.~\eqref{eq:E_EP}, Eq.~\eqref{eq:phi_2x2} can be recast as +%\begin{equation}\label{eq:phi_EP} +%\phi_{\pm}= +%\end{equation} +As branch point (square-root) singularities, EPs inherit their topology from the multi-valued function $\sqrt{\lambda - \lambda_\text{EP}}$. +Then, if the eigenvectors are properly normalized, they behave as $(\lambda - \lambda_\text{EP})^{-1/4}$, which gives the following pattern when looping around one EP: \begin{align} \phi_{\pm}(2\pi) & = \phi_{\mp}(0), & - \phi_{\pm}(4\pi) & = -\phi_{\pm}(0) \\ + \phi_{\pm}(4\pi) & = -\phi_{\pm}(0), + \\ \phi_{\pm}(6\pi) & = -\phi_{\mp}(0), & \phi_{\pm}(8\pi) & = \phi_{\pm}(0). \end{align} In plain words, four loops around the EP are necessary to recover the initial state. We can also see that looping the other way around leads to a different pattern. -\antoine{Is this a bit better like this ?} %============================================================% \section{Perturbation theory} @@ -345,7 +354,7 @@ K_j(1)\phi_i(1)=\left[\int\dd\vb{x}_2\phi_j^*(2)\frac{1}{r_{12}}\phi_j(2) \right \subsection{M{\o}ller-Plesset perturbation theory} -The Hartree-Fock Hamiltonian \eqref{eq:HFHamiltonian} can be used as the zeroth-order Hamiltonian of the equation \eqref{eq:SchrEq-PT}. This partitioning of the Hamiltonian leads to the so-called M{\o}ller-Plesset (MP) perturbation theory \cite{Moller_1934}. The discovery of a partitioning of the Hamiltonian that allowed chemists to recover a part of the correlation energy (i.e. the difference between the exact energy and the Hartree-Fock energy) using perturbation theory has been a major step in the development of post-Hartree-Fock methods. This yields the Hamiltonian $\bH(\lambda)$ of the equation \eqref{eq:MPHamiltonian} where the two-electron part of the Fock operator $f(i)$ has been written $V_i^{HF}$ for convenience. +The Hartree-Fock Hamiltonian \eqref{eq:HFHamiltonian} can be used as the zeroth-order Hamiltonian of the equation \eqref{eq:SchrEq-PT}. This partitioning of the Hamiltonian leads to the so-called M{\o}ller-Plesset (MP) perturbation theory \cite{Moller_1934}. The discovery of a partitioning of the Hamiltonian that allowed chemists to recover a part of the correlation energy (i.e., the difference between the exact energy and the Hartree-Fock energy) using perturbation theory has been a major step in the development of post-Hartree-Fock methods. This yields the Hamiltonian $\bH(\lambda)$ of the equation \eqref{eq:MPHamiltonian} where the two-electron part of the Fock operator $f(i)$ has been written $V_i^{HF}$ for convenience. \begin{equation}\label{eq:MPHamiltonian} \bH(\lambda)=\sum\limits_{i=1}^{n}\left[-\frac{1}{2}\grad_i^2 - \sum\limits_{k=1}^{N} \frac{Z_k}{|\vb{r}_i-\vb{R}_k|} + (1-\lambda)V_i^{\text{HF}}+\lambda\sum\limits_{i