\documentclass[11pt,a4paper]{article} % définition des marges du document \setlength{\topmargin}{0cm} \setlength{\headheight}{0.4cm} \setlength{\headsep}{0.8cm} \setlength{\footskip}{1cm} \setlength{\textwidth}{17cm} \setlength{\textheight}{25cm} \setlength{\voffset}{-1.5cm} \setlength{\hoffset}{-0.5cm} \setlength{\oddsidemargin}{0cm} \setlength{\evensidemargin}{0cm} \usepackage[ backend=biber, style=phys, mcite=true, subentry, hyperref=true, ]{biblatex} \addbibresource{Rapport.bib} \usepackage{graphicx} % inclusion des figures \usepackage{tabularx} % gestion avancée des tableaux \usepackage{physics} \usepackage{wrapfig} \usepackage{amsmath} % collection de symboles mathématiques \usepackage{amssymb} % collection de symboles mathématiques \usepackage[utf8]{inputenc} \usepackage[T1]{fontenc} \DeclareUnicodeCharacter{2212}{-} \usepackage[english]{babel} \usepackage{siunitx} \usepackage[version=4]{mhchem} \usepackage{xcolor} % gestion de différentes couleurs \definecolor{linkcolor}{rgb}{0,0,0.6} % définition de la couleur des liens pdf \newcommand{\titou}[1]{\textcolor{red}{#1}} \newcommand{\antoine}[1]{\textcolor{orange}{#1}} \usepackage[ pdftex,colorlinks=true, pdfstartview=FitV, linkcolor= linkcolor, citecolor= linkcolor, urlcolor= linkcolor, hyperindex=true, hyperfigures=false] {hyperref} % fichiers pdf 'intelligents', avec des liens entre les références, etc. \usepackage{fancyhdr} % entêtes et pieds de pages personnalisés % définition de l'entête et du pied de page \pagestyle{fancy} \fancyhead[L]{\scriptsize \textsc{Perturbation theory in the complex plane}} \fancyhead[R]{\scriptsize \textsc{Antoine \textsc{MARIE}}} \fancyfoot[C]{ \thepage} \newcommand{\hH}{\Hat{H}} \newcommand{\hV}{\Hat{V}} \newcommand{\hI}{\Hat{I}} \newcommand{\hS}{\Hat{S}} \newcommand{\bH}{\mathbf{H}} \newcommand{\bV}{\mathbf{V}} \newcommand{\pt}{$\mathcal{PT}$} \begin{document} % Pour faciliter la mise en forme de la page du titre, on supprime l'indentation automatique en début de paragraphe \setlength{\parindent}{0pt} % Pas d'en-tête ni de pied pour la première page \thispagestyle{empty} \includegraphics[height=2cm]{logoens.eps} \hfill \includegraphics[height=2cm]{logoucbl.eps} \hfill \includegraphics[height=2cm]{logounivlyon.eps} \vspace{0.5cm} \begin{tabularx}{\textwidth}{@{} l X l @{} } {\sc Licence / Master Science de la matière} & & Stage 2019-2020 \\ {\it École Normale Supérieure de Lyon} & & Antoine \textsc{MARIE} \\ {\it Université Claude Bernard Lyon I} & & M1 Chimie \end{tabularx} \begin{center} \vspace{1.5cm} \rule[11pt]{5cm}{0.5pt} \textbf{\huge Pertubation theory in the complex plane} \rule{5cm}{0.5pt} \vspace{1.5cm} \parbox{15cm}{\small \textbf{Abstract} : \it In this work, we explore the extension of quantum chemistry in the complex plane. We observe that the physics of a quantum system is intimately connected to the position of the energy singularities in the complex plane. After a brief presentation of the fundamental notions of quantum chemistry and perturbation theory in the complex plane, we provide a historical overview of the various research activities that have been performed on the physic of singularities. Then we connect and further discuss these different aspects using the spherium model (i.e., two opposite-spin electrons restricted to remain on the surface of a sphere) as a theoretical playground. In particular, we explore various perturbative partitioning strategies and the effects of symmetry breaking on the singularity structure of the electronic energy. This provides fundamental insights on the location of these singularities in the complex plane (that one calls exceptional points) and, specifically, on the magnitude of the radius of convergence associated with the perturbative treatment. } \vspace{0.5cm} \parbox{15cm}{ \textbf{Keywords} : \it quantum chemistry, perturbation theory, complex plane, spherium, exceptional points, radius of convergence, symmetry breaking } %fin de la commande \parbox des mots clefs \vspace{0.5cm} \parbox{15cm}{ Internship supervisor: {\bf Pierre-François \textsc{LOOS}} \href{mailto:loos@irsamc.ups-tlse.fr}{\tt loos@irsamc.ups-tlse.fr} / tél. (+33) 5 61 55 73 39 Laboratoire de Chimie et Physique Quantiques {\it 118, route de Narbonne 31062 Toulouse - France} \url{https://www.irsamc.ups-tlse.fr/loos} } %fin de la commande \parbox encadrant / laboratoire d'accueil \vspace{0.5cm} \includegraphics[height=2cm]{LCPQ_logo.pdf} \hfill \includegraphics[height=2cm]{LogoCNRS.eps} \hfill \includegraphics[height=2cm]{UPS_logo.jpg} \end{center} \vfill \hfill \today \newpage \thispagestyle{empty} \setlength{\parindent}{17pt} \section*{Acknowledgments} \tableofcontents \newpage \setcounter{page}{1} %============================================================% \section{Introduction} \label{sec:intro} %============================================================% \subsection{Background} Due to the ubiquitous influence of processes involving electronic excited states in physics, chemistry, and biology, their faithful description from first principles has been one of the grand challenges faced by theoretical chemists since the dawn of computational chemistry. Accurately predicting ground- and excited-state energies (hence excitation energies) is particularly valuable in this context, and it has concentrated most of the efforts within the community. An armada of theoretical and computational methods have been developed to this end, each of them being plagued by its own flaws. The fact that none of these methods is successful in every chemical scenario has encouraged chemists to carry on the development of new excited-state methodologies, their main goal being to get the most accurate excitation energies (and properties) at the lowest possible computational cost in the most general context. One common feature of all these methods is that they rely on the notion of quantized energy levels of Hermitian quantum mechanics, in which the different electronic states of a molecule or an atom are energetically ordered, the lowest being the ground state while the higher ones are excited states. Within this quantized paradigm, electronic states look completely disconnected from one another. Many current methods study excited states using only ground-state information, creating a ground-state bias that leads to incorrect excitation energies. However, one can gain a different perspective on quantization extending quantum chemistry into the complex domain. In a non-Hermitian complex picture, the energy levels are \textit{sheets} of a more complicated topological manifold called \textit{Riemann surface}, and they are smooth and continuous \textit{analytic continuation} of one another. In other words, our view of the quantized nature of conventional Hermitian quantum mechanics arises only from our limited perception of the more complex and profound structure of its non-Hermitian variant. 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 \cite{Yarkony_1996}. Conical intersections are ubiquitous in non-adiabatic processes and play a key role in photo-chemical mechanisms. In the case of auto-ionizing resonances, EPs have a role in deactivation processes similar to conical intersections in the decay of bound excited states. Although Hermitian and non-Hermitian Hamiltonians are closely related, the behavior of their eigenvalues near degeneracies is starkly different. For example, encircling non-Hermitian degeneracies at EPs leads to an interconversion of states, and two loops around the EP are necessary to recover the initial energy (see Fig.~\ref{fig:TopologyEP} for a graphical example). Additionally, the wave function picks up a geometric phase (also known as Berry phase \cite{Berry_1984}) and four loops are required to recover the initial wave function. In contrast, encircling Hermitian degeneracies at conical intersections only introduces a geometric phase while leaving the states unchanged. More dramatically, whilst eigenvectors remain orthogonal at conical intersections, 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}. \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 \begin{equation} \label{eq:H_2x2} \bH = \begin{pmatrix} \epsilon_1 & \lambda \\ \lambda & \epsilon_2 \end{pmatrix}, \end{equation} which represents two states of energies $\epsilon_1$ and $\epsilon_2$ coupled with a strength of magnitude $\lambda$. This Hamiltonian could represent, for example, a minimal-basis configuration interaction with doubles (CID) for the hydrogen molecule \cite{SzaboBook}. \begin{figure}[h!] \centering \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} For real $\lambda$, the Hamiltonian \eqref{eq:H_2x2} is clearly Hermitian, and it becomes non-Hermitian for any complex $\lambda$ value. Its eigenvalues are \begin{equation} \label{eq:E_2x2} E_{\pm} = \frac{\epsilon_1 + \epsilon_2}{2} \pm \frac{1}{2} \sqrt{(\epsilon_1 - \epsilon_2)^2 + 4\lambda^2}, \end{equation} and they are represented as a function of $\lambda$ in Fig.~\ref{fig:2x2}. One notices that the two states become degenerate only for a pair of complex conjugate values of $\lambda$ \begin{equation} \label{eq:lambda_EP} \lambda_\text{EP} = \pm i\,\frac{\epsilon_1 - \epsilon_2}{2}, \quad \text{with energy} \quad E_\text{EP} = \frac{\epsilon_1 + \epsilon_2}{2}, \end{equation} which correspond to square-root singularities in the complex-$\lambda$ plane (see Fig.~\eqref{fig:2x2}). These two $\lambda$ values, given by Eq.~\eqref{eq:lambda_EP}, are the so-called EPs and one can clearly see that they connect the ground and excited states. Starting from $\lambda_\text{EP}$, two square-root branch cuts run on the imaginary axis towards $\pm i \infty$. In the real $\lambda$ axis, the point for which the states are the closest ($\lambda = 0$) is called an avoided crossing and this occurs at $\lambda = \Re(\lambda_\text{EP})$. The ``shape'' of the avoided crossing is linked to the magnitude of $\Im(\lambda_\text{EP})$: the smaller $\Im(\lambda_\text{EP})$, the sharper the avoided crossing is. Around $\lambda = \lambda_\text{EP}$, Eq.~\eqref{eq:E_2x2} behaves as \cite{MoiseyevBook} \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 \begin{equation} E_{\pm}(\theta) = E_\text{EP} \pm \sqrt{2\lambda_\text{EP} R} \exp(i\theta/2), \end{equation} and we have \begin{align} E_{\pm}(2\pi) & = E_{\mp}(0), & E_{\pm}(4\pi) & = E_{\pm}(0). \end{align} This evidences that encircling non-Hermitian degeneracies at EPs leads to an interconversion of states, and two loops around the EP are necessary to recover the initial energy. The eigenvectors associated to the eigenenergies \eqref{eq:E_2x2} are \begin{equation}\label{eq:phi_2x2} \begin{split} \phi_{\pm} & = \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{split} \end{equation} and, for $\lambda=\lambda_\text{EP}$, they become \begin{align} \phi_{\pm} & = \begin{pmatrix} -i \\ 1\end{pmatrix}, & \phi_{\pm} & = \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}(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. %============================================================% \section{Perturbation theory} %============================================================% \subsection{Rayleigh-Schr\"odinger perturbation theory} Within the Born-Oppenheimer approximation, \begin{equation}\label{eq:ExactHamiltonian} \hH = - \frac{1}{2} \sum_{i}^{n} \grad_i^2 - \sum_{i}^{n} \sum_{A}^{N} \frac{Z_A}{\abs{\vb{r}_i-\vb{R}_A}} + \sum_{i3/2$ the sb-UHF solution is the global minimum of the HF equations and the RHF solution presented before is a local minimum. This solution corresponds to the configuration with the spin-up electron in an orbital on one side of the sphere and the spin-down electron in a miror-image orbital on the opposite side and the configuration the other way round. The electrons can be on opposite sides of the sphere because the choice of p\textsubscript{z} as a basis function induced a privileged axis on the sphere for the electrons. For $R>3/2$, this solution has the energy \begin{equation}\label{eq:EsbUHF} E_{\text{sb-UHF}}=-\frac{75}{112R^3}+\frac{25}{28R^2}+\frac{59}{84R}. \end{equation} The exact solution for the ground state is a singlet. The spherical harmonics are eigenvectors of $\hS^2$ (the spin operator) and they are associated to different eigenvalues. Yet, the symmetry-broken orbitals are linear combinations of $Y_0$ and $Y_1$. Hence, the symmetry-broken orbitals are not eigenvectors of $\hS^2$. However, this solution gives lower energies than the RHF one at large $R$ as shown in Table \ref{tab:ERHFvsEUHF} even if it does not have the exact spin symmetry. In fact, at the Coulson-Fischer point, it becomes more effective to minimize the Coulomb repulsion than the kinetic energy in order to minimize the total energy. Thus, within the HF approximation, the variational principle is allowed to break the spin symmetry because it yields a more effective minimization of the Coulomb repulsion. This type of symmetry breaking is called a spin density wave because the system oscillates between the two symmetry-broken configurations \cite{GiulianiBook}. \begin{table}[h!] \centering \caption{RHF and sb-UHF energies in the minimal basis and exact energies (in the complete basis) for various $R$.} \begin{tabular}{ccccccccc} \hline \hline $R$ & 0.1 & 1 & 2 & 3 & 5 & 10 & 100 & 1000 \\ \hline RHF & 10.00000 & 1.000000 & 0.500000 & 0.333333 & 0.200000 & 0.100000 & 0.010000 & 0.001000 \\ UHF & 10.00000 & 1.000000 & 0.490699 & 0.308532 & 0.170833 & 0.078497 & 0.007112 & 0.000703 \\ Exact & 9.783874 & 0.852781 & 0.391959 & 0.247898 & 0.139471 & 0.064525 & 0.005487 & 0.000515 \\ \hline \hline \end{tabular} \label{tab:ERHFvsEUHF} \end{table} \begin{wrapfigure}{R}{0.5\textwidth} \centering \includegraphics[width=\linewidth]{EsbHF.pdf} \caption{Energies of the five solutions of the HF equations (multiplied by $R^2$). The dotted curves correspond to the analytic continuation of the symmetry-broken solutions.} \label{fig:SpheriumNrj} \end{wrapfigure} There is also another symmetry-broken solution for $R>75/38$ but this one corresponds to a maximum of the HF equations. This solution is associated with another type of symmetry breaking somewhat less known. It corresponds to a configuration where both electrons are on the same side of the sphere, in the same spatial orbital. This solution is called symmetry-broken RHF (sb-RHF). \titou{At a critical value of $R$, placing two electrons in the same orbital on the same side of the sphere increases the repulsion energy more than the kinetic energy of the two electrons in the p\textsubscript{z} orbital.} This configuration breaks the spatial symmetry of charge. Hence this symmetry breaking is associated with a charge density wave, the system oscillates between the situations with the two electrons on each side \cite{GiulianiBook}. The energy associated with this sb-RHF solution reads \begin{equation} E_{\text{sb-RHF}}=\frac{75}{88R^3}+\frac{25}{22R^2}+\frac{91}{66R}. \end{equation} We can also consider negative values of $R$, which corresponds to the situation where one of the electrons is replaced by a positron (the anti-particle of the electron) as readily seen in Eq.~\eqref{eq:H-sph-omega}. For negative $R$ values, there are also a sb-RHF ($R<-3/2$) and a sb-UHF ($R<-75/38$) solution for negative values of $R$ (see Fig.~\ref{fig:SpheriumNrj}) but in this case the sb-RHF solution is a minimum and the sb-UHF is a maximum of the HF equations. Indeed, the sb-RHF state minimizes the attraction energy by placing the electron and the positron on the same side of the sphere. And the sb-UHF state maximizes the energy because the two attracting particles are on opposite sides of the sphere. In addition, we can also consider the symmetry-broken solutions beyond their respective Coulson-Fischer points by analytically continuing their respective energies leading to the so-called holomorphic solutions \cite{Hiscock_2014, Burton_2019, Burton_2019a}. All those energies are plotted in Fig.~\ref{fig:SpheriumNrj}. The dotted curves corresponds to the holomorphic domain of the energies. \section{Radius of convergence and exceptional points} \subsection{Evolution of the radius of convergence} In this subsection, we investigate how \titou{parameters} of $\hH(\lambda)$ influence the radius of convergence of the perturbation series. Let us remind the reader that the radius of convergence is equal to the distance of the closest singularity to the origin of $E(\lambda)$. Hence, we have to determine the locations of the EPs to obtain information on the convergence properties of the perturbative series. To find them we solve simultaneously the following equations: \begin{subequations} \begin{align} \label{eq:PolChar} \det[E\hI-\hH(\lambda)] & = 0, \\ \label{eq:DPolChar} \pdv{E}\det[E\hI-\hH(\lambda)] & = 0, \end{align} \end{subequations} where $\hI$ is the identity operator. Equation \eqref{eq:PolChar} is the well-known secular equation giving the energies of the system. If an energy is also solution of Eq.~\eqref{eq:DPolChar} then this energy is degenerate. In this case the energies obtained are dependent of $\lambda$ so solving those equations with respect to $E$ and $\lambda$ gives the value of $\lambda$ where two energies are degenerate. These degeneracies can be conical intersections between two states with different symmetry for real value of $\lambda$ \cite{Yarkony_1996} or EPs between two states with the same symmetry for complex values of $\lambda$. Let us assume that electron 1 is spin-up and electron 2 is spin-down. Hence, we can forget the spin part of the spin-orbitals and from now on we will work with spatial orbitals. In the restricted formalism the spatial orbitals are the same so the two-electron basis set can be defined as \begin{align}\label{eq:rhfbasis} \psi_1 & =Y_{0}(\theta_1)Y_{0}(\theta_2), & \psi_2 & =Y_{0}(\theta_1)Y_{1}(\theta_2),\\ \psi_3 & =Y_{1}(\theta_1)Y_{0}(\theta_2), & \psi_4 & =Y_{1}(\theta_1)Y_{1}(\theta_2). \end{align} The Hamiltonian $\hH(\lambda)$ is block diagonal in this basis because of its symmetry, i.e., $\psi_1$ only interacts with $\psi_4$ and $\psi_2$ with $\psi_3$. The two singly-excited states yield after diagonalization a spatially anti-symmetric singlet sp\textsubscript{z} and a spatially symmetric triplet sp\textsubscript{z} state. Hence those states do not have the same symmetry as the spatially symmetric singlet ground state. Thus these states can not be involved in an avoided crossing with the ground state as can be seen in Fig.~\ref{fig:RHFMiniBas} and, \textit{a fortiori} cannot be involved in an EP with the ground state. However there is an avoided crossing between the s\textsuperscript{2} and p\textsubscript{z}\textsuperscript{2} states which gives two EPs in the complex plane. \begin{wrapfigure}{R}{0.5\textwidth} \centering \includegraphics[width=\linewidth]{EMP_RHF_R10.pdf} \caption{Energies $E(\lambda)$ in the restricted basis set \eqref{eq:rhfbasis} with $R=10$.} \label{fig:RHFMiniBas} \end{wrapfigure} To simplify the problem, it is convenient to only consider basis functions with the symmetry of the exact wave function, such basis functions are called configuration state functions (CSFs). It simplifies the problem because with such a basis set, one only gets the degeneracies of interest for the convergence properties, i.e., the EPs between states with the same symmetry as the ground state. In this case the ground state is a totally symmetric singlet. According to the angular momentum theory \cite{AngularBook, SlaterBook, Loos_2009}, we expand the exact wave function in the following two-electron basis: \begin{equation} \Phi_\ell(\omega)=\frac{\sqrt{2\ell+1}}{4\pi R^2}P_\ell(\cos\omega), \end{equation} where $P_\ell$ are the Legendre polynomial and $\omega$ is the interelectronic angle. Then, using this basis set we can compare the different partitioning of Sec.~\ref{sec:AlterPart}. Figure \ref{fig:RadiusPartitioning} shows the evolution of the radius of convergence $R_{\text{CV}}$ as a function of $R$ for the MP, the EN, the WC and the SC partitioning in a minimal basis (i.e., $P_0$ and $P_1$) and in the same basis augmented with $P_2$. We see that $R_{\text{CV}}$ for the SC partitioning increases with $R$ whereas it is decreasing for the three others partitioning. This result was expected because the three decreasing partitioning use a weakly correlated reference so $\hH^{(0)}$ is a good approximation for small $R$. On the contrary, the strong coupling one uses a strongly correlated reference so the SC series converges better when the electron are strongly correlated, i.e., when $R$ is large in the spherium model. The MP partitioning is always better than WC in Fig.~\ref{fig:RadiusPartitioning}. In the weak correlation partitioning the powers of $R$ are well-separated so each term of the series is a different power of $R$. Whereas the MP reference is proportional to $R^{-1}$ and $R^{-2}$ so the MP series is not well-defined in terms of powers of $R$. Moreover it can be proved that the $m$th order energy of the weak correlation series can be obtained as a Taylor approximation of MP$m$ respective to $R$. It seems that the EN partitioning is better than the MP one for very small $R$ in the minimal basis. In fact, it is just an artefact of the minimal basis because in the minimal basis augmented with $P_2$ (and in larger basis set) the MP series has a greater radius of convergence for all values of $R$. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{PartitioningRCV2.pdf} \includegraphics[width=0.45\textwidth]{PartitioningRCV3.pdf} \caption{Radius of convergence $R_{\text{CV}}$ for two (left) and three (right) basis functions for various partitionings.} \label{fig:RadiusPartitioning} \end{figure} Figure \ref{fig:RadiusBasis} shows that the radius of convergence is not very sensitive to the size of the basis set. The CSFs have all the same spin and spatial symmetries so we expect that the singularities obtained within this basis set will be $\alpha$ singularities. Table \ref{tab:SingAlpha} shows that the singularities considered in this case are indeed $\alpha$ singularities. This is consistent with the observation of Goodson and Sergeev \cite{Goodson_2004} who stated that $\alpha$ singularities are relatively insensitive to the basis set size. The discontinuities observed in Fig.~\ref{fig:RadiusBasis} for the MP partitioning are due to changes in dominant singularity. We can observe this change in Table \ref{tab:SingAlpha}, the value for $R=1$ and $R=2$ are respectively in the positive and negative plane. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{MPlargebasis.pdf} \includegraphics[width=0.45\textwidth]{WCElargebasis.pdf} \caption{Radius of convergence $R_{\text{CV}}$ in the CSF basis with $K$ basis functions for the MP (left) and WC (right) partitioning.} \label{fig:RadiusBasis} \end{figure} \begin{table}[h!] \centering \footnotesize \caption{Dominant singularity in the CSF basis set ($K=8$) for various value of $R$ in the MP and WC partitioning.} \begin{tabular}{cccccccc} \hline \hline $R$ & 0.1 & 1 & 2 & 3 & 5 & 10 & 100 \\ \hline MP & $+14.1-10.9\,i$ & $+2.38-1.47\,i$ & $-0.67-1.30\,i$ & $-0.49-0.89\,i$ & $-0.33-0.55\,i$ & $-0.22-0.31\,i$ & $+0.03-0.05\,i$ \\ WC & $-9.6-10.7\,i$ & $-0.96-1.07\,i$ & $-0.48-0.53\,i$ & $-0.32-0.36\,i$ & $-0.19-0.21\,i$ & $-0.10-0.11\,i$ & $-0.01-0.01\,i$ \\ \hline \hline \end{tabular} \label{tab:SingAlpha} \end{table} Now, we investigate the differences in the singularity structure between the RHF and UHF formalism. To do so, we use the symmetry-broken orbitals discussed in Sec.~\ref{sec:spherium}. Thus, the UHF two-electron basis is \begin{align}\label{eq:uhfbasis} \psi_1 & =\phi_{\alpha,1}(\theta_1)\phi_{\beta,1}(\theta_2), & \psi_2 & =\phi_{\alpha,1}(\theta_1)\phi_{\beta,2}(\theta_2),\\ \psi_3 & =\phi_{\alpha,2}(\theta_1)\phi_{\beta,1}(\theta_2), & \psi_4 & =\phi_{\alpha,2}(\theta_1)\phi_{\beta,2}(\theta_2). \end{align} with the symmetry-broken orbitals \begin{subequations} \begin{align}\label{eq:uhforbitals} \phi_{\alpha,1}(\theta) & =\frac{\sqrt{75+62R}Y_{0}(\theta)+5\sqrt{-3+2R}Y_{1}(\theta)}{4\sqrt{7R}}, \\ \phi_{\beta,1}(\theta) & =\frac{\sqrt{75+62R}Y_{0}(\theta)-5\sqrt{-3+2R}Y_{1}(\theta)}{4\sqrt{7R}}, \\ \phi_{\alpha,2}(\theta) & =\frac{-5\sqrt{-3+2R}Y_{0}(\theta)+\sqrt{75+62R}Y_{1}(\theta)}{4\sqrt{7R}}, \\ \phi_{\beta,2}(\theta) & =\frac{5\sqrt{-3+2R}Y_{0}(\theta)+\sqrt{75+62R}Y_{1}(\theta)}{4\sqrt{7R}}. \end{align} \end{subequations} In the UHF formalism the Hamiltonian $\hH(\lambda)$ is no more block diagonal, $\psi_4$ can interact with $\psi_2$ and $\psi_3$. The matrix elements of the Hamiltonian corresponding to this interaction are \begin{equation}\label{eq:MatrixElem} H_{24}=H_{34}=H_{42}=H_{43}=\sqrt{-3+2R}\sqrt{75+62R}\frac{25+2R}{280R^3} \end{equation} For $R=3/2$ the Hamiltonian is block diagonal and this is equivalent to the RHF case but for $R>3/2$ the matrix elements become real. This interaction corresponds to the spin contamination of the wave function. For $R<3/2$ the matrix elements are complex, this corresponds to the holomorphic solution of Fig.~\ref{fig:SpheriumNrj}, the singularities in this case will be treated in Sec.~\ref{sec:uhfSing}. The matrix elements become real again for $R<-75/62$, this corresponds to the sb-UHF solution for negative value of $R$ observed in Sec.~\ref{sec:spherium}. We will refer to the domain where the matrix elements are complex as the holomorphic domain. \begin{wrapfigure}{R}{0.5\textwidth} \centering \includegraphics[width=\linewidth]{EMP_UHF_R10.pdf} \caption{Energies $E(\lambda)$ in the unrestricted basis set \eqref{eq:uhfbasis} with $R=10$.} \label{fig:UHFMiniBas} \end{wrapfigure} The singularity structure in this case is more complex because of the spin contamination of the wave function. We can not use CSFs in this case. So when we compute all the degeneracies using Eqs.~\eqref{eq:PolChar} and \eqref{eq:DPolChar} some correspond to EPs and some correspond to conical intersections. The numerical distinction of those singularities is very difficult so we will first look at the energies $E(\lambda)$ obtained with this basis set. Figure \ref{fig:UHFMiniBas} is the analog of Fig.~\ref{fig:RHFMiniBas} in the UHF formalism. We see that in this case the sp\textsubscript{z} triplet interacts with the s\textsuperscript{2} and the p\textsubscript{z}\textsuperscript{2} singlets. Those avoided crossings are due to the spin contamination of the wave function. The EPs resulting from those avoided crossings will be discussed in Sec.~\ref{sec:uhfSing}. In this study we have used spherical harmonics (or combination of spherical harmonics) as basis functions which have a delocalized nature. It would also be interesting to investigate the use of localized basis functions \cite{Seidl_2018} (for example gaussians) because these functions would be more adapted to describe the strongly correlated regime. \subsection{Exceptional points in the UHF formalism}\label{sec:uhfSing} In the RHF case there are only $\alpha$ singularities and large avoided crossings but we can see in Fig.~\ref{fig:UHFMiniBas} that in the UHF case there are sharp avoided crossings which are known to be connected to $\beta$ singularities. For example at $R=10$ the pair of singularities connected to the avoided crossing between s\textsuperscript{2} and sp\textsubscript{z} $^{3}P$ is $0.999\pm0.014\,i$. And the one between sp\textsubscript{z} $^{3}P$ and p\textsubscript{z}\textsuperscript{2} is connected with the singularities $2.207\pm0.023\,i$. However, in the spherium the electrons can't be ionized so those singularities cannot be the same $\beta$ singularities as the ones highlighted by Sergeev and Goodson \cite{Sergeev_2005}. We can see in Fig.~\ref{fig:UHFEP} that the s\textsuperscript{2} singlet and the sp\textsubscript{z} triplet states are degenerated for $R=3/2$. For $R>3/2$, it becomes an avoided crossing on the real axis and the degeneracies are $moved$ in the complex plane. The wave function is spin contaminated by $Y_1$ for $R>3/2$ this why the s\textsuperscript{2} singlet energy can not cross the sp\textsubscript{z} triplet curves anymore. When $R$ increases this avoided crossing becomes sharper. As presented before $\beta$ singularities are linked to quantum phase transition so it seems that this singularity is linked to the spin symmetry breaking of the UHF wave function. The fact that a similar pair of $\beta$ singularities appears for $R<-75/62$ confirms this assumption. A second pair of $\beta$ singularities appear for $R\gtrsim 2.5$, this is probably due to an excited-state quantum phase transition but this still need to be investigated. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{UHFCI.pdf} \includegraphics[width=0.45\textwidth]{UHFEP.pdf} \caption{Energies $E(\lambda)$ in the unrestricted basis set \eqref{eq:uhfbasis} for $R=1.5$ (left) and $R=1.51$ (right).} \label{fig:UHFEP} \end{figure} As shown before, some matrix elements of the Hamiltonian become complex in the holomorphic domain. Therefore the Hamiltonian becomes non-Hermitian for those value of $R$. In \cite{Burton_2019a} Burton \textit{et al.}~proved that for the \ce{H_2} molecule the unrestricted Hamiltonian is not \pt -symmetric in the holomorphic domain. An analog reasoning can be done with the spherium model to prove the same result. The \pt -symmetry (invariance with respect to combined space reflection $\mathcal{P}$ and time reversal $\mathcal{T}$) is a property which ensures that a non-Hermitian Hamiltonian has a real energy spectrum. Thus \pt -symmetric Hamiltonian can be seen as an intermediate between Hermitian and non-Hermitian. Figure \ref{fig:UHFPT} shows that for the spherium model a part of the energy spectrum becomes complex when R is in the holomorphic domain. The parameter domain of value where the energy becomes complex is called the broken \pt symmetry region. This is consistent with the fact that in the holomorphic domain the Hamiltonian is no more \pt -symmetric. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{ReNRJPT.pdf} \includegraphics[width=0.45\textwidth]{ImNRJPT.pdf} \caption{\centering Real part (left) and imaginary part (right) of $E(\lambda)$ in the unrestricted basis set \eqref{eq:uhfbasis} for $R=1$.} \label{fig:UHFPT} \end{figure} For a non-Hermitian Hamiltonian the EPs can lie on the real axis. In particular, at the point of PT transition (the point where the energies become complex) the two energies are degenerate resulting in such an EP on the real axis. This degeneracy can be seen in Fig.~\ref{fig:UHFPT}. \section{Conclusion} In order to model accurately chemical systems, one must choose, in a ever growing zoo of methods, which computational protocol is adapted to the system of interest. This choice can be, moreover, motivated by the type of properties that one is interested in. That means that one must understand the strengths and weaknesses of each method, i.e., why one method might fail in some cases and work beautifully in others. We have seen that for methods relying on perturbation theory, their successes and failures are directly connected to the position of EPs in the complex plane. Exhaustive studies have been performed on the causes of failure of MP perturbation theory. First, it was understood that, for chemical systems for which the HF Slater determinant is a poor approximation to the exact wave function, MP perturbation theory fails too. Such systems can be, for example, molecules where the exact ground-state wave function is dominated by more than one configuration, i.e., multi-reference systems. More preoccupying cases were also reported. For instance, it has been shown that systems considered as well-understood (e.g., \ce{Ne}) can exhibit divergent behavior when the basis set is augmented with diffuse functions. Later, these erratic behaviors of the perturbation series were investigated and rationalized in terms of avoided crossings and singularities in the complex plane. It was shown that the singularities can be classified in two families. The first family includes $\alpha$ singularities resulting from a large avoided crossing between the ground state and a low-lying doubly-excited states. The $\beta$ singularities, which constitutes the second family, are artifacts generated by the incompleteness of the Hilbert space, and they are directly connected to an ionization phenomenon occurring in the complete Hilbert space. These singularities are close to the real axis and connected with sharp avoided crossing between the ground state and a highly diffuse state. We have found that the $\beta$ singularities modeling the ionization phenomenon described by Sergeev and Goodson are actually part of a more general class of singularities. Indeed, those singularities close to the real axis are connected to quantum phase transition and symmetry breaking, and theoretical physics have demonstrated that the behavior of the EPs depends of the type of transitions from which the EPs result (first or higher orders, ground state or excited state transitions). In this work, we have shown that $\beta$ singularities are involved in the spin symmetry breaking of the UHF wave function. This confirms that $\beta$ singularities can occur for other types of transition and symmetry breaking than just the formation of a bound cluster of electrons. It would be interesting to investigate the difference between the different type of symmetry breaking and how it affects the singularity structure. Moreover the singularity structure in the non-Hermitian case still need to be investigated. In the holomorphic domain, some singularities lie on the real axis and it would also be interesting to look at the differences between the different symmetry breaking and their respective holomorphic domain. To conclude, this work shows that our understanding of the singularity structure of the energy is still incomplete but we hope that it opens new perspectives for the understanding of the physics of EPs in electronic structure theory. \newpage \printbibliography \end{document}