\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{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{\bH}{\mathbf{H}} \newcommand{\bV}{\mathbf{V}} \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} %============================================================% \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 quantised 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 quantised 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 quantisation 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 quantised 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 realisation 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. 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) \cite{Heiss_1990, Heiss_1999, Heiss_2012, Heiss_2016} are non-Hermitian analogs of conical intersections (CIs) \cite{Yarkony_1996} where two states become exactly degenerate. CIs 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 CIs 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 \autoref{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 CIs only introduces a geometric phase while leaving the states unchanged. 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 \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=0.45\textwidth]{2x2.pdf} \includegraphics[width=0.45\textwidth]{i2x2.pdf} \caption{\centering Energies, as given by Eq.~\eqref{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.~\ref{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} \phi_{\pm}= \begin{pmatrix} E_{\pm}/\lambda \\ 1 \end{pmatrix}, \end{equation} and, for $\lambda=\lambda_\text{EP}$, they become \begin{equation} \phi_{\pm}=\begin{pmatrix} \mp i \\ 1 \end{pmatrix}, \end{equation} which are clearly self-orthogonal. \titou{what do you mean by self-orthogonal?} Using Eq.~\eqref{eq:E_EP}, Eq.~\eqref{eq:phi_2x2} can be recast as \begin{equation} \phi_{\pm}= \begin{pmatrix} E_\text{EP}/\lambda \pm \sqrt{2\lambda_\text{EP}/\lambda} \sqrt{1 - \lambda_\text{EP}/\lambda} \\ 1 \end{pmatrix}, \end{equation} where one clearly see that, if normalised, they behave as $(\lambda - \lambda_\text{EP})^{-1/4}$ resulting in 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} \titou{I don't think it is too obvious that the eigenvectors behave as $(\lambda - \lambda_\text{EP})^{-1/4}$.} 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 (time-independent) Rayleigh-Schr\"odinger perturbation theory, the Schr\"odinger equation \begin{equation} \label{eq:SchrEq} \bH \Psi = E \Psi \end{equation} is recast as \begin{equation} \label{eq:SchrEq-PT} \bH(\lambda) \Psi(\lambda) = (\bH^{(0)} + \lambda \bV ) \Psi(\lambda) = E(\lambda) \Psi(\lambda), \end{equation} where $\bH^{(0)}$ is the zeroth-order Hamiltonian and $\bV = \bH - \bH^{(0)}$ is the so-called perturbation. The ``physical'' system of interest is recovered by setting the coupling parameter $\lambda$ to unity. This decomposition is obviously non-unique and motivated by several factors as discussed below. Accordingly to Eq.~\eqref{eq:SchrEq-PT}, the energy can then be written as a power series of $\lambda$ \begin{equation} \label{eq:Elambda} E(\lambda) = \sum_{k=0}^\infty \lambda^k E^{(k)} \end{equation} However it is not guaranteed that the series \eqref{eq:Elambda} has a radius of convergence $\abs{\lambda_0} < 1$. In other words, the series might well be divergent for the physical system at $\lambda = 1$. One can prove that the actual value of the radius of convergence $\abs{\lambda_0}$ can be obtained by looking for the singularities of $E(\lambda)$ in the complex $\lambda$ plane. This is due to the following theorem \cite{Goodson_2012}: \begin{quote} \textit{``The Taylor series about a point $z_0$ of a function over the complex $z$ plane will converge at a value $z_1$ if the function is non-singular at all values of $z$ in the circular region centered at $z_0$ with radius $\abs{z_1 − z_0}$. If the function has a singular point $z_s$ such that $\abs{z_s − z_0} < \abs{z_1 − z_0}$, then the series will diverge when evaluated at $z_1$.''} \end{quote} This theorem means that the radius of convergence of the perturbation series is equal to the distance to the origin of the closest singularity of $E(\lambda)$. To illustrate this result we consider the simple function \eqref{eq:DivExample}. This function is smooth for $x \in \mathbb{R}$ and infinitely differentiable in $\mathbb{R}$. One would expect that the Taylor series of such a function would be convergent for all $x \in \mathbb{R}$, however this series is divergent for $x\geq1$. This is because the function has four singularities in the complex plane ($x = e^{i\pi/4}, e^{-i\pi/4}, e^{i3\pi/4}, e^{-i3\pi/4}$) with a modulus equal to 1. This simple example shows the importance of the singularities in the complex plane to understand the convergence properties on the real axis. \begin{equation} \label{eq:DivExample} f(x)=\frac{1}{1+x^4} \end{equation} \subsection{The Hartree-Fock Hamiltonian} In the Born-Oppenheimer approximation, the equation \eqref{eq:ExactHamiltonian} gives the exact electronic Hamiltonian for a chemical system with $n$ electrons and $N$ nuclei. The first term is the kinetic energy of the electrons, the two following terms account respectively for the electron-nuclei attraction and the electron-electron repulsion. \begin{equation}\label{eq:ExactHamiltonian} \bH=\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|}+\sum\limits_{i3/2$ an other stationary UHF solution appear, this solution is a minimum of the Hartree-Fock equations. This solution corresponds to the configuration with the electron $\alpha$ on one side of the sphere and the electron $\beta$ 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. This solution have the energy \eqref{eq:EsbUHF} for $R>3/2$. \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 so this wave function does not have the true symmetry. Indeed, the spherical harmonics are eigenvectors of $S^2$ but the symmetry-broken solution is a linear combination of the two eigenvectors and is not an eigenvector of $S^2$. However this solution gives more accurate results for the energy at large R as shown in \autoref{tab:ERHFvsEUHF}. In fact at the Coulson-Fischer point, it becomes more efficient to minimize the Coulomb repulsion than the kinetic energy in order to minimize the total energy. Thus the wave function break the spin symmetry because it allows a more efficient 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{\centering RHF and UHF energies in the minimal basis and exact energies for various R.} \begin{tabular}{c c c c c c c c c} $R$ & 0.1 & 1 & 2 & 3 & 5 & 10 & 100 & 1000 \\ \hline RHF & 10 & 1 & 0.5 & 0.3333 & 0.2 & 0.1 & 0.01 & 0.001 \\ \hline UHF & 10 & 1 & 0.490699 & 0.308532 & 0.170833 & 0.078497 & 0.007112 & 0.000703 \\ \hline Exact & 9.783874 & 0.852781 & 0.391959 & 0.247898 & 0.139471 & 0.064525 & 0.005487 & 0.000515 \\ \end{tabular} \label{tab:ERHFvsEUHF} \end{table} 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. At the critical value of R, the repulsion of the two electrons on the same side of the sphere maximizes more the energy than the kinetic energy of the p\textsubscript{z} orbitals. This symmetry breaking is associated with a charge density wave: the system oscillates between the situations with the electrons on each side \cite{GiulianiBook}. \begin{equation} E_{\text{sb-RHF}}=\frac{75}{88R^3}+\frac{25}{22R^2}+\frac{91}{66R} \end{equation} We can also consider negative value of R. This corresponds to the situation where one of the electrons is replaced by a positron. 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. Indeed, the sb-RHF states minimize the energy by placing the electron and the positron on the same side of the sphere. And the sb-UHF states maximize the energy because the two 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 \autoref{fig:SpheriumNrj}. The dotted curves corresponds to the holomorphic domain of the energies. \begin{figure}[h!] \centering \includegraphics[width=0.9\textwidth]{EsbHF.pdf} \caption{\centering 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{figure} \section{Radius of convergence and exceptional points} \subsection{Evolution of the radius of convergence} In this part, we will try to investigate how some parameters of $\bH(\lambda)$ influence the radius of convergence of the perturbation series. The radius of convergence is equal to the closest singularity to the origin of $E(\lambda)$ so we need to determine the locations of the exceptional points. To find them we solve simultaneously the equations \eqref{eq:PolChar} and \eqref{eq:DPolChar}. The equation \eqref{eq:PolChar} is the well-known secular equation giving the energies of the system, if an energy is also solution of \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 i.e. the exceptional points. \begin{equation}\label{eq:PolChar} \text{det}[E-\bH(\lambda)]=0 \end{equation} \begin{equation}\label{eq:DPolChar} \pdv{E}\text{det}[E-\bH(\lambda)]=0 \end{equation} The electron 1 have a spin $\alpha$ and the electron 2 a spin $\beta$. 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 is: \begin{align}\label{eq:rhfbasis} \psi_1 & =Y_{00}(\theta_1)Y_{00}(\theta_2), & \psi_2 & =Y_{00}(\theta_1)Y_{10}(\theta_2),\\ \psi_3 & =Y_{10}(\theta_1)Y_{00}(\theta_2), & \psi_4 & =Y_{10}(\theta_1)Y_{10}(\theta_2). \end{align} The Hamiltonian $\bH(\lambda)$ is block diagonal because of the symmetry of the basis set ($\psi_1$ only interacts with $\psi_4$ and $\psi_2$ with $\psi_3$). The two singly excited states yields a singlet and a triplet sp\textsubscript{z} but they don't have the appropriate symmetry so these states can't form exceptional points with the ground state. However there is an avoided crossing (see Fig. \ref{fig:RHFMiniBas}) between the s\textsuperscript{2} and the p\textsubscript{z}\textsuperscript{2} which is connected to two exceptional points in the complex plane. \begin{figure}[h!] \centering \includegraphics[width=0.7\textwidth]{EMP_RHF_R10.pdf} \caption{\centering Energies $E(\lambda)$ in the basis set \eqref{eq:rhfbasis} with $R=10$.} \label{fig:RHFMiniBas} \end{figure} To simplify the problem, it is convenient to only consider basis function with the symmetry of the exact wave function, such basis functions are called Configuration State Function (CSF). In this case the ground state is a totally symmetric singlet. According to the angular-momentum theory we expand the exact wave function in the following two-electron basis: \begin{equation} \Phi_l(\theta)=\frac{\sqrt{2l+1}}{4\pi R^2}P_l(\cos\theta) \end{equation} where $P_l$ are the Legendre polynomial and $\theta$ is the interelectronic angle. Then using this basis set we can compare the different partitioning of \autoref{sec:AlterPart}. The \autoref{fig:RadiusPartitioning} shows the evolution of the radius of convergence $R_{\text{CV}}$ in function of $R$ for the M{\o}ller-Plesset, the Epstein-Nesbet, the Weak Correlation and the Strong Coupling partitioning in a minimal basis i.e. $P_0$ and $P_1$ and in the three function basis. We see that the radius of convergence of the strong coupling partitioning is growing with R whereas it is decreasing for the three other partitioning. This result was expected because the three decreasing partitioning use a weakly correlated reference so $\bH^{(0)}$ is a good approximation for small $R$. On the contrary the strong coupling one uses a strongly correlated reference so this series converge better when the electron are strongly correlated i.e. when $R$ is large for the spherium model. The M{\o}ller-Plesset partitioning is always better than the weak correlation in \autoref{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 M{\o}ller-Plesset reference is proportionnal 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 n-th order energy of the weak correlation series can be obtained as a limited development of MP$n$ respective to $R$. It seems that the Epstein-Nesbet partitioning is better than the M{\o}ller-Plesset one for very small R in the minimal basis. In fact, it is just an artifact of the minimal basis because in the minimal basis augmented with $P_2 $ (and in larger basis set) the M{\o}ller-Plesset series converge faster for all value of $R$. \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{PartitioningRCV2.pdf} \includegraphics[width=0.45\textwidth]{PartitioningRCV3.pdf} \caption{\centering Radius of convergence in the minimal basis (left) and in the minimal basis augmented with $P_2$ (right) for different partitioning of the Hamiltonian$\bH(\lambda)$.} \label{fig:RadiusPartitioning} \end{figure} The \autoref{fig:RadiusBasis} shows that the radius of convergence is not very sensitive to the expansion of the basis set. The CSF basis function have all the same spin and spatial symmetry so we expect that the singularities obtained within this basis set will be $\alpha$ singularities. The \autoref{tab:SingAlpha} proves that the singularities considered in this case are $\alpha$ singularities. This is consistent with the observation of Goodson and Sergeev \cite{Goodson_2004} on $\alpha$ singularities. The discontinuities observed in \autoref{fig:RadiusBasis} with the M{\o}ller-Plesset partitioning are due to a change of the dominant singularity. \\ \begin{figure}[h!] \centering \includegraphics[width=0.45\textwidth]{MPlargebasis.pdf} \includegraphics[width=0.45\textwidth]{WCElargebasis.pdf} \caption{\centering Radius of convergence in the CSF basis with $n$ basis function for the MP partitioning (left) and the WC partitioning (right).} \label{fig:RadiusBasis} \end{figure} \begin{table}[h!] \centering \caption{\centering } \begin{tabular}{c} a \end{tabular} \label{tab:SingAlpha} \end{table} Now we will investigate the differences in the singularity structure between the RHF and UHF formalism. To do this we use the symmetry-broken orbitals obtained in \autoref{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{align*}\label{eq:uhforbitals} \phi_{\alpha,1}(\theta) & =\frac{\sqrt{75+62R}Y_{00}(\theta)+5\sqrt{-3+2R}Y_{10}(\theta)}{4\sqrt{7R}}, & \phi_{\beta,1}(\theta) & =\frac{\sqrt{75+62R}Y_{00}(\theta)-5\sqrt{-3+2R}Y_{10}(\theta)}{4\sqrt{7R}},\\ \phi_{\alpha,2}(\theta) & =\frac{-5\sqrt{-3+2R}Y_{00}(\theta)+\sqrt{75+62R}Y_{10}(\theta)}{4\sqrt{7R}}, & \phi_{\beta,2}(\theta) & =\frac{5\sqrt{-3+2R}Y_{00}(\theta)+\sqrt{75+62R}Y_{10}(\theta)}{4\sqrt{7R}}. \end{align*} In the UHF formalism the Hamiltonian $\bH(\lambda)$ is no more block diagonal, $\psi_4$ can interact with $\psi_2$ and $\psi_3$. The matrix element for this interaction is proportional to $\sqrt{-3+2R}$. For $R=3/2$ the Hamitonian is block diagonal and this is equivalent to the RHF case but for R>3/2 the matrix element become real. This interaction corresponds to the spin contamination of the wave function. For $R<3/2$ the matrix element is complex, this corresponds to the holomorphic solution of \autoref{fig:SpheriumNrj}, the singularities in this case will be treated in \autoref{sec:uhfSing}. The singularity structure in this case is more complex because of this spin contamination because we can't use configuration state function. So when we compute all the degeneracies using \eqref{eq:PolChar} and \eqref{eq:DPolChar} some corresponds to EPs and some 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. The \autoref{fig:UHFMiniBas} is the analog of \autoref{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 exceptional points resulting from those avoided crossings will be discussed in \autoref{sec:uhfSing}. \\ \begin{figure}[h!] \centering \includegraphics[width=0.7\textwidth]{EMP_UHF_R10.pdf} \caption{\centering Energies $E(\lambda)$ in the basis set \eqref{eq:rhfbasis} with $R=10$.} \label{fig:UHFMiniBas} \end{figure} In this study we have used spherical harmonics (or combination of spherical harmonics) as basis function which are diffuse wave functions. It would also be interesting to investigate the use of localized basis function \cite{Seidl_2018} (for example gaussians) because those functions would be more adapted to describe the correlated regime. \\ \subsection{Exceptional points in the UHF formalism}\label{sec:uhfSing} RHF vs UHF UHF spin contamination -> Riemann surfaces ?? UHF: CF point = QPT, ESQPT ??? PT broken symmetry sb UHF \section{Conclusion} \newpage \printbibliography \newpage \appendix \section{ERHF and EUHF} \end{document}