EFFECTIVE BETA–DECAY OPERATORS WITH COUPLED CLUSTER THEORY By Samuel John Novario A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics—Doctor of Philosophy 2018 ABSTRACT EFFECTIVE BETA–DECAY OPERATORS WITH COUPLED CLUSTER THEORY By Samuel John Novario Coupled Cluster theory is a powerful ab initio framework for solving the many-body the Schrödinger equation and has been utilized successfully to describe the highly-correlated systems found in quantum chemistry and nuclear physics. This method uses a special similarity transformation to decouple a system’s ground state from excitations from it. This transformation contains significant correlations that can be used to extend coupled cluster theory to excited states and open-shell systems with the equation-of-motions method. Additionally, properties of these states can be obtained by consistently transforming relevant operators using the coupled cluster similarity transformation. The coupled cluster method is systematically improvable and scales polynomially with the system size. With this flexibility and reach, coupled cluster theory can be applied across the nuclear chart to contribute to many important open problems in physics. Several fundamental questions in modern physics involve electroweak interactions within nuclei, including the search for the elusive neutrinoless double-beta decay. Often the largest uncertainty within these experiments is due to nuclear-structure-dependent quantities that are calculated within some many-body framework. The main focus of this thesis is to apply the coupled cluster method to calculate effective Fermi and Gamow-Teller beta-decay operators between open shell states. By confirming the validity of this method, it can be extended to double-beta decay and other electroweak processes. ACKNOWLEDGMENTS First and foremost, I would like to thank my parents, Amy and Steve, for their unconditional support of my educational aspirations and passionate cultivation of my curiosities in addition to their persistent emotional, financial, and logistical aid they have provided during my long journey up till this point. Also, I would like to thank my siblings, Maggie, Lucy, and Charlie, for their friendship and inspiration, as well as the examples they set which I was able to follow. Moreover, I would like to thank my grandparents, aunts, uncles and cousins for their words of encouragement and praise. Lastly, I would like to thank my friends in East Lansing, Illinois, and around the country for helping me have fun and maintain my peace of mind, and never failing to ask me when I would graduate. Next, I would like to thank my all my teachers throughout my long tenure as a student. I would especially like to acknowledge my science and math teachers that fostered and encouraged my interests. In particular, I would like to credit Bryon Leonard for opening my mind to the beauty of modern physics and the path that leads to where I am now. Additionally, I am thankful for Philippe Collon and Michael Thoennessen for giving me my first research opportunities as an undergraduate student when I’m sure I was more of a burden than a blessing. I am grateful to Dr. Georg Bollen and Dr. Ryan Ringle for inviting me into the LEBIT collaboration, my first academic family. Along with my fellow graduate students, David Lincoln, Scott Bustabad, and Brad Barquest, my LEBIT colleagues helped me find my place as a graduate student, guided me as a scientist and collaborator, provided opportunities to travel and present at national conferences, and gave me research experience that few theorists hold. While in LEBIT, I was fortunate to share an office with Jenna Smith, who looked over iii me as a young graduate student and gave me valuable advice and encouragement when it was greatly needed. When I mentioned that I decided to change my research direction, my current advisor, Morten Hjorth-Jensen, without hesitation and without any solicitation from me, took a risk by taking me on as a student. For this, I am eternally grateful, and because Morten has become an invaluable mentor, colleague, and friend, I feel very lucky for finding myself in that situation. Along with my advisor, I would like to thank the members of my committee– Scott Bogner, Piotr Piecuch, Carlo Piermarocchi, and Remco Zegers–for agreeing to join my adventure through graduate school and guiding me along the way. I’m also indebted to Heiko Hergert for fielding my relentless questions, Ragnar Stroberg for unknowingly helping with his clean and well-commented code, and Kim Crosslan for shouldering most of my blunders for more than 7 years and for making sure I was aware of them. Lasly, I would like to thank my fellow graduate students. To Justin Lietz, my office mate, world traveling partner, coupled-cluster colleague, and friend, I’m thankful for lending his advice, his expertise, and his ear. To Nathan Parzchuowki, Titus Morris, and Fei Yuan, my fellow nuclear theory colleagues and friends, I’m thankful for their constant help and the pleasant work environment that they created. And to all of the people that I haven’t mentioned, I realize that I am just a product of the circumstances that unfolded around me, especially the people with which I’ve crossed paths. So I’d like to give thanks to everyone, and to no one in particular, who helped shape my success in completing this dissertation. iv TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii KEY TO SYMBOLS AND ABBREVIATIONS . . . . . . . . . . . . . . . . . xiii Chapter 1 Introduction . . . . . . . . . . . . . . 1.1 A Brief History of Nuclear Structure Theory 1.2 Electroweak Theory and Nuclear Structure . 1.3 Ab-Initio Descriptions of Beta Decay . . . . 1.4 Thesis Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 2 4 6 7 Chapter 2 Many-Body Quantum Mechanics 2.1 Independent-Particle Model . . . . . . . . 2.2 Second Quantization . . . . . . . . . . . . 2.3 Normal Ordering . . . . . . . . . . . . . . 2.4 Wick’s Theorem . . . . . . . . . . . . . . . 2.5 Hartree–Fock Method . . . . . . . . . . . . 2.6 Configuration-Interaction . . . . . . . . . . 2.7 Many-Body Perturbation Theory . . . . . 2.7.1 Factorization Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 8 10 12 15 20 25 27 31 Chapter 3 Coupled-Cluster Theory . . . . . . 3.1 Exponential Ansatz . . . . . . . . . . . . . . 3.1.1 The Coupled Cluster Equations . . . 3.2 Linked-Cluster Theorem and MBPT . . . . 3.3 Example: Pairing Model . . . . . . . . . . . 3.4 Solving the Coupled Cluster Equations . . . 3.4.1 Symmetry Channels . . . . . . . . . 3.4.2 Matrix Structures and Intermediates 3.5 Example: Homogeneous Electron Gas . . . . 3.6 Coupled Cluster for Finite Nuclei . . . . . . 3.6.1 Harmonic Oscillator Basis . . . . . . 3.6.2 The Nuclear Interaction . . . . . . . 3.6.3 Ground-State Results for Nuclei . . . 3.7 Ground-State Center-of-Mass Factorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 34 37 38 43 48 49 51 57 62 62 65 68 71 Chapter 4 Equation-of-Motion Method . 4.1 Equation-of-Motion States . . . . . . . 4.2 Dual Solutions . . . . . . . . . . . . . . 4.2.1 Induced Three-Body Interaction 4.3 Solving the EOM equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 77 80 82 83 v . . . . . . . . . . . . . . . 4.4 4.5 4.6 EOM-CC for 2D Quantum Dots 4.4.1 Quantum Dot Formalism 4.4.2 Quantum Dot Results . Quality of EOM Solutions . . . EOM-CC for Finite Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 90 92 93 96 Chapter 5 Beta–Decay Effective Operators . . . . . . . . . . . . . . . . . . . 103 5.1 Beta-Decay Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 5.2 Coupled Cluster Effective Operators . . . . . . . . . . . . . . . . . . . . . . 108 Chapter 6 Conclusions and Perspectives . . . . . . . . . . . . . . . . . . . . . 114 6.1 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 APPENDICES . . . . . . . . . . . . . . . . . Appendix A CCSD Diagrams . . . . . . . . Appendix B Effective Hamiltonian Diagrams Appendix C Computational Implementation Appendix D Angular Momentum Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 117 124 131 142 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 vi LIST OF TABLES Table 5.1: Summary of the selection rules for allowed beta decays according to the angular momentum J and parity π of the initial (I) and final (F ) states. vii 106 LIST OF FIGURES Figure 1.1: Figure 1.2: Figure 2.1: Figure 2.2: Figure 2.3: Figure 3.1: Figure 3.2: Figure 3.3: Nuclear chart of nuclei with ground-state energies which have been calculated with ab-initio methods and NN+3N interactions. Figure taken from [1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 Progress of ab-initio nuclear structure from calculations of ground-state energies with NN+3N interactions. Early progress was approximately linear as the problem size scaled with Moore’s law while more recent progress has taken advantage of new algorithms which have outpaced Moore’s law. Data taken from [1]. . . . . . . . . . . . . . . . . . . . . . 5 A depiction of the closed-shell reference state in the independent particle model. Each horizontal line represents a shell of single-particle orbits, represented by circles, and the dotted line represents the Fermi level which separates the unoccupied particle states from the occupied hole states. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 A depiction of 1p -1h, 2p -2h, 1p -0h, and 0p -1h Slater determinants defined relative to the reference state in the independent particle model. . 14 Scaling of the matrix size and number of non-zero matrix elements for nuclear CI calculations of light nuclei. Even for modestly-sized model spaces, the memory requirements approach the limit of petascale supercomputers (∼ 1010 ). Figure taken from [2]. . . . . . . . . . . . . . . . . 27 Schematic representation of the pairing model space. The shells are equally spaced and doubly degenerate with one spin-up and one spindown state. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 Correlation energy for the pairing model with exact diagonalization, CCD, and perturbation theory to third (MBPT3) and fourth order (MBPT4) for a range of interaction values, g. . . . . . . . . . . . . . . . . . . . . . 46 Visualization of the CCD similarity transform on the pairing Hamiltonian for four particles and six shells. This shows the main function of CCD, which is to decouple 2p -2h excitations from the ground state, shown by the suppression of matrix elements on the first column. In the pairing model, this also has the effect of decoupling 2p -2h excitations from 4p -4h excitations. Also, the non-unitary nature of the transformation is obvious given the asymmetry of the resulting Hamiltonian. . . . . . . . . . . . . 48 viii Figure 3.4: Figure 3.5: Visulization of the Fourier transform of a finite box. This transformation characterizes the construction of the single-particle basis for infinite matter, mapping plane waves in coordinate space onto finitely-spaced points in momentum space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 CCD energy per electron in Hartrees for the 3D homogeneous electron gas as function of the Wigner-Seitz radius in units of Bohr radii. The calculation used periodic boundary conditions and a basis with 25 shells, resulting in a total of 1238 single-particle states. Also plotted are the variational quantum Monte Carlo (VMC) results from [3]. . . . . . . . . 61 Figure 3.6: A schematic illustration of the harmonic oscillator basis used for calculations of nuclei. Shown is an example of a initial reference state for carbon14, with 6 protons filled to the p3/2-subshell closure and 8 neutrons filled to the p1/2-shell closure. See text for details on the single-particle states. 64 Figure 3.7: Diagrammatic form of the chiral EFT expansion up to N3 LO. The solid lines represent nucleons and the dashed lines represent pions. The different vertices represent higher-order interactions. Figure taken from [4]. . 67 Ground-state energies for 16 O for the EM N3 LO NN only interaction and with the added 3N interaction from Navrátil, both SRG softened with λSRG = 1.88, 2.24 fm−1 . The energies are plotted for emax = 10, 12. The most obvious difference is between the NN and NN+3N calculations, showing the importance of including 3N forces. The differences between the cutoff parameters are resolved within ∼ 1% with the inclusion of 3N forces and can be rectified further by including additional correlations or full 3N forces. The experimental binding energy is shown with the grey dashed line. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 Ground-state energies for doubly magic nuclei as a function of the harmonic oscillator energy ~ω with the NN+3N(400) interaction, SRG softened with λSRG = 2fm−1 . The energies are plotted for emax = 8, 10, 12, showing the convergence as the model space increases. The results are independent of the underlying oscillator frequency to ∼ 1% for emax = 12. The grey dashed line is the experimental binding energy. The overbinding of this interaction becomes apparent as the system size increases. . . . . 70 Figure 3.10: Ground-state energies for singly magic nuclei as a function of the harmonic oscillator energy ~ω with the NN+3N(400) interaction, SRG softened with λSRG = 2fm−1 . The energies are plotted for different emax . The results are independent of the underlying oscillator frequency to ∼ 1% for emax = 12. The grey dashed line is the experimental binding energy. These results underbind with respect to their doubly-magic counterparts in Fig. 3.9. . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 Figure 3.8: Figure 3.9: ix Figure 3.11: Ground-state COM energies, Eq. (3.68), for 16 O and 40 Ca at various harmonic oscillator frequencies with the NN+3N(400)-induced interaction with λSRG = 2.0 fm−1 at emax = 12. Using the proper COM oscillator frequencies shows the approximate factorization of Eq. (3.67). . . . . . . Figure 4.1: Figure 4.2: 75 The 42 lowest single-particle states (the first 5 shells) in the 2D harmonic oscillator basis. Each box represents a single-particle state arranged by m` , ms , and energy, and the up/down arrows indicate the spin of the states. Within each column, the principal quantum number n increases as one traverses upward. . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 Ground state energy (in Hartrees) of quantum dots with N particles and an oscillator frequency of ω calculated with several different methods. . 93 Figure 4.3: Particle-attached energy (in Hartrees) of quantum dots with N particles and an oscillator frequency of ω calculated with several different methods. 94 Figure 4.4: Particle-removed energy (in Hartrees) of quantum dots with N particles and an oscillator frequency of ω calculated with several different methods. 94 Figure 4.5: Energy difference of particle-attached and particle-removed states between the EOM-CC method and the exact FCI method for a quantum dot with various parameters plotted against the single-particle overlap of the FCI state, n1-particle or n1-hol , see Eqs. (4.42) and (4.43). The strong correlation shows that the quality of EOM states can be judged by this metric. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 Ground-state energies for the particle-attached nuclei 17 O, 17 F, 23 O, and 23 F as a function of the harmonic oscillator energy ~ω with the NN+3N(400) interaction, SRG softened with λSRG = 2fm−1 . The energies are plotted for emax = 8, 10, 12, showing the convergence as the model space increases. The grey dashed line is the experimental binding energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 Ground-state energies for the particle-removed nuclei 15 N, 15 O, 21 N, and 21 O as a function of the harmonic oscillator energy ~ω with the NN+3N(400)-induced interaction, SRG softened with λSRG = 2fm−1 . The energies are plotted for emax = 8, 10, 12, showing the convergence as the model space increases. The grey dashed line is the experimental binding energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 Figure 4.6: Figure 4.7: x Figure 4.8: Ground-state COM energies, Eq. (3.68), for open-shell nuclei at varies harmonic oscillator frequencies with the NN+3N(400)-induced interaction with λSRG = 2.0 fm−1 at emax = 12. The top row shows the results for the particle-attached nuclei 17 O, 17 F, 23 O, and 23 F, and the bottom row shows the results for the particle-removed nuclei 15 N, 15 O, 21 N, and 21 O. The right column shows that the COM wave function practically vanishes according to Eq. (3.67). . . . . . . . . . . . . . . . . . . . . . . 100 Figure 4.9: Low-lying PA-EOM states for 17 O and 17 F with and without a LawsonGloeckner term, along with the experimentally determined spectra. The negative-partity states are COM contaminants and are removed by artificially raising the COM excitation energy with the parameter β according to Eq. (3.72). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 Figure 4.10: Low-lying PA-EOM states for 23 O and 23 F with and without a LawsonGloeckner term. The negative-partiy states in 17 O are COM contaminants and are removed by artificially raising the COM excitation energy with the Lawson-Gloeckner method, Eq. (3.72). The excited states of these nuclei have not been experimentally detemined. . . . . . . . . . . 101 Figure 4.11: Low-lying PR-EOM states for 15 N and 15 O with and without a lawsongloeckner term, and the experimentally determined spectra. The 1/2+ state is a COM contaminant and is removed by artificially raising the COM excitation energy with the parameter β according to Eq. (3.72). . 102 Figure 4.12: Low-lying PA-EOM states for 21 N and 21 O with and without a LawsonGloeckner term, Eq. (3.72), and the experimentally determined spectra. The negative-partiy states in 17 O are COM contaminants and are removed by artificially raising the COM excitation energy. The excited states of these nuclei have not been experimentally detemined. . . . . . 102 Figure 5.1: Schematic representations of the three free-space weak processes in this work: β − decay (a), β + decay (b), and electron-capture (c). The coupling constant for the point interaction vertex is the weak-interaction coupling constant gW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 Figure 5.2: Schematic representations of a higher-order weak interactions involving pion-exchange that occur within a nucleus, for β − decay (a), β + decay (b), and electron-capture (c). These processes are not included in the impulse approximation. The coupling constant for the point interaction vertex is the effective weak-interaction coupling constant GF . . . . . . . 105 xi Figure 5.3: Schematic representations of the impulse approximations to the different weak processes within an A-body nucleus: β − decay (a), β + decay (b), and electron-capture (c). The active nucleon doesn’t interact with the initial and final nuclei during the weak process. The coupling constant for each interation vertex is the weak-interaction coupling constant GF . xii 106 KEY TO SYMBOLS AND ABBREVIATIONS Ap -Bh – A-particle, B-hole excitation or de-excitation from the reference state |0i – vacuum state |Φ0 i – reference state |Ψi – correlated ground state a ···a |Φi 1···i A i – specific Ap -Bh state 1 B {· · · } – normal-ordered with respect to the reference state ~ω – harmonic oscillator energy scale Ĥ – Hamiltonian H – similarity-transformed Hamiltonian ĤN – normal-ordered Hamiltonian HN – normal-ordered similarity-transformed Hamiltonian T̂ – cluster operator aA i a ···a a i εi 1···i A – energy denominator, fi 1 + · · · + fi B − fa11 − · · · − faA 1 1 B B CC – coupled cluster CCD – coupled cluster with doubles CCSD – coupled cluster with singles and doubles CCSDT – coupled cluster with singles, doubles, and triples CI – configuration interaction FCI – full configuration interaction COM – center of mass EOM – equation-of-motion xiii PA – particle-attached PR – particle-removed HF – Hartree-Fock IM-SRG – in-medium similarity renormalization group HO – harmonic oscillator MBPT – many-body perturbation theory DIIS – direct-inversion of the iterative subspace QCD – quantum chromodynamics EFT – effective field theory NN – nucleon-nucleon 3N – three-nucleon NLO – next-to leading order N2 LO – next-to-next-to leading order N3 LO – next-to-next-to-next-to leading order xiv Chapter 1 Introduction Steady progress in any modern scientific endeavor requires a strong, dynamic relationship between experimental data to paint an accurate picture of some natural phenomena and theoretical models to interpret those phenomena with respect to the growing network of other scientific models. Conversely, the predictive capability of theoretical models can highlight blurry or unfinished areas of that picture which can be clarified or completed by new or improved experimental techniques. In the pursuit to understand and describe the atomic nucleus and the corresponding implications from quarks to neutron stars, this push-andpull coordination between theory and experiment makes progress in modern nuclear physics robust and persistent. An integral component of modern nuclear physics is describing the structure and emergent properties of self-bound systems of protons and neutrons. The systems in questions can be stable nuclei, rare isotopes far from stability, and even infinite nuclear matter which can be used to model neutron stars. Relevant properties to nuclear structure include ground-state energies–for determining nuclear masses, excited-state energies–for identification in gamma or neutron spectroscopy, and transition or decay amplitudes–for calculating the respective rates for those processes. This wide array of emergent properties inserts both nuclear structure theory and experiment into a prominent role within every other subfield of modern nuclear physics, from lattice quantum chromodynamics (QCD) to nuclear astrophysics, and beyond, to questions about fundamental symmetries and dark matter. However, two inextricable characteristics of a comprehensive model of nuclear structure–the increasingly large 1 size of many-body nuclear systems and the complexity and strength of the nucleon-nucleon interactions–have been imposing hurdles for theorists to overcome. 1.1 A Brief History of Nuclear Structure Theory A major step in the project to solve the correlation problem in many-fermion systems was taken with the work of Brueckner, Bethe, and Goldstone [5, 6, 7] with the reformulation of the nuclear interaction by accounting for two-body correlations from the nuclear medium. This work continued with the work of Coester [8, 9] and Kummel [10], amongst many others, with a further resummation of nuclear correlations in the form of an exponential ansatz into what would become coupled-cluster (CC) theory. However, there were two major obstacles that hindered the progress in this area for decades. First, while these methods were systematically improvable by including progressively higher-level correlations, the highly nonperturbative nature of the nuclear force required computationally infeasible summations. Second, there wasn’t a reliable and consistent theory to model nucleon-nucleon interactions. On the other hand, with the well-known and highly-perturbative Coulomb force, which underlies the many-electron systems in atoms and molecules, the field of quantum chemistry made consistent advances in ab-initio quantum chemistry possible since the 1950s. Along with the quasi-exact method of configuration interaction (CI) theory [11, 12, 13, 14] which have been utilized since the formulation of quantum mechanics, chemists successfully employed approximate methods like many-body perturbation theory (MBPT) [15, 16, 17, 18] and coupled-cluster theory [19, 20, 21, 22, 18]. Fortunately, within the past decade, two breakthroughs have allowed ab initio nuclear structure to resurface and thrive the way that quantum chemistry had done in the previous decades. First was the invention of chiral effective field theory (EFT) [23, 24] which gave 2 Figure 1.1: Nuclear chart of nuclei with ground-state energies which have been calculated with ab-initio methods and NN+3N interactions. Figure taken from [1]. theorists the ability to construct nucleon-nucleon interactions consistent with the underlying QCD of the strong nuclear force. Second was the application of renormalization group (RG) methods to the nuclear force [25, 26]. This procedure can “soften” the NN interaction, to decouple the high- and low-momentum components of the nuclear force and generate lesscorrelated systems that can be calculated at a reasonable computational cost. These major changes to nuclear structure theory made it possible to merge the field with the progress of quantum chemistry and open a new area for additional developments in ab initio descriptions 3 of many-fermion systems, see Fig. 1.1. Along with exponential improvements to high-performance computing, these novel techniques have allowed modern many-body methods to extend their reach and deepen their applicability across the nuclear chart, see Fig. 1.2. The no-core shell model (NCSM), an exact method for a given model space, has been useful in calculating the radii, transition strengths, and effective interactions of light nuclei and has been extended to nuclei in the sd shell [27, 28, 29]. A quasi-exact technique which follows a completely different methodology than NCSM, quantum Monte Carlo (QMC), has also progressed and is now capable of calculating properties of light nuclei with modern chiral forces [30, 31, 32]. In addition to these exponentially scaling techniques’ successes with lighter nuclei, polynomially scaling techniques–such as the in-medium similarity renormalization group (IMSRG) [33, 34, 35, 36, 37, 38, 39, 40], self-consistent Green’s functions (SCGF) [41, 42, 43], and coupled cluster theory [44, 45, 46, 47, 48, 49, 50, 51]–have been able to reach open-shell nuclei through the pf shell and even up to the chain of even tin isotopes with equations-of-motion and multi-reference techniques [52]. 1.2 Electroweak Theory and Nuclear Structure Nuclear structure is implicated in performing and analyzing experiments to probe fundamental symmetries and physics beyond the Standard Model. One example is determining the Vud component of the Cabibbo-Kobayashi-Maskawa (CKM) matrix, which relates quark eigenstates of the weak interaction to their mass eigenstates [53, 54]. This component can be determined from by measuring the half-lives of superallowedbeta decays [55] and applying a nucleus-dependent structure correction [56, 57, 58, 59, 60]. The value of |Vud | is used to test the unitarity of the CKM matrix and the conserved-vector current hypothesis, which 4 140 Mass Number (A) 120 100 80 60 40 20 0 1985 1990 1995 2000 Year 2005 2010 2015 Figure 1.2: Progress of ab-initio nuclear structure from calculations of ground-state energies with NN+3N interactions. Early progress was approximately linear as the problem size scaled with Moore’s law while more recent progress has taken advantage of new algorithms which have outpaced Moore’s law. Data taken from [1]. relates the f t-values of superallowed Fermi β decays of different nuclei, both predicted by the standard model [61]. Another example of physics beyond the standard model is the neutrinoless double-beta decay (0νββ) [62, 63]. The extremely-rare, two-neutrino double-beta decay (2νββ) has been observed in many experiments [64, 65], which has motivated the search for its neutrinoless counterpart, in which two Majorana neutrinos, being their own antiparticles, annihilate one another, which is not possible in the standard electro-weak theory. The long half-lives of these theoretical decays depend on a phase-space factor, which is highly dependent on the decay Q-value, and a nuclear matrix element. The Q-value can be determined from high-precision mass measurements of the relevant nuclei [66, 67, 68, 69], while the nuclear matrix element, which contributes the largest source of uncertainty, must be calculated with 5 a reliable many-body theory. The weak interaction and nuclear structure can also be exploited for supernova neutrino detection and spectroscopy. While these original detectors were based on electron-neutrino scattering [70, 71], more recent experiments utilize correlated nucleon effects of large nuclei to enhance the scattering cross section and therefore the ability to resolve energies and distinguish neutrino flavors [72, 73, 74, 75]. Supernova models predict distinct distributions for different neutrino flavors based on the temperatures at which they are emitted [76, 77]. With nuclear structure calculations that include sufficient nuclear correlations, these highresolution detectors can be used to verify specific models. 1.3 Ab-Initio Descriptions of Beta Decay Since Enrico Fermi’s originally rejected paper describing beta decay in 1934 [78, 79], theorists have worked to refine this description within the ever-growing library of knowledge concerning the nature of the weak force, the characteristics of the neutrino, and the structure of nuclei. With the success of ab initio calculations for nuclear properties such as masses, radii, and electromagnetic phenomena, these techniques also seem promising ways to calculate relevant quantities involved in nuclear beta decay. Because the kinematics of the decay and the underlying weak process are well understood, the remaining task for nuclear theory to tackle is calculating the transition amplitudes between the initial and final nuclei. Modern calculations of these beta-decay matrix elements were originally performed using phenomenological interactions in the shell model framework [80, 81, 82, 83]. Also, predecessors to current ab-initio techniques like the random-phase approximation (RPA) [84] included core-correlation effects in these early descriptions. These methods were able to successfully reproduce experimental lifetime data and address technical issues such as the quenching of 6 the axial-vector coupling constant. More recently, the success of the shell model has inspired an extension to the new method, known as the ab-initio shell model [29], where an effective interaction is constructed within a certain valence space using a many-body method such as CC [85] or IMSRG [86]. However, these techniques are computationally expensive and cannot currently reach heavy nuclei of interest. The most common method used in their place is known as the quasiparticle random-phase approximation (QRPA) [87, 88]. While these calculations can be performed for heavy nuclei in large spaces, they also rely on phenomenological effective interactions. Therefore, there is a demand for computationally-economical, ab initio techniques that can capture the relevant many-body correlations needed to accurately describe the nuclear structure aspects of electro-weak processes. 1.4 Thesis Structure The main goal of this work is to explore the ab initio description of nuclear beta decay within the coupled-cluster theory framework of the equation-of-motion coupled cluster with singles and doulbes (EOM-CCSD) method using renormalized chiral NN and 3N interactions. The organization of the thesis builds from a general description of the many-body problem of quantum mechanics in chapter 2. Then, in chapter 3, this many-body framework is applied within the coupled-cluster theory and applied to various systems including, atomic nuclei. In chapter 4, coupled-cluster theory is extended to the equation-of-motion method to describe open-shell systems. Then, in chapter 5, different types of beta decay are described in detail then outlines the procedure to express beta-decay observables as effective coupled-cluster operators and how to calculate those observables in the equation-of-motion framework. Lastly, conclusions and future perspectives are given in chapter 6 while technical details concerning the formalism and implementation are given in the appendices A – D. 7 Chapter 2 Many-Body Quantum Mechanics Ab initio structure calculations of many-fermion systems such as those in nuclear and electronic structure aim to describe emergent phenomena from the constituent particles subject to the underlying microscopic Hamiltonian. This amounts to finding the solution to the many-body Schrödinger equation. However, a calculation of the exact solution needs to account for all possible correlations among the particles and thus scales factorially. This motivates the need for approximations to the exact solution that account for the most important correlations. This chapter first establishes the formalism necessary to define the many-body problem then illustrates several successive approximations to its solution. Because the type of fermions and the underlying Hamiltonian can be kept generic until specific systems are considered, the formalism and many-body methods can be kept generic as well. 2.1 Independent-Particle Model The nonrelativistic A-body quantum problem begins with the Schrödinger equation, ĤΨν (r1 , · · · , rA ) = Eν Ψν (r1 , · · · , rA ) , (2.1) for the correlated wave function Ψν (r1 , · · · , rA ) and the corresponding energy Eν . The Hamiltonian can be written generically as a sum of k-body pieces which, in principle, can 8 contain up to A-body interactions, Ĥ = (1)Ĥ + (2)Ĥ + (3)Ĥ + · · · = A X i (1)Ĥ(r ) + i A X (2)Ĥ  ri , rj + i 0 and JI > 0) +1 +1 +1 Table 5.1: Summary of the selection rules for allowed beta decays according to the angular momentum J and parity π of the initial (I) and final (F ) states. stated that while Gamow-Teller transitions can occur without a change in the nuclear spin (∆J = 0 for JF > 0 and JI > 0), the Gamow-Teller operator carries an angular momentum of J = 1. The half-life of these processes T1/2 can be calculated using a combination of both the 106 Fermi and Gamow-Teller type transitions in the form of their reduced transition amplitudes, BF and BGT , respectively, gV2 |MF |2 BF = 2Ji + 1 2 |M 2 gA GT | BGT = , 2Ji + 1 (5.1) where Ji is the final state angular momentum. The factor gV is the vector coupling constant, and its value can be shown to be exactly gV = 1.0. The axial-vector coupling constant gA has a free-space value of gA = −gV , but is altered within nuclei due to nucleon-nucleon correlations. The exact problem of how to treat the value of the axial-vector constant has been a widely studied topic for decades [157, 158, 159, 160], but this work will use the value gA /gV = 1.261(8) [161]. The transition matrix elements MF and MGT are measures of the overlap integral between the initial and final nuclear states for the different transitions and will be discussed in the next section. Inserting these reduced transition amplitudes into the standard result from time-dependent perturbation theory gives the decay half-life, T1/2 = f . K0 (BF + BGT ) (5.2) The factor f represents a phase-space integral over the final nuclear and lepton states and depends on the decay Q-value. The factor K0 encodes the relevant constants involved, K0 = 2π 3 ~7 ln 2 ≈ 6147 s, m5e c4 G2F (5.3) where me is the electron mass, and GF is the effective coupling constant. The next section describes the process of improving upon the impulse approximation by using coupled cluster theory to include higher-body interactions with the CC similarity transformation. 107 5.2 Coupled Cluster Effective Operators The main step in calculating dynamic properties within an ab initio framework is to calculate the transition matrix elements of the relevant operator between two correlated many-body states. In the cases considered in this thesis, the correlated many-body states are given by right and left expansions for the PA-EOM(2) and PR-EOM(2) operators in Eqs. (4.3) and (4.3), respectively. Beta-decay properties are computed with the Fermi and Gamow-Teller operators which are one-body operators that change a neutron to a proton or vice versa and carry the quantum numbers that correspond to the rules in table 5.1. In the impulse approximation, the Fermi operator, which has no spin component, is simply equivalent to the isospin raising operator for the β − Fermi transition, which changes a neutron to a proton, or the lowering operator for the β + transition, which changes a proton to a neutron, ÔF ∓ = X pq n o † hpk τ̂± kqi âp âq . (5.4) The reduced matrix element, see appendix D, hpk τ̂± kqi, is given by, see [162], hpk τ̂± kqi = p 2jp + 1 δnp nq δlp lq δjp jq δtp tq ±1 , (5.5) where the quantum numbers n and l can refer to the quantum numbers of any spherical basis, such as the harmonic oscillator basis, and t is the states isosping projection. Similarly, the Gamow-Teller operator also includes the isospin raising/lowering operator in addition to the spin operator σ̂, ÔGT ∓ = X pq n o † hpk σ̂τ̂± kqi âp âq . 108 (5.6) Here, the reduced matrix element, hpk σ̂τ̂± kqi, is also given by, see [162], hpk σ̂τ̂± kqi = q ( 6(2jp + 1)(2jq + 1) 1 2 jp 1 2 jq 1 lp ) 3 (−1)lp +jp + 2 δnp nq δlp lq δtp tq ±1 . (5.7) These operators can now be used to calculate the transition matrix elements in Eq. (5.1) by applying them between an initial state, |Ψi i, and a final state, hΨf |. The Fermi reduced transition amplitude is, MF = hΨf k ÔF ∓ kΨi i = δJ Ji f X pq n o † hpk τ̂± kqi hΨf k âp âq kΨi i . (5.8) Similarly, the Gamow-Teller reduced transition amplitude is, MGT = hΨf k ÔGT ∓ kΨi i = X pq n o † hpk σ̂τ̂± kqi hΨf k âp âq kΨi i . (5.9) Within the coupled cluster framework, the initial and final states take the form of PAEOM or PR-EOM states given generically in Eq. (4.1). According to Eqs. (4.7) and (4.12), the left and right eigenstates of the bare Hamiltonian can be given by hΨf | = hΦ0 | L̂f e−T̂ and |Ψi i = eT̂ R̂i |Φ0 i. Inserting these EOM states into Eqs. (5.8) and (5.9), gives, MF = δJ Ji f X pq n o † hpk τ̂± kqi hΦ0 k L̂f e−T̂ âp âq eT̂ R̂i kΦ0 i . (5.10) Similarly, the Gamow-Teller matrix element becomes, MGT = X pq n o † hpk σ̂τ̂± kqi hΦ0 k L̂f e−T̂ âp âq eT̂ R̂i kΦ0 i . (5.11) The resemblance of these equations to Eq. (3.6) motivates the construction of an effective 109 operator, λ Ō. The Fermi and Gamow-Teller effective operators have the form, ŌF ∓ = X ŌGT ∓ = X pq pq hpk τ̂± kqi e−T̂ hpk σ̂τ̂± n o † âp âq eT̂ , kqi e−T̂ n o † âp âq eT̂ . (5.12) (5.13) n o † The similarity-transformed component, e−T̂ âp âq eT̂ , is known as the one-body density matrix. Using the Baker-Campbell-Housedorf expansion like Eq. (3.7) and the reduction to connected diagrams like Eq. (3.8), the one-body density matrix can be reduced to the form, n o n o  † † e−T̂ âp âq eT̂ = âp âq eT̂ . c (5.14) The effective operators are best calculated using diagrammatic techniques like those used for the effective Hamiltonian. In this case, the bare one-body operators are depicted by while the effective one- and two-body operators are depicted by the vertex type and , respectively. The one-body beta-decay operators can be split into hp, hh, pp, and ph components which are analogous to the one-body components of the effective Hamiltonian, see appendix B. The hp component has no connected terms in Eq. (5.14), and so it is unchanged by the similarity transformation, = i a i λOi a a = λ Oai . 110 (5.15) The pp component is augmented by single excitations from the reference state. This component’s diagrammatic and algebraic expressions are, a a a = b + k b b λOa b = λ Oba − X λ O k ta . b k (5.16) k Similarly, the hh component is also augmented by single excitations from the reference state, j j j = i + c i λOi j = λ Oji + X i λ O i tc . c j (5.17) c The ph component of the effective beta-decay operator includes effects from single and double excitations from the reference state. The diagrammatic and algebraic expressions for this component are, a i a a i = + i i + a c λOa i = λ Oia + X c λ O a tc c i − k X λ O k ta i k k a + + X c λ O k tac . c ik i k (5.18) kc Like the higher-body interactions induced by the CC similarity transformation, two-body effective operators are induced from the one-body bare beta-decay operator. Calculating properties using PA-EOM(2) and PR-EOM(2) states requires only two components of the 111 effective two-body operator. The pphp component is the results of double excitations from the reference state and is given by the following diagram and the corresponding expression, b a b i a = c c λ O ab ic =− X i k λ O k tab . c ik (5.19) k Similarly, the hphh component is represented by the following diagram and its corresponding algebraic expression, j k j a k = i i λ O ia jk = X a c λ O i tca . c jk (5.20) c After constructing the effective operators, the reduced transition amplitudes in Eq. (5.1) can be calculated. Because of the ambiguity in the bi-orthonormalization discussed in section 4.2, the square-norm of the matrix elements, |M |2 , must be written using both the left and right solutions for each of the initial and final states. Expanded in terms of the EOM states, these reduced amplitudes for the Fermi and Gamow-Teller operators become, BF = X gV2 δJ J i hΦ0 k L̂i ŌF ± R̂f kΦ0 i hΦ0 k L̂f ŌF ∓ R̂i kΦ0 i , 2Ji + 1 f (5.21) f BGT = 2 gA X 2Ji + 1 f hΦ0 k L̂i ŌGT ± R̂f kΦ0 i hΦ0 k L̂f ŌGT ∓ R̂i kΦ0 i . (5.22) Using the machinery developed in this thesis and the techniques discussed in this chapter, 112 the calculated half-lives of various nuclei will be presented in an upcoming paper. 113 Chapter 6 Conclusions and Perspectives 6.1 Summary and Conclusions Accurate ab initio calculations of beta-decay transition amplitudes are necessary for answering many open questions from a wide range of areas in modern physics from nucleosynthesis to fundamental symmetries. The accuracy and scope needed to answer such questions necessitate a technique that is widely applicable, systematically improvable, and scalable to large systems. In this thesis, we have developed the formalism for and achieved the application of techniques based on coupled cluster theory that fulfill these requirements. The large and versatile program that implements these techniques can be used for many different fermionic systems including the homogeneous electron gas, quantum dots, and finite nuclei. Because of the program’s modular form, additional systems like neutron drops, infinite nuclear matter, and atomic systems can easily be added in subsequent updates. Importantly, we extended the single-determinant coupled-cluster method to open-shell systems using the equation-of-motion method which grants a broader reach across the nuclear chart. Also, we added the crucial ability to perform calculations with and without three-body forces which ensures accurate results. Also modular, these components of the program can be easily extended to higher-order EOM approximations, two-particle-attached and two-particleremoved EOM states, and the inclusion of full three-body forces. Lastly, we’ve implemented the ability to construct any effective one-body operator and calculate the corresponding 114 observables using ground and excited EOM states. In future iterations of this code, higherorder effective operators, like those required for double-beta-decay experiments, can also be implemented. Also, these higher-order operators can be constructed from two-body chiral weak currents and used to investigate the quenching of the axial vector coupling constant. In addition to developing and implementing these various techniques, we performed calculations at each step to verify the results. In particular, we provided a proof of principle by comparing our results with those from other ab initio methods for various different systems. Also, we calculated ground-state energies as well as particle-attached and particle-removed spectra for various light nuclei, focusing mainly on the oxygen chain. In future publications, this machinery will be extended up the nuclear chart to heavier nuclei and out to the limits of stability to calculate beta-decay properties of nuclei around 78 Ni and 100 Sn, which are important to future experiments at FRIB. For example, consistent beta-decay lifetime calculations from ab initio methods will be invaluable for astrophysical simulations of different nucleosynthesis processes. 115 APPENDICES 116 Appendix A CCSD Diagrams The following diagrams and their corresponding algebraic expressions comprise the different contributions to the CCSD cluster amplitudes without directly building the effective Hamiltonian, H . The boxed diagrams are automatically zero in a Hartree-Fock basis. The contributions to the CCSD singles equation, (3.10), are given by Eqs. (A.1)-(A.8). a i i fˆN t̂1 |Φ0 ic = = fia − a a + + k X fik tak + X i c fca tci (A.1) c k a V̂N t̂1 |Φ0 ic = c k =− X 117 kc c Vka ic tk i (A.2) fˆN t̂2 |Φ0 ic = = a c i k P k ac fc tki (A.3) kc a V̂N t̂2 |Φ0 ic = = i i c k a + d c k l 1 X ka cd 1 X kl ca Vcd tki + Vic tkl 2 2 kcd (A.4) klc i fˆN t̂21 |Φ0 ic a = = l d P l ad fd tl ti (A.5) kcd a V̂N t̂21 |Φ0 ic = = i i c X k d c d Vka cd tk ti + kcd c X klc 118 a + k c a Vkl ic tk tl l (A.6) i V̂N t̂1 t̂2 |Φ0 ic = a a =− l i + d c d k a + l c k c k d 1 X kl cd a 1 X kl ca d X kl ad c Vcd tki tl − Vcd tkl ti + Vcd til tk 2 2 klcd klcd i l (A.7) klcd i V̂N t̂31 |Φ0 ic = a l =− d X c k c d a Vkl cd tk ti tl (A.8) klcd The contributions to the CCSD doubles equation, (3.10), are given by Eqs. (A.9)-(A.19). a i b j V̂N |Φ0 ic = = Vab ij (A.9) b fˆN t̂2 |Φ0 ic = i a = P̂ (ab) j j + a i b c X c fcb tac ij − P̂ (ij) 119 k X k fjk tab ik (A.10) i V̂N t̂1 |Φ0 ic = b V̂N t̂2 |Φ0 ic = + = X X c Vab cj ti b i + (A.11) c l i j c + b a j d k c X 1 X kl ab 1 X ab cd ac Vkb Vij tkl + Vcd tij − P̂ (ij|ab) ic tkj 2 2 j a X kl a b + cd i j c X b i l a b Vkl ij tk tl + (A.12) kc cd k = a Vkb ij tk + P̂ (ij) a b i = b k kl V̂N t̂21 |Φ0 ic j c j a k a i k = − P̂ (ab) i j a + d c d Vab cd ti tj − P̂ (ij|ab) 120 b a j k X kc c a c Vkb ic tk tj (A.13) a V̂N t̂22 |Φ0 ic = b i c k l a c k a + d j i d l c b i b + k a c d k j l i j + = j l b d X X 1 X kl ab cd ac tbd − P̂ (ij) 1 Vcd tkl tij + P̂ (ab) Vkl t Vkl tab tcd cd lj ki cd lj ki 4 2 klcd klcd klcd 1 X kl db ca − P̂ (ab) Vcd tij tkl 2 (A.14) klcd a fˆN t̂1 t̂2 |Φ0 ic = i i b k = − P̂ (ab) c j + a j c P k a cb P fc tk tij − P̂ (ij) fck tci tab kj kc kc 121 b k (A.15) a i i b V̂N t̂1 t̂2 |Φ0 ic = l c i + j j k = P̂ (ij|ab) + d c X kcd k j a j c X klc b + k d bc d Vka cd tjk ti − P̂ (ij|ab) c i b c d j + k + l b i i a b c j k a a a k b l bc a Vkl ci tjk tl − P̂ (ab) 1 X kb cd a Vcd tij tk 2 kcd X X 1 X kl ab c tck tab Vkl − P̂ (ij) + P̂ (ij) Vka tck tdb Vcj tkl ti + P̂ (ab) ci ij lj cd 2 a V̂N t̂31 |Φ0 ic = i i d j k c = − P̂ (ij|ab) + b X a j b c k l a c d Vkb cd tk ti tj + P̂ (ij|ab) kcd X klc 122 (A.16) klc kcd klc c a b Vkl cj ti tk tl (A.17) a V̂N t̂21 t̂2 |Φ0 ic = b i c k l i a + d c k b d i c k a j l j l a + d a b + = j c l k l i d c b k i j + j b d X 1 X kl ab c d 1 X kl cd a b Vcd tkl ti tj + Vcd tij tk tl + P̂ (ij|ab) Vkl tac tbk tdi cd lj 2 2 klcd klcd klcd 1 X kl ab c d 1 X kl db c a − P̂ (ij) Vcd tlj tk ti − P̂ (ab) Vcd tij tk tl 2 2 klcd a V̂N t̂41 |Φ0 ic (A.18) klcd = = b k X i c l a b c d Vkl cd tk tl ti tj klcd 123 j d (A.19) Appendix B Effective Hamiltonian Diagrams The following diagrams and their corresponding algebraic expressions comprise the different components of the CCSD effective Hamiltonian, H . Because some components are used as intermediates to build other components, they must be built in the order written. Some intermediate components overcount some diagrams which motivates the need for further intermediates, denoted by X 0 , X 00 , and X 000 . The boxed diagrams are automatically zero in a Hartree-Fock basis. The one-body components to the CCSD effective Hamiltonian, Eq. (3.8), are given by Eqs. (B.1)-(B.4). = i a + i i a Xai = fai + X a c k c Vik ac tk (B.1) kc a a a = b + b a a k c l + + b b c 1 X kl ac X ka c X k a Xba = fba − Vbc tkl + Vcb tk − Xb tk 2 klc kc 124 k k k b (B.2) j j j = i j + i c d k + c i i k 1 X ik cd X ik c X 0ij = fji + Vcd tjk + Vjc tk 2 kcd j kc j j = + i (B.3) c i i Xji = X 0ij + X Xci tcj (B.4) c Once the one-body components have been constructed, the pseudo-linear form for the CCSD singles equation, Eq. (3.9), can be evaluated, a i a a i =0= + i i + k a + Xia = 0 = fia + c + 1X 2 kcd c Xca tci − cd Vka cd tki − a + d X c k i i k i + c c a a X k c a X 0k i tk + X ac Vkl ic tkl + X k 1X 2 l a + i k c Vka ci tk kc klc Xck tac ik . (B.5) kc During a CC iteration, it’s possible to use these updated singles amplitues T̂ 1 when evaluating the doubles amplitudes T̂ 2 in the same iteration, which can accelerate the convergence. 125 The two-body components to the CCSD effective Hamiltonian, Eq. (3.8), are given by Eqs. (B.6)-(B.20). a a a = c b i c b i b ia = Via − Xbc bc X a i 0ij ij j a i ij ij c i j a 1 X ij c Vca tk 2 c (B.8) k + a (B.7) 1 + 2 k j i k = i b k X ka = Vka + k c a Vik bc tk k j k i = i i (B.6) + c k b a a b c k = c k 1 X ik a Vbc tk 2 ia X 0ia bc = Vbc − a i 1 + 2 j Xka = Vka + a X ij Vca tck c 126 c i j a (B.9) a b a a b = c + d c b a X c (B.10) a d ab = X 0ab + Xcd cd d a X 0kb cd tk + d c k b = c k d ab X 0ab cd = Vcd − P̂ (ab) a b b k l c d 1 X kl ab Vcd tkl 2 (B.11) kl k l k = i j + i Xkl = Vkl + l c j ij ij k l l k + d i i j j c X 0ij 1 X ij cd Vcd tkl + P̂ (kl) X kc tcl 2 c (B.12) cd j a j = i b j a + i ia X 0ia jb = Vjb + b X c c X 0ia cb tj − 127 a c i b 1 X ik a Vjb tk 2 k a j 1 + 2i b k (B.13) j a j a i b j = i b ia X 00ia jb = Vjb + a 1 + 2 c i a j 1 + 2i b k b 1 X 0ia c 1 X ik a X cb tj − Vjb tk 2 c 2 (B.14) k j a j j a = i b i b ia X 000ia jb = Vjb + a 1 + 2 a j + c i b i k b 1 X ia c X ik a Vjb tk Xcb tj − 2 c (B.15) k j a j  j 1 + 2 a = i b i ia Xjb The factor of   1 2 = b X 000ia jb a c j 1 + 2 k i a c i b b  X X 1 ia tc ca + 1 Vik Xcb t − j cb jk 2 2 c kc (B.16) is applied when solving the CCSD equations but omitted when applying the effective Hamiltonian in post-CC methods. b a i b a i = c c ab X 0ab ic = Vic + b 1 + 2c a b i d a + c X 1 X ab d a Vdc ti − P̂ (ab) X 00kb ic tk 2 k d 128 i k (B.17) b a i b a b i = c a + b ab = Vab + Xic ic b a + X d X k a j X j k a k = i j a a k i k + c k a + j X c a Vil jk tl + P̂ (jk) d i lc X a c c X 000ia jc tk c l X k + l i ia = Via − Xjk jk + P̂ (jk) a a i j c (B.19) j l i + a l l k j k 1 + 2i + i (B.18) 1 X il a Vjk tl 2 j a 1 X kl ab Xic tkl 2 kl i j k a X 0kb ic tk kb tad + Xdc ik k ia X 0ia jk = Vjk − l k = i + k kd k i a i c d Vab dc ti − P̂ (ab) Xck tab ik + P̂ (ab) j k c c d X i k c a b a + i + d c c − b i il tca + Xjc lk 1X 2 cd 129 ia tcd + Xcd jk X c Xci tca jk (B.20) a i b j a i b a j =0= i i + b j a + c a + b i i j c + j a d i b k + l ab = 0 = Vab + P̂ (ab) Xij ij X c + b k a b a j k j + c Xca tcb ij − P̂ (ij) i j a i j b + k X b c Xik tab kj k X X X 1 X 0ab cd 1 X kl ab kb tac − P̂ (ab) c X cd tij + Xij tkl − P̂ (ab|ij) Xic X 0kb tak + P̂ (ij) X 0ab ij ic tj kj 2 2 c cd kl kc k (B.21) 130 Appendix C Computational Implementation The sums involved in building the CC effective Hamiltonian, solving the CC equations, solving the EOM-CC equations, and building effective operators can all be reformulated as matrix-matrix multiplications and thus performed with efficient LAPACK and BLAS routines. To take advantage of this efficiency, the various cluster amplitudes and matrix elements must be grouped into structures with similar index organization so that summed indices map to the same states and matrix elements. An additional benefit to these structures is that angular-momentum-coupling coefficients are automatically removed by summing over Clebsch-Gordon coefficients, see chapter D. Symmetry Channels Each matrix structure is separated into different symmetry channels for different perutations of its indices. The channels are denoted as Σξ~, where ξ~ represents the relevant quantum numbers of a certain channel. There are four different channel types that are relevant for the structures used in this work. The direct two-body channel categorizes the vector sum of two single-particle-state quantum numbers, Σξ~ , 1 ξ~pq = ξ~p + ξ~q −→ |pqi ∈ Σξ~ =ξ~ . 1 pq (C.1) The cross two-body channel categorizes the vector difference of two single-particle-state 131 quantum numbers or, equivalently, the vector sum of a the quantum numbers of a singleparticle state and a time-reversed single-particle state, Σξ~ , 2 ξ~pq̄ = ξ~p − ξ~q = ξ~p + ξ~q̄ −→ |pq̄i ∈ Σξ~ . 2 (C.2) The one-body channel categorizes single-particle states by their vector quantum numbers, Σξ~ , 3 ξ~p −→ |pi ∈ Σξ~ . (C.3) 3 The cross three-body state categorizes the vector difference between the quantum numbers of a direct two-body state and a single-particle state or, equivalently, the vector sum of the quantum numbers of a two-body direct state and a time-reversed single-particle state, Σξ~ , 3 ξ~pqr̄ = ξ~p + ξ~q − ξ~r = ξ~p + ξ~q + ξ~r̄ −→ |pqr̄i ∈ Σξ~ =ξ~ . 3 pqr̄ Channel-Partitioned Structures Different matrix structures are indexed by their channel type: 1 for direct channels, 2 for cross channels, and 3 for one-/three-body channels. For matrices with more than one structure of the same type, there is an additional index that depends on the specific permutation involved. n o p † For a one-body operator Aq âp âq , there is a direct-channel matrix element and a 132 cross-channel matrix element, p A2 = Apq̄ . A1 = Aq , (C.5) n o pq † † For a two-body operator Ars âp âq âs âr , there is a direct-channel matrix element, four cross-channel matrix elements, and four one-channel matrix elements, pq A1 = Ars , qr̄ ps̄ A22 = Asp̄ , A31 = Arsq̄ , A32 = Arsp̄ , A21 = Arq̄ , p q pq For an EOM operator of the form Ar pr̄ A23 = Asq̄ , pqs̄ A33 = Ar , qs̄ A24 = Arp̄ , pqr̄ A34 = As . (C.6) n o † † âp âq âr , there is a direct-channel matrix ele- ment, a one-channel matrix element, and two cross-channel matrix elements, pq A3 = Apqr̄ , p A22 = Arp̄ . A1 = Ar , q A21 = Arq̄ , (C.7) n o p † EOM operators of the form Aqr âp âr âq have similar structures, p A3 = Aqrp̄ , pr̄ A22 = Ar . A1 = Aqr , pq̄ A21 = Aq , 133 (C.8)   T̂ Matrix Form of H = ĤN e c tck̄ Xai = fai + Viā ck̄ hp hp hhpp X2 ←− f2 + V2 3 · t2 (C.9) 1 ab̄ ck̄ a k Xba = fba − taklc̄ Vklc̄ b + Vck̄ t − tk X b 2 1 pp hp hhpp X3 ←− − t3 · V3 − t3 · X 3 3 2 1 pp pp hppp X2 ←− f2 + V2 4 · t2 (C.10) 1 k̄ + Vij̄ tck̄ X 0ij = fji + Vicdk̄ tcd j ck̄ 2 1 hhpp hh X0 3 ←− V3 · t3 3 2 1 hh hhhp X0 2 ←− f2hh + V2 3 · t2 (C.11) 1 ij̄ ck̄ k̄ i c Xji = fji + Vicdk̄ tcd j + Vck̄ t + X c tj 2 1 hhpp hp Xhh 3 ←− 2 V31 · t33 + X 3 · t3 hh hhhp 3 X0 2 ←− f2hh + V2 134 · t2 (C.12) 1 a cdk̄ 1 a klc̄ aī ck̄ aī kc̄ Xia = fia + Xca tci − tak X 0k i − Vck̄ t + 2 Vcdk̄ ti − 2 tklc̄ Vi + tkc̄ X 1 hppp 1 hh ph pp hhhp X3 ←− X3 · t3 − t3 · X0 3 + V3 · t3 − t3 · V3 4 3 2 2 2 1 ph ph hphp X2 ←− f2 − V2 2 hp · t2 + t2 · X 2 (C.13) 3 1 a k ia X 0ia bc = Vbc − 2 tk Vbcī 1 hppp hppp hhpp − t3 · V 3 X0 3 ←− V3 2 2 2 2 (C.14) ia = Via − ta Vk Xbc bc k bcī hppp 2 X3 hppp 2 ←− V3 hhpp 2 − t3 · V3 1 ijā 0ij ij X ka = Vka + Vc tck 2 1 hhpp hhhp ij · t3 X0 3 ←− Vka + V3 3 2 3 ij ij (C.15) (C.16) ijā Xka = Vka + Vc tck hhhp 3 X3 ij hhpp 3 ←− Vka + V3 135 · t3 (C.17) ab a 0k X 0ab cd = Vcd − P̂ (ab) tk X cdb̄ pppp X0 1 pppp 1(2) X0 3 pppp ←− V1 hppp 1 ←− ∓t3 · X0 3 ab = X 0ab + 1 tab Vkl Xcd cd 2 kl cd 1 pppp hhpp pppp X1 ←− X0 1 + t1 · V1 2 (C.18) (C.19) 1 ij ij ij 0ij k̄ Xkl = Vkl + Vcd tcd + P̂ (kl) X c tcl kl 2 1 hhpp · t1 Xhhhh ←− Vhhhh + V1 1 1 2 hhhp 4 Xhhhh ←− ∓X0 3 3 3(4) · t3 (C.20) 1 a k ia 0iab̄ c X 0ia jb = Vjb + X c tj − 2 tk Vjbī hphp 1 ←− V2 hphp 3 ←− X0 3 hphp 2 1 hhhp ←− − t3 · V3 2 2 X0 2 X0 3 X0 3 hphp 1 hppp 3 136 · t3 (C.21) 1 0iab̄ c 1 a k ia X 00ia jb = Vjb + 2 X c tj − 2 tk Vjbī hphp hphp X00 2 ←− V2 1 1 1 hppp hphp X00 3 ←− X0 3 · t3 3 3 2 1 hphp hhhp X00 3 ←− − t3 · V3 2 2 2 (C.22) 1 iab̄ c a k ia X 000ia jb = Vjb + 2 X c tj − tk Vjbī hphp hphp X000 2 ←− V2 1 1 1 hppp hphp X000 3 ←− X0 3 · t3 3 3 2 hphp hhhp X000 3 ←− −t3 · V3 2 2 ia Xjb hphp X2 1 = Via jb ←− + Xciab̄ tcj hphp V2 1 hppp 3 ←− X3 hphp 2 ←− −t3 · V3 X3   1 − Vicb̄k̄ tcjāk̄ 2   1 hhpp − V 2 t2 1 1 2 hphp 3 X3 − tak Vkjbī (C.23) · t3 hhhp 2 (C.24) 1 abc̄ d ab a 00k X 0ab ic = Vic + 2 Vd ti − P̂ (ab) tk X icb̄ 1 pppp pphp pphp X0 3 ←− V3 + V3 · t3 3 3 2 3 pphp 1(2) X0 3 hphp ←− ∓t3 · X00 3 1 137 (C.25) ab = Vab + Vabc̄ td − P̂ (ab) ta X 0k − tabī X k + P̂ (ab) taī X k d¯ + 1 tab X kl Xic c ic k icb̄ k d i k d¯ cb̄ 2 kl ic pphp 3 ←− V3 pphp 1(2) ←− ∓t3 · X0 3 pphp 4 ←− −t3 · X3 pphp 2(3) ←− ∓t2 · X2 pphp 1 hhhp ←− t1 · X1 2 X3 pppp 3 + V3 · t3 hphp 1 X3 X3 X2 X1 pphp 3 hp 4 hppp 3 3 (C.26) 1 a l ia X 0ia jk = Vjk − 2 tl Vjk ī 1 hphh hphh − t3 · Vhhhh X0 3 ←− V3 32 2 2 2 000iaj̄ c tk ia = Via − ta Vl + P̂ (jk) X Xjk c jk l jk ī hphh 2 hphh 2 ←− V3 hphh 3(4) ←− ∓X000 3 · t3 4 hphh 1(3) ←− ∓X2 hphh 1 hppp ←− X1 · t1 2 hphh 1 ←− X3 · t3 X3 X3 X2 X1 X3 (C.27) 1 ia cd ij̄ ¯l tjk + Xci tcjkā + P̂ (jk) X ¯tckā + Xcd cl 2 − t3 · Vhhhh 3 2 hphp hhhp 3 · t2 3 hp (C.28) 1 138 ab = Vab + P̂ (ab) X a tc − P̂ (ij) tabj̄ X k + 1 X 0ab tcd + 1 tab X kl Xij c ij b̄ ij i k 2 cd ij 2 kl ij aj̄ ī c − P̂ (ab|ij) tkc̄ Xikc̄ − P̂ (ab) tak X 0k + P̂ (ij) X 0ab c tj b̄ ij b̄ ←− V1 pphh 1(2) ←− ±X3 · t3 ∓ t3 X0 3 1 1 pphh 3(4) 0 ←− ∓t3 · Xhh 3 ±X 3 pphh 1(2) ←− −t2 · X2 pphh 3(4) ←− t2 · X2 X3 X3 X2 X2 pphh 1 pppp 1 + X0 1 · t1 + t1 · Xhhhh 1 2 2 pphh X1 hphh pp pphp t 4 3 3 hphp 1 1 hphp 1 1 Matrix Form of  (C.29) H NR̂A+1 µ  c = ωµR̂A+1 µ a X kc̄ − 1 X a r cdk̄ ωk ra = Xca rc + rkc̄ 2 cdk̄ 1 hppp hp pp · r3 ωk r ←− X3 · r + r2 · X2 − X3 1 2 2 (C.30) c − r ab X k + 1 X ab r cd − P̂ (ab) r b X kc̄ − 1 tabī Vk r cd¯l ωk riab = −Xcabī rc + P̂ (ab) Xcb riā i k kc̄ iā 2 cd i 2 k cd¯l 1 hhpp pphp ωk r3 ←− −X3 · r − t3 · V3 · r3 1 4 2 3 ωk r2 1(2) pp hphp ←− ∓X3 · r2 ± r2 · X2 2 2 1 1 pppp ωk r1 ←− −r1 · Xhh · r1 3 + 2 X1 (C.31) 139 Matrix Form of L̂A+1 H N = EµL̂A+1 µ µ One main difference for the left eigenproblem is that the disconnected term is computed as an outer product rather than with matrix-matrix multiplication. 1 Ek la = lc Xac − lcdk̄ Xacdk̄ 2 1 pp pphp Ek l ←− l · X3 − l3 · X3 4 2 (C.32) i = P̂ (ab) l X ib̄ − l X c − X i lk + P̂ (ab) liā X c + 1 li X cd Ek lab a c abī c k ab b 2 cd ab iā lkc̄ − 1 l tcd¯l Vk − P̂ (ab) Xkc̄ b 2 cd¯l k abī 1 hhpp hppp Ek l3 ←− −l · X3 − l3 · t3 · V3 3 1 2 2 Ek l2 1(2) pp hp hphp ←− ±l ⊗ X2 ∓ l2 · X3 ± X2 2 1 · l2 2 1 pppp Ek l1 ←− −Xhh 3 · l1 + 2 l1 · X 1 Matrix Form of  H NR̂µA−1  c (C.33) = ωµR̂µA−1 1 ωk ri = −rk Xik + Xck̄ rick̄ − rklc̄ Xiklc̄ 2 1 hp hhhp ωk r ←− −r · Xhh 3 + X 20 · r21 − 2 r3 · X 33 140 (C.34) a = −r X k − P̂ (ij) r aī X k + X a r c + 1 r a X kl − P̂ (ij) X aī r ck̄ − 1 r Vkld¯tc ωk rij c ij k j k ijā ck̄ j 2 kl ij 2 kld¯ c ijā 1 hphh hhpp ωk r3 ←− −r · X3 − r3 · V3 · t3 1 1 3 2 ωk r2 1(2) hphp 2 ←− ±r2 · Xhh 3 ± X2 2 · r2 2 1 pp ωk r1 ←− X3 · r1 + r1 · Xhhhh 1 2 (C.35) Matrix Form of L̂A−1 H N = EµL̂µA−1 µ Again, the disconnected term is computed as an outer product rather than with matrixmatrix multiplication. 1 i klc̄ Ek li = −Xki lk − Xklc̄ l 2 1 hphh · l3 Ek l ←− −XHH · l − X3 3 2 1 (C.36) 1 ij ij ij ijā j Ek la = P̂ (ij) li Xaj̄ − Xk lk + lc Xac − P̂ (ij) Xk lakī + Xkl lakl 2 1 ijā ¯ j − P̂ (ij) l Xack̄ī − Vc tckld¯lkld ck̄ 2 1 hhpp hhhp Ek l3 ←− −X3 · l − V3 · t3 · l3 1 3 2 3 Ek l2 1(2) hp hphp ←− ±l ⊗ X 0 ± Xhh 3 · l22 ± l22 · X 22 2 1 pp Ek l1 ←− l1 · X3 + Xhhhh · l1 2 1 141 (C.37) Appendix D Angular Momentum Coupling Before deriving useful equations for J-scheme angular momentum coupling, it’s necessary to list some shorthand notations, definitions, and useful relationships: p̂ ≡ X {m} p 2jp + 1 ≡ sum over all m (D.1) (D.2) Clebsch-Gordan coefficients: hpq|Ji ≡ hjp mp ; jq mq |JM i (D.3) hpq̄|Ji ≡ hjp , mp ; jq , −mq |JM i(−1)(q−mq ) (D.4) X JM hpq|Jihp0 q 0 |Ji = δmp m 0 δmq m 0 p q X mp mq hpq|Jihpq|J 0 i = δJJ 0 δM M 0 (D.5) (D.6) (D.7) 142 Clebsch-Gordan coefficient symmetries: hjp mp ; jq mq |JM i = (−1)jp +jq −J hjp mp ; jq mq |JM i (D.8) = (−1)jp +jq −J hjq mq ; jp mp |JM i (D.9) Jˆ = (−1)jp −mp hjp mp ; J − M |jq − mq i q̂ Jˆ = (−1)jq +mq hJ − M ; jq mq |jp − mp i p̂ Jˆ = (−1)jp −mp hJM ; jp − mp |jq mq i q̂ Jˆ = (−1)jq +mq hjq − mq ; JM |jp mp i p̂ (D.10) (D.11) (D.12) (D.13) Six-J symbols: ( p s ( X j ĵ32 1 j4 ) ( jp jq jr ≡ js jt ju )( ) δj j 0 j2 j3 j1 j2 j3 6 = 26 0 j5 j6 j4 j5 j6 ĵ3 j3 ( ) X X p q J hps̄|J 0 ihrq̄|J 0 i = Jˆ02 hpq|Jihrs|Ji r s J0 0 JM M ( ) X X p q J 0 0 hpq|Jihrs|Ji = Jˆ2 0 hps̄|J ihrq̄|J i r s J 0 0 M q r t u ) (D.14) (D.15) (D.16) (D.17) J M Two-body, scalar J-scheme matrix elements (T̂ , Ĥ, H ), in terms of M-scheme matrix 143 elements: pq J X J = rs 0 ps̄J X 0 rq̄ J = p pq J s̄ {m} X {m} X J = rs q̄ Xr X = X {m} X {m} pmp qmq Xrmr sms hpq|Jihrs|Ji pmp qmq Xrmr sms hps̄|J 0 ihrq̄|J 0 i pmp qmq Xrmr sms hrs|JihJ q̄|pi pmp qmq Xrmr sms hpq|JihJ s̄|ri (D.18) (D.19) (D.20) (D.21) Two-body M-scheme matrix elements in terms of J-scheme, scalar matrix elements: pmp qmq Xrmr sms = X JM = pq J X J hpq|Jihrs|Ji rs X J 0M 0 = X = X JM 0 ps̄J X 0 hps̄|J 0 ihrq̄|J 0 i rq̄ J p X J hrs|JihJ q̄|pi rs q̄ pq J s̄ Xr JM hpq|JihJ s̄|ri (D.22) (D.23) (D.24) (D.25) To find the relationship between the scalar matrix elements of T̂ , Ĥ, and H with different couplings, the M-scheme expressions are written in terms of their different couplings, then the Clebsch-Gordon coefficients are reorganized using Eqs. (D.8)–(D.17) so that they have the same form. A few examples of this recoupling are shown below with the relevant equation pmp qmq used at each step. The shorthand X ≡ Xrmr sms is used for clarity. The relationship 144 0 ps̄J pq J between X J and X 0 is, rs rq̄ J pq J X J rs ) ( p q J 0 0 = Xhpq|Jihrs|Ji = X Jˆ2 0 hps̄|J ihrq̄|J i r s J M {m} J 0 M 0 {m} ) ( X ps̄J 0 p q J (D.19) = X 0 Jˆ2 J r s J0 rq̄ X X (D.17) (D.26) J pq J p As another example, the relationship between X J and X J is, rs rs q̄ pq J X J = rs = X {m} X {m} Xhpq|Jihrs|Ji = X Xhqp|Jihrs|Ji(−1)jp +jq −J {m} Jˆ X hJ q̄|pihrs|Ji(−1)jp +jq −J p̂ Jˆ p = X J (−1)jp +jq −J rs q̄ p̂ (D.9) (D.19) (D.12) (D.27) When sums are formulated in the terms of channel-partitioned matrices like those in sections 3.4.2 and C, the factors related to a structure’s angluar momentum coupling are automatically summed with the identity Eq. (D.6). To demonstrate this, an example is shown here. First, from the CCSD equations the sum in Eq. (3.44) is rewritten in terms of the J-scheme structures. The indices represent single-particle states in the M-scheme 145 expression but represent degenerate shells in J-scheme, J2 c̄ 1 X kl db ca 1X a Vcd tij tkl −→ t J hkl|J1 ihJ1 c̄|aiVkl hkl|J2 ihJ2 c̄|ditd J hij|J3 ihJ3 b̄|di d 1 kl c̄ ij 3 b̄ 2 2 klcd = klcd J1 J2 J3 {m}   1X X J2 c̄ d  ta J Vkl t hkl|J1 ihkl|J2 i hJ1 c̄|aihJ2 c̄|dihij|J3 ihJ3 b̄|di kl 1 c̄ d ij J3 b̄ 2 mk ml klcd J1 J2 J3 {m} = h i J2 c̄ d 1X a t δ t J Vkl δ J1 J2 M1 M2 hJ1 c̄|aihJ2 c̄|dihij|J3 ihJ3 b̄|di kl 1 c̄ d ij J3 b̄ 2 klcd J1 J2 J3 {m} =  1X 2 klcd J1 J3 {m} = J1 c̄ d ta J Vkl t J  d kl 1 c̄ ij 3 b̄  X M1 mc hJ1 c̄|aihJ1 c̄|di hij|J3 ihJ3 b̄|di h i J1 c̄ d 1X a t J Vkl t δ δ ja jd ma md hij|J3 ihJ3 b̄|di kl 1 c̄ d ij J3 b̄ 2 klcd J1 J3 {m} = J1 c̄ d 1X a t J Vkl t J hij|J3 ihJ3 b̄|ai kl 1 c̄ d ij 3 b̄ 2 (D.28) klcd J1 J3 {m} This final form has the same structure as Eq. (D.24) so that the sum can be collected into the following structure with no angular-momenutm coupling coefficients, ta J ij b̄ ← J1 c̄ d 1X a t J . t J Vkl ij 3 b̄ kl 1 c̄ d 2 (D.29) klcd J1 J3 Amplitudes of different coupling structures are then gathered using relationships like those in Eq. (D.26) and (D.27). 146 REFERENCES 147 REFERENCES [1] H. Hergert, “Private communication.” [2] M. Shao, H. M. Aktulga, C. Yang, E. G. Ng, P. Maris, and J. P. Vary, “Accelerating nuclear configuration interaction calculations through a preconditioned block iterative eigensolver,” CoRR, vol. abs/1609.01689, 2016. [3] P. López Rı́os, A. Ma, N. D. Drummond, M. D. Towler, and R. J. Needs, “Inhomogeneous backflow transformations in quantum monte carlo calculations,” Phys. Rev. E, vol. 74, p. 066701, Dec 2006. [4] R. Machleidt, “Chiral symmetry and the nucleon-nucleon interaction,” Symmetry, vol. 8, no. 4, 2016. [5] K. A. Brueckner and C. A. Levinson, “Approximate reduction of the many-body problem for strongly interacting particles to a problem of self-consistent fields,” Phys. Rev., vol. 97, pp. 1344–1352, Mar 1955. [6] H. A. Bethe, “Nuclear many-body problem,” Phys. Rev., vol. 103, pp. 1353–1390, Sep 1956. [7] J. Goldstone, “Derivation of the brueckner many-body theory,” Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, vol. 239, no. 1217, pp. 267–279, 1957. [8] F. Coester, “Bound states of a many-particle system,” Nuclear Physics, vol. 7, no. Supplement C, pp. 421 – 424, 1958. [9] F. Coester and H. Kmmel, “Short-range correlations in nuclear wave functions,” Nuclear Physics, vol. 17, no. Supplement C, pp. 477 – 485, 1960. [10] H. Kümmel, K. Lührmann, and J. Zabolitzky, “Many-fermion theory in exps- (or coupled cluster) form,” Physics Reports, vol. 36, no. 1, pp. 1 – 63, 1978. [11] J. C. Slater, “The theory of complex spectra,” Phys. Rev., vol. 34, pp. 1293–1322, Nov 1929. [12] E. U. Condon, “The theory of complex spectra,” Phys. Rev., vol. 36, pp. 1121–1133, Oct 1930. [13] R. F. Bacher, “The interaction of configurations: sd−p2 ,” Phys. Rev., vol. 43, pp. 264– 269, Feb 1933. [14] C. W. Ufford, “Configuration interaction in complex spectra,” Phys. Rev., vol. 44, pp. 732–739, Nov 1933. 148 [15] J. Hubbard, “The description of collective motions in terms of many-body perturbation theory,” Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, vol. 240, no. 1223, pp. 539–560, 1957. [16] N. Hugenholtz, “Perturbation theory of large quantum systems,” Physica, vol. 23, no. 1, pp. 481 – 532, 1957. [17] H. Schaefer, Quantum chemistry: the development of ab initio methods in molecular electronic structure theory. Oxford science publications, Clarendon Press, 1984. [18] I. Shavitt and R. J. Bartlett, Many-Body Methods in Chemistry and Physics: MBPT and Coupled-Cluster Theory. Cambridge Molecular Science, Cambridge University Press, 2009. [19] J. Čı́žek, “On the correlation problem in atomic and molecular systems. calculation of wavefunction components in urselltype expansion using quantumfield theoretical methods,” The Journal of Chemical Physics, vol. 45, no. 11, pp. 4256–4266, 1966. [20] J. Čı́žek and J. Paldus, “Correlation problems in atomic and molecular systems iii. rederivation of the coupled-pair many-electron theory using the traditional quantum chemical methodst,” International Journal of Quantum Chemistry, vol. 5, no. 4, pp. 359– 379, 1971. [21] J. Čı́žek and J. Paldus, “Coupled cluster approach,” Physica Scripta, vol. 21, no. 3-4, p. 251, 1980. [22] P. Piecuch, K. Kowalski, I. S. O. Pimienta, and M. J. Mcguire, “Recent advances in electronic structure theory: Method of moments of coupled-cluster equations and renormalized coupled-cluster approaches,” International Reviews in Physical Chemistry, vol. 21, no. 4, pp. 527–655, 2002. [23] E. Epelbaum, H.-W. Hammer, and U.-G. Meißner, “Modern theory of nuclear forces,” Rev. Mod. Phys., vol. 81, pp. 1773–1825, Dec 2009. [24] R. Machleidt and D. Entem, “Chiral effective field theory and nuclear forces,” Physics Reports, vol. 503, no. 1, pp. 1 – 75, 2011. [25] S. Bogner, R. Furnstahl, and A. Schwenk, “From low-momentum interactions to nuclear structure,” Progress in Particle and Nuclear Physics, vol. 65, no. 1, pp. 94 – 147, 2010. [26] R. Roth, J. Langhammer, A. Calci, S. Binder, and P. Navrátil, “Similarity-transformed chiral N N + 3N interactions for the ab initio description of 12 C and 16 O,” Phys. Rev. Lett., vol. 107, p. 072501, Aug 2011. [27] P. Navrátil, J. P. Vary, and B. R. Barrett, “Large-basis ab initio no-core shell model and its application to 12 C,” Phys. Rev. C, vol. 62, p. 054311, Oct 2000. 149 [28] P. Navrátil, S. Quaglioni, I. Stetcu, and B. R. Barrett, “Recent developments in nocore shell-model calculations,” Journal of Physics G: Nuclear and Particle Physics, vol. 36, no. 8, p. 083101, 2009. [29] B. R. Barrett, P. Navrátil, and J. P. Vary, “Ab initio no core shell model,” Progress in Particle and Nuclear Physics, vol. 69, no. Supplement C, pp. 131 – 181, 2013. [30] B. S. Pudliner, V. R. Pandharipande, J. Carlson, S. C. Pieper, and R. B. Wiringa, “Quantum monte carlo calculations of nuclei with a ≤ 7,” Phys. Rev. C, vol. 56, pp. 1720–1750, Oct 1997. [31] S. C. Pieper, , and R. B. Wiringa, “Quantum monte carlo calculations of light nuclei,” Annual Review of Nuclear and Particle Science, vol. 51, no. 1, pp. 53–90, 2001. [32] J. Carlson, S. Gandolfi, F. Pederiva, S. C. Pieper, R. Schiavilla, K. E. Schmidt, and R. B. Wiringa, “Quantum monte carlo methods for nuclear physics,” Rev. Mod. Phys., vol. 87, pp. 1067–1118, Sep 2015. [33] K. Tsukiyama, S. K. Bogner, and A. Schwenk, “In-medium similarity renormalization group for nuclei,” Phys. Rev. Lett., vol. 106, p. 222502, Jun 2011. [34] K. Tsukiyama, S. K. Bogner, and A. Schwenk, “In-medium similarity renormalization group for open-shell nuclei,” Phys. Rev. C, vol. 85, p. 061304, Jun 2012. [35] H. Hergert, S. Binder, A. Calci, J. Langhammer, and R. Roth, “Ab initio calculations of even oxygen isotopes with chiral two-plus-three-nucleon interactions,” Phys. Rev. Lett., vol. 110, p. 242501, Jun 2013. [36] S. K. Bogner, H. Hergert, J. D. Holt, A. Schwenk, S. Binder, A. Calci, J. Langhammer, and R. Roth, “Nonperturbative shell-model interactions from the in-medium similarity renormalization group,” Phys. Rev. Lett., vol. 113, p. 142501, Oct 2014. [37] H. Hergert, S. K. Bogner, T. D. Morris, S. Binder, A. Calci, J. Langhammer, and R. Roth, “Ab initio multireference in-medium similarity renormalization group calculations of even calcium and nickel isotopes,” Phys. Rev. C, vol. 90, p. 041302, Oct 2014. [38] H. Hergert, “In-medium similarity renormalization group for closed and open-shell nuclei,” Physica Scripta, vol. 92, no. 2, p. 023002, 2017. [39] S. R. Stroberg, H. Hergert, J. D. Holt, S. K. Bogner, and A. Schwenk, “Ground and excited states of doubly open-shell nuclei from ab initio valence-space hamiltonians,” Phys. Rev. C, vol. 93, p. 051301, May 2016. [40] S. R. Stroberg, A. Calci, H. Hergert, J. D. Holt, S. K. Bogner, R. Roth, and A. Schwenk, “Nucleus-dependent valence-space approach to nuclear structure,” Phys. Rev. Lett., vol. 118, p. 032502, Jan 2017. 150 [41] V. Somà, C. Barbieri, and T. Duguet, “Ab initio gorkov-green’s function calculations of open-shell nuclei,” Phys. Rev. C, vol. 87, p. 011303, Jan 2013. [42] V. Somà, C. Barbieri, and T. Duguet, “Ab initio self-consistent gorkov-green’s function calculations of semi-magic nuclei: Numerical implementation at second order with a two-nucleon interaction,” Phys. Rev. C, vol. 89, p. 024323, Feb 2014. [43] V. Somà, A. Cipollone, C. Barbieri, P. Navrátil, and T. Duguet, “Chiral two- and threenucleon forces along medium-mass isotope chains,” Phys. Rev. C, vol. 89, p. 061301, Jun 2014. [44] M. Wloch, D. J. Dean, J. R. Gour, M. Hjorth-Jensen, K. Kowalski, T. Papenbrock, and P. Piecuch, “Ab-initio coupled-cluster study of 16 O,” Phys. Rev. Lett., vol. 94, p. 212501, Jun 2005. [45] M. Woch, J. R. Gour, P. Piecuch, D. J. Dean, M. Hjorth-Jensen, and T. Papenbrock, “Coupled-cluster calculations for ground and excited states of closed- and open-shell nuclei using methods of quantum chemistry,” Journal of Physics G: Nuclear and Particle Physics, vol. 31, no. 8, p. S1291, 2005. [46] G. R. Jansen, J. Engel, G. Hagen, P. Navratil, and A. Signoracci, “Ab-initio coupledcluster effective interactions for the shell model: Application to neutron-rich oxygen and carbon isotopes,” Phys. Rev. Lett., vol. 113, p. 142502, Oct 2014. [47] G. R. Jansen, M. D. Schuster, A. Signoracci, G. Hagen, and P. Navrátil, “Open sd-shell nuclei from first principles,” Phys. Rev. C, vol. 94, p. 011301, Jul 2016. [48] G. Hagen, A. Ekström, C. Forssén, G. R. Jansen, W. Nazarewicz, T. Papenbrock, K. A. Wendt, S. Bacca, N. Barnea, B. Carlsson, C. Drischler, K. Hebeler, M. Hjorth-Jensen, M. Miorelli, G. Orlandini, A. Schwenk, and J. Simonis, “Neutron and weak-charge distributions of the 48 Ca nucleus,” Nature Physics, vol. 12, pp. 186 EP –, Nov 2015. Article. [49] K. Kowalski, D. J. Dean, M. Hjorth-Jensen, T. Papenbrock, and P. Piecuch, “Coupled cluster calculations of ground and excited states of nuclei,” Phys. Rev. Lett., vol. 92, p. 132501, Apr 2004. [50] J. R. Gour, P. Piecuch, M. Hjorth-Jensen, M. Wloch, and D. J. Dean, “Coupled-cluster calculations for valence systems around 16 O,” Phys. Rev. C, vol. 74, p. 024310, Aug 2006. [51] S. Binder, P. Piecuch, A. Calci, J. Langhammer, P. Navrátil, and R. Roth, “Extension of coupled-cluster theory with a noniterative treatment of connected triply excited clusters to three-body hamiltonians,” Phys. Rev. C, vol. 88, p. 054319, Nov 2013. [52] T. D. Morris, J. Simonis, S. R. Stroberg, C. Stumpf, G. Hagen, J. D. Holt, G. R. Jansen, T. Papenbrock, R. Roth, and A. Schwenk, “Structure of the lightest tin isotopes,” 2017. 151 [53] N. Cabibbo, “Unitary symmetry and leptonic decays,” Phys. Rev. Lett., vol. 10, pp. 531–533, Jun 1963. [54] M. Kobayashi and T. Maskawa, “Cp-violation in the renormalizable theory of weak interaction,” Progress of Theoretical Physics, vol. 49, no. 2, pp. 652–657, 1973. [55] I. S. Towner and J. C. Hardy, “The evaluation of v ud , experiment and theory,” Journal of Physics G: Nuclear and Particle Physics, vol. 29, no. 1, p. 197, 2003. [56] I. S. Towner and J. C. Hardy, “An Improved calculation of the isospin-symmetrybreaking corrections to superallowed Fermi beta decay,” Phys. Rev., vol. C77, p. 025501, 2008. [57] I. Towner, “Quenching of spin operators in the calculation of radiative corrections for nuclear beta decay,” Physics Letters B, vol. 333, no. 1, pp. 13 – 16, 1994. [58] I. Towner, “The nuclear-structure dependence of radiative corrections in superallowed fermi beta-decay,” Nuclear Physics A, vol. 540, no. 3, pp. 478 – 500, 1992. [59] F. Barker, B. Brown, W. Jaus, and G. Rasche, “Determination of —vud— from fermi decays and the unitarity of the km-mixing matrix,” Nuclear Physics A, vol. 540, no. 3, pp. 501 – 519, 1992. [60] W. Jaus and G. Rasche, “Nuclear-structure dependence of o(α) corrections to fermi decays and the value of the kobayashi-maskawa matrix element Vud ,” Phys. Rev. D, vol. 41, pp. 166–176, Jan 1990. [61] J. C. Hardy and I. S. Towner, “Superallowed 0+ → 0+ nuclear β decays: A critical survey with tests of the conserved vector current hypothesis and the standard model,” Phys. Rev. C, vol. 71, p. 055501, May 2005. [62] J. Suhonen and O. Civitarese, “Weak-interaction and nuclear-structure aspects of nuclear double beta decay,” Physics Reports, vol. 300, no. 3, pp. 123 – 214, 1998. [63] F. T. Avignone, S. R. Elliott, and J. Engel, “Double beta decay, majorana neutrinos, and neutrino mass,” Rev. Mod. Phys., vol. 80, pp. 481–516, Apr 2008. [64] S. R. Elliott, A. A. Hahn, and M. K. Moe, “Direct evidence for two-neutrino doublebeta decay in 82 Se,” Phys. Rev. Lett., vol. 59, pp. 2020–2023, Nov 1987. [65] H. S. Miley, F. T. Avignone, R. L. Brodzinski, J. I. Collar, and J. H. Reeves, “Suggestive evidence for the two-neutrino double-β decay of 76 Ge,” Phys. Rev. Lett., vol. 65, pp. 3092–3095, Dec 1990. [66] D. L. Lincoln, J. D. Holt, G. Bollen, M. Brodeur, S. Bustabad, J. Engel, S. J. Novario, M. Redshaw, R. Ringle, and S. Schwarz, “First direct double-β decay q-value measurement of 82 Se in support of understanding the nature of the neutrino,” Phys. Rev. Lett., vol. 110, p. 012501, Jan 2013. 152 [67] K. Gulyuz, J. Ariche, G. Bollen, S. Bustabad, M. Eibach, C. Izzo, S. J. Novario, M. Redshaw, R. Ringle, R. Sandler, S. Schwarz, and A. A. Valverde, “Determination of the direct double-β-decay q value of 96 Zr and atomic masses of 90−92,94,96 Zr and 92,94−98,100 Mo,” Phys. Rev. C, vol. 91, p. 055501, May 2015. [68] M. Redshaw, G. Bollen, M. Brodeur, S. Bustabad, D. L. Lincoln, S. J. Novario, R. Ringle, and S. Schwarz, “Atomic mass and double-β-decay q value of 48 Ca,” Phys. Rev. C, vol. 86, p. 041306, Oct 2012. [69] S. Bustabad, G. Bollen, M. Brodeur, D. L. Lincoln, S. J. Novario, M. Redshaw, R. Ringle, S. Schwarz, and A. A. Valverde, “First direct determination of the 48 Ca double-β decay q value,” Phys. Rev. C, vol. 88, p. 022501, Aug 2013. [70] K. Hirata, T. Kajita, M. Koshiba, M. Nakahata, Y. Oyama, N. Sato, A. Suzuki, M. Takita, Y. Totsuka, T. Kifune, T. Suda, K. Takahashi, T. Tanimori, K. Miyano, M. Yamada, E. W. Beier, L. R. Feldscher, S. B. Kim, A. K. Mann, F. M. Newcomer, R. Van, W. Zhang, and B. G. Cortez, “Observation of a neutrino burst from the supernova sn1987a,” Phys. Rev. Lett., vol. 58, pp. 1490–1493, Apr 1987. [71] R. M. Bionta, G. Blewitt, C. B. Bratton, D. Casper, A. Ciocio, R. Claus, B. Cortez, M. Crouch, S. T. Dye, S. Errede, G. W. Foster, W. Gajewski, K. S. Ganezer, M. Goldhaber, T. J. Haines, T. W. Jones, D. Kielczewska, W. R. Kropp, J. G. Learned, J. M. LoSecco, J. Matthews, R. Miller, M. S. Mudan, H. S. Park, L. R. Price, F. Reines, J. Schultz, S. Seidel, E. Shumard, D. Sinclair, H. W. Sobel, J. L. Stone, L. R. Sulak, R. Svoboda, G. Thornton, J. C. van der Velde, and C. Wuest, “Observation of a neutrino burst in coincidence with supernova 1987a in the large magellanic cloud,” Phys. Rev. Lett., vol. 58, pp. 1494–1496, Apr 1987. [72] C. Hargrove, I. Batkin, M. Sundaresan, and J. Dubeau, “A lead astronomical neutrino detector: Land,” Astroparticle Physics, vol. 5, no. 2, pp. 183 – 196, 1996. [73] D. B. Cline, G. M. Fuller, W. P. Hong, B. Meyer, and J. Wilson, “Prospects for detection of a cosmologically significant neutrino mass from a galactic supernova neutrino burst using a neutral-current-based detector,” Phys. Rev. D, vol. 50, pp. 720–729, Jul 1994. [74] G. Ewan, “The sudbury neutrino observatory,” Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, vol. 314, no. 2, pp. 373 – 379, 1992. [75] K. Langanke, P. Vogel, and E. Kolbe, “Signal for supernova νµ and ντ neutrinos in water Čerenkov detectors,” Phys. Rev. Lett., vol. 76, pp. 2629–2632, Apr 1996. [76] E. Kolbe, K. Langanke, G. Martnez-Pinedo, and P. Vogel, “Neutrinonucleus reactions and nuclear structure,” Journal of Physics G: Nuclear and Particle Physics, vol. 29, no. 11, p. 2569, 2003. 153 [77] O. Benhar, N. Farina, H. Nakamura, M. Sakuda, and R. Seki, “Electron- and neutrinonucleus scattering in the impulse approximation regime,” Phys. Rev. D, vol. 72, p. 053005, Sep 2005. [78] E. Fermi, “Versuch einer Theorie der β-Strahlen. I,” Zeitschrift fur Physik, vol. 88, pp. 161–177, Mar. 1934. [79] F. L. Wilson, “Fermi’s theory of beta decay,” American Journal of Physics, vol. 36, no. 12, pp. 1150–1160, 1968. [80] B. H. Wildenthal, M. S. Curtin, and B. A. Brown, “Predicted features of the beta decay of neutron-rich sd-shell nuclei,” Phys. Rev. C, vol. 28, pp. 1343–1366, Sep 1983. [81] B. A. Brown and B. H. Wildenthal, “Corrections to the free-nucleon values of the singleparticle matrix elements of the m1 and gamow-teller operators, from a comparison of shell-model predictions with sd-shell data,” Phys. Rev. C, vol. 28, pp. 2397–2413, Dec 1983. [82] E. K. Warburton and B. A. Brown, “Effective interactions for the 0p1s0d nuclear shell-model space,” Phys. Rev. C, vol. 46, pp. 923–944, Sep 1992. [83] W. E. Ormand and B. A. Brown, “Isospin-mixing corrections for fp-shell fermi transitions,” Phys. Rev. C, vol. 52, pp. 2455–2460, Nov 1995. [84] I. S. Towner and F. C. Khanna, “Quenching of allowed gamow-teller β transitions in mirror nuclei,” Phys. Rev. Lett., vol. 42, pp. 51–54, Jan 1979. [85] A. Ekström, G. R. Jansen, K. A. Wendt, G. Hagen, T. Papenbrock, S. Bacca, B. Carlsson, and D. Gazit, “Effects of three-nucleon forces and two-body currents on gamowteller strengths,” Phys. Rev. Lett., vol. 113, p. 262504, Dec 2014. [86] S. K. Bogner, H. Hergert, J. D. Holt, A. Schwenk, S. Binder, A. Calci, J. Langhammer, and R. Roth, “Nonperturbative shell-model interactions from the in-medium similarity renormalization group,” Phys. Rev. Lett., vol. 113, p. 142501, Oct 2014. [87] J. Suhonen and O. Civitarese, “Probing the quenching of ga by single and double beta decays,” Physics Letters B, vol. 725, no. 1, pp. 153 – 157, 2013. [88] J. Engel, “Nuclear matrix elements for double-b decay,” XXVI International Conference on Neutrino Physics and Astrophysics (Neutrino 2014), vol. 1666, 2015. [89] J. C. Slater, “The theory of complex spectra,” Phys. Rev., vol. 34, pp. 1293–1322, Nov 1929. [90] A. Fetter and J. Walecka, Quantum Theory of Many-particle Systems. Dover Books on Physics, Dover Publications, 2003. [91] G. C. Wick, “The evaluation of the collision matrix,” Phys. Rev., vol. 80, pp. 268–272, Oct 1950. 154 [92] D. R. Hartree, “The wave mechanics of an atom with a non-coulomb central field. part i. theory and methods,” Mathematical Proceedings of the Cambridge Philosophical Society, vol. 24, no. 1, p. 89110, 1928. [93] V. Fock, “Näherungsmethode zur lösung des quantenmechanischen mehrkörperproblems,” Zeitschrift für Physik, vol. 61, pp. 126–148, Jan 1930. [94] D. Thouless, “Stability conditions and nuclear rotations in the hartree-fock theory,” Nuclear Physics, vol. 21, pp. 225 – 232, 1960. [95] L. Brillouin, “Les problèmes de perturbations et les champs self-consistents,” Le journal de physique et le radium, vol. 3, no. 9, pp. 373–389, 1932. [96] P. Pulay, “Convergence acceleration of iterative sequences. the case of scf iteration,” Chem. Phys. Lett., vol. 73, no. 2, pp. 393 – 398, 1980. [97] P. Pulay, “Improved scf convergence acceleration,” J. Comput. Chem., vol. 3, no. 4, pp. 556–560, 1982. [98] C. G. Broyden, “A class of methods for solving nonlinear simultaneous equations,” Math. Comput., vol. 19, no. 92, pp. 577–593, 1965. [99] E. U. Condon, “The theory of complex spectra,” Phys. Rev., vol. 36, pp. 1121–1133, Oct 1930. [100] C. D. Sherrill and H. F. Schaefer, “The configuration interaction method: Advances in highly correlated approaches,” vol. 34 of Advances in Quantum Chemistry, pp. 143 – 269, Academic Press, 1999. [101] C. Møller and M. S. Plesset, “Note on an approximation treatment for many-electron systems,” Phys. Rev., vol. 46, pp. 618–622, Oct 1934. [102] J. Rayleigh, The Theory of Sound. The Theory of Sound, Macmillan, 1894. [103] E. Schrödinger, “Quantification of the eigen value problem,” Annalen der Physik, vol. 80, no. 13, pp. 437–490, 1926. [104] L. M. Frantz and R. L. Mills, “Many-body basis for the optical model,” Nuclear Physics, vol. 15, pp. 16 – 32, 1960. [105] B. H. Brandow, “Linked-cluster expansions for the nuclear many-body problem,” Rev. Mod. Phys., vol. 39, pp. 771–828, Oct 1967. [106] T. P. Živković, “Existence and reality of solutions of the coupled-cluster equations,” International Journal of Quantum Chemistry, vol. 12, no. S11, pp. 413–420, 1977. [107] P. Piecuch, S. Zarrabian, J. Paldus, and J. Čı́žek, “Coupled-cluster approaches with an approximate account of triexcitations and the optimized-inner-projection technique. ii. coupled-cluster results for cyclic-polyene model systems,” Phys. Rev. B, vol. 42, pp. 3351–3379, Aug 1990. 155 [108] P. Piecuch and K. Kowalski, “In search of the relationship between multiple solutions characterizing coupled-cluster theories,” in Computational Chemistry: Reviews of Current Trends (J. Leszczynski, ed.), vol. 5, pp. 1–104, Singapore: World Scientific, 2000. [109] N. Moiseyev, Non-Hermitian Quantum Mechanics. Cambridge University Press, 2011. [110] R. J. Bartlett and M. Musial, “Coupled-cluster theory in quantum chemistry,” Rev. Mod. Phys., vol. 79, pp. 291–352, Feb 2007. [111] G. Hagen, T. Papenbrock, M. Hjorth-Jensen, and D. J. Dean, “Coupled-cluster computations of atomic nuclei,” Rep. Prog. in Phys., vol. 77, p. 096302, Sep 2014. [112] P. Szakács and P. R. Surján, “Stability conditions for the coupled cluster equations,” International Journal of Quantum Chemistry, vol. 108, no. 12, pp. 2043–2052, 2008. [113] “An updated set of basic linear algebra subprograms (blas),” ACM Trans. Math. Softw., vol. 28, p. 135, 2002. [114] E. K. U. Gross, E. Runge, and O. Heinonen, Many-particle theory / E.K.U. Gross, E. Runge, O. Heinonen. A. Hilger Bristol ; Philadelphia, 1991. [115] N. Michel, W. Nazarewicz, M. Ploszajczak, and J. Rotureau, “Antibound states and halo formation in the gamow shell model,” Phys. Rev. C, vol. 74, p. 054305, Nov 2006. [116] G. Hagen, D. Dean, M. Hjorth-Jensen, and T. Papenbrock, “Complex coupled-cluster approach to an ab-initio description of open quantum systems,” Physics Letters B, vol. 656, no. 4, pp. 169 – 173, 2007. [117] E. Wigner and J. Griffin, Group Theory and Its Application to the Quantum Mechanics of Atomic Spectra. Pure and applied Physics, Academic Press, 1959. [118] C. Eckart, “The application of group theory to the quantum dynamics of monatomic systems,” Rev. Mod. Phys., vol. 2, pp. 305–380, Jul 1930. [119] T. Hatsuda and T. Kunihiro, “Qcd phenomenology based on a chiral effective lagrangian,” Physics Reports, vol. 247, no. 5, pp. 221 – 367, 1994. [120] G. P. Lepage and S. J. Brodsky, “Exclusive processes in perturbative quantum chromodynamics,” Phys. Rev. D, vol. 22, pp. 2157–2198, Nov 1980. [121] S. R. Beane, E. Chang, W. Detmold, H. W. Lin, T. C. Luu, K. Orginos, A. Parreño, M. J. Savage, A. Torok, and A. Walker-Loud, “Deuteron and exotic two-body bound states from lattice qcd,” Phys. Rev. D, vol. 85, p. 054511, Mar 2012. [122] J. Beringer, J. F. Arguin, R. M. Barnett, K. Copic, O. Dahl, D. E. Groom, C. J. Lin, J. Lys, H. Murayama, C. G. Wohl, W. M. Yao, P. A. Zyla, C. Amsler, M. Antonelli, D. M. Asner, H. Baer, H. R. Band, T. Basaglia, C. W. Bauer, J. J. Beatty, V. I. Belousov, E. Bergren, G. Bernardi, W. Bertl, S. Bethke, H. Bichsel, O. Biebel, 156 E. Blucher, S. Blusk, G. Brooijmans, O. Buchmueller, R. N. Cahn, M. Carena, A. Ceccucci, D. Chakraborty, M. C. Chen, R. S. Chivukula, G. Cowan, G. D’Ambrosio, T. Damour, D. de Florian, A. de Gouvêa, T. DeGrand, P. de Jong, G. Dissertori, B. Dobrescu, M. Doser, M. Drees, D. A. Edwards, S. Eidelman, J. Erler, V. V. Ezhela, W. Fetscher, B. D. Fields, B. Foster, T. K. Gaisser, L. Garren, H. J. Gerber, G. Gerbier, T. Gherghetta, S. Golwala, M. Goodman, C. Grab, A. V. Gritsan, J. F. Grivaz, M. Grünewald, A. Gurtu, T. Gutsche, H. E. Haber, K. Hagiwara, C. Hagmann, C. Hanhart, S. Hashimoto, K. G. Hayes, M. Heffner, B. Heltsley, J. J. HernándezRey, K. Hikasa, A. Höcker, J. Holder, A. Holtkamp, J. Huston, J. D. Jackson, K. F. Johnson, T. Junk, D. Karlen, D. Kirkby, S. R. Klein, E. Klempt, R. V. Kowalewski, F. Krauss, M. Kreps, B. Krusche, Y. V. Kuyanov, Y. Kwon, O. Lahav, J. Laiho, P. Langacker, A. Liddle, Z. Ligeti, T. M. Liss, L. Littenberg, K. S. Lugovsky, S. B. Lugovsky, T. Mannel, A. V. Manohar, W. J. Marciano, A. D. Martin, A. Masoni, J. Matthews, D. Milstead, R. Miquel, K. Mönig, F. Moortgat, K. Nakamura, M. Narain, P. Nason, S. Navas, M. Neubert, P. Nevski, Y. Nir, K. A. Olive, L. Pape, J. Parsons, C. Patrignani, J. A. Peacock, S. T. Petcov, A. Piepke, A. Pomarol, G. Punzi, A. Quadt, S. Raby, G. Raffelt, B. N. Ratcliff, P. Richardson, S. Roesler, S. Rolli, A. Romaniouk, L. J. Rosenberg, J. L. Rosner, C. T. Sachrajda, Y. Sakai, G. P. Salam, S. Sarkar, F. Sauli, O. Schneider, K. Scholberg, D. Scott, W. G. Seligman, M. H. Shaevitz, S. R. Sharpe, M. Silari, T. Sjöstrand, P. Skands, J. G. Smith, G. F. Smoot, S. Spanier, H. Spieler, A. Stahl, T. Stanev, S. L. Stone, T. Sumiyoshi, M. J. Syphers, F. Takahashi, M. Tanabashi, J. Terning, M. Titov, N. P. Tkachenko, N. A. Törnqvist, D. Tovey, G. Valencia, K. van Bibber, G. Venanzoni, M. G. Vincter, P. Vogel, A. Vogt, W. Walkowiak, C. W. Walter, D. R. Ward, T. Watari, G. Weiglein, E. J. Weinberg, L. R. Wiencke, L. Wolfenstein, J. Womersley, C. L. Woody, R. L. Workman, A. Yamamoto, G. P. Zeller, O. V. Zenin, J. Zhang, R. Y. Zhu, G. Harper, V. S. Lugovsky, and P. Schaffner, “Review of particle physics,” Phys. Rev. D, vol. 86, p. 010001, Jul 2012. [123] D. R. Entem and R. Machleidt, “Accurate charge-dependent nucleon-nucleon potential at fourth order of chiral perturbation theory,” Phys. Rev. C, vol. 68, p. 041001, Oct 2003. [124] P. Navrátil, “Local three-nucleon interaction from chiral effective field theory,” FewBody Systems, vol. 41, pp. 117–140, Dec 2007. [125] R. Roth, S. Binder, K. Vobig, A. Calci, J. Langhammer, and P. Navrátil, “Mediummass nuclei with normal-ordered chiral N N + 3N interactions,” Phys. Rev. Lett., vol. 109, p. 052501, Jul 2012. [126] S. Levit and D. Orgad, “Statistics of hartree-fock levels in small disordered systems,” Physical review. B, Condensed matter, vol. 60, no. 8, pp. 5549–5560, 1999. [127] H. J. Lipkin, “Center-of-mass motion in the nuclear shell model,” Phys. Rev., vol. 110, pp. 1395–1397, Jun 1958. [128] D. Gloeckner and R. Lawson, “Spurious center-of-mass motion,” Physics Letters B, vol. 53, no. 4, pp. 313 – 318, 1974. 157 [129] R. F. Bishop, M. F. Flynn, M. C. Boscá, E. Buenda, and R. Guardiola, “Translationally invariant coupled cluster theory for simple finite systems,” Phys. Rev. C, vol. 42, pp. 1341–1360, Oct 1990. [130] A. Nogga, H. Kamada, W. Glöckle, and B. R. Barrett, “The α particle based on modern nuclear forces,” Phys. Rev. C, vol. 65, p. 054003, May 2002. [131] C. M. Vincent, “Optimal separation of center-of-mass motion,” Phys. Rev. C, vol. 8, pp. 929–937, Sep 1973. [132] G. Hagen, T. Papenbrock, and D. J. Dean, “Solution of the center-of-mass problem in nuclear structure calculations,” Phys. Rev. Lett., vol. 103, p. 062503, Aug 2009. [133] G. R. Jansen, “Spherical coupled-cluster theory for open-shell nuclei,” Phys. Rev. C, vol. 88, p. 024305, Aug 2013. [134] G. H. Diercksen, B. O. Roos, and A. J. Sadlej, “Legitimate calculation of first-order molecular properties in the case of limited ci functions. dipole moments,” Chemical Physics, vol. 59, no. 1, pp. 29 – 39, 1981. [135] M. Ernzerhof, C. Marian, and S. Peyerimhoff, “Energy derivative versus expectation value approach: the dipole moment of co,” Chemical Physics Letters, vol. 204, no. 1, pp. 59 – 64, 1993. [136] D. J. Rowe, “Equations-of-motion method and the extended shell model,” Rev. Mod. Phys., vol. 40, pp. 153–166, Jan 1968. [137] C. Lanczos, “An iteration method for the solution of the eigenvalue problem of linear differential and integral operators,” Journal of research of the National Bureau of Standards (1934), vol. 45, pp. 255–282, 1950 10. [138] W. E. Arnoldi, “The principle of minimized iterations in the solution of the matrix eigenvalue problem,” Quarterly of applied mathematics, vol. 9, no. 1, pp. 17–29, 1951. [139] R. Lehoucq, D. Sorensen, and C. Yang, ARPACK Users’ Guide. Society for Industrial and Applied Mathematics, 1998. [140] S. M. Reimann and M. Manninen, “Electronic structure of quantum dots,” Rev. Mod. Phys., vol. 74, p. 1238, 2002. [141] H.-A. Engel, V. N. Golovach, D. Loss, L. M. K. Vandersypen, J. M. Elzerman, R. Hanson, and L. P. Kouwenhoven, “Measurement Efficiency and n-Shot Readout of Spin Qubits,” Phys. Rev. Lett., vol. 93, p. 106804, 2004. [142] J. Birman, R. Nazmitdinov, and V. Yukalov, “Effects of symmetry breaking in finite quantum systems,” Phys. Rep., vol. 526, no. 1, pp. 1 – 91, 2013. Effects of symmetry breaking in finite quantum systems. 158 [143] S. Tarucha, D. G. Austing, T. Honda, R. J. van der Hage, and L. P. Kouwenhoven, “Shell Filling and Spin Effects in a Few Electron Quantum Dot,” Phys. Rev. Lett., vol. 77, p. 3613, 1996. [144] M. P. Lohne, “Coupled-cluster studies of quantum dots,” Master’s thesis, University of Oslo, 2010. [145] “NIST Digital Library of Mathematical Functions.” http://dlmf.nist.gov/, Release 1.0.14 of 2016-12-21, 2016. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller and B. V. Saunders, eds. [146] I. Lindgren, “The rayleigh-schrodinger perturbation and the linked-diagram theorem for a multi-configurational model space,” J. Phys. Pt. B Atom. M. P., vol. 7, no. 18, p. 2441, 1974. [147] J. Høgberget, “Quantum Monte-Carlo Studies of Generalized Many-body Systems,” Master’s thesis, University of Oslo, 2013. [148] V. K. B. Olsen, “Full Configuration Interaction Simulation of Quantum Dots,” Master’s thesis, University of Oslo, 2013. [149] N. M. Parzuchowski, T. D. Morris, and S. K. Bogner, “Ab initio excited states from the in-medium similarity renormalization group,” Phys. Rev. C, vol. 95, p. 044304, Apr 2017. [150] K. Emrich, “An extension of the coupled cluster formalism to excited states (i),” Nuclear Physics A, vol. 351, no. 3, pp. 379 – 396, 1981. [151] K. Emrich, “An extension of the coupled cluster formalism to excited states: (ii). approximations and tests,” Nuclear Physics A, vol. 351, no. 3, pp. 397 – 438, 1981. [152] K. Emrich and J. Zabolitzky, “Negative parity states in 16 O from coupled-cluster equations,” Nuclear Physics A, vol. 351, no. 3, pp. 439 – 455, 1981. [153] J. F. Stanton and R. J. Bartlett, “The equation of motion coupledcluster method. a systematic biorthogonal approach to molecular excitation energies, transition probabilities, and excited state properties,” The Journal of Chemical Physics, vol. 98, no. 9, pp. 7029–7039, 1993. [154] G. Hagen, M. Hjorth-Jensen, G. R. Jansen, R. Machleidt, and T. Papenbrock, “Continuum effects and three-nucleon forces in neutron-rich oxygen isotopes,” Phys. Rev. Lett., vol. 108, p. 242501, Jun 2012. [155] A. Ekström, G. R. Jansen, K. A. Wendt, G. Hagen, T. Papenbrock, B. D. Carlsson, C. Forssén, M. Hjorth-Jensen, P. Navrátil, and W. Nazarewicz, “Accurate nuclear radii and binding energies from a chiral interaction,” Phys. Rev. C, vol. 91, p. 051301, May 2015. 159 [156] F. Halzen and A. D. Martin, Quarks and Leptons: An Introductory Course in Modern Particle Physics. 1984. [157] J. C. Hardy and I. S. Towner, “Superallowed 0+ → 0+ nuclear β decays: A critical survey with tests of the conserved vector current hypothesis and the standard model,” Phys. Rev. C, vol. 71, p. 055501, May 2005. [158] I. S. Towner and J. C. Hardy, “Improved calculation of the isospin-symmetry-breaking corrections to superallowed fermi β decay,” Phys. Rev. C, vol. 77, p. 025501, Feb 2008. [159] K. Kubodera, J. Delorme, and M. Rho, “Axial currents in nuclei,” Phys. Rev. Lett., vol. 40, pp. 755–758, Mar 1978. [160] E. Oset and M. Rho, “Axial currents in nuclei: The gamow-teller matrix element,” Phys. Rev. Lett., vol. 42, pp. 47–50, Jan 1979. [161] D. Wilkinson, “Analysis of neutron beta decay,” Nuclear Physics A, vol. 377, no. 2, pp. 474 – 504, 1982. [162] J. Suhonen, From Nucleons to Nucleus. Theoretical and Mathematical Physics, Berlin, Germany: Springer, 2007. 160