DEVELOPMENT AND APPLICATIONS OF SEMI-STOCHASTIC COUPLED-CLUSTER METHODS FOR OPEN-SHELL SYSTEMS AND ELECTRONIC EXCITATIONS IN MOLECULES By Arnab Chakraborty A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Chemistry—Doctor of Philosophy 2024 ABSTRACT It is well established that the exponential wave function ansatz of coupled-cluster (CC) theory and its equation-of-motion (EOM) extension to excited states are among the most appealing ways to describe the electronic structure of molecules. One of the key challenges in the development of the CC and EOMCC methodologies is the incorporation of many-electron correlation effects due to higher–than–two-body components of the cluster and EOM ex- citation operators, needed to achieve a quantitative description, without running into the usually prohibitive computational costs of the CC approach with singles, doubles, and triples (CCSDT) and its excited-state EOMCCSDT extension, the CC method with singles, doubles, triples, and quadruples (CCSDTQ) and its excited-state EOMCCSDTQ extension, etc., and without resorting to perturbative CCSD(T)-type ideas that fail in multireference situations, such as bond breaking and excited states dominated by two-electron transitions. One of the most promising approaches in this area is the semi-stochastic CC(P ;Q) methodology that identifies the most important higher–than–doubly excited determinants needed in the high- level CC/EOMCC calculations with the help of the stochastic configuration interaction (CI) and CC Quantum Monte Carlo (QMC) wave function propagations and uses the suitably designed deterministic iterative and noniterative steps of the CC(P;Q) formalism to converge the desired CCSDT/EOMCCSDT, CCSDTQ/EOMCCSDTQ, etc., energetics. In this dis- sertation, we first discuss our recent work on extending the semi-stochastic, CIQMC-driven, particle-conserving CC(P ;Q) framework to excited electronic states and open-shell systems. We tested performance of the resulting methods by examining their ability to recover the high-level CCSDT/EOMCCSDT energetics in calculations of the electronic excitation spec- tra of the CH+, CH, and CNC molecules and singlet–triplet gaps in a few biradical systems, including methylene, (HFH)−, cyclobutadiene, cyclopentadienyl cation, and trimethylen- emethane. The second part of this dissertation focuses on an alternative way of determining ground and excited states of open-shell systems within the EOMCC framework by turning to the single and double electron attachment (EA) and single and double ionization potential (IP) EOMCC schemes. By generating ground and excited states of open-shell species, such as radicals and biradicals, with the help of suitably designed operators that can formally add electrons to or remove electrons from the parent closed-shell cores (an operation pro- ducing the appropriate multi-configurational reference space within a single-reference frame- work, while relaxing the remaining electrons), the resulting EA/IP- and DEA/DIP-EOMCC methods offer several advantages over the particle-conserving CC/EOMCC treatments, in- cluding rigorous spin and symmetry adaptation of the calculated electronic states and the ability of handling high- and low-spin states in an accurate and well-balanced manner. We demonstrate how to utilize the stochastic CIQMC wave function propagations, which are of particle-conserving character, in identifying the dominant higher-order 3-particle-2-hole (3p- 2h)/3-hole-2-particle (3h-2p) and 4-particle-2-hole (4p-2h)/4-hole-2-particle (4h-2p) compo- nents of the respective particle-nonconserving electron attaching/ionizing operators, needed to obtain a quantitative description, without having to resort to the previously exploited user- and system-dependent active-space concepts. The effectiveness of the semi-stochastic, CIQMC-driven, EA/IP/DEA/DIP-EOMCC approaches will be illustrated by examining the adiabatic excitations in the C2N, CNC, N3, and NCO radicals and by revisiting the singlet– triplet gaps in methylene and trimethylenemethane. Copyright by ARNAB CHAKRABORTY 2024 To my parents, teachers, and friends. v ACKNOWLEDGMENTS First and foremost, I would like to thank my doctoral advisor, Professor Piotr Piecuch, for his invaluable advice, guidance, and support throughout my Ph.D. journey. He introduced me to the world of coupled-cluster and other many-body theories, which I sincerely appreciate. Professor Piecuch’s readiness to engage in scientific discussions and his inspiring presence have been fundamental to my academic growth. I am grateful to him for sharing his deep knowledge, his patient mentoring, and his constant motivation to approach research with extreme care. I would also like to acknowledge the current and previous members of my guidance com- mittee: Professor Katharine Hunt, Professor James Jackson, Professor Angela Wilson, and Professor Benjamin Levine, for their support, time, and advice. I would like to thank Pro- fessor Marcos Dantus for the opportunity to work on the extensive study of H3 + generation from the double ionization of organic halogens and pseudo halogens. This collaboration has taught me how to communicate ideas to a diverse group of scientists. I would also like to thank Professors Madhav Ranganathan, Srihari Keshavamurthy, and Nishanth Nair from the Indian Institute of Technology Kanpur; Professor Satrajit Adhikari from the Indian Associ- ation for the Cultivation of Science, and Professor Pinaki Chaudhury from the University of Calcutta for sparking my interest in theoretical chemistry. Furthermore, I would like to thank Dr. Kathryn Severin and Dr. Elizabeth McGaw, under whom I have had so many wonderful semesters of teaching opportunities for CEM 395 and CEM 495, and Professor Katharine Hunt, under whom I had the great opportunity to teach CEM 484. These teaching assignments gave me the scope to develop my communication skills and helped me understand the fundamental concepts of physical chemistry more deeply. I am also grateful to the current and former Piecuch group members: Dr. Jun Shen, Dr. Suhita Basumallick, Dr. J. Emiliano Deustua, Dr. Ilias Magoulas, Dr. Stephen Yuwono, Mr. Karthik Gururangan, Mr. Tiange Deng, Ms. Swati Priyadarsini, and Mr. Agnibha Hanra. Their support, help, and encouragement made my Ph.D. journey unforgettable. vi I would also like to thank all my friends from the East Lansing area. My Ph.D. experience would not have been so enjoyable without you all. I want to thank my parents, Mr. Hari Sadhan Chakraborty and Mrs. Mira Chakraborty, for their continuous support throughout my life. Thank you all! vii TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 CHAPTER 2 BACKGROUND INFORMATION . . . . . . . . . . . . . . . . . . 15 2.1 Single-Reference Coupled-Cluster Theory and Its Equation-of-Motion Extension to Excited Electronic States . . . . . . . . . . . . . . . . . . . 2.2 The CC(P;Q) Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Particle Nonconserving Equation-of-Motion Coupled-Cluster Theories: The Single and Double Electron Attachment and Ionization Potential Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Overview of Configuration Interaction Quantum Monte Carlo CHAPTER 3 THE SEMI-STOCHASTIC CC(P;Q) METHODOLOGY FOR GROUND AND EXCITED STATES . . . . . . . . . . . . . . . . 3.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Electronic Excitations in CH+, CH, and CNC . . . . . . . . . . . . . . . 3.3 Singlet–Triplet Gaps in Methylene, (HFH)−, Cyclobutadiene, Cyclopen- 15 19 22 30 34 34 39 tadienyl Cation, and Trimethylenemethane . . . . . . . . . . . . . . . . . 61 CHAPTER 4 THE SEMI-STOCHASTIC EXTENSIONS OF PARTICLE NONCONSERVING EQUATION-OF-MOTION COUPLED-CLUSTER THEORIES . . . . . . . . . . . . . . . . . 111 4.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 4.2 Adiabatic Excitations in C2N, CNC, N3, and NCO . . . . . . . . . . . . 116 4.3 Singlet–Triplet Gaps in Methylene and Trimethylenemethane . . . . . . . 137 CHAPTER 5 CONCLUDING REMARKS AND FUTURE OUTLOOK . . . . . 155 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 viii LIST OF TABLES Table 3.1 Convergence of the CC(P )/EOMCC(P ) and CC(P;Q) energies toward CCSDT/EOMCCSDT for CH+, calculated using the [5s3p1d/3s1p] ba- sis set of Ref. [164], at the C–H internuclear distance R = Re = 2.13713 bohr. The P spaces used in the CC(P ) and EOMCC(P ) calculations were defined as all singles, all doubles, and subsets of triples extracted from i-FCIQMC propagations for the lowest states of the relevant sym- metries. Each i-FCIQMC run was initiated by placing 1500 walkers on the appropriate reference function [the RHF determinant for the 1Σ+ g states, the 3σ → 1π state of the 1B1(C2v) symmetry for the 1Π states, and the 3σ2 → 1π2 state of the 1A2(C2v) symmetry for the 1∆ states], set- ting the initiator parameter na at 3, and the time step ∆τ at 0.0001 a.u. The Q spaces used in constructing the CC(P;Q) corrections consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [100]. . . . . . 48 Table 3.2 Same as Table 3.1 for the stretched C–H internuclear distance R = 2Re = 4.27426 bohr. Adapted from Ref. [100]. . . . . . . . . . . . . . . . . . . . 49 Table 3.3 Convergence of the CC(P )/EOMCC(P ) and CC(P;Q) energies toward CCSDT/EOMCCSDT for CH, calculated using the aug-cc-pVDZ basis set. The P spaces used in the CC(P ) and EOMCC(P ) calculations were defined as all singles, all doubles, and subsets of triples extracted from i-FCIQMC propagations for the lowest states of the relevant symme- tries. Each i-FCIQMC run was initiated by placing 1500 walkers on the appropriate reference function [the ROHF 2B2(C2v) determinant for the X 2Π state, the 1π → 4σ state of the 2A1(C2v) symmetry for the A 2∆ and C 2Σ+ states, and the 3σ → 1π state of the 2A2(C2v) symmetry for the B 2Σ− state], setting the initiator parameter na at 3, and the time step ∆τ at 0.0001 a.u. The Q spaces used in constructing the CC(P;Q) corrections consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [100]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.4 Convergence of the CC(P )/EOMCC(P ) and CC(P;Q) energies toward CCSDT/EOMCCSDT for CNC, calculated using DZP[4s2p1d] basis set. The P spaces used in the CC(P ) and EOMCC(P ) calculations were defined as all singles, all doubles, and subsets of triples extracted from i-FCIQMC propagations for the lowest states of the relevant symmetries. Each i-FCIQMC run was initiated by placing 1500 walkers on the appro- priate reference function [the ROHF 2B2g(D2h) determinant for the X 2Πg state and the 3σu → 1πg state of the 2B1u(D2h) symmetry for the A 2∆u and B 2Σ+ u states], setting the initiator parameter na at 3, and the time step ∆τ at 0.0001 a.u. The Q spaces used in constructing the CC(P;Q) corrections consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [100]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 59 ix Table 3.5 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3B1 and A 1A1 states of methylene, as described by the aug-cc-pVTZ basis set, and of the corresponding adiabatic singlet–triplet gaps toward their parent CCSDT values. The geometries of the X 3B1 and A 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were de- fined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) correc- tions consisted of the triply excited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the ROHF (X 3B1 state) and RHF (A 1A1 state) reference determinants and the na parameter of the initiator algorithm was set at 3. In all post- Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d and f orbitals were employed throughout. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.6 The total numbers of walkers, reported as percentages of the total walker populations at 200000 MC iterations, characterizing the i-FCIQMC prop- agations with δτ = 0.0001 a.u. that were needed to generate the CC(P) and CC(P;Q) results for methylene reported in Table 3.5. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.7 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1Σ+ g state of (HFH)−, as described by the 6-31G(d,p) basis set, at selected H–F distances RH-F toward their parent CCSDT values. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the RHF reference determinant and the na parameter of the initiator algorithm was set at 3. In all post- Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 69 80 x Table 3.8 Convergence of the CC(P ) and CC(P ;Q) energies of the A 3Σ+ u state of (HFH)−, as described by the 6-31G(d,p) basis set, at selected H–F dis- tances RH-F toward their parent CCSDT values. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply ex- cited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the ROHF reference determi- nant and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.9 Convergence of the CC(P ) and CC(P ;Q) singlet–triplet gaps of (HFH)−, as described by the 6-31G(d,p) basis set, at selected H–F distances RH-F toward their parent CCSDT values. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the RHF (X 1Σ+ g state) and ROHF (A 3Σ+ u state) reference determinants and the na parameter of the ini- tiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. . . . . . . . Table 3.10 The total numbers of walkers, reported as percentages of the total walker populations at 200000 MC iterations, characterizing the i-FCIQMC prop- agations with δτ = 0.0001 a.u. that were needed to generate the CC(P) and CC(P;Q) results for the X 1Σ+ u states of (HFH)− reported in Tables 3.7 and 3.8. Adapted from Ref. [102]. . . . . . . . . . . . . . . . g and A 3Σ+ xi 81 82 83 Table 3.11 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1B1g and A 3A2g states of cyclobutadiene, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet–triplet gaps toward their parent CCSDT values. All calculations were performed at the D4h-symmetric transition-state geometry of the X 1B1g state optimized in the MR-AQCC calculations reported in Ref. [202]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited de- terminants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited de- terminants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the RHF (X 1B1g state) and ROHF (A 3A2g state) reference determinants and the na parameter of the ini- tiator algorithm was set at 3. In all post-Hartree–Fock calculations, the four lowest core orbitals were kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. . . . . Table 3.12 The total numbers of walkers, reported as percentages of the total walker populations at 80000 MC iterations, characterizing the i-FCIQMC prop- agations with δτ = 0.0001 a.u. that were needed to generate the CC(P ) and CC(P ;Q) results for cyclobutadiene reported in Table 3.11. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 3.13 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 and A 1E′ 2 states of cyclopentadienyl cation, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet–triplet gaps toward their parent CCSDT values. All calculations were performed at the D5h-symmetric geometry of the X 3A′ 2 state optimized using the unre- stricted CCSD/cc-pVDZ approach reported in Ref. [203]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determi- nants extracted from the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections con- sisted of the triply excited determinants not captured by the correspond- ing i-CISDTQ-MC runs. The i-CISDTQ-MC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the ROHF (X 3A′ 2 state) reference determinants and In all post- the na parameter of the initiator algorithm was set at 3. Hartree–Fock calculations, the five lowest core orbitals were kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. 2 state) and RHF (A 1E′ . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 92 99 xii Table 3.14 The total numbers of walkers, reported as percentages of the total walker populations at 80000 MC iterations, characterizing the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. that were needed to generate the CC(P ) and CC(P ;Q) results for the cyclopentadienyl cation reported in Table 3.13. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . 100 Table 3.15 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 and B 1A1 states of trimethylenemethane, as described by the cc-pVDZ basis set, and of the corresponding adiabatic singlet–triplet gaps toward their parent CCSDT values. The D3h- and C2v-symmetric geometries of the 2 and B 1A1 states, respectively, optimized in the SF-DFT/6-31G(d) X 3A′ calculations, were taken from Ref. [231]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and dou- bly excited determinants and subsets of triply excited determinants ex- tracted from the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-CISDTQ-MC runs. The i-CISDTQ-MC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on 2 state) and RHF (B 1A1 state) reference determinants the ROHF (X 3A′ and the na parameter of the initiator algorithm was set at 3. In all post- Hartree–Fock calculations, the four lowest core orbitals were kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Table 3.16 The total numbers of walkers, reported as percentages of the total walker populations at 80000 MC iterations, characterizing the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. that were needed to generate the CC(P ) and CC(P ;Q) results for trimethylenemethane reported in Table 3.15. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . 109 Table 4.1 Convergence of the EA-EOMCC(P ) energies [abbreviated as EA(P )] of the X 2Π, A 2∆, B 2Σ−, and C 2Σ+ states of C2N, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and of the corresponding adia- batic excitation energies toward their parent EA-EOMCC(3p-2h) values. The geometries of the X 2Π, A 2∆, B 2Σ−, and C 2Σ+ states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the EA-EOMCC(P ) calculations were defined as all 1p and 2p-1h determinants and subsets of 3p-2h deter- minants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the EA-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determi- nants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the carbon and nitrogen atoms were kept frozen. . . . . . 121 xiii Table 4.2 Convergence of the EA-EOMCC(P ) energies [abbreviated as EA(P )] of the X 2Πg, A 2∆u, and B 2Σ+ u states of CNC, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and of the corresponding adia- batic excitation energies toward their parent EA-EOMCC(3p-2h) values. The geometries of the X 2Πg, A 2∆u, and B 2Σ+ u states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the EA-EOMCC(P ) calculations were defined as all 1p and 2p-1h determinants and subsets of 3p-2h deter- minants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the EA-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determi- nants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the carbon and nitrogen atoms were kept frozen. . . . . . 126 Table 4.3 Convergence of the IP-EOMCC(P ) energies [abbreviated as IP(P )] of the X 2Πg and B 2Σ+ u states of N3, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and of the corresponding adiabatic excitation energy toward their parent IP-EOMCC(3h-2p) values. The geometries of the X 2Πg and B 2Σ+ u states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the IP-EOMCC(P ) calculations were defined as all 1h and 2h-1p determinants and subsets of 3h-2p determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the IP-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determinants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the nitrogen atoms were kept frozen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 Table 4.4 Convergence of the IP-EOMCC(P ) energies [abbreviated as IP(P )] of the X 2Π, A 2∆, and B 2Π states of NCO, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and the corresponding adiabatic excitation energies toward their parent IP-EOMCC(3h-2p) values. The geometries of the X 2Π, A 2∆u, and B 2Π states, optimized in the SAC-CI SDT- R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the IP-EOMCC(P ) calculations were defined as all 1h and 2h-1p determinants and subsets of 3h-2p determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the IP-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determinants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the carbon, nitrogen, and oxygen atoms were kept frozen. . . . . . . . . . . . . . . . . 135 xiv Table 4.5 Convergence of the DEA-EOMCC(P ) energies [abbreviated as DEA(P )] of the X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set of Ref. [180], and of the corresponding adiabatic singlet–triplet gaps toward their parent DEA-EOMCC(4p-2h) values. The geometries of the X 3B1, A 1A1, B 1B1, and C 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces used in the DEA-EOMCC(P ) calculations were defined as all 2p and 3p-1h determinants and subsets of 4p-2h determi- nants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations were initiated by placing 500 walkers on the corresponding ROHF reference determinant for the X 3B1 state and the corresponding RHF reference determinants for the remaining states and the na parameter of the initiator algorithm was set at 3. In all the post- Hartree–Fock calculations, the lowest core orbital of the carbon atom was kept frozen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 Table 4.6 Convergence of the DIP-EOMCC(P ) energies [abbreviated as DIP(P )] of the X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set of Ref. [180], and of the corresponding adiabatic singlet–triplet gaps toward their parent DIP-EOMCC(4h-2p) values. The geometries of the X 3B1, A 1A1, B 1B1, and C 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces used in the DIP-EOMCC(P ) calculations were defined as all 2h and 3h-1p determinants and subsets of 4h-2p determinants ex- tracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations were initiated by placing 500 walkers on the cor- responding ROHF reference determinant for the X 3B1 state and the corresponding RHF reference determinants for the remaining states and the na parameter of the initiator algorithm was set at 3. In all the post- Hartree–Fock calculations, the lowest core orbital of the carbon atom was kept frozen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146 xv Table 4.7 Convergence of the DEA-EOMCC(P ) and DIP-EOMCC(P ) energies [ab- breviated as DEA(P ) and DIP(P ), respectively] of the X 3A′ 2 and B 1A1 states of TMM, as described by the 6-31G(d) basis set of Ref. [178], and of the corresponding adiabatic singlet–triplet (S–T) gaps toward their parent DEA-EOMCC(4p-2h) and DIP-EOMCC(4h-2p) values. The ge- ometries of the X 3A′ 2 and B 1A1 states, optimized using the SF-DFT/6- 31G(d) calculations, were taken from Ref. [231]. The P spaces used in the DEA-EOMCC(P ) calculations were defined as all 2p and 3p-1h determinants and subsets of 4p-2h determinants extracted from the i- CISDTQ-MC propagations with δτ = 0.0001 a.u. The P spaces used in the DIP-EOMCC(P ) calculations were defined as all 2h and 3h-1p determinants and subsets of 4h-2p determinants extracted from the i- CISDTQ-MC propagations with δτ = 0.0001 a.u. In all the post-SCF calculations, the lowest core orbitals of the carbon atoms were kept frozen and the spherical components of the carbon d orbitals were employed throughout. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152 xvi LIST OF FIGURES Figure 2.1 A schematic representation of the CIQMC algorithm. |Φ⟩ represents the reference determinant, while {|ΦS1⟩, |ΦS2⟩, |ΦS3⟩, |ΦS4⟩, · · · } denote singly excited determinants, {|ΦD1⟩, |ΦD2⟩, |ΦD3⟩, |ΦD4⟩, · · · } doubly excited determinants, {|ΦT1⟩, |ΦT2⟩, |ΦT3⟩, |ΦT4⟩, · · · } triples, {|ΦQ1⟩, |ΦQ2⟩, |ΦQ3⟩, |ΦQ4⟩, · · · } quadruples, etc. The number of walkers on a particular determinant is indicated as a superscript within parenthesis. Green and red rectangles distinguish positive and negative walkers, re- spectively, with darker shades indicating a higher number of walkers on the determinants. Here, the simulation starts with one walker placed on the reference determinant |Φ⟩. Panel (a) illustrates spawning steps, panel (b) depicts the death of a walker on the determinant |ΦD4⟩ with a gray box, panel (c) shows more spawning events and panel (d) displays the annihilation step on |ΦQ3⟩ with an orange rectangular box. . . . . . . Figure 3.1 A schematic illustration depicting the construction of P -spaces in CC(P) and EOMCC(P) computations. Panel (a) showcases the stabilization of correlation energy (green line) and the corresponding increase in the total number of walkers is shown in panel (b) (red line). On the right four snapshots from a QMC calculation are presented, featuring the lists of determinants picked up by the QMC algorithm at various time steps (green for 2000, orange for 20000, violet for 50000, and magenta for 100000 QMC iterations). It is evident that QMC deems some determi- nants more important than others by placing more walkers on them. . . . Figure 3.2 Convergence of the EOMCC(P ) [panels (a) and (c)] and CC(P;Q) [pan- els (b) and (d)] energies toward EOMCCSDT for the three lowest-energy excited states of the 1Σ+ symmetry, two lowest states of the 1Π sym- metry, and two lowest 1∆ states of the CH+ ion, as described by the [5s3p1d/3s1p] basis set of Ref. [164], at the C–H internuclear distance R set at Re = 2.13713 bohr [panels (a) and (b)] and 2Re = 4.27426 bohr [panels (c) and (d)]. Adapted from Ref. [100]. . . . . . . . . . . . . . . . Figure 3.3 The distributions of the differences between the R(MC) µ,3 amplitudes and their EOMCCSDT counterparts resulting from the EOMCC(P ) com- putations at (a) 4000, (b) 10,000, and (c) 50,000 MC iterations using τ = 0.0001 a.u. for the 21Σ+ state of CH+ at R = 2Re with the analogous distribution characterizing the Rµ,3 amplitudes obtained with the EOM- CCSDt approach employing the 3σ, 1πx, 1πy, and 4σ active orbitals to define the corresponding triples space [panel (d)]. All vectors Rµ needed to construct panels (a)–(d) were normalized to unity. Adapted from Ref. [100]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 36 50 51 xvii Figure 3.4 Convergence of the EOMCC(P) and CC(P;Q) energies of the A 2∆ [panel (a)], B 2Σ− [panel (b)], and C 2Σ+ [panel (c)] states of CH toward EOMCCSDT for the three lowest-energy excited states of CH calculated as described by the aug-cc-pVDZ basis set. The geometries used are the equilibrium C–H distances reported in Refs. [62, 68, 161], which are 1.1031 ˚A for the A 2∆ state [175], 1.1640 ˚A for the B 2Σ− state [176], and 1.1143 ˚A for the C 2Σ+ state [177]. . . . . . . . . . . . . . . . . . . . Figure 3.5 Convergence of the EOMCC(P) and CC(P;Q) energies of the A 2∆u [panel (a)] and B 2Σ+ u [panel (b)] states of CNC toward EOMCCSDT for the two lowest-energy doublet excited states of CNC calculated as described by the DZP[4s2p1d] basis set. The geometries used are the equilibrium C–N distances reported in Refs. [68,162], which are 1.256 ˚A for the A 2∆u state and 1.259 ˚A for the B 2Σ+ u state. . . . . . . . . . . . Figure 3.6 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3B1 [panel (a)] and A 1A1 [panel (b)] states of methylene, as described by the aug- cc-pVTZ basis set, and of the corresponding adiabatic singlet–triplet gaps [panel (c)] toward their parent CCSDT values. The geometries of the X 3B1 and A 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i- FCIQMC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [102]. . . . Figure 3.7 Comparison of convergences of the CC(P), CC(P;Q), and the underly- ing i-FCIQMC calculations toward their respective limits for the X 3B1 and A 1A1 states of the CH2 molecule at their respective geometries op- timized in the FCI calculations using the TZ2P basis set are taken from Ref. [181]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.8 Total electronic energies of the X 1Σ+ g (open circles and solid line) and A 3Σ+ u (filled circles and dotted line) states of (HFH)− with increase in the H–F distance, from 1.5 ˚A to 4.0 ˚A, obtained from the FCI (red cir- cles), CCSD (blue circles), and CCSDT (green circles) methods. Recre- ated from the data reported in Refs. [66, 84, 88, 197]. . . . . . . . . . . . xviii 56 60 69 70 84 Figure 3.9 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1Σ+ g [pan- els (a) and (b)] and A 3Σ+ u [panels (c) and (d)] states of (HFH)−, as described by the 6-31G(d,p) basis set, and of the corresponding singlet– triplet gaps [panels (e) and (f)] toward their parent CCSDT values. The H–F distances RH-F used are 1.50 ˚A, 1.75 ˚A, 2.00 ˚A, 2.50 ˚A, and 4.00 ˚A. The P spaces consisted of all singles and doubles and subsets of triples identified during i-FCIQMC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.10 π molecular orbital network of the cyclobutadiene molecule, obtained at the HF/cc-pVDZ level, at the D4h-symmetric transition-state geometry of the X 1B1g state optimized in the MR-AQCC calculations in Ref. [202]. The orbital irreducible representations in the D4h symmetry are shown in black and the corresponding labels in the C2v symmetry are shown in the parenthesis in orange. This electronic configuration refers to the triplet state. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.11 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1B1g [panel (a)] and A 3A2g [panel (b)] states of cyclobutadiene, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet–triplet gaps [panel (c)] toward their parent CCSDT values. All calculations were performed at the D4h-symmetric transition-state geometry of the X 1B1g state optimized in the MR-AQCC calculations in Ref. [202]. The P spaces consisted of all singles and doubles and subsets of triples iden- tified during the i-FCIQMC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 92 93 Figure 3.12 π molecular orbital network of the cyclopentadienyl cation molecule, obtained at the HF/cc-pVDZ level, at the D5h-symmetric geometry of the X 3A′ 2 state optimized using the unrestricted CCSD/cc-pVDZ ap- proach in Ref. [203]. The orbital irreducible representations in the D5h symmetry are shown in black and the corresponding labels in the C2v symmetry are shown in the parenthesis in orange. This electronic con- figuration refers to the triplet state. . . . . . . . . . . . . . . . . . . . . . 100 xix Figure 3.13 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 [panel (a)] and A 1E′ 2 [panel (b)] states of cyclopentadienyl cation, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet– triplet gaps [panel (c)] toward their parent CCSDT values. All calcula- tions were performed at the D5h-symmetric geometry of the X 3A′ 2 state optimized using the unrestricted CCSD/cc-pVDZ approach in Ref. [203]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-CISDTQ- MC. Adapted from Ref. [102]. . . . . . . . . . . . . . . . . . . . . . . . . 101 Figure 3.14 π molecular orbital network of the trimethylenemethane molecule, ob- tained at the HF/cc-pVDZ level, at the D3h-symmetric geometry of the X 3A′ 2 state optimized in the SF-DFT/6-31G(d) calculations and taken from Ref. [231]. The orbital irreducible representations in the D3h symmetry are shown in black and the corresponding labels in the C2v symmetry are shown in the parenthesis in orange. This electronic configuration refers to the triplet state. . . . . . . . . . . . . . . . . . . . 109 Figure 3.15 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 [panel (a)] and B 1A1 [panel (b)] states of trimethylenemethane, as described by the cc-pVDZ basis set, and of the corresponding adiabatic singlet–triplet gaps [panel (c)] toward their parent CCSDT values. The geometries of the X 3A′ 2 and B 1A1 states, optimized in the SF-DFT/6-31G(d) calcu- lations, were taken from Ref. [231]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-CISDTQ-MC. Adapted from Ref. [102]. . . . . . 110 Figure 4.1 A schematic illustration depicting the construction of P -spaces in the EA/IP/DEA/DIP-EOMCC(P ) computations. Panel (a) showcases the stabilization of correlation energy (green line) and the corresponding increase in the total number of walkers is shown in panel (b) (red line). On the right four snapshots from a QMC calculation are presented, featuring the lists of determinants picked up by the QMC algorithm at various time steps (light blue for 2000, dark blue for 20000, violet for 40000, and magenta for 100000 QMC iterations). It is evident that QMC deems some determinants more important than others by placing more walkers on them. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 Figure 4.2 Convergence of (a) the EA-EOMCC(P ) energies of the X 2Π, A 2∆, B 2Σ−, and C 2Σ+ states of C2N, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energies toward their parent EA-EOMCC(3p-2h) values. . . . . . . . . . . . . . . . . . . . . . 122 xx Figure 4.3 Convergence of (a) the EA-EOMCC(P ) energies of the X 2Πg, A 2∆u, and B 2Σ+ u states of CNC, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energies toward their parent EA-EOMCC(3p-2h) values. . . . . . . . . . . . . . . . . . . . . . 127 Figure 4.4 Convergence of (a) the IP-EOMCC(P ) energies of the X 2Πg and B 2Σ+ u states of N3, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energy toward their parent IP- EOMCC(3h-2p) values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 Figure 4.5 Convergence of (a) the IP-EOMCC(P ) energies of the X 2Π, A 2∆, and B 2Π states of NCO, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energies toward their parent IP-EOMCC(3h-2p) values. . . . . . . . . . . . . . . . . . . . . . . . . . . 136 Figure 4.6 Convergence of (a) DEA-EOMCC(P ) energies of X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set, and (b) of the corresponding adiabatic singlet–triplet gaps towards their parent DEA-EOMCC(4p-2h) values. . . . . . . . . . . . . . . . . . . . . 143 Figure 4.7 Convergence of (a) DIP-EOMCC(P ) energies of X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set, and (b) of the corresponding adiabatic singlet–triplet gaps towards their parent DIP-EOMCC(4h-2p) values. . . . . . . . . . . . . . . . . . . . . . 147 Figure 4.8 Jahn-Teller distortion in the trimethylenemethane molecule. At the geometry of the D3h-symmetric triplet ground state (shown in blue), trimethylenemethane has a doubly degenerate singlet excited state. Due to Jahn–Teller distortion these states split into an open-shell singlet state A 1B1 (shown in green) and a multi-configurational singlet state B 1A1 (shown in red). Although the A 1B1 state becomes the first ex- cited state, it is not observed experimentally due to unfavorable Franck– Condon factors. Thus, we calculate the singlet–triplet gap between the ground triplet and the second excited singlet state B 1A1. . . . . . . . . . 153 Figure 4.9 Convergence of DEA-EOMCC(P ) energies of (a) X 3A′ 2 and (b) B 1A1 states of TMM, as described by the 6-31G(d) basis set, and (c) of the corresponding adiabatic singlet–triplet gap towards their parent DEA- EOMCC(4p-2h) values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 Figure 4.10 Convergence of DIP-EOMCC(P ) energies of (a) X 3A′ 2 and (b) B 1A1 states of TMM, as described by the 6-31G(d) basis set, and (c) of the corresponding adiabatic singlet–triplet gap towards their parent DIP- EOMCC(4h-2p) values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 xxi CHAPTER 1 INTRODUCTION Electronic structure theory is a major branch of quantum mechanics that aims to elucidate the behavior of electrons in atoms, molecules, and condensed matter systems. It provides a fundamental framework for understanding and predicting the properties of matter at the microscopic level. At its core, lies the time-independent electronic Schr¨odinger equation, which, within the Born–Oppenheimer approximation, has the form HΨµ(X; R) = Eµ(R)Ψµ(X; R), (1.1) where H is the electronic Hamiltonian, Ψµ, which depends on the coordinates of all electrons X and parametrically on the coordinates of the nuclei R that provide the external poten- tial for the motion of electrons, is the many-electron wave function characterizing the µth electronic state, with µ = 0 denoting the ground state and µ > 0 designating excited states, and Eµ that parametrically depends on the nuclear coordinates R is the corresponding total electronic energy. In the non-relativistic description, the electronic Hamiltonian operator for a system containing N electrons and M nuclei can be expressed in atomic units as H = N X (− i=1 1 2 ∇2 i − M X A=1 ZA riA ) + N X i 0 designates excited states). Typically, we truncate the excitation operator Rµ at the same level of truncation as the cluster operator T and, as a result, we obtain the hierarchy of EOMCC approxima- tions similar to the ground state problem. Thus, mA = 2 gives rise to the basic EOM- CCSD approximation [R(EOMCCSD) µ = rµ,01 + Rµ,1 + Rµ,2], mA = 3 produces EOMCCSDT [R(EOMCCSDT) µ = rµ,01 + Rµ,1 + Rµ,2 + Rµ,3], mA = 4 results in the EOMCCSDTQ approach [R(EOMCCSDTQ) µ = rµ,01 + Rµ,1 + Rµ,2 + Rµ,3 + Rµ,4], etc. To obtain the equations for the excitation amplitudes ri1...in µ,a1...an and energies E(A) µ , one inserts Eq. (2.7) into the Schr¨odinger equation and projects the resulting equation onto the set of excited Slater determinants |Φa1...an i1...in ⟩ = Ea1...an i1...in |Φ⟩ with mA ≤ N , to obtain an eigenvalue problem ⟨Φa1···an i1···in |( ¯H (A) openR(A) µ,open)C|Φ⟩ = ω(A) µ ri1···in µ,a1···an, (2.10) i1 < · · · < in, a1 < · · · < an, n = 1, . . . , mA, where ¯H (A) open = ¯H (A) − ¯H (A) closed = ¯H (A) − E(A) 0 1 and R(A) µ,open = R(A) parts of ¯H (A) and R(A) µ having external Fermion lines and ω(A) µ = E(A) µ − rµ,01 are the open µ − E(A) is the vertical 0 excitation energy corresponding to the excited state. After solving Eq. (2.10) (typically, using the iterative Davidson-type diagonalization procedure generalized to non-Hermitian eigenvalue problems) for amplitudes ri1...in µ,a1...an and excitation energies ω(A) µ , we determine the coefficient rµ,0 entering Rµ, which is nonzero for excited states of the same symmetry as the ground state, by using the expression rµ,0 = ⟨Φ|( ¯H (A) openR(A) µ,open)C|Φ⟩/ω(A) µ . (2.11) It is important to point out that the similarity-transformed Hamiltonian ¯H (A) is not Hermitian. This implies that for the CC and EOMCC theories, the “bra” and “ket” eigen- 17 vectors of ¯H (A) corresponding to a given eigenvalue are completely different. Therefore, we have to distinguish between the right or ket CC and EOMCC states defined by Eqs. 2.1 and 2.7, respectively, and their left or bra counterparts. The appropriate ansatz for the left CC (µ = 0) and EOMCC (µ > 0) states matching their |Ψ0⟩ and |Ψµ⟩ is as follows: ⟨ ˜Ψµ| = ⟨Φ|Lµe−T , (2.12) where Lµ is a hole-particle de-excitation operator that satisfies the biorthonormality condi- tion ⟨ ˜Ψµ|Ψν⟩ = ⟨Φ|LµRν|Φ⟩ = δµν. Here, δµν is the Kronecker delta and the de-excitation operator Lµ is defined as where Lµ = δµ01 + N X n=1 Lµ,n, Lµ,n = X µ,i1...in(Ea1...an la1...an i1...in )† i1<··· 0) states, one solves the CC/EOMCC equations in the subspace of the N - electron Hilbert space referred to as the P space, denoted as H(P ). This space is spanned by the excited determinants |ΦK⟩ = EK|Φ⟩ which, together with the reference determinant |Φ⟩, dominate the ground- and excited-state wave functions |Ψµ⟩ of interest (EK is the elementary particle–hole excitation operator generating |ΦK⟩ from |Φ⟩; for the sake of brevity of this description, we assume that ground and excited states have the same symmetry; excited states having different symmetries than the ground state are addressed later). This is done in a usual way adopted in all single-reference CC and EOMCC calculations, i.e., one starts with obtaining the cluster operator T (P ) = X tKEK, |ΦK ⟩∈H(P ) (2.17) with tK representing the corresponding cluster amplitudes by solving the system of equations ⟨ΦK| ¯H (P )|Φ⟩ = 0, |ΦK⟩ ∈ H(P ), (2.18) where ¯H (P ) = e−T (P )HeT (P ) = (HeT (P ))C is the relevant similarity-transformed Hamiltonian, and determining the corresponding ground-state energy E(P ) 0 = ⟨Φ| ¯H (P )|Φ⟩. (2.19) Then, the similarity-transformed Hamiltonian ¯H (P ) is diagonalized in the P space H(P ) to determine the excited-state EOMCC(P ) energies E(P ) µ and the corresponding EOM excita- tion and de-excitation operators, µ = rµ,01 + X R(P ) rµ,KEK |ΦK ⟩∈H(P ) (2.20) 19 and µ = δµ01 + X L(P ) lµ,K(EK)†, |ΦK ⟩∈H(P ) (2.21) respectively, where rµ,K and lµ,K designate the relevant amplitudes, which define the EOMCC(P ) ket states |Ψ(P ) µ ⟩ = R(P ) µ eT (P )|Φ⟩ and the CC(P )/EOMCC(P ) bra states ⟨ ˜Ψ(P ) µ | = ⟨Φ|L(P ) µ e−T (P ) (2.22) (2.23) satisfying ⟨ ˜Ψ(P ) µ |Ψ(P ) ν ⟩ = δµν. Once all of this is done, one proceeds to the second step, which is the calculation of the noniterative corrections δµ(P ;Q) to the CC(P) and EOMCC(P) energies E(P ) µ that account for the remaining many-electron correlation effects of interest captured with the help of the another subspace of the N -electron Hilbert space, referred to as the Q space and designated as H(Q) [H(Q) ⊆ (H(0) ⊕ H(P ))⊥, where H(0) is a one- dimensional subspace spanned by |Φ⟩]. The expression for these corrections is where when µ = 0, and δµ(P ;Q) = X ℓµ,K(P )Mµ,K(P ), |ΦK ⟩∈H(Q) M0,K(P ) = ⟨ΦK| ¯H (P )|Φ⟩, Mµ,K(P ) = ⟨ΦK| ¯H (P )R(P ) µ |Φ⟩, (2.24) (2.25) (2.26) when µ > 0, are the generalized moments of the CC(P) and EOMCC(P) equations that correspond to projections of the P space CC/EOMCC equations on to the complementary Q- space determinants |ΦK⟩ ∈ H(Q) and the coefficients ℓµ,K(P ) multiplying moments Mµ,K(P ) in Eq. (2.24) are calculated with the expression ℓµ,K(P ) = ⟨Φ|L(P ) µ ¯H (P )|ΦK⟩/Dµ,K(P ), (2.27) 20 in which the Dµ,K(P ) denominators are given by Dµ,K(P ) = E(P ) µ − ⟨ΦK| ¯H (P )|ΦK⟩ (2.28) (one could replace the Epstein–Nesbet Dµ,K(P ) denominators entering Eq. (2.27) ℓµ,K(P ) by their Møller–Plesset analogs, but, as shown in the past, for example in Refs. [63, 64, 66, 68, 84, 88, 101], the Epstein–Nesbet form is generally more effective). The final CC(P;Q) electronic energies for the ground (µ = 0) and excited (µ > 0) states are determined as E(P +Q) µ = E(P ) µ + δµ(P ; Q). (2.29) Now, a question arises as to how to define the P and Q spaces entering the CC(P;Q) framework to obtain accurate ground- and excited-state energetics that match the quality of the target high-level CC/EOMCC calculations without incurring the substantial computa- tional costs of methods, such as CCSDT/EOMCCSDT, CCSDTQ/EOMCCSDTQ, etc. The simplest possibility is to rely on the conventional choices, where the P and Q spaces are defined based on the many-body ranks of the excited determinants |ΦK⟩ included in them. For example, if we want to correct the CCSD and EOMCCSD energies for triples using the CC(P;Q) formulas, the P space H(P ) is spanned by the singly and doubly excited determi- nants |Φa i ⟩ and |Φab ij ⟩, respectively, and the Q space H(Q) by the triply excited determinants |Φabc ijk⟩. As already alluded to above, the resulting CR-CC(2,3) and CR-EOMCC(2,3) correc- tions to the CCSD and EOMCCSD energies have been very successful, but, by decoupling the low-order Tn and Rµ,n components with n ≤ 2 from their higher-order, such as T3 and Rµ,3 counterparts, they may not be as accurate as desired, for example in situations where T3 and Rµ,3 components become large, resulting in substantial inaccuracies and difficulties in balancing ground- and excited-state energies. One can address this problem by using active orbitals to enrich the relevant P spaces with the dominant higher–than–doubly ex- cited determinants, as in the CC(t;3), CC(t,q;3), CC(t,q;3,4), etc., methodologies mentioned in the Introduction, but the resulting approaches are no longer computational black boxes. The semi-stochastic CC(P;Q) approach to ground- and excited-state calculations [99–102], 21 which is described in detail in Chapter 3, exploits the CIQMC or CCMC propagations to identify the leading higher–than–doubly excited determinants pertinent to the CC/EOMCC calculations of interest in a black-box manner, while using corrections δµ(P ;Q) to capture the remaining correlations that the CC(P)/EOMCC(P) energies at a given QMC propaga- tion time do not describe, eliminates the above concerns. Since one of the objectives of this dissertation project is to extend the semi-stochastic CC(P;Q) ideas to the EA/IP/DEA/DIP EOMCC methods, these particle nonconserving approaches are discussed next. 2.3 Particle Nonconserving Equation-of-Motion Coupled-Cluster Theories: The Single and Double Electron Attachment and Ionization Potential Methods Radicals, biradicals, and other open-shell species and their low-lying excited states constitute a major challenge to CC and other ab initio methods due to, in most cases, their inherent multiconfigurational character that is difficult to capture using the low-rank approximations, such as CCSD, EOMCCSD, and their linear response analogs. One can always think of using genuine MRCC and other MR methodologies, but as pointed out in the Introduction, their routine use is not straightforward and they have their own, often severe, challenges. The failures of CCSD/EOMCCSD in the presence of electronic quasi-degeneracies in open-shell systems imply the need to incorporate higher–than–double excitations in the CC/EOMCC wave functions, but this results in several computational and formal difficulties. As already discussed, the full inclusion of higher–than–two-body components in the cluster and EOM excitation operators improves the results, but the associated computer costs become very (often prohibitively) large as we increase the system size. One can think of resorting to the active-space ideas, giving rise to methods such as CCSDt/EOMCCSDt, CCSDtq/EOMCCS- Dtq etc., but with the usual spin-integrated spin-orbital formulation, neither these methods nor any other particle conserving CC/EOMCC methodology properly account for the spin symmetry of the calculated states of open-shell systems. There is, however, an interesting, underappreciated, alternative discussed in this section and pursued in my doctoral work. One can deal with the above problems by utilizing the flexibility of the EOMCC formal- 22 ism and resorting to the particle-nonconserving EA/IP and DEA/DIP EOMCC frameworks. These approaches provide an elegant way of obtaining orthogonally spin-adapted results for the ground and excited states of radicals and biradicals by diagonalizing the similarity- transformed Hamiltonian of a closed-shell CC theory in the subspaces of the Fock space obtained by adding one or two electrons to (EA, DEA) or removing one or two electrons from (IP, DIP) a related closed-shell core. In these kinds of methods, one starts with a CC ground-state wave function |Ψ0⟩ = eT |Φ⟩ of an N -electron closed-shell system for which the corresponding correlation energy is defined as ∆E(N ) 0 = E(N ) 0 − ⟨Φ|H|Φ⟩ = (HN eT )C,closed, (2.30) where ⟨Φ|H|Φ⟩ is the reference energy, HN = H −⟨Φ|H|Φ⟩ is the Hamiltonian operator in the normal ordered form relative to the N -electron Fermi vacuum |Φ⟩, the superscript (N ) indi- cates the number of electrons in the closed-shell system, and T is the corresponding cluster operator. In the EA and IP-EOMCC theories, the similarity-transformed Hamiltonian, ¯HN,open = (HN eT )C,open = e−T HN eT − ∆E(N ) 0 , (2.31) corresponding to the N -electron closed-shell CC theory, is diagonalized in the (N +1)-electron (EA) and (N − 1)-electron (IP) subspaces, H (N +1) and H (N −1), respectively, of the Fock space. In doing so, the ground and excited-state wave functions of the target (N + 1)- and (N − 1)-electron systems of interest are defined as |Ψ(N ±1) µ ⟩ = R(N ±1) µ |Ψ(N ) 0 ⟩ = R(N ±1) µ eT |Φ⟩, (2.32) where the cluster operator T is obtained by solving the usual ground-state CC equations for the N -electron closed-shell system. R(N +1) µ is an electron-attaching operator defined as R(N +1) µ = MRX n=0 Rµ,(n+1)p-nh, (2.33) 23 where Rµ,(n+1)p-nh = X µ,aa1...anEaa1...an ri1...in i1...in (n ≥ 1). (2.34) i1>···>in ai1>···in a1<···j,b In these two cases, we end up with diagonalizing the similarity-transformed Hamiltonian of CCSD, (HN eT1+T2)C,open, in the space spanned by |Φa⟩ = aa|Φ⟩ and |Φab j ⟩ = aaabaj|Φ⟩ determinants for EA-EOMCC(2p-1h) and the space spanned by the |Φi⟩ = ai|Φ⟩ and |Φb ij⟩ = abajai|Φ⟩ determinants when the IP-EOMCC(2h-1p) calculations are performed. Although the EA-EOMCC(2p-1h) and IP-EOMCC(2h-1p) methods can be useful in obtain- ing the lowest electron affinities and ionization potentials, they struggle to describe electron attachment and ionization energies corresponding to excited states of anions and cations, what is particularly relevant for this dissertation, the electronic excitations in radicals, es- pecially when the target states have inherently multiconfigurational character. The good news is that one can usually address this issue by incorporating the higher-rank 3p-2h/3h-2p components in the R(N ±1) µ operator expressions i.e., by using the and R(N +1) µ = Rµ,1p + Rµ,2p-1h + Rµ,3p-2h, R(N −1) µ = Rµ,1h + Rµ,2h-1p + Rµ,3h-2p 25 (2.42) (2.43) operators in the EA/IP-EOMCC computations. Now, there are two options: either one uses Eqs. (2.42) and (2.43) for the electron attaching and ionizing operators and the similarity- transformed Hamiltonian of CCSD, ¯H (CCSD) N,open = (HN eT1+T2)C,open, resulting in the EA-EOMCC(3p-2h) and IP-EOMCC(3h-2p) approaches, respectively, or, one keeps the same definitions of the R(N ±1) µ operators as in Eqs. (2.42) and (2.43), but replaces ¯H (CCSD) N,open by the similarity- transformed Hamiltonian of the CCSDT theory, ¯H (CCSDT) N,open = (HN eT1+T2+T3)C,open, resulting in the EA-EOMCCSDT and IP-EOMCCSDT methods. Both classes of methods are similarly effective at describing the electronic spectra of the (N ± 1) electron radical systems, so we will focus on EA-EOMCC(3p-2h) and IP-EOMCC(3h-2p), which do not require solving the CCSDT equations, but they may still be too expensive in practice due to the relatively high computational costs of diagonalizing ¯H (CCSD) N,open in spaces containing |Φabc jk ⟩ = aaabacakaj|Φ⟩ and |Φbc ijk⟩ = abacakajai|Φ⟩ determinants that scale as N 7 with the system size. One way to deal with the problem of high costs of EA-EOMCC(3p-2h) and IP-EOMCC(3h- 2p) computations is to resort to the active-space ideas, similar to the ground-state CC and excited-state EOMCC cases, by including only the leading components of the Rµ,3p-2h/Rµ,3h-2p operators defined by active orbitals. These active-space components, defined in Refs. [78, 79, 140–142], will be denoted as rµ,3p-2h/rµ,3h-2p. Therefore, one incorporates all singly and doubly excited amplitudes of the cluster operator T1 and T2, all attaching Rµ,1p and Rµ,2p-1h amplitudes, and all electron ionizing Rµ,1h and Rµ,2h-1p amplitudes, but only small subsets of 3p-2h/3h-2p amplitudes of R(N ±1) µ defined by active orbitals, giving rise to the active-space EA/IP-EOMCC methods known as EA-EOMCCSDt and IP-EOMCCSDt or EA-EOMCC(3p-2h){Nu} and IP-EOMCC(3h-2p){No}, where No and Nu are the numbers of active occupied and active unoccupied orbitals in |Φ⟩, respectively [78, 79, 140–142]. In the EA-EOMCCSDt = EA-EOMCC(3p-2h){Nu} approach, the cluster operator T of the N -electron closed-shell core obtained with CCSD is defined as T = T1 + T2 (2.44) 26 and the electron attaching operator given by Eq. (2.42) is replaced by R(N +1) µ {Nu} = Rµ,1p + Rµ,2p-1h + rµ,3p-2h, where rµ,3p-2h = X rjk µ,AbcaAabacakaj. j>k Aj>k b···>in aj>i1>···>in a1<···l AJ>k>l c 0, the walker is removed from the simulation, and if pd < 0, the walker is cloned with the probability |pd|. The cloning (or birth) step is usually rare. 3. The annihilation step: Subsequent to the previous two steps, the annihilation step is performed, wherein all pairs of walkers with opposite sign are removed from each determinant. A schematic representation of how the CIQMC algorithm works is depicted in Fig. 2.1. After reaching a sufficiently large number of walkers, the correlation energy and the walker populations stabilize using a suitable energy shift S. Since the propagation is performed in the space of Slater determinants, the resulting wave function has the proper Fermionic symmetry. As already alluded to above, if the spawning of walkers is restricted to a cer- tain level of truncation, we can obtain truncated CIQMC approaches, such as CISDT-MC and CISDTQ-MC to mention two representative examples, where spawning of walkers on determinants higher than triples, in the CISDT-MC case, or higher than quadruples in the CISDTQ-MC case, is forbidden. By replacing the time-dependent CI expansion, Eq. (2.61), by the analogous CC ansatz, one obtains the CCMC approach introduced in Ref. [95] (see, 32 Figure 2.1 A schematic representation of the CIQMC algorithm. |Φ⟩ represents the reference determinant, while {|ΦS1⟩, |ΦS2⟩, |ΦS3⟩, |ΦS4⟩, · · · } denote singly excited determinants, {|ΦD1⟩, |ΦD2⟩, |ΦD3⟩, |ΦD4⟩, · · · } doubly excited determinants, {|ΦT1⟩, |ΦT2⟩, |ΦT3⟩, |ΦT4⟩, · · · } triples, {|ΦQ1⟩, |ΦQ2⟩, |ΦQ3⟩, |ΦQ4⟩, · · · } quadruples, etc. The number of walkers on a particular determinant is indicated as a superscript within parenthesis. Green and red rectangles distinguish positive and negative walkers, respectively, with darker shades indicating a higher number of walkers on the determinants. Here, the simulation starts with one walker placed on the reference determinant |Φ⟩. Panel (a) illustrates spawning steps, panel (b) depicts the death of a walker on the determinant |ΦD4⟩ with a gray box, panel (c) shows more spawning events and panel (d) displays the annihilation step on |ΦQ3⟩ with an orange rectangular box. also, Refs. [96–98]). Several ideas have been explored to improve the convergence of the QMC algorithm. One of them is the initiator CIQMC (i-CIQMC) approach of Ref. [92], which we use in our work, and its i-CCMC counterpart, where spawning is allowed as long as there is at least a minimum walker population, na, on the parent determinant. In the next chapters, we discuss how we can take advantage of the CIQMC algorithm to accelerate our CC(P;Q) and EA/IP/DEA/DIP-EOMCC computations, while making them fully au- tomated and free from the user- and system-dependent active orbitals when identifying the appropriate P spaces [CC(P;Q)] or the leading 3p-2h/3h-2p/4p-2h/4h-2p amplitudes. 33 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|i(1)|S1i(+11)|D1i(+11)|T1i(+11)|Q1i(+11)···|S2i(1)+1|D2i(1)+1|T2i(+11)|Q2i(+11)···|S3i(+11)|D3i(+11)|T3i(+11)|Q3i(+11)···|S4i(+11)|D4i(1)+1|T4i(+11)|Q4i(+11)···...............SpawningBirth & DeathAnnihilationSpawningPositive walkerNegative 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|i(4)|S1i(+11)|D1i(2)+1|T1i(+11)|Q1i(1)+1···|S2i(1)+1|D2i(2)+1|T2i(1)+1|Q2i(+11)···|S3i(+11)|D3i(1)+1|T3i(1)+1|Q3i(+11)···|S4i(2)+1|D4i(+11)|T4i(1)+1|Q4i(+11)···...............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|i(1)|S1i(+11)|D1i(1)+1|T1i(+11)|Q1i(1)+1···|S2i(1)+1|D2i(2)+1|T2i(+11)|Q2i(+11)···|S3i(+11)|D3i(+11)|T3i(1)+1|Q3i(+11)···|S4i(1)+1|D4i(0)+1|T4i(+11)|Q4i(+11)···...............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|i(3)|S1i(+11)|D1i(2)+1|T1i(+11)|Q1i(1)+1···|S2i(1)+1|D2i(2)+1|T2i(1)+1|Q2i(+11)···|S3i(+11)|D3i(1)+1|T3i(1)+1|Q3i(+11)···|S4i(2)+1|D4i(+11)|T4i(1)+1|Q4i(+11)···...............(a)(b)(d)(c) CHAPTER 3 THE SEMI-STOCHASTIC CC(P;Q) METHODOLOGY FOR GROUND AND EXCITED STATES In the previous chapters, we have discussed how the deterministic CC(P;Q) theory that combined the CR-CC/EOMCC and active space ideas emerged as a very robust tool for studying challenging chemical systems in a computationally affordable way. The only draw- back, however, is the need to choose a suitable set of active orbitals that can generate an ensemble of determinants, which define the proper active space, and this selection is system and user dependent. On the other hand, the QMC methods based on propagating the CI ansatz according to the imaginary time Schr¨odinger equation in the many electron Hilbert space spanned by Slater determinants are very efficient in identifying the most important de- terminants, based on probabilistic arguments, in an automated manner, but they spend a lot of time balancing the corresponding coefficients. In this chapter we discuss a novel hybrid approach, namely the semi-stochastic CC(P;Q) formalism, that merges the deterministic CC(P;Q) framework with the stochastic wave function sampling of CIQMC methodologies [99–102]. In particular, we focus on my work on extending the semi-stochastic CC(P;Q) methodology to ground and excited states of open-shell systems and non-singlet excited states. In doing so, we follow the results presented in Refs. [100] and [102], with a focus on my contributions to these papers. 3.1 Theory The CIQMC algorithm needs a very long propagation time to converge a stable wave func- tion, but the leading determinants are identified much sooner, and on the other hand the deterministic CC(P;Q) framework allows for very accurate energetics even in challenging MR situations as long as the lower clusters, such as T1 and T2, are allowed to relax in the presence of the higher order T3, T4, etc., clusters. This observation was utilized in Ref. [99], where it was demonstrated that one could use the information about the leading determi- nants captured during the early stages of i-CIQMC to create lists of determinants defining 34 P spaces for CC(P ) calculations and then use the deterministic, noniterative, CC(P;Q) corrections to capture the correlations effects missing in the P -space CC(P ) calculations. Later in Ref. [137], the semi-stochastic CC(P ) approach was extended to the excited state EOMCC(P ) method that provided a fast convergence to target energetics in a very straight- forward manner, without resorting to more-complex excited state i-CIQMC frameworks of Refs. [159, 160]. In this work we discuss the extension of the semi-stochastic ground-state CC(P;Q) [99] and EOMCC(P ) [137] approaches to open-shell systems and non-singlet ex- cited states and the CC(P;Q) corrections to EOMCCC(P ), as described in Refs. [100] and [102]. The key steps of the semi-stochastic CC(P;Q) [99–102, 137] algorithm for ground and excited states that resulted from the merger of deterministic CC(P;Q) algorithm and the stochastic CIQMC wave function sampling, are described below. 1. Initiate a CIQMC run for the ground state and, if the system of interest has spin, spatial, or other symmetries, the analogous QMC propagation for the lowest state of each irreducible representation (irrep) to be considered in the CC(P;Q) calculations by placing a certain number of walkers on the appropriate reference function(s) |Φ⟩ (e.g., the restricted Hartree–Fock (RHF) or restricted open- shell Hartree–Fock (ROHF) determinants). 2. At some propagation time τ > 0, i.e., after a certain number of CIQMC time steps, called MC iterations, extract a list or, if states belonging to multiple irreps are targeted, lists of determinants relevant to the desired CC(P;Q) computations from the QMC propagation(s) initiated in step 1 to determine the P space or spaces needed to set up the ground-state CC(P ) and excited state EOMCC(P ) calculations (for example, cf. Fig. 3.1). If the goal is to converge the CCSDT/EOMCCSDT-level energetics, the P space for the CC(P ) calculations and the EOMCC(P ) calculations for excited states belonging to the same irrep as the ground state is defined as all singly and 35 Figure 3.1 A schematic illustration depicting the construction of P -spaces in CC(P) and EOMCC(P) computations. Panel (a) showcases the stabilization of correlation energy (green line) and the corresponding increase in the total number of walkers is shown in panel (b) (red line). On the right four snapshots from a QMC calculation are presented, featuring the lists of determinants picked up by the QMC algorithm at various time steps (green for 2000, orange for 20000, violet for 50000, and magenta for 100000 QMC iterations). It is evident that QMC deems some determinants more important than others by placing more walkers on them. doubly excited determinants and a subset of triply excited determinants, where each triply excited determinant in the subset is populated by a minimum of nP positive or negative walkers (in this work, nP = 1). For the excited states belonging to other irreps, the P space defining the CC(P ) problem is the same as that used in the case of the ground state, but the lists of triply excited determinants defining the EOMCC(P ) diagonalizations are provided by the CIQMC propagations for the lowest-energy states of these irreps. One proceeds in a similar way when the goal is to converge other types of high-level CC/EOMCC energetics. For example, if we want to obtain the results of the CCSDTQ/EOMCCSDTQ quality, we also have to extract the lists of quadruples, in addition to the triples, from the CIQMC runs to define the corresponding P spaces. 36 (a)(b)# walkersdeterminants 3. Solve the CC(P ) and EOMCC(P ) equations in the P space or spaces obtained in the previous step. If we are targeting the CCSDT/EOMCCSDT-level energetics and the excited states of interest belong to the same irrep as the ground state, we define T (P ) = T1 + T2 + T (MC) 3 , R(P ) µ = rµ,01 + Rµ,1 + Rµ,2 + R(MC) µ,3 , and L(P ) µ = δµ,01 + Lµ,1 + Lµ,2 + L(MC) µ,3 , where the list of triples T (MC) 3 , R(MC) µ,3 , and L(MC) µ,3 is extracted from the ground-state CIQMC propagation at time τ . For the excited states belonging to other irreps, we construct the similarity-transformed Hamiltonian ¯H (P ), to be diagonalized in the EOMCC steps, in the same way as in the ground-state computations, but then use the CIQMC propagations for the lowest states of these irreps to define the lists of triples in R(MC) µ,3 and L(MC) µ,3 . We follow a similar procedure when targeting the CCSDTQ/EOMCCSDTQ-level energetics in which case T (P ) = T1+T2+T (MC) µ = δµ,01+Lµ,1+Lµ,2+L(MC) µ = rµ,01+Rµ,1+Rµ,2+R(MC) µ,3 +R(MC) , and L(P ) R(P ) µ,4 3 µ,3 +L(MC) µ,4 . +T (MC) 4 , 4. Correct the CC(P ) and EOMCC(P ) energies for the missing correlations of interest that were not captured by the CIQMC propagations at the time τ the lists of the P -space excitations were created (the remaining triples if the goal is to recover the CCSDT/EOMCCSDT energetics, the remaining triples and quadruples if one targets CCSDTQ/EOMCCSDTQ, etc.) using the CC(P;Q) corrections δµ(P ;Q) defined by Eq. (2.24). 5. Check the convergence of the resulting E(P +Q) µ energies calculated using Eq. (2.29) by repeating steps 2–4 at some later CIQMC propagation time τ ′ < τ . If the E(P +Q) µ energies do not change within a given convergence threshold, we can stop the calcula- tions. One can also stop them if τ in steps 2–4 is chosen such that the stochastically determined P space(s) contain sufficiently large fraction(s) of higher–than–doubly ex- cited determinants relevant to the target CC/EOMCC level. Our unpublished tests using the CC(P;Q)-based CC(t;3) corrections to the EOMCCSDt energies, the ground- state semi-stochastic CC(P;Q) calculations reported in Ref. [99], and the excited-state 37 CC(P;Q) calculations using i-FCIQMC to generate the underlying P spaces performed in this work indicate that one should be able to reach millihartree or sub-millihartree accuracies relative to the parent CC/EOMCC computations, when the stochastically determined P spaces contain as little as ∼5–10% and no more than ∼30–40% of higher– than–double excitations of interest, although this may need further study. Similarly to the semi-stochastic form of the ground-state CC(P;Q) methodology intro- duced in Ref. [99], the above algorithm offers significant savings in the computational effort compared to the fully deterministic, high-level, EOMCC approaches it targets. These savings originate from three factors. First, the computational times associated with the early stages of the i-CIQMC walker propagations are very short compared to the corresponding converged runs. Second, the CC(P ) calculations and the subsequent EOMCC(P ) diagonalizations of- fer significant speedups compared to their CC/EOMCC parents, when the corresponding excitation manifolds contain small fractions of higher–than–doubly excited determinants. For example, as pointed out in Refs. [99, 137], when the most expensive ⟨Φabc ⟨Φabc ijk|[ ¯H (2), T3]|Φ⟩, where ¯H (2) = exp(−T1−T2)Hexp(T1+T2)) and ⟨Φabc ijk|[H, T3]|Φ⟩ (or ijk|[ ¯H (P ), Rµ,3]|Φ⟩ terms in the CCSDT and EOMCCSDT equations are isolated and reprogrammed using techniques similar to implementing selected CI approaches, combined with sparse matrix multiplica- tion and index rearrangement routines (rather than conventional many-body diagrams that assume continuous excitation manifolds labelled by occupied and unoccupied orbitals from the respective ranges of indices; generally, the stochastically determined lists of excitations do not form continuous manifolds that could be a priori identified), one can speed up their determination by a factor of up to (D/d)2, where D is the number of all triples and d is the number of triples included in the stochastically determined P space. Other terms, such as ⟨Φabc ijk|[H, T2]|Φ⟩ and ⟨Φabc ijk|[ ¯H (P ), Rµ,2]|Φ⟩ or ⟨Φab ij |[H, T3]|Φ⟩ and ⟨Φab ij |[ ¯H (P ), Rµ,3]|Φ⟩, when treated in a similar manner, may offer additional speedups, on the order of (D/d), too. Our current CC(P ) and EOMCC(P ) routines are not as efficient yet, but the speedups that scale linearly with (D/d) in the most expensive ⟨Φabc ijk|[H, T3]|Φ⟩ and ⟨Φabc ijk|[ ¯H (P ), Rµ,3]|Φ⟩ contri- 38 butions are attainable. The third factor contributing to major savings in the computational effort offered by the semi-stochastic CC(P;Q) approach is the observation that the deter- mination of the noniterative correction δµ(P ;Q) for a given electronic state µ is much less expensive than the time required to complete a single iteration of the target CC/EOMCC calculation (in the case of the calculations aimed at the CCSDT/EOMCCSDT energetics, the computational time associated with each δµ(P ;Q) scales no worse than ∼2n3 on4 u, as opposed to the n3 on5 u scaling of every CCSDT and EOMCCSDT iteration). Before going to the next section, we must discuss an interesting aspect of the semi- stochastic CC(P )/EOMCC(P ) and CC(P;Q) methodologies. The CC(P ) and EOMCC(P ) energies at τ = 0 are identical to the energies obtained in the CCSD and EOMCCSD cal- culations and that the corresponding τ = 0 CC(P;Q) corrections are equivalent to those of CR-CC(2,3) (the ground state) and CR-EOMCC(2,3) (excited states). It should also be noted that the CC(P ) and EOMCC(P ) energies at τ = ∞ are identical to the energies obtained in the full CCSDT and EOMCCSDT calculations. The semi-stochastic CC(P;Q) calculations recover the CCSDT and EOMCCSDT energetics in this limit too, although the τ = ∞ values of the δµ(P ;Q) corrections are zero in this case, since the τ = ∞ P spaces contain all the triples, i.e., the corresponding Q-space triples lists are empty. These relation- ships between the semi-stochastic CC(P ), EOMCC(P ), and CC(P;Q) approaches and the fully deterministic CCSD/EOMCCSD, CR-CC(2,3)/CREOMCC(2,3), and CCSDT/EOM- CCSDT methodologies were helpful in examining the correctness of our codes. They also point to the ability of the CC(P ), EOMCC(P ), and CC(P;Q) calculations driven by the information extracted from CIQMC to offer a systematically improvable description as τ approaches ∞. 3.2 Electronic Excitations in CH+, CH, and CNC In order to explore the performance of the semi-stochastic CC(P;Q) approach to excited states proposed in this work and examine, in particular, the ability of the noniterative δµ(P ;Q) corrections to accelerate the convergence of the CIQMC-driven EOMCC(P ) calcu- 39 lations toward the desired EOMCC energetics, represented in this study by EOMCCSDT, we carried out benchmark calculations for the frequently studied vertical excitations in CH+ ion at the equilibrium [Table 3.1 and Fig. 3.2(a) and (b)] and stretched [Table 3.2 and Fig. 3.2(c) and (d)] geometries, which were previously used to test the EOMCC(P) framework [137] and which was a useful system to check the correctness of our codes, and the adiabatic excitations in the challenging open-shell CH (Table 3.3) and CNC (Table 3.4) systems, which have low lying excited states dominated by two-electron transitions that require at least the EOMCCSDT theory level to obtain a reliable description [62, 68, 78, 79, 141, 161–163]. The CH+ ion was described by the [5s3p1d/3s1p] basis set of Ref. [164] and we used the aug- cc-pVDZ [165, 166] and DZP[4s2p1d] [167, 168] bases for the CH and CNC species, respec- tively. Following Refs. [99, 137] (cf., also, Ref. [169]), we used the HANDE software package [170, 171] to execute the stochastic i-FCIQMC runs, needed to generate the lists of triply excited determinants included in the CC(P) and EOMCC(P) calculations. Our standalone CC/EOMCC codes, interfaced with the RHF, ROHF, and integral routines in the GAMESS program suite [172–174], were used to carry out the required CC(P ), EOMCC(P ), CC(P;Q), and fully deterministic (CCSD/EOMCCSD and CCSDT/EOMCCSDT) computations (the Q spaces used to construct the CC(P;Q) corrections to the CC(P ) and EOMCC(P ) energies consisted of the triples not captured by the i-FCIQMC runs at the corresponding propaga- tion times τ ). Each i-FCIQMC run was initiated by placing 1500 walkers on the relevant reference function (see Tables 3.1–3.4 for the details) and we set the initiator parameter na at 3. All of the i-FCIQMC propagations used the time step τ of 0.0001 a.u. In the post-ROHF computations for the CH and CNC species, the core electrons corresponding to the 1s shells of the carbon and nitrogen atoms were kept frozen. In the case of CH+, we correlated all electrons. 3.2.1 CH+ We begin our discussion of the numerical results with the CH+ ion, where we investigated the three lowest excited states of the 1Σ+ symmetry (labelled as 2 1Σ+, 3 1Σ+, and 4 1Σ+; 40 the ground state is designated as 1 1Σ+), two lowest states of the 1Π symmetry (1 1Π and 2 1Π), and two lowest 1∆ states (1 1∆ and 2 1∆). While these results were obtained mainly by Dr. Stephen Yuwono, with minor contributions from me, I will begin with them, since they are good for setting the stage for the subsequent discussion of the electronic excitation spectra of open-shell CH and CNC species and singlet–triplet gaps in biradicals, where I was a lead contributor. Two C–H internuclear separations were considered, the equilibrium distance R = Re = 2.13713 bohr [Table 3.1 and Fig. 3.2(a) and (b)] and the stretched R = 2Re geometry [Table 3.2 and Fig. 3.2(c) and (d)]. Following the semi-stochastic CC(P;Q) algorithm, as described above, and our interest in converging the CCSDT/EOMCCSDT energetics, the cluster and right and left EOM operators used in the calculations for the 1Σ+ states were approximated by T (P ) = T1 + T2 + T (MC) 3 , R(P ) µ = rµ,01 + Rµ,1 + Rµ,2 + R(MC) µ,3 , and L(P ) µ = δµ,01 + Lµ,1 + Lµ,2 + L(MC) , respectively, where the list of triples defining the three-body components T (MC) , R(MC) µ,3 state i-FCIQMC propagation at the same value of τ . The T (MC) 3 3 at a given time τ was obtained from the ground- component of T (P ) used µ,3 , L(MC) µ,3 in the CC(P;Q) computations of the 1Π and 1∆ states, needed to determine the similarity- transformed Hamiltonian ¯H (P ) to be diagonalized in the subsequent EOMCC steps, was defined in the same way as in the case of the 1Σ+ states, but the lists of triples entering the R(MC) µ,3 component of R(P ) µ and the L(MC) µ,3 component of L(P ) µ were obtained differently. They were extracted from the i-FCIQMC runs for the lowest states within the irreps of C2v relevant to the symmetries of interest, meaning the 1B1 (C2v) component of 1 1Π for the 1Π states and the 1A2 (C2v) component of 1 1∆ for the 1∆ states (C2v is the largest Abelian subgroup of the true point group of CH+, C∞v; our codes cannot handle non-Abelian symmetries). As implied by Eq. 2.24, the δµ(P ;Q) corrections to the CC(P ) and EOMCC(P ) energies at a given time τ were computed using the Mµ,K(P ) and ℓµ,K(P ) amplitudes corresponding to the triply excited determinants |ΦK⟩ not captured by i-FCIQMC at the same τ . As pointed out in Refs. [51, 52, 137], the 2 1Σ+, 2 1Π, 1 1∆, and 2 2∆ states of CH+ at R = Re and all of the excited states of the stretched CH+/R = 2Re system, which we 41 calculated in this work, are characterized by substantial MR correlations that originate from large two-electron excitation contributions (the 2 2∆ state at R = 2Re also has significant triple excitations [51, 52, 137]). It is therefore not surprising that the basic EOMCCSD level, equivalent to the EOMCC(P) calculations at τ = 0, performs poorly for all of these states, producing very large errors relative to EOMCCSDT that are about 12, 20, and 34–35 millihartree for the 2 1Σ+, 2 1Π, and both 1∆ states, respectively, at R = Re and ∼14– 144 millihartree when the excited state at R = 2Re are considered (see Tables 3.1 and 3.2). The EOMCCSD energies for the 3 1Σ+, 4 1Σ+, and 1Π, states at the equilibrium geometry, which are dominated by one-electron transitions, are more accurate, but errors on the order of 3–6 millihartree still remain. As shown in Tables 3.1 and 3.2, the CR- EOMCC(2,3) triples correction to EOMCCSD, equivalent to the CC(P;Q) calculations at τ = 0, offers substantial improvements, as exemplified by the small errors, on the order of 1–3 millihartree, for the majority of excited states of CH+ considered in this subsection, but there are cases, especially the 4 1Σ+ and 2 1∆ states at R = Re, where the differences between the CR-EOMCC(2,3) and parent EOMCCSDT energies, which are about 12 millihartree in the former case and more than 63 millihartree in the case of the latter state, remain very large. This is related to the substantial coupling of the one- and two-body components of the cluster and EOM excitation and deexcitation operators with their three-body counterparts, which the CR-EOMCC(2,3) corrections to EOMCCSD neglect. Our group’s older active- space EOMCCSDt calculations for CH+ reported in Refs. [51, 52] and the more recent semi- stochastic EOMCC(P) calculations for the same system described in Ref. [137] are indicating that the incorporations of the leading triples in the relevant P spaces, which allows the one- and two-body components of T , Rµ, and Lµ to relax in the presence of their three-body counterparts, is the key to improve the results of the CR-EOMCC(2,3) calculations. This is exactly what we observe in Tables 3.1 and 3.2 and Fig. 3.2. In agreement with our previous work [137], by enriching the P spaces used in the CC(P) and EOMCC(P) computations with the subsets of triples captured during i-FCIQMC propagations, the results 42 greatly improve, allowing us to reach the millihartree or sub-millihartree accuracy levels for all the calculated excited states of CH+ at both nuclear geometries considered in this work when the stochastically determined P spaces contain about 20–30% of all triples. The CC(P;Q) corrections to the EOMCC(P) energies based on Eq. (2.24) accelerate the convergence toward EOMCCSDT even further. As shown in Tables 3.1 and 3.2 and Fig. 3.2, these corrections are so effective that we reach the millihartree or sub-millihartree accuracy levels relative to the parent EOMCCSDT energetics almost instantaneously, i.e., out of early stages of the i-FCIQMC propagations, when no more that 5–10% of all triples are included in the relevant P spaces. This is true even when the highly complex 4 1Σ+ and 2 1∆ states at R = Re, for which the EOMCCSD calculations produce the massive, ∼33 and ∼144 millihartree, errors, which remain large (about 13 and 63 millihartree, respectively) at the CR-EOMCC(2,3) level. As shown in Table 3.2, the CC(P;Q) corrections to the EOMCC(P) energies, which account for the missing triples that the i-FCIQMC propagations at a given time τ did not capture, allow us to reach the sub-millihartree accuracy levels with less than 5% (the 2 1∆ state) or ∼10% (the 4 1Σ+ state) of triples in the relevant P spaces. The uncorrected EOMCC(P) calculations display the relatively fast convergence toward EOMCCSDT as well, but they reach similar accuracies at later propagation time τ , when about 15% (the 2 1∆ state) or 25% (the 4 1Σ+ state) of triples are captured by i-FCIQMC. Obviously, the details of the rate of convergence of the semi-stochastic CC(P;Q) calculations toward EOMCCSDT, especially when one wants to tighten it, depend on the specific excited state being calculated, but, as shown in Tables 3.1 and 3.2, once about 20% of triples are captured by the i-FCIQMC propagations, we recover the EOMCCSDT energetics for all the calculated excited states of CH+ at both geometries examined in this study to within 0.1 millihartree or better. Interestingly, there is a great deal of consistency between the behavior of the uncorrected semi-stochastic EOMCC(P) approach, in which the lists of triples defining the relevant P spaces are extracted from i-FCIQMC propagations, and the fully deterministic EOMCCSDt 43 calculations for CH+ reported in Refs. [51, 52], in which the leading triples were identified using active orbitals. Indeed, once the stochastically determined P spaces extracted from i-FCIQMC capture about 20–30% of all triples, which in the case of CH+ system examined here is achieved after 50000 or fewer ∆τ = 0.0001 a.u. MC iterations, the energies resulting from the EOMCC(P) computations become very similar to those obtained with the EOM- CCSDt method using the active space that consists of the highest-energy occupied (3σ) and three lowest-energy unoccupied (1πx, 1πy, and 4σ) orbitals. Following, the definitions of the ‘little t’ T3 and Rµ,3 operators adopted in EOMCCSDt, for the state symmetries considered in this work, the active spaces consisting of 3σ, 1πx, 1πy, and 4σ valence orbitals amount to about 26–29% of all triples included in the respective EOMCC diagonalization spaces [51, 52]. This suggests that the types and values of the triply excited amplitudes defining the Rµ,3 components of the EOM operators Rµ, which characterize the EOMCCSDt compu- tations reported in Refs. [51, 52], and those that define the R(MC) µ,3 components obtained in the i-FCIQMC-driven EOMCC(P) calculations performed after 50000 MC iterations using ∆τ = 0.0001 a.u. should be similar too. This is illustrated in Fig. 3.3, where we compare the distributions of the differences between the R(MC) µ,3 amplitudes and their full EOMCCSDT counterparts resulting from the EOMCC(P) computations at 4000 [Fig. 3.3(a)], 10000 [Fig. 3.3(b)], and 50000 [Fig. 3.3(c)] MC iterations for the 2 1Σ+ state of CH+ at R = 2Re with the analogous distribution characterizing the Rµ,3 amplitudes obtained with the EOMCCSDt approach using the 3σ, 1πx, 1πy, and 4σ active orbitals to define the corresponding triples space [Fig. 3.3(d)]; all EOM vectors Rµ needed to construct Fig. 3.3, corresponding to the EOMCC(P) EOMCCSDt, and EOMCCSDT calculations, were normalized to unity). As shown in Fig. 3.3 [cf. Fig. 3.3(c) and 3.3(d)], the small differences between the R(MC) µ,3 ampli- tudes resulting from the EOMCC(P) calculations performed after 50000 MC iterations and the Rµ,3 amplitudes obtained with EOMCCSDT, including their numerical values and dis- tribution, closely resemble those characterizing the active-space EOMCCSDt computations reported in Refs. [51, 52]. This is in perfect agreement with the small errors relative to EOM- 44 CCSDT characterizing the two calculations, which are 0.302 millihartree in the former case (cf. Table 3.2) and 0.576 millihartree in the case of EOMCCSDt [51, 52]. When we start using considerably smaller fractions of triples and, as a consequence, significantly smaller P spaces in the EOMCC(P) calculations, which is what happens when the underlying i- FCIQMC propagation is terminated too soon, the differences between the R(MC) µ,3 amplitudes resulting from the EOMCC(P) calculations and their EOMCCSDT counterparts, including their values and distribution, and the errors in the EOMCC(P) energies relative to EOM- CCSDT increase. This can be seen in Fig. 3.3, especially when one compares panel (a), which corresponds to the EOMCC(P) calculations performed after 4000 MC iterations that use only 7% of triples, with panel (d) representing EOMCCSDt, which uses a much larger fraction of triple excitations (∼30%), and in Table 3.2, where the error in EOMCC(P) energy of the 2 1Σ+ state of CH+ at R = 2Re relative to EOMCCSDT obtained after 4000 MC iterations, of 4.263 millihartree, is ∼14 times larger than the analogous error obtained after 50000 MC steps. The above analysis, which could be repeated for the remaining states of CH+, reach- ing similar conclusions, has several interesting implications for the semi-stochastic CC(P;Q) methodology pursued in this study, which will be examined by us in the future. It sug- gests, for example, that the CC(P)/EOMCC(P) and CC(P;Q) approaches using CIQMC propagations to determine the lists of higher–than–double excitations in the corresponding P spaces can be regarded as natural alternatives to the fully deterministic active-space EOMCC methods, such as EOMCCSDt, and their CC(P;Q)-corrected counterparts, such as CC(t;3) [70, 84, 87], whose performance in excited-state calculations will be an interesting thing for a future study. It also suggests that the fractions of higher–than–double excitations used to define the stochastically determined P spaces, needed to achieve high accuracies observed in the semi-stochastic CC(P;Q) calculations discussed in this work, should decrease with the basis set. It was already observed in the previous ground-state semi-stochastic work [99], and we anticipate that the same will remain true in the CIQMC-driven excited-state 45 CC(P;Q) calculations. While this remark requires a separate thorough study, beyond the scope of this initial work on the excited-state CC(P;Q) we can rationalize it by referring to the analogies between the semi-stochastic CC(P)/EOMCC(P) and CC(P;Q) approaches and their deterministic CCSDt/EOMCCSDt and CC(t;3) counterparts. Indeed, the afore- mentioned (D/d) ratio that controls the speedups offered by the CC(P)/EOMCC(P) and CC(P;Q) calculations becomes (no/No)(nu/Nu) when the active-space CCSDt/EOMCCSDt and CC(t;3) calculations, based on the ideas laid down in Refs. [51, 52, 70, 77, 79, 87], are considered, where No and Nu are the numbers of active occupied and active unoccupied orbitals, respectively, which either do not grow with the basis set or grow with it very slowly compared to no and nu. Finally, before moving to the next molecular example, we would like to point out that, in analogy to the CC(P;Q)-based CC(t;3), CC(t,q;3), and CC(t,q;3,4) calculations using active orbitals to define the underlying P spaces (see, e.g., Ref. [84]), one is better off by using smaller P spaces in the semi-stochastic CC(P)/EOMCC(P) considerations, which can be extracted out of the early stages of CIQMC propagations, and capturing the remain- ing correlations using noniterative CC(P;Q) corrections, than by running long-time CIQMC simulations to generate larger P spaces for the uncorrected CC(P)/EOMCC(P) calculations. This can be seen in Table 3.1 and 3.2 for CH+ and in the remaining Tables 3.3 and 3.4 dis- cussed in the next two subsections. We illustrate this remark by inspecting the EOMCC(P) and CC(P;Q) calculations for the 4 1Σ+ state of CH+. As shown in Table 3.1, one needs to capture about 50% of triples in the P space to reach 0.1 millihartree accuracy relative to EOMCCSDT at R = Re using the uncorrected EOMCC(P) approach. When the CC(P;Q) correction is employed, only 15% of triples are needed to reach the same accuracy level. At the more challenging R = 2Re geometry (Table 3.2), one reaches a ∼0.1 millihartree accuracy level with about 40% of triples in the P space when using the uncorrected EOMCC(P) ap- proach. This fraction reduces to about 20%, without any accuracy loss, when the CC(P;Q) correction is added to the EOMCC(P) energy. Based on the information provided in Sec- 46 tion 3.1, running the EOMCC(P) calculations with a smaller fraction of triples in the P space offers much larger savings in the computational effort than the additional time spent on determining the CC(P;Q) correction, which is, as pointed out above, considerably less expensive than a single EOMCCSDT iteration. For example, in the pilot implementation of the excited-state EOMCC(P) and CC(P;Q) approaches aimed at recovering EOMCCSDT energetics, employed in this work, the uncorrected EOMCC(P) run using 50% of triples in the P space, needed to reach a ∼0.1 millihartree accuracy relative to EOMCCSDT for the 4 1Σ+ state of CH+ at R = Re, is about twice as fast as the corresponding EOMCCSDT cal- culation. The EOMCC(P) diagonalization that forms part of the analogous CC(P;Q) run, which needs only 15% of triples in the P space to reach the same accuracy level, is about 6 times faster than EOMCCSDT. The noniterative CC(P;Q) correction is so inexpensive here that one can largely ignore the computational costs associated with its determination in this context [cf. Ref. [90] for the analogous comments made in the context of comparing costs of the CCSDt computations with those of CC(t;3)]. 47 Table 3.1 Convergence of the CC(P )/EOMCC(P ) and CC(P;Q) energies toward CCSDT/EOMCCSDT for CH+, calculated using the [5s3p1d/3s1p] basis set of Ref. [164], at the C–H internuclear distance R = Re = 2.13713 bohr. The P spaces used in the CC(P ) and EOMCC(P ) calculations were defined as all singles, all doubles, and subsets of triples extracted from i-FCIQMC propagations for the lowest states of the relevant symmetries. Each i-FCIQMC run was initiated by placing 1500 walkers on g states, the 3σ → 1π state of the 1B1(C2v) symmetry for the appropriate reference function [the RHF determinant for the 1Σ+ the 1Π states, and the 3σ2 → 1π2 state of the 1A2(C2v) symmetry for the 1∆ states], setting the initiator parameter na at 3, and the time step ∆τ at 0.0001 a.u. The Q spaces used in constructing the CC(P;Q) corrections consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [100]. MC iter. (103) 0d 2 4 6 8 10 50 100 150 200 ∞e P a 1.845 1.071 0.423 0.249 0.181 0.172 0.077 0.044 0.015 0.006 1 1Σ+ g (P ;Q)b 0.063 %Tc 0 2 1Σ+ g (P ;Q)b P a 1.373 3.856 3 1Σ+ g (P ;Q)b P a 0.787 5.537 4 1Σ+ g (P ;Q)b P a 0.954 3.080 P a 19.694 0.024 0.009 0.003 0.003 0.004 0.001 0.000 0.000 0.000 7 15 20 23 24 37 48 59 69 11.004 0.909 3.248 0.587 4.826 −4.469 0.772 5.474 4.712 1.371 1.572 0.755 0.277 0.085 0.024 0.090 1.893 0.047 1.980 0.100 0.513 0.111 1.268 0.046 1.077 0.068 0.213 0.112 0.643 0.067 0.702 0.075 0.170 0.061 0.295 0.044 0.385 0.026 0.118 0.026 0.139 0.037 0.208 0.032 0.053 0.009 0.007 0.013 0.155 0.017 0.021 0.005 0.058 0.006 0.041 0.007 0.008 0.002 0.014 0.002 0.002 0.003 0.004 1 1Π (P ;Q)b 0.792 0.179 0.102 0.054 0.058 0.046 0.027 0.013 0.005 0.003 2 1Π %Tc 0 P a 11.656 (P ;Q)b 2.805 P a 34.304 13 20 25 27 29 43 57 71 82 3.746 1.852 0.957 0.743 0.411 0.157 0.063 0.020 0.008 0.530 0.128 0.073 0.060 0.047 0.027 0.012 0.004 1.492 0.525 0.471 0.240 0.198 0.039 0.014 0.004 −0.001 0.003 1 1∆ (P ;Q)b −0.499 0.151 0.051 0.028 0.021 0.017 0.008 0.005 0.002 0.002 2 1∆ %Tc 0 P a 34.685 (P ;Q)b 0.350 10 16 18 22 24 42 56 71 82 5.951 2.542 1.892 0.940 0.877 0.133 0.043 0.008 0.003 0.432 0.128 0.094 0.057 0.041 0.011 0.005 0.003 0.002 −38.019516 −37.702621 −37.522457 −37.386872 −37.900921 −37.498143 −37.762113 −37.402308 g ground state) and EOMCC(P ) (excited states) energies relative to the corresponding CCSDT and EOMCCSDT aErrors in the CC(P ) (the 1 1Σ+ data, in millihartree. bErrors in the CC(P;Q) energies relative to the corresponding CCSDT and EOMCCSDT data, in millihartree. cThe %T values are the percentages of triples captured during the i-FCIQMC propagations for the lowest state of a given symmetry [the 1 1Σ+ 1 1A1(C2v) ground state for the 1Σ+ state for the 1∆ states]. dThe CC(P ) and EOMCC(P ) energies at τ = 0.0 a.u. are identical to the energies obtained in the CCSD and EOMCCSD calculations. The τ = 0.0 a.u. CC(P;Q) energies are equivalent to the CR-CC(2,3) (the ground state) and the CR-EOMCC(2,3) (excited states) energies. eThe CC(P ) and EOMCC(P ) energies at τ = ∞ a.u. are identical to the energies obtained in the CCSDT and EOMCCSDT calculations. g = g states, the 1B1(C2v) component of the 1 1Π state for the 1Π states, and the 1A2(C2v) component of the 1 1∆ 48 Table 3.2 Same as Table 3.1 for the stretched C–H internuclear distance R = 2Re = 4.27426 bohr. Adapted from Ref. [100]. MC iter. (103) 0 2 4 6 8 10 50 100 150 200 ∞ 1 1Σ+ g (P ;Q) %T 2 1Σ+ g (P ;Q) P 3 1Σ+ g (P ;Q) P 4 1Σ+ g P (P ;Q) P 1 1Π (P ;Q) %T 2 1Π P (P ;Q) P 1 1∆ (P ;Q) 2 1∆ %T P (P ;Q) 0.012 0.031 0.015 0.002 0.004 0.003 0.000 0.000 0.000 0.000 0 3 7 11 12 14 26 39 52 63 17.140 1.646 19.929 −2.871 32.639 12.657 13.552 5.209 4.263 1.405 1.543 0.792 0.302 0.103 0.031 0.024 0.478 12.524 −2.079 33.400 14.297 1.398 −1.741 6.383 −0.760 12.671 2.178 0.712 0.047 0.065 0.094 0.002 0.003 0.000 0.000 1.352 1.173 0.613 0.339 0.119 0.035 0.019 0.051 0.020 0.047 0.007 0.006 0.003 5.870 4.406 2.331 0.457 0.110 0.076 0.593 0.409 0.699 0.436 0.342 0.227 0.013 0.061 0.011 0.013 0.006 0.005 0.000 −0.006 0.001 0.002 2.303 0.306 0.058 0.033 0.050 0.039 0.007 0.002 0.002 0.001 0 7 12 14 16 17 30 41 52 65 21.200 −1.429 44.495 −4.526 1.644 0.724 0.612 0.457 0.220 0.079 0.016 0.007 0.001 −0.060 1.372 0.050 0.031 0.451 0.422 −0.002 0.253 0.014 0.060 0.004 0.002 0.000 0.122 0.047 0.013 0.005 0.001 0.046 0.014 0.022 0.007 −0.001 0.005 0.004 0.001 0.000 0 6 9 12 13 14 26 36 47 57 144.414 −63.405 13.363 3.338 2.340 2.088 0.862 0.288 0.038 0.014 0.003 0.368 0.130 0.063 0.021 0.038 0.005 0.000 0.000 0.000 P 5.002 1.588 0.504 0.275 0.263 0.148 0.030 0.009 0.004 0.001 −37.900394 −37.704834 −37.650242 −37.495275 −37.879532 −37.702345 −37.714180 −37.494031 49 Figure 3.2 Convergence of the EOMCC(P ) [panels (a) and (c)] and CC(P;Q) [panels (b) and (d)] energies toward EOMCCSDT for the three lowest-energy excited states of the 1Σ+ symmetry, two lowest states of the 1Π symmetry, and two lowest 1∆ states of the CH+ ion, as described by the [5s3p1d/3s1p] basis set of Ref. [164], at the C–H internuclear distance R set at Re = 2.13713 bohr [panels (a) and (b)] and 2Re = 4.27426 bohr [panels (c) and (d)]. Adapted from Ref. [100]. 50 Figure 3.3 The distributions of the differences between the R(MC) amplitudes and their EOMCCSDT counterparts resulting from the EOMCC(P ) computations at (a) 4000, (b) 10,000, and (c) 50,000 MC iterations using τ = 0.0001 a.u. for the 2 1Σ+ state of CH+ at R = 2Re with the analogous distribution characterizing the Rµ,3 amplitudes obtained with the EOMCCSDt approach employing the 3σ, 1πx, 1πy, and 4σ active orbitals to define the corresponding triples space [panel (d)]. All vectors Rµ needed to construct panels (a)–(d) were normalized to unity. Adapted from Ref. [100]. µ,3 51 3.2.2 CH This subsection and Section 3.3 focus on the results of semi-stochastic CC(P;Q) calcula- tions for open-shell species and biradicals reported in Refs. [100] and [102] and obtained by me. In this subsection, we discuss the results for the CH radical that I contributed to Ref. [100]. Similar convergence patterns in the semi-stochastic EOMCC(P) and CC(P;Q) calcula- tions are observed for the CH radical (see Table 3.3). In this case, following our earlier deterministic EOMCC work, including the CR-EOMCC [62, 68] and electron-attachment (EA) EOMCC [68, 78, 141] approaches, and a wide range of EOMCC computations, in- cluding the high EOMCCSDT and EOMCCSDTQ levels, published by Hirata [161], along with the X 2Π ground state, we examined the three low-lying doublet excited states, des- ignated as A 2∆, B 2Σ−, and C 2Σ+, which belong to different irreducible representations than that of the ground state. In analogy to the aforementioned EOMCC studies of CH [62, 68, 78, 141, 161], the relevant CC(P) (the X 2Π state) and EOMCC(P) (excited states) electronic energies and their CC(P;Q) counterparts were determined at the corresponding experimentally derived equilibrium C–H distances, which are 1.1197868 ˚A for the X 2Π state [175], 1.031 ˚A for the A 2∆ state [175], 1.1640 ˚A for the B 2Σ− state [176], and 1,1143 ˚A for the C 2Σ+ state [177] (cf. Table 3.3). SInce all of our CC(P)/EOMCC(P) and CC(P;Q) calculations, starting from the τ = 0 CCSD/EOMCCSD and CR-EOMCC(2,3) levels and ending up with the larger values of τ needed to examine the convergence toward the parent CCSDT/EOMCCSDT energetics, were performed using the ROHF reference determinant, we also computed the ROHF-based CCSDT/EOMCCSDT energies, which formally corre- spond to the τ = ∞ CC(P)/EOMCC(P) and CC(P;Q) results. We had to do it, since the previously published CCSDT/EOMCCSDT results [161] relied on the unrestricted Hartree– Fock rather than the ROHF reference. In analogy to CH+, the lists of triples defining the T (MC) 3 component of the cluster op- erator T (P ) and the R(MC) µ,3 and the L(MC) µ,3 components of the EOM excitation and deexcita- 52 tion operators, R(P ) µ and L(P ) µ , respectively, used in the CC(P), EOMCC(P), and CC(P;Q) calculations for the CH radical, were extracted from the i-FCIQMC propagations for the lowest-energy states of the relevant irreps of C2v, namely, the 2B2(C2v) component of the X 2Π state, the lowest state of the 2A1(C2v) symmetry in the case of the A 2∆ and C 2Σ+ states, and the lowest 2A2(C2v) state when considering the B 2Σ− state (again, we used C2v, which is the largest Abelian subgroup of the true point group of CH, C∞v). As explained in the Piecuch group’s earlier papers [62, 68, 78, 141] and as shown in Ref. [161], all three excited states of the CH radical considered here, especially B 2Σ− and C 2Σ+, which are dominated by two-electron excitations (cf. the reduced excitation level (REL) diagnostic values in Tables II and III of Ref. [68] or Table II of Ref. [62]), constitute a significant challenge, requiring the full EOMCCSDT treatment to obtain a reliable adiabatic excitation spectrum. This can be seen by inspecting the τ = 0 EOMCC(P) i.e., EOM- CCSD, energies for the A 2∆, B 2Σ−, and C 2Σ+ states of CH shown in Table 3.3, which are characterized by the ∼13, ∼39, and ∼44 millihartree errors relative to EOMCCSDT, respectively. The CR-EOMCC(2,3) triples corrections to EOMCCSD, represented in Table 3.3 by the τ = 0 CC(P;Q) values, help, especially in the case of the C 2Σ+ state, but the situation is far from ideal, since errors on the order of 8 and 5 millihartree for the A 2∆ and B 2Σ− states, respectively, remain. The situation considerably improves when we turn to the semi-stochasticCC(P;Q) calculations, which incorporate the leading triples in the relevant P spaces by extracting them from the corresponding i-FCIQMC propagations and correct the resulting energies for the remaining triple excitations that were not captured by i-FCIQMC at a given time τ . As shown in Table 3.3, in the case of the A 2∆ and B 2Σ− states, which are not only challenging to EOMCCSD, but also to CR-EOMCC(2,3), we can reach com- fortable 1–2 millihartree errors relative to EOMCCSDT using the semi-stochasticCC(P;Q) corrections developed in this work once the relevant P spaces contain about 20–40% of all triples. With ∼50% triples in the same P space, the CC(P;Q) energies of the A 2∆ and B 2Σ− states are within fractions of a millihartree from EOMCSDT. These are consider- 53 able improvements relative to the purely deterministic EOMCCSD and CR-EOMCC(2,3) computations, which give ∼13–39 and ∼5–8 millihartree errors, respectively, for the same two states, and the semi-stochasticEOMCC(P) calculations that reach 1–2 millihartree ac- curacy levels with about 70–80% triples in the respective P spaces. In the case of the C 2Σ+ state, which is a major challenge to EOMCCSD, but not to CR-EOMCC(2,3), the behavior of the EOMCC(P) and CC(P;Q) approaches is different, since the CC(P;Q) cor- rections obtained with the help of some triples in the P space captured by i-FCIQMC are no longer needed to obtain the well-converged energetics, i.e., the τ = 0 CC(P;Q) result, where the P space is spanned by singles and doubles only, is sufficiently accurate, but it is still interesting to observe that one can tighten the convergence further, reaching sta- ble < 0.1 millihartree errors relative to EOMCCSDT with about 50% of all triples in the P space. In analogy to the A 2∆ and B 2Σ− states, it is also interesting to observe a reasonably smooth convergence of the uncorrected EOMCC(P) energies toward EOMCCSDT. It is clear from the results presented in Table 3.3 that the CC(P;Q) corrections to the semi-stochastic CC(P) and EOMCC(P) energies offer considerable speedups compared to the uncorrected CC(P)/EOMCC(P)calculations, not only for the closed-shell molecules, such as CH+, but also when examining open-shell species. 54 Table 3.3 Convergence of the CC(P )/EOMCC(P ) and CC(P;Q) energies toward CCS- DT/EOMCCSDT for CH, calculated using the aug-cc-pVDZ basis set. The P spaces used in the CC(P ) and EOMCC(P ) calculations were defined as all singles, all doubles, and sub- sets of triples extracted from i-FCIQMC propagations for the lowest states of the relevant symmetries. Each i-FCIQMC run was initiated by placing 1500 walkers on the appropriate reference function [the ROHF 2B2(C2v) determinant for the X 2Π state, the 1π → 4σ state of the 2A1(C2v) symmetry for the A 2∆ and C 2Σ+ states, and the 3σ → 1π state of the 2A2(C2v) symmetry for the B 2Σ− state], setting the initiator parameter na at 3, and the time step ∆τ at 0.0001 a.u. The Q spaces used in constructing the CC(P;Q) corrections consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [100]. MC iter. (103) 0 2 4 6 8 10 12 14 16 18 20 50 100 150 200 ∞e X 2Π (P ;Q)b 0.231 0.170 0.086 0.035 0.022 0.019 0.015 0.013 0.008 0.008 0.006 0.002 0.002 0.000 0.000 P a 2.987 2.405 1.413 0.883 0.603 0.495 0.445 0.389 0.309 0.292 0.243 0.150 0.055 0.025 0.010 P a %Tc 0.0 13.474 13.8 13.009 41.7 10.907 58.9 10.119 66.8 7.764 72.6 6.987 76.5 6.640 77.5 7.040 79.2 6.047 80.3 4.646 82.2 3.809 89.1 1.367 95.3 0.177 A 2∆ (P ;Q)b 7.727 7.395 5.288 4.577 2.436 2.170 1.981 1.887 1.667 0.875 0.754 0.112 0.017 98.1 0.042 −0.003 %Tc 0.0 P a 38.620 9.8 10.602 19.3 27.2 34.6 38.1 42.3 45.7 48.3 49.8 52.6 74.1 91.7 98.0 7.066 3.452 2.309 1.965 1.832 1.180 1.303 1.349 0.796 0.298 0.144 0.010 B 2Σ− (P ;Q)b −4.954 −1.848 −1.259 −0.371 −0.149 −0.024 −0.081 0.030 0.012 P a %Tc 0.0 43.992 18.5 40.700 38.9 31.017 53.2 26.364 61.4 20.545 64.8 17.180 69.5 16.929 72.2 13.114 C 2Σ+ (P ;Q)b 0.087 −0.689 −0.319 −0.508 −0.412 0.435 0.029 0.253 75.6 7.646 −0.041 −0.062 77.5 5.312 0.038 0.038 0.014 0.007 79.5 4.691 91.6 1.436 98.3 0.204 99.6 0.063 0.011 0.108 0.070 0.013 0.010 0.001 %Tc 0.0 9.8 19.7 28.8 34.3 38.3 42.5 45.1 48.7 50.1 52.2 74.0 91.3 98.2 99.7 99.2 0.007 0.001 99.7 −0.001 −0.001 99.9 0.010 −38.387749 −38.276770 −38.267544 −38.238205 aErrors in the CC(P ) (the X 2Π ground state) and EOMCC(P ) (excited states) energies relative to the corresponding CCSDT and EOMCCSDT data, in millihartree, calculated at the experimentally obtained equilibrium C–H distances used in Refs. [62, 68, 161], which are 1.1197868 ˚A for the X 2Π state [175], 1.1031 ˚A for the A 2∆ state [175], 1.1640 ˚A for the B 2Σ− state [176], and 1.1143 ˚A for the C 2Σ+ state [177]. The lowest-energy core orbital was frozen in all correlated calculations. bErrors in the CC(P;Q) energies relative to the corresponding CCSDT and EOMCCSDT data, in milli- hartree, calculated at the experimentally determined equilibrium C–H distances as used in Refs. [62, 68, 161] (see footnote a for the C–H distances). cThe %T values are the percentages of triples captured during the i-FCIQMC propagations for the lowest state of a given symmetry [the 2B2(C2v) component of the X 2Π ground state, the lowest 2A1(C2v) state for the A 2∆ and C 2Σ+ states, and the lowest 2A2(C2v) state for the B 2Σ− state]. dThe CC(P ) and EOMCC(P ) energies at τ = 0.0 a.u. are identical to the energies obtained in the CCSD and EOMCCSD calculations. The τ = 0.0 a.u. CC(P;Q) energies are equivalent to the CR-CC(2,3) (the ground state) and the CR-EOMCC(2,3) (excited states) energies. eThe CC(P ) and EOMCC(P ) energies at τ = ∞ a.u. are identical to the energies obtained in the ROHF- based CCSDT and EOMCCSDT calculations. 55 Figure 3.4 Convergence of the EOMCC(P) and CC(P;Q) energies of the A 2∆ [panel (a)], B 2Σ− [panel (b)], and C 2Σ+ [panel (c)] states of CH toward EOMCCSDT for the three lowest-energy excited states of CH calculated as described by the aug-cc-pVDZ basis set. The geometries used are the equilibrium C–H distances reported in Refs. [62, 68, 161], which are 1.1031 ˚A for the A 2∆ state [175], 1.1640 ˚A for the B 2Σ− state [176], and 1.1143 ˚A for the C 2Σ+ state [177]. 56 050100150200MC Iterations (103)−1012345678Error rel. (o EOMCCSDT (mE )A2Δ(a)EOMCC(P)CC(P;Q)050100150200MC I(era(ions (103)−1012345678B2Σ−(b)EOMCC(P)CC(P;Q)050100150200MC I(era(ions (103)−1012345678C2Σ+(c)EOMCC(P)CC(P;Q) 3.2.3 CNC Our last example, which is also the largest many-electron system considered in the present study, is the linear, D∞h symmetric, CNC molecule. Following our earlier CR-CC(2,3)/CR- EOMCC(2,3) and EA-EOMCC calculations for this challenging open-shell molecular species [68, 162, 163], we considered the X 2Πg ground state and the two low-lying doublet excited states, A 2∆u and B 2Σ+ u . The i-FCIQMC-driven CC(P) ground-state and EOMCC(P) excited-state energies and the corresponding CC(P;Q) corrections, along with their deter- ministic EOMCCSD, CR-EOMCC(2,3), and EOMCCSDT counterparts, were calculated us- ing the equilibrium C–N distances optimized in Ref. [162] with EA-SAC-CI. They are 1.253 ˚A for the X 2Πg state, 1.256 ˚A for the A 2∆u state, and 1.259 ˚A for the B 2Σ+ u state. As in the case of the CH radical, we used the ROHF reference determinant. Following the com- putational protocol adopted in this study, and in analogy to the CH+ and CH species, the lists of triples defining the T MC 3 , R(MC) µ,3 , and L(MC) µ,3 components used in the semi-stochastic CC(P),EOMCC(P), and CC(P;Q) calculations for CNC were obtained using the i-FCIQMC propagations for the lowest-energy states of the relevant irreps of the largest Abelian sub- group of D∞h, i.e., D2h, meaning the 2B2g (D2h) component of the X 2Πg state and the lowest state of the 2B1u (D2h) symmetry in the case of the A 2∆u and B 2Σ+ u states. As shown in Table 3.4 and in agreement with one of our previous studies [68], all three states of CNC considered in this work, especially A 2∆u and B 2Σ+ u , are poorly described by CCSD and EOMCCSD, which produce more than 18, 31, and 111 millihartree errors, respec- tively, relative to the target EOMCCSDT energetics (see the τ = 0 CC(P) and EOMCC(P) energies in Table 3.4). The excessively large, > 111 millihartree, error in the EOMCCSD energy of the B 1Σ+ u state is related to its strongly multireference character dominated by two- electron excitations (cf. the REL values characterizing the excited states of CNC in Table IV of Ref. [68]). In the case of the ground state and the B 2Σ+ u excited state, the CR-CC(2,3) and CR-EOMCC(2,3) corrections to CCSD and EOMCCSD seem to be quite effective, reducing the large errors relative to CCSDT/EOMCCSDT observed in the CCSD and EOMCCSD 57 calculations to a sub-millihartree level, but the ∼16 millihartree error resulting from the CR-EOMCC(2,3) calculations for the A 2∆u state, while considerably lower than the > 31 millihartree error obtained with EOMCCSD, is still rather large (see the τ = 0 CC(P;Q) energies in Table 3.4). By incorporating the dominant triply excited determinants captured by the i-FCIQMC propagations in the respective P spaces, the semi-stochasticCC(P) and EOMCC(P) approaches help, allowing us to reach stable ∼1–2 millihartree accuracy levels for the X 2Πg and A 2∆u states relative to the target CCSDT/EOMCCSDT energetics with about 50–60% triples, but the CC(P;Q) corrections that account for the remaining triples, missing in the i-FCIQMC wave functions, are considerably more effective. In the case of the A 2∆u state, which poses problems to both EOMCCSD and CR-EOMCC(2,3), which give about 31 and 16 millihartree errors relative to EOMCCSDT, respectively, we reach a stable ∼1–2 millihartree accuracy level with about 30–40% triples in the corresponding P space, as opposed to the aforementioned 50–60% needed in the uncorrected EOMCC(P) run. The benefits of using the semi-stochastic CC(P;Q) vs. deterministic CR-EOMCC(2,3) corrections for the X 2Πg and B 2Σ+ u states are less obvious, but it is encouraging to observe the rapid convergence toward the target CCSDT and EOMCCSDT energetics in the former calculations. In particular, they allow us to lower the 0.4–0.5 millihartree Errors obtained with CR-EOMCC(2,3) to a 0.1 millihartree level with about 10% of all triples, identified by i-FCIQMC, in the case of the X 2Πg state and with ∼30–40% triples in the P space when the B 2Σ+ u state is considered. Once again, the CC(P;Q) corrections to the energies resulting from the semi-stochastic CC(P) and EOMCC(P) calculations speed up the uncor- rected CC(P)/EOMCC(P) computations, while allowing us to improve the CR-CC(2,3) and CR-EOMCC(2,3) energetics by bringing them very close to the CCSDT and EOMCCSDT levels at the fraction of the cost. 58 Table 3.4 Convergence of the CC(P )/EOMCC(P ) and CC(P;Q) energies toward CCS- DT/EOMCCSDT for CNC, calculated using DZP[4s2p1d] basis set. The P spaces used in the CC(P ) and EOMCC(P ) calculations were defined as all singles, all doubles, and sub- sets of triples extracted from i-FCIQMC propagations for the lowest states of the relevant symmetries. Each i-FCIQMC run was initiated by placing 1500 walkers on the appropriate reference function [the ROHF 2B2g(D2h) determinant for the X 2Πg state and the 3σu → 1πg state of the 2B1u(D2h) symmetry for the A 2∆u and B 2Σ+ u states], setting the initiator pa- rameter na at 3, and the time step ∆τ at 0.0001 a.u. The Q spaces used in constructing the CC(P;Q) corrections consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [100]. MC iter. (103) 0 2 4 6 8 10 12 14 16 18 20 50 100 150 ∞e X 2Πg (P ;Q)b −0.495 −0.043 −0.029 −0.011 −0.013 −0.006 −0.003 −0.005 −0.003 −0.003 −0.001 0.000 0.000 0.000 P a 18.458 10.331 4.424 2.824 1.818 1.306 1.092 0.911 0.820 0.651 0.610 0.077 0.002 0.000 P a %Tc 0.0 31.157 13.2 18.835 33.2 10.637 44.1 7.555 49.9 6.181 53.3 5.187 56.5 4.162 58.7 3.529 60.6 3.106 62.5 2.510 63.9 2.395 79.7 0.172 94.5 0.002 99.3 0.000 A 2∆u (P ;Q)b 16.017 9.114 5.717 4.199 3.090 2.441 1.778 1.418 1.149 0.811 0.785 0.058 0.001 0.000 P a %Tc 0.0 111.307 6.5 81.493 16.1 53.677 22.7 35.539 27.5 26.767 30.8 21.337 34.0 17.056 37.0 12.843 39.5 41.7 44.4 70.9 92.3 99.1 9.197 8.879 7.548 0.732 0.005 0.000 B 2Σ+ u (P ;Q)b −0.433 −2.496 −2.526 −1.254 −0.864 −0.284 0.196 0.046 0.134 −0.034 0.151 0.055 0.003 0.000 %Tc 0.0 6.5 16.0 22.8 27.9 31.5 34.3 37.5 39.9 42.4 44.7 70.7 91.9 99.1 −130.421932 −130.276946 −130.252999 aErrors in the CC(P ) (X 2Πg state) and EOMCC(P ) (the remaining states) energies relative to the cor- responding CCSDT and EOMCCSDT data, in millihartree, calculated at the experimentally obtained equilibrium C–N distances optimized in Ref. [162], which are 1.253 ˚A for the X 2Πg state, 1.256 ˚A for the A 2∆u state, and 1.259 ˚A for the B 2Σ+ u state. The three lowest-energy core orbital was frozen in all correlated calculations. bErrors in the CC(P;Q) energies relative to the corresponding CCSDT and EOMCCSDT data, in milli- hartree, calculated at the equilibrium C–N distances optimized in Ref. [162] (see footnote a for these C–N distances). cThe %T values are the percentages of triples captured during the i-FCIQMC propagations for the lowest state of a given symmetry [the 2B2g(D2h) component of the X 2Πg ground state and the lowest 2B1u(D2h) state for the A 2∆ and B 2Σ+ dThe CC(P ) and EOMCC(P ) energies at τ = 0.0 a.u. are identical to the energies obtained in the CCSD and EOMCCSD calculations. The τ = 0.0 a.u. CC(P;Q) energies are equivalent to the CR-CC(2,3) (the ground state) and the CR-EOMCC(2,3) (excited states) energies. eThe CC(P ) and EOMCC(P ) energies at τ = ∞ a.u. are identical to the energies obtained in the ROHF- based CCSDT and EOMCCSDT calculations. u states]. 59 Figure 3.5 Convergence of the EOMCC(P) and CC(P;Q) energies of the A 2∆u [panel (a)] and B 2Σ+ u [panel (b)] states of CNC toward EOMCCSDT for the two lowest-energy doublet excited states of CNC calculated as described by the DZP[4s2p1d] basis set. The geometries used are the equilibrium C–N distances reported in Refs. [68, 162], which are 1.256 ˚A for the A 2∆u state and 1.259 ˚A for the B 2Σ+ u state. 60 050100150MC Iterations (103)−20246810121416Error re . to EOMCCSDT (mEh)A2Δu(a)EOMCC(P)CC(P;Q)050100150MC Iterations (103)−20246810121416B2Σ+u(b)EOMCC(P)CC(P;Q) 3.3 Singlet–Triplet Gaps in Methylene, (HFH)−, Cyclobutadiene, Cyclopenta- dienyl Cation, and Trimethylenemethane In order to assess the performance of our semi-stochastic, CIQMC-driven, CC(P ;Q) methodology in converging the full CCSDT data for the singlet–triplet gaps and the corre- sponding singlet- and triplet-state energies of biradical systems, we applied it to methylene, (HFH)−, cyclobutadiene, cyclopentadienyl cation, and trimethylenemethane. The results discussed in this section, which were all generated by me as part of this dissertation project, were reported in Ref. [102]. Following our earlier studies of the singlet–triplet gaps in the same systems using the deterministic CC(P ;Q) [88] and DEA/DIP-EOMCC [80, 153–155] approaches, we used the aug-cc-pVTZ basis set [165, 166] for methylene, the 6-31G(d,p) basis [178, 179] for the (HFH)− ion, and the cc-pVDZ basis set [165] for cyclobutadiene, cyclopenta- dienyl cation, and trimethylenemethane. In the case of methylene and trimethylenemethane, we focused on the ability of the semi-stochastic CC(P ;Q) algorithm to converge the adiabatic ∆ES-T data obtained with CCSDT. When executing the semi-stochastic CC(P ;Q) calcula- tions for (HFH)−, cyclobutadiene, and cyclopentadienyl cation, we focused on recovering the CCSDT values of the vertical singlet–triplet gaps. Throughout this work, we define ∆ES-T as ES − ET, where ES and ET are the electronic energies of the corresponding singlet and triplet states, i.e., the positive ∆ES-T value implies that triplet is lower in energy. All of the CC calculations reported in this section were performed using our group’s standalone codes, interfaced with the RHF, ROHF, and integral transformation routines in the GAMESS package [172, 173] which were originally developed in Refs. [70, 84, 87, 88], and extended to the stochastically generated P spaces for the use in CC(P ) and CC(P ;Q) computations in Refs. [99–101, 137]. The i-FCIQMC [methylene, (HFH)−, and cyclobuta- diene] and i-CISDTQ-MC (cyclopentadienyl cation and trimethylenemethane) calculations, needed to generate the lists of triples for the semi-stochastic CC(P ) and CC(P ;Q) runs, were carried out with the HANDE software [170, 171]. Each of the i-FCIQMC and i-CISDTQ-MC propagations was initiated by placing 1500 walkers on the relevant reference determinant. 61 The CIQMC time step δτ and the initiator parameter na were set at 0.0001 a.u. and 3, respectively. In all post-Hartree–Fock calculations, the core MOs correlating with the 1s orbitals of the C and F atoms were kept frozen. If the true point group of the biradical sys- tem of interest was not Abelian, we used its largest Abelian subgroup, since our CC codes interfaced with GAMESS and the CIQMC routines in HANDE cannot handle non-Abelian symmetries. 3.3.1 Methylene We begin the discussion of our results by analyzing the performance of the semi-stochastic CC(P ;Q) approach in converging the CCSDT energies of the ground (X 3B1) and first- excited (A 1A1) states of the methylene/aug-cc-pVTZ system and the adiabatic gap between them. The C2v-symmetric geometries of CH2 in the two states, optimized using FCI and the [5s3p/3s] triple zeta basis set of Dunning [180] augmented with two sets of polarization func- tions (TZ2P), were taken from Ref. [181]. The electronically nondegenerate triplet ground state has a predominantly SR nature dominated by the (1a1)2(2a1)2(1b2)2(3a1)1(1b1)1 config- uration, whereas the first-excited singlet state exhibits a significant MR character requiring a linear combination of the (1a1)2(2a1)2(1b2)2(3a1)2 and doubly excited (1a1)2(2a1)2(1b2)2(1b1)2 closed-shell determinants for a proper zeroth-order description. Because of these fundamen- tally different characteristics of the X 3B1 and A 1A1 states, a well-balanced and accurate treatment of dynamical and nondynamical correlation effects is the key to a reliable descrip- tion of the singlet–triplet gap in methylene. It is, therefore, unsurprising that one usually resorts to methods of the MRCI [181–186] or MRCC [187–190] type, or to the high-level SRCC theories that account for higher–than–doubly excited clusters in an iterative manner, such as full CCSDT used in Refs. [88, 191], to accomplish this goal (for other examples of high-level ab initio calculations for the X 3B1 and A 1A1 states of methylene, see Refs. [80, 153, 154, 192] and references therein). The CCSDT results for the adiabatic singlet– triplet gap in methylene, which are of interest in the present study, are indeed very accurate. As shown, for example, in Ref. [88], the difference between the adiabatic ∆ES-T value ob- 62 tained in the CCSDT/TZ2P calculations and the corresponding FCI result of 11.14 kcal/mol [181] is only 0.11 kcal/mol or 38 cm−1. As demonstrated in Ref. [88] as well, the purely elec- tronic A 1A1 − X 3B1 separation resulting from the CCSDT computations using the aug-cc- pVTZ basis employed in this work is only about 0.15 kcal/mol (∼50 cm−1) higher than the experimentally derived value of 9.37 kcal/mol reported in Ref. [188], obtained by correcting the vibrationless adiabatic singlet–triplet gap determined in Ref. [193] for the relativistic and nonadiabatic (Born–Oppenheimer diagonal correction) effects estimated in Refs. [194] and [195], respectively. It is, therefore, interesting to examine if the semi-stochastic CC(P ;Q) approach advocated in this work is capable of reproducing the high-quality CCSDT/aug- cc-pVTZ data for the X 3B1 and A 1A1 states of methylene and the adiabatic separation between them. The results of our FCIQMC-driven CC(P ;Q) calculations for the methylene/aug-cc-pVTZ system, reported as errors relative to the parent CCSDT data, and their CC(P ) counterparts are shown in Table 3.5 and Fig. 3.6. The reference determinants |Φ⟩ used to initiate the i-FCIQMC propagations and to carry out the CC(P ), CC(P ;Q), CCSD, CR-CC(2,3), and CCSDT calculations were the ROHF determinant in the case of the X 3B1 state and the RHF determinant for the A 1A1 state. The subsets of triply excited determinants needed to construct the P spaces used in the CC(P ) and CC(P ;Q) computations at various propagation times τ were the Sz = 1 triples of the B1 symmetry captured during the i-FCIQMC run for the X 3B1 state and the Sz = 0 triples of the A1 symmetry captured during the analogous run for the A 1A1 state. Following the semi-stochastic CC(P ;Q) algorithm described in Section 3.1, the Q spaces needed to determine corrections δ(P ; Q) were defined as the remaining triples not captured by i-FCIQMC. Let us start our analysis by examining the CC(P ) and CC(P ;Q) data at τ = 0, where the P spaces do not contain any triply excited determinants. As shown in Table 3.5, the CC(P ) energies of the X 3B1 and A 1A1 states at τ = 0, which are equivalent to those obtained using conventional CCSD, are above their CCSDT [i.e., τ = ∞ CC(P )] counterparts by 63 4.187 and 5.918 millihartree, respectively. This translates into a 380 cm−1 or ∼11% error in the adiabatic ∆ES-T value when compared to the 3328 cm−1 singlet–triplet gap obtained with CCSDT. The situation improves when the CC(P ;Q) corrections δ(P ; Q) due to T3 correlation effects, calculated by placing all triply excited determinants in the respective Q spaces, are added to the CC(P ) energies. The τ = 0 CC(P ;Q) or CR-CC(2,3) energy characterizing the X 3B1 state is only 0.177 millihartree above the parent CCSDT value, which is an error reduction relative to CCSDT compared to the underlying CC(P ) result by a factor of ∼24. The δ(P ; Q) correction improves the τ = 0 CC(P ) energy of the more challenging A 1A1 state as well, although the difference between the resulting CR-CC(2,3) energy and its CCSDT counterpart, of 0.656 millihartree, is almost 4 times larger than the analogous difference obtained for the X 3B1 state. As a result, the 105 cm−1 error relative to CCSDT characterizing the adiabatic A 1A1 − X 3B1 separation obtained in the τ = 0 CC(P ;Q) or CR-CC(2,3) calculations, while considerably smaller than the 380 cm−1 obtained in the underlying CC(P) (i.e., CCSD) runs, leaves room for further improvements. One can improve the CR-CC(2,3) energies of the X 3B1 and A 1A1 states and the gap between them by enriching the P spaces defining the CC(P ) calculations with the leading triply excited determinants identified using active orbitals and correcting the resulting CCSDt energies for the remaining T3 correlations that have not been captured by CCSDt [88], but our objective here is to examine if one can accomplish the same, or improve the CC(t;3) results reported in Ref. [88] even further, by turning to the more black-box semi-stochastic CC(P ;Q) methodology, in which the dominant triply excited determinants are identified with CIQMC. The results in Table 3.5 and Fig. 3.6 show that when the τ = 0 P spaces are augmented with the subsets of triply excited determinants captured in the i-FCIQMC runs at τ > 0 and, following the CC(P ;Q) recipe, the resulting CC(P ) energies are corrected for the remaining T3 correlations, the convergence of the total electronic energies of the X 3B1 and A 1A1 states and the adiabatic separation between them toward their CCSDT parents is rapid. We can see this already in the early stages of the i-FCIQMC propagations. For example, at τ = 0.8 64 a.u., i.e., after only 8000 δτ = 0.0001 a.u. MC iterations, the errors in the CC(P ;Q) energies of the X 3B1 and A 1A1 states and the corresponding ∆ES-T value relative to CCSDT are 0.049 millihartree, 0.106 millihartree, and 13 cm−1, respectively, substantially improving the CR-CC(2,3) [i.e., τ = 0 CC(P ;Q)] calculations, which give 0.177 millihartree for the X 3B1 state, 0.656 millihartree for the A 1A1 state, and 105 cm−1 for ∆ES-T. This confirms our expectation that the main source of errors in the CR-CC(2,3) computations, especially in the case of the more MR A 1A1 state, which is characterized by larger T3 effects, is the use of the unrelaxed T1 and T2 amplitudes obtained with CCSD in constructing the correction due to triples. The FCIQMC-based CC(P ;Q) calculations at τ = 0.8 a.u., which use as little as 16% of all triply excited determinants to define the P space for the X 3B1 state and only 25% of all triples in the P space for the A 1A1 state, are also more accurate than the CC(t;3) computations reported in Ref. [88], which produced the 0.130 millihartree, 0.409 millihartree, and 61 cm−1 errors relative to CCSDT for the X 3B1 and A 1A1 energies and ∆ES-T, respectively. This is all very promising, especially if we realize that the i-FCIQMC propagations used to generate the lists of triples for our semi-stochastic CC(P;Q) runs, which work so well, are very far from convergence when τ = 0.8 a.u. Indeed, as seen in Table 3.6, the total numbers of walkers at 8000 δτ = 0.0001 a.u. MC iterations, which are 132689 in the case of the X 3B1 state and 165564 for the A 1A1 state, represent tiny fractions, 2.17% and 1.11%, respectively, of the total walker populations at τ = 20.0 a.u., where we stopped our i-FCIQMC propagations (see Fig. 3.7 for a comparison of the rate of convergence of the CC(P), CC(P;Q), and the underlying i-FCIQMC calculations). As demonstrated in Table 3.5 and Fig. 3.6, the convergence of the energies of the X 3B1 and A 1A1 states and the gap between them resulting from the FCIQMC-driven CC(P;Q) calculations remains fast at the larger propagation times τ as well. For example, if we allow i-FCIQMC to populate the respective P spaces with about 26–38% of all triples, which happens after 20000 δτ = 0.0001 a.u. MC iterations, the CC(P ;Q) energies of the X 3B1 and A 1A1 states and the resulting singlet–triplet gap become practically indistinguishable from 65 the parent CCSDT data, with errors in the total electronic energies and ∆ES-T being only ∼20 microhartree and 2 cm−1, respectively. As shown in Table 3.6 of the, walker populations characterizing the X 3B1 and A 1A1 states produced by i-FCIQMC at 20000 δτ = 0.0001 a.u. MC time steps are still very small compared to the total numbers of walkers at τ = 20.0 a.u., where we terminated our i-FCIQMC propagations (4.11% for the X 3B1 state and 2.17% in the case of the A 1A1 state). It is also interesting to note that the more MR A 1A1 state requires a higher fraction of triply excited determinants in the P space than its SR X 3B1 counterpart to achieve similar accuracy levels in the semi-stochastic CC(P ;Q) computations for both states. For example, the i-FCIQMC propagation has to capture about 25% of all triples, for the inclusion in the P space, if we are to reduce errors relative to CCSDT in the CC(P ;Q) calculations for the A 1A1 state to ∼0.1 millihartree. In the case of the X 3B1 state, the analogous fraction of triples is about 10% (cf. Table 3.5). This highlights the importance of balancing the SR triplet state with the more MR singlet state in obtaining accurate ∆ES-T estimates, which is not a problem for the semi-stochastic CC(P;Q) methodology because the underlying i-FCIQMC wave function sampling is very effective in identifying the dominant higher–than–doubly excited determinants, to be included in the relevant P spaces, and the δ(P ; Q) corrections to the CC(P ) energies take care of the remaining correlation effects of interest. Before concluding this subsection and discussing other molecular examples, we would like to comment on the effectiveness of the noniterative corrections δ(P ; Q), adopted in the CC(P ;Q) formalism, in accelerating convergence of the underlying CC(P ) calculations toward CCSDT. The CC(P ) and CC(P ;Q) error curves shown in Fig. 3.6 illustrate this best. It is clear from this figure that the CC(P ;Q) energies of the X 3B1 and A 1A1 states [Fig. 3.6 (a) and (b)] and the corresponding ∆ES-T values [Fig. 3.6 (c)] converge to the parent CCSDT data much faster than in the case of the uncorrected CC(P ) computations. One can see the same by inspecting the numerical data shown in Table 3.5. In this context, it is worth commenting on the CC(P) and CC(P;Q) results obtained after 8000 MC iterations. In 66 that case, the CC(P;Q) calculations reduce the ∼2.4 millihartree errors relative to CCSDT characterizing the CC(P ) energies of the X 3B1 and A 1A1 states to 0.1 millihartree or less, which is a desired behavior, but the CC(P ;Q) ∆ES-T value is less accurate than that obtained with the uncorrected CC(P). One should not read too much into this though. The fact that the CC(P;Q) calculations at 8000 MC iterations increase the very small 3 cm−1 error obtained with CC(P) to 13 cm−1 is a coincidence arising from the accidental cancellation of errors in the CC(P) total electronic energies obtained at this particular propagation time. Indeed, when the later stages of the i-FCIQMC propagations are considered, the differences between the CC(P) and CCSDT values of ∆ES-T become increasingly negative, reaching −107 cm−1 at 50000 MC iterations, before eventually converging to the CCSDT limit, whereas the corresponding CC(P;Q) results display a smooth behavior, rapidly approaching CCSDT. In particular, they reduce the relatively large negative error value obtained for ∆ES-T in the CC(P) calculations at 50000 MC iterations to a numerical 0 cm−1. This highlights, once again, the ability of the CC(P;Q) corrections δ(P ; Q) to offer a well-balanced description of the lowest singlet and triplet states in methylene, in addition to improving the individual state energies. 67 Table 3.5 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3B1 and A 1A1 states of methylene, as described by the aug-cc-pVTZ basis set, and of the corresponding adiabatic singlet–triplet gaps toward their parent CCSDT values. The geometries of the X 3B1 and A 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corre- sponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the ROHF (X 3B1 state) and RHF (A 1A1 state) reference determinants and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d and f orbitals were employed throughout. Adapted from Ref. [102]. MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 100000 150000 200000 ∞ P a 4.187d 3.948 3.281 2.749 2.428 2.192 1.703 1.133 0.532 0.218 0.076 X 3B1 (P ; Q)a %Tb 0.177e 0.162 0.111 0.072 0.049 0.038 0.018 0.005 0.000 0.000 0.000 −39.080575f 0 1.8 7.1 12.4 16.3 19.0 26.3 39.1 59.5 76.8 88.7 P a 5.918d 5.361 3.908 2.993 2.444 2.093 1.358 0.644 0.171 0.037 0.006 A 1A1 (P ; Q)a %Tb 0.656e 0.549 0.304 0.190 0.106 0.080 0.025 0.004 0.000 0.000 0.000 −39.065411f A 1A1 − X 3B1 (P ; Q)c P c 105e 380d 85 310 42 138 26 53 13 3 9 2 0 0 0 0 0 3.0 11.9 19.7 24.9 28.7 −22 37.7 −76 54.8 −107 76.5 −79 90.7 −40 97.2 −15 3328g a Unless otherwise stated, all energies are reported as errors relative to CCSDT in millihartree. b The %T values are the percentages of triples captured during the i-FCIQMC propagations (the Sz = 1 triply excited determinants of the B1 symmetry in the case of the X 3B1 state and the Sz = 0 triply excited determinants of the A1 symmetry in the case of the A 1A1 state). c Unless otherwise specified, the values of the singlet–triplet gap are reported as errors relative to CCSDT in cm−1. d Equivalent to CCSD. e Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR- CC(2,3),D or CR-CC(2,3)D]. f Total CCSDT energy in hartree. g The CCSDT singlet–triplet gap in cm−1. 68 Table 3.6 The total numbers of walkers, reported as percentages of the total walker popula- tions at 200000 MC iterations, characterizing the i-FCIQMC propagations with δτ = 0.0001 a.u. that were needed to generate the CC(P) and CC(P;Q) results for methylene reported in Table 3.5. Adapted from Ref. [102]. MC Iterations X 3B1 A 1A1 0.01a 0.19 0.51 0.83 1.11 1.35 2.17 4.64 13.13 36.96 100c 0 2000 4000 6000 8000 10000 20000 50000 100000 150000 200000 0.02a 0.39 1.00 1.65 2.17 2.58 4.11 7.69 18.59 43.10 100b a The initial walker population, meaning 1500 walkers on the ROHF (X 3B1 state) and RHF (A 1A1 state) reference determinants. b The total number of walkers at 200000 MC iterations is 6118222. c The total number of walkers at 200000 MC iterations is 14878766. Figure 3.6 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3B1 [panel (a)] and A 1A1 [panel (b)] states of methylene, as described by the aug-cc-pVTZ basis set, and of the corresponding adiabatic singlet–triplet gaps [panel (c)] toward their parent CCSDT values. The geometries of the X 3B1 and A 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i-FCIQMC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [102]. 69 0501001502000246(a)X 3B1CC(P)CC(P;Q)050100150200MC Ite ations (103)0246E o el. to CCSDT (mEh)(b)A 1A1CC(P)CC(P;Q)050100150200MC Ite ations (103)−1000100200300400E o el. to CCSDT (cm−1)(c)ΔES–TCC(P)CC(P;Q) Figure 3.7 Comparison of convergences of the CC(P), CC(P;Q), and the underlying i- FCIQMC calculations toward their respective limits for the X 3B1 and A 1A1 states of the CH2 molecule at their respective geometries optimized in the FCI calculations using the TZ2P basis set are taken from Ref. [181]. 70 3.3.2 (HFH)− Our next example is the linear, D∞h-symmetric, (HFH)− anion, a prototype magnetic system in which unpaired spins of terminal hydrogen atoms couple to singlet and triplet states via a polarizable diamagnetic bridge of F− [196]. The energies of the lowest two electronic states of the (HFH)− system, including the singlet ground state X 1Σ+ g and the first- excited triplet state A 3Σ+ u , and the vertical gap between them, which is proportional to the magnetic exchange coupling constant J and which should approach zero as both H–F bonds are stretched to infinity, were used in the past to test various quantum chemistry approaches [66, 80, 84, 88, 153, 196–199]. Among them were methods developed in the Piecuch group, including CR-CC(2,3) [66, 197], CR-CC(2,4)[84], the DIP-EOMCC approaches with full and active-space treatments of 4h-2p correlations of top of CCSD [80, 153], and the active-orbital- based CC(t;3), CC(t,q;3), and CC(t,q;3,4) hierarchy [84, 88]. Here, we test the alternative to CC(t;3) offered by the semi-stochastic, FCIQMC-driven, CC(P ;Q) algorithm aimed at the CCSDT energetics. As in our previous studies [66, 80, 84, 88, 153, 197], we used the 6-31G(d,p) basis set and several stretches of both H–F bonds, including RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A, where RH-F is the distance between the hydrogen and fluorine nuclei. An accurate computation of the singlet–triplet gap in the (HFH)− system is compli- cated by the fact that, unlike the A 3Σ+ u state, which is weakly correlated and well rep- resented by a single ROHF determinant, its ground-state counterpart X 1Σ+ g displays a substantial MR character that includes a significant contribution from the doubly excited (HOMO)2 → (LUMO)2 determinant, in addition to the RHF configuration. The MR char- acter of the X 1Σ+ g state, which is already noticeable at shorter H–F separations and which substantially strengthens as RH-F increases, can be illustrated by the ratio of the FCI ex- pansion coefficients at the (HOMO)2 → (LUMO)2 and RHF determinants or the equivalent T2 cluster amplitude extracted from FCI, which increases, in absolute value, from 0.38 at RH-F = 1.50 ˚A to 1.17 at RH-F = 4.00 ˚A, when the 6-31G(d,p) basis is employed [66, 197] (the HOMO and LUMO have different symmetries, σg and σu, respectively, so that the 71 HOMO → LUMO T1 amplitude is zero). As a result of all this, it is difficult to balance the lowest two states of the (HFH)− system in a single quantum chemistry calculation, especially when the SRCC framework using the RHF reference determinant for the X 1Σ+ g state and the ROHF reference for the A 3Σ+ u state is employed. Indeed, as shown in Refs. [66, 88, 197], the differences between the energies obtained in the CCSD/6-31G(d,p) computations and their FCI counterparts at RH-F = 1.50 ˚A are 12.674 millihartree for the X 1Σ+ millihartree when the A 3Σ+ g state and only 2.628 u state is considered. The analogous differences at RH-F = 2.00 ˚A are 19.398 and 2.068 millihartree, respectively. The observed large discrepancies between the errors in the CCSD energies for the X 1Σ+ g and A 3Σ+ u states translate into a poor descrip- tion of the singlet–triplet gaps. One can see this by comparing the ∆ES-T values resulting from the RHF/ROHF-based CCSD/6-31G(d,p) computations at RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A with the corresponding FCI data. CCSD/6-31G(d,p) gives −7320, −1838, 1656, 3605, and 230 cm−1, respectively, as opposed to −9525, −4911, −2147, −277, and 0 cm−1 obtained with FCI [66, 88, 197]. If we are to improve the CCSD results within the SRCC framework, we must turn to higher-level theories, such as the CCSDT approach that interests us in this study [84, 88, 198], CCSDTQ [84], or the DIP-EOMCC methodology, especially after incorporating 4h-2p correlations [80, 153]. The CCSDT method is indeed very accurate, reducing the 2205, 3073, 3804, 3882, and 230 cm−1 errors relative to FCI in the ∆ES-T values obtained with CCSD/6-31G(d,p) at RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A to 198, 270, 341, 420, and 58 cm−1, respectively [84, 88, 198]. It also greatly improves the total electronic energies. Indeed, the differences between the CCSDT and FCI energies of the X 1Σ+ g and A 3Σ+ u states in the entire RH-F = 1.50 − 4.00 ˚A region obtained using the 6-31G(d,p) basis set do not exceed 2.276 and 0.389 millihartree, respectively [84, 88]. The analogous differences between the CCSD and FCI energies are as large as 20.546 millihartree for the former state and 2.628 millihartree when the latter state is considered (see Fig. 3.8 for a comparison of CCSD, CCSDT, and FCI energetics throughout the 1.500 ˚A–4.00˚A range). One can reduce the remaining small errors in the CCSDT results even further or practi- 72 cally eliminate them by using CCSDTQ [84] or the DIP-EOMCC approaches with 4h-2p contributions, [80, 153] but the objective of this study is to assess the performance of our semi-stochastic CC(P ;Q) methodology in converging the CCSDT data. The results of our FCIQMC-driven CC(P ;Q)/6-31G(d,p) computations for the X 1Σ+ g and A 3Σ+ u states of the linear (HFH)− system at the H–F distances RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A and the corresponding ∆ES-T values, along with the underlying CC(P ) data, are reported in Tables 3.7–3.9 and Fig. 3.9. In all of our CC(P ) and CC(P ;Q) computations and the underlying i-FCIQMC runs for the D∞h-symmetric (HFH)− system, we used the D2h Abelian subgroup of D∞h. In particular, the i-FCIQMC calculations for the X 1Σ+ A 3Σ+ u states were set up to converge the lowest-energy states of the 1Ag(D2h) and 3B1u g and (D2h) symmetries. As a result, the subsets of triply excited determinants used to construct the P spaces for the subsequent CC(P ) and CC(P ;Q) computations for the X 1Σ+ g state at the various RH-F and τ values considered in this work were defined as the Sz = 0 triples of the Ag (D2h) symmetry captured in the underlying i-FCIQMC propagations. Similarly, the subsets of triply excited determinants used to design the P spaces for the CC(P ) and CC(P ;Q) calculations for the A 3Σ+ u state were the Sz = 1 triples of the B1u (D2h) symmetry extracted from i-FCIQMC. In analogy to all other CC(P ;Q) computations performed in this work, the Q spaces used to determine the δ(P ; Q) corrections to the CC(P ) energies were defined as the remaining triples not captured by the respective i-FCIQMC runs. As shown in Table 3.7, and in line with our earlier CC(P ;Q) work [88] and the above remarks, the CC(P ) energies of the X 1Σ+ g state of (HFH)− obtained at τ = 0, which are identical to those resulting from the conventional CCSD calculations reported in Refs. [66, 88, 197], are characterized by large errors relative to their τ = ∞, i.e., CCSDT, parents. Indeed, the differences between the τ = 0 and τ = ∞ CC(P ) energies for the X 1Σ+ g state increase from 11.412 millihartree at RH-F = 1.50 ˚A to more than 17 millihartree at RH-F = 2.00 and 2.50 ˚A. These differences become smaller at large H–F separations, represented in our calculations by RH-F = 4.00 ˚A, where the D∞h-symmetric (HFH)− system is essentially 73 dissociated into the stretched hydrogen molecule, which has only two electrons, so that CCSD becomes exact, and the closed-shell fluoride ion, which has the electronic structure of the neon atom and which is characterized by small Tn correlations with n > 2, but they remain large when the RH-F values are smaller. This should be contrasted with the small, ∼1–2 millihartree, differences between the τ = 0 and τ = ∞ CC(P ) energies obtained at all values of RH-F for the predominantly SR A 3Σ+ above, and as shown in Table 3.9, this imbalance in the description of the X 1Σ+ u state (see Table 3.8). As already alluded to g and A 3Σ+ u states by the CCSD, i.e., τ = 0 CC(P ), calculations gives rise to large errors in the resulting ∆ES-T values relative to their τ = ∞ (CCSDT) counterparts, which range from 2007 cm−1 to 3462 cm−1 in the RH-F = 1.50–2.50 ˚A region. Once again, these errors become small at large H–F separations, such as RH-F = 4.00 ˚A used in this work, where (HFH)− is more or less equivalent to the stretched H2 and F−, resulting in the nearly degenerate singlet and triplet states and the 172 cm−1 difference between the τ = 0 and τ = ∞ CC(P ) values of ∆ES-T, but at shorter H–F distances they are large and comparable to or even larger than the singlet–triplet gap values provided by CCSDT or FCI. The situation dramatically changes, when the τ = 0 CC(P ) or CCSD energies are cor- rected for T3 correlations with the help of the noniterative correction δ(P ; Q), as in the τ = 0 CC(P ;Q) calculations, which are equivalent to the purely deterministic CR-CC(2,3) runs reported in Refs. [66, 84, 88, 197]. As shown in Tables 3.7–3.9, the τ = 0 CC(P ;Q), i.e., CR-CC(2,3), energies of the X 1Σ+ g and A 3Σ+ u states at the various H–F distances considered in this study and the gaps between them are substantially more accurate than their uncor- rected CC(P ) (i.e., CCSD) counterparts. For example, the CR-CC(2,3) approach reduces the large, more than 17 millihartree, errors in the CCSD energies of the X 1Σ+ g state relative to their CCSDT [τ = ∞ CC(P ) or CC(P ;Q)] parents at RH-F = 2.00 and 2.50 ˚A to ∼1–3 millihartree. We see similarly significant improvements in the CCSD energies of the X 1Σ+ g state by CR-CC(2,3) at other H–F distances, even at the “easiest” RH-F = 4.00 ˚A value, where the triples correction δ(P ; Q) is capable of reducing the already small, 1.907 milli- 74 hartree, difference between the CCSD and CCSDT energies to the much smaller (in absolute value) 0.291 millihartree (see Table 3.7). Consistent with our earlier studies, [66, 84, 88, 197] the CR-CC(2,3) method performs even better when the weakly correlated A 3Σ+ u state is examined, reducing the ∼1–2 millihartree errors in the underlying CCSD energetics relative to our CCSDT target to about 0.2 millihartree (see Table 3.8). As a result of all of these accuracy improvements, the singlet–triplet gap values obtained using CR-CC(2,3) are much closer to their CCSDT parents than their CCSD counterparts, reducing the 2007, 2803, 3462, 3462, and 172 cm−1 errors relative to CCSDT obtained with CCSD at RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A, respectively, by factors ranging from 6 at RH-F = 2.50 ˚A to 72 at RH-F = 1.50 ˚A, but, as shown in Table 3.9 [see, also, Ref. [88], where one can find a comparison of the CCSD, CR-CC(2,3), and CCSDT ∆ES-T data for additional H–F dis- tances], the differences on the order of (−600)–(−300) cm−1 between the CR-CC(2,3) and CCSDT singlet–triplet separations in the intermediate RH-F = 2.00–3.00 ˚A region remain. The question arises if one can refine the CR-CC(2,3) results by enriching the P spaces used in the CC(P ;Q) calculations, which in CR-CC(2,3) consist of only singles and doubles, with the subsets of triply excited determinants identified by i-FCIQMC propagations. As shown in Tables 3.7–3.9 and Fig. 3.9, once the leading triply excited determinants, captured using i-FCIQMC at τ > 0, are included in the respective P spaces and the δ(P ; Q) corrections due to the remaining T3 correlations are added to the energies obtained in the CC(P ) calculations, the resulting CC(P ;Q) values of the X 1Σ+ g and A 3Σ+ u energies and vertical gaps between them display very fast convergence toward their CCSDT counterparts. This is already observed when the i-FCIQMC propagation times are short, engaging tiny walker populations that are orders of magnitude smaller than those required to converge the i-FCIQMC runs, and the fractions of the triply excited determinants captured by i-FCIQMC are small. For example, after as few as 2000 δτ = 0.0001 a.u. MC iterations, where τ is only 0.2 a.u. and where, as shown in Table 3.10, the total walker populations characterizing the underlying i-FCIQMC runs are 0.01–0.11% of the respective numbers of walkers at τ = 20.0 75 a.u. [the termination time for our i-FCIQMC propagations for (HFH)−], the differences between the CC(P ;Q) and CCSDT energies obtained for the strongly correlated X 1Σ+ g state are −0.035 millihartree for RH-F = 1.50 ˚A, −0.056 millihartree for RH-F = 1.75 ˚A, −0.110 millihartree for RH-F = 2.00 ˚A, −0.583 millihartree for RH-F = 2.50 ˚A, and −0.025 millihartree for RH-F = 4.00 ˚A. In spite of using only about 10–30% of all triply excited determinants in the underlying P spaces, the FCIQMC-based CC(P ;Q) energies of the X 1Σ+ g state obtained after 2000 MC iterations reduce the errors relative to CCSDT characterizing the CR-CC(2,3) [i.e., τ = 0 CC(P ;Q)] computations in the RH-F = 1.50–4.00 ˚A region by factors ranging from 5 to 13 (see Table 3.7). In fact, with an exception of RH-F = 2.00 and 2.50 ˚A, they are much more accurate than the results produced by the purely deterministic CC(t;3) analog of the semi-stochastic CC(P ;Q) methodology, reported in Refs. [84, 88]. One can observe even more dramatic improvements over CR-CC(2,3) offered by the FCIQMC- driven CC(P ;Q) approach, when the propagation time τ increases. For example, after 4000 MC iterations, where the i-FCIQMC propagations are still far from being converged (cf. the total walker populations used by our i-FCIQMC runs relative to the termination time τ = 20.0 a.u. in Table 3.10) and the fractions of triples included in the stochastically determined P spaces, which range from 12% at RH-F = 4.00 ˚A to 56% at RH-F = 1.50 ˚A, remain relatively small, the differences between the CC(P ;Q) and CCSDT energies obtained for the X 1Σ+ g state at RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A are −28, −9, −17, −50, and −4 microhartree, respectively, reducing the errors relative to CCSDT that characterize the corresponding CR-CC(2,3) calculations by factors ranging from 12 to 86 [2 to 61 when compared to the CC(t;3) results reported in Refs. [84, 88]]. As shown in Table 3.8, the performance of the FCIQMC-driven CC(P ;Q) approach becomes even more impressive when the A 3Σ+ u state, which has a SR character, is examined. After 2000 δτ = 0.0001 a.u. MC time steps, the errors in the CC(P ;Q) energies relative to their CCSDT parents obtained for the A 3Σ+ u state at RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A are only −40, −24, −38, −29, and −14 microhartree, respectively. After 4000 MC iterations, they become −10, −10, −12, 76 −9, and −2 microhartree, respectively. Once again, these are considerable improvements compared to CR-CC(2,3) and CC(t;3) that both give errors on the order of −0.2 millihartree [84, 88], especially if we realize that the fractions of triples captured by the i-FCIQMC runs after 2000 and 4000 MC iterations are relatively small (5–28% and 5–49%, respectively) and, as shown in Table 3.10, the corresponding numbers of walkers represent only about 1–2% of the total numbers of walkers at τ = 20.0 a.u., where we stopped our i-FCIQMC propagations. As a consequence of the small errors in the CC(P ;Q) total energies characterizing the X 1Σ+ g and A 3Σ+ u states in the early stages of the i-FCIQMC propagations, the resulting singlet–triplet gap values are very accurate as well. This is demonstrated in Table 3.9, where one can see that after 2000 δτ = 0.0001 a.u. MC iterations, which is, as already explained, a very short propagation time engaging tiny walker populations and small fractions of triples, most of the differences between the CC(P ;Q) and CCSDT ∆ES-T values in the RH-F = 1.50–4.00 ˚A region are on the order of a few reciprocal centimeter. The only exception is the semi-stochastic CC(P ;Q) run at RH-F = 2.50 ˚A, where the −122 cm−1 error relative to CCSDT characterizing the singlet–triplet gap obtained after 2000 MC time steps, while representing a five-fold error reduction compared to CR-CC(2,3), is comparable, in magnitude, to the CCSDT value of ∆ES-T. This happens because the CC(P ;Q) energy of the strongly correlated X 1Σ+ g state obtained after 2000 MC iterations at RH-F = 2.50 ˚A differs from its CCSDT parent by −0.583 millihartree, whereas the analogous difference between the CC(P ;Q) and CCSDT energies for its weakly correlated A 3Σ+ u companion is only −29 microhartree. This is not a problem though, since by running i-FCIQMC a little longer and capturing about 20% of all triply excited determinants in the relevant P spaces, as is the case when 4000 δτ = 0.0001 a.u. MC time steps are considered, one reduces the differences between the CC(P ;Q) and CCSDT energies of the X 1Σ+ g and A 3Σ+ u states to −50 and −9 microhartree, respectively (cf. Tables 3.7 and 3.8), so that the 122 cm−1 unsigned error in the CC(P ;Q) value of ∆ES-T relative to CCSDT obtained after 2000 MC iterations 77 decreases to less than 10 cm−1. This is yet another illustration of the ability of the semi- stochastic CC(P ;Q) methodology pursued in this work to balance the more MR singlet and weakly correlated triplet states of biradical systems in a single computation at the fraction of the cost of the parent high-level CC calculations. As shown in Table 3.9, at τ = 0.4 a.u., where the i-FCIQMC propagations are still far from being converged, the FCIQMC-driven CC(P ;Q) calculations recover the CCSDT values of the singlet–triplet gaps in (HFH)− at all H–F distances considered in this study to within a few reciprocal centimeter, reaching a 1–2 cm−1 or better accuracy after 6000 MC iterations. Last, but not least, the results reported in Tables 3.7–3.9 and Fig. 3.9 also demonstrate the remarkable efficiency of the δ(P ; Q) corrections in accelerating the convergence of the CC(P ) energies of the X 1Σ+ g and A 3Σ+ u states and the vertical gaps between them toward CCSDT, independent of the H–F distance considered. Let us, for example, compare the un- corrected CC(P ) and corrected CC(P ;Q) energies of the X 1Σ+ g and A 3Σ+ u states of (HFH)− at the five H–F separations considered in this work obtained after 2000 MC iterations. In the case of the former, more MR, state, the CC(P ;Q) corrections δ(P ; Q) reduce the positive 2.601, 3.998, 3.511, 6.586, and 0.412 millihartree errors relative to CCSDT resulting from the CC(P ) computations at RH-F = 1.50, 1.75, 2.00, 2.50, and 4.00 ˚A to the much smaller neg- ative error values of −0.035, −0.056, −0.110, −0.583, and −0.025 millihartree, respectively. When the latter state, which is characterized by much weaker correlations, is considered, the 0.995, 0.826, 0.834, 0.502, and 0.239 millihartree errors obtained with CC(P ) are reduced to −40, −24, −38, −29, and −14 microhartree, respectively, when the CC(P ;Q) approach is employed. It is interesting to notice that while the errors characterizing the CC(P ) calcu- lations for the A 3Σ+ u state are generally much smaller than their X 1Σ+ g counterparts, and the two states have a substantially different character, the error reductions offered by the CC(P ;Q) corrections δ(P ; Q), by at least one order of magnitude, apply to both states. As already alluded to above, and as shown in Table 3.9 and Fig. 3.9 (e) and (f), where we examine the convergence of the CC(P ) and CC(P ;Q) ∆ES-T values toward their CCSDT 78 parents, the noniterative corrections δ(P ; Q) are also very effective in improving the balance in the description of the X 1Σ+ g and A 3Σ+ u states by the CC(P ) approach and smoothing the convergence of the resulting singlet–triplet gaps toward their CCSDT limits. This can be illustrated by comparing the behavior of the error values relative to CCSDT character- izing the CC(P ) calculations of ∆ES-T at RH-F = 2.00 ˚A with their CC(P ;Q) counterparts, shown in Table 3.9. In the former case, the 3462 cm−1 error at τ = 0 decreases, in absolute value, to 8 cm−1 at τ = 0.8 a.u. (8000 MC iterations), to increase to 17 cm−1 at τ = 2.0 a.u. (20000 MC iterations), to decrease again to a numerical 0 cm−1 at τ = 20.0 a.u. (200000 MC iterations). Once the CC(P ) energies of the X 1Σ+ g and A 3Σ+ u states are corrected using the δ(P ; Q) corrections, the unsigned errors in the resulting CC(P ;Q) values of ∆ES-T relative to their CCSDT parent monotonically and rapidly decrease, from 282 cm−1 at τ = 0 to a numerical 0 cm−1 at τ ≥ 0.8 a.u. It is clear from Tables 3.7–3.9 and Fig. 3.9 that while both the CC(P ) and CC(P ;Q) energies converge to the parent CCSDT limit, the latter energies and the gaps between them converge to CCSDT a lot faster. 79 Table 3.7 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1Σ+ g state of (HFH)−, as described by the 6-31G(d,p) basis set, at selected H–F distances RH-F toward their parent CCSDT values. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the RHF reference determinant and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. MC Iters. 0 2000 4000 6000 8000 10000 20000 50000 100000 150000 200000 ∞ RH-F = 1.50 ˚A P a (P ; Q)a %Tb 0 11.412c −0.343d 2.601 −0.035 34.2 0.843 −0.028 56.1 0.595 −0.004 63.9 0.225 −0.003 68.6 0.258 −0.003 70.9 77.2 0.000 0.112 88.4 0.000 0.017 97.7 0.000 0.002 99.5 0.000 0.000 99.9 0.000 0.000 −100.588130e RH-F = 1.75 ˚A P a (P ; Q)a %Tb RH-F = 2.00 ˚A P a (P ; Q)a %Tb RH-F = 2.50 ˚A P a (P ; Q)a %Tb P a RH-F = 4.00 ˚A (P ; Q)a %Tb 0 0 17.453c −1.455d 14.738c −0.686d 3.511 −0.110 22.6 3.998 −0.056 30.5 1.979 −0.017 40.5 1.078 −0.009 49.6 0.432 −0.010 46.7 0.434 −0.003 58.1 0.187 −0.003 50.2 0.477 −0.007 61.4 0.136 −0.003 54.5 0.161 −0.002 63.3 0.079 −0.002 61.1 0.056 −0.001 71.0 77.5 0.000 0.005 85.8 0.000 0.019 94.4 0.000 0.001 96.3 0.000 0.001 99.2 0.000 99.4 0.000 0.000 0.000 99.9 0.000 100.0 0.000 0.000 0.000 −100.576056e −100.561110e 17.051c −2.800d 1.907c −0.291d 0 0 7.6 6.586 −0.583 15.2 0.412 −0.025 0.973 −0.050 25.6 0.141 −0.004 11.7 0.459 −0.012 30.2 0.076 −0.003 13.2 0.225 −0.003 33.9 0.037 −0.001 14.4 0.167 35.4 0.025 −0.001 15.3 0.042 −0.001 41.8 0.026 −0.001 19.0 58.8 0.002 −0.001 28.6 0.009 54.8 81.8 0.000 0.000 73.7 94.1 0.000 0.000 86.9 99.2 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 −100.525901e −100.539783e 0.000 a Unless otherwise stated, all energies are reported as errors relative to CCSDT in millihartree. b The %T values are the percentages of triples captured during the i-FCIQMC propagations [the Sz = 0 triply excited determinants of the Ag (D2h) symmetry]. c Equivalent to CCSD. d Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR-CC(2,3),D or CR-CC(2,3)D]. e Total CCSDT energy in hartree. 80 Table 3.8 Convergence of the CC(P ) and CC(P ;Q) energies of the A 3Σ+ u state of (HFH)−, as described by the 6-31G(d,p) basis set, at selected H–F distances RH-F toward their parent CCSDT values. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the ROHF reference determinant and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. RH-F = 1.50 ˚A RH-F = 1.75 ˚A RH-F = 2.00 ˚A RH-F = 2.50 ˚A RH-F = 4.00 ˚A MC Iters. P a (P ; Q)a %Tb P a (P ; Q)a %Tb P a (P ; Q)a %Tb P a (P ; Q)a %Tb P a (P ; Q)a %Tb 0 2000 4000 6000 8000 10000 20000 50000 100000 150000 200000 ∞ 0 0 0 0 1.967c −0.181d 1.277c −0.167d 2.268c −0.217d 1.123c −0.180d 1.678c −0.172d 0.995 −0.040 27.8 0.826 −0.024 24.2 0.834 −0.038 19.1 0.502 −0.029 10.7 0.239 −0.014 0.456 −0.010 49.4 0.477 −0.010 41.7 0.475 −0.012 33.7 0.236 −0.009 17.2 0.079 −0.002 0.338 −0.005 56.4 0.266 −0.001 50.5 0.321 −0.005 41.2 0.174 −0.003 21.5 0.070 −0.003 0.290 −0.003 60.1 0.254 −0.003 54.2 0.225 −0.003 44.7 0.195 −0.006 23.9 0.064 −0.002 0.271 −0.003 61.1 0.267 −0.004 56.6 0.201 −0.002 45.7 0.064 −0.003 25.0 0.056 −0.002 0.201 −0.002 67.9 0.151 −0.001 62.1 0.157 −0.002 52.2 0.078 −0.003 28.6 0.025 −0.001 0.000 0.082 0.000 0.021 0.000 0.007 0.000 0.001 −100.526164e 76.3 0.069 −0.001 66.1 0.049 −0.001 37.4 0.012 52.9 0.002 82.8 0.014 89.4 0.015 68.6 0.001 92.9 0.002 95.8 0.003 81.8 0.000 97.1 0.000 98.4 0.001 0.000 0.000 0.000 0.000 −100.545633e 0.000 0.000 0.000 0.000 −100.554908e 0.000 0.000 0.000 −100.552882e 0.000 0.000 0.000 −100.540435e 80.0 0.056 91.8 0.016 96.7 0.003 98.8 0.001 0 4.5 5.4 5.9 6.0 6.4 7.4 8.4 11.7 16.8 23.8 a Unless otherwise stated, all energies are reported as errors relative to CCSDT in millihartree. b The %T values are the percentages of triples captured during the i-FCIQMC propagations [the Sz = 1 triply excited determinants of the B1u (D2h) symmetry]. c Equivalent to CCSD. d Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR-CC(2,3),D or CR-CC(2,3)D]. e Total CCSDT energy in hartree. 81 Table 3.9 Convergence of the CC(P ) and CC(P ;Q) singlet–triplet gaps of (HFH)−, as described by the 6-31G(d,p) basis set, at selected H–F distances RH-F toward their parent CCSDT values. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the RHF (X 1Σ+ g state) and ROHF (A 3Σ+ u state) reference determinants and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbital was kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. RH-F = 1.50 ˚A RH-F = 1.75 ˚A RH-F = 2.00 ˚A RH-F = 2.50 ˚A RH-F = 4.00 ˚A MC Iters. P a (P ; Q)a P a 2007b −28c 0 353 2000 85 4000 6000 56 8000 −14 10000 −3 20000 −20 50000 −14 100000 −4 150000 −2 200000 0 ∞ 1 −4 0 0 0 0 0 0 0 0 −9327d (P ; Q)a P a (P ; Q)a P a (P ; Q)a (P ; Q)a P a 2803b −111c 3462b −282c 3462b −578c 172b −24c −2 696 0 132 0 37 0 49 0 −23 0 −21 0 −8 0 −3 0 −1 0 0 58d 1335 −122 −9 162 −2 62 1 7 0 23 1 −8 0 −9 0 −3 0 0 0 0 143d −16 −1 −1 0 0 0 0 0 0 0 −1806d −7 0 0 −1 0 0 0 0 0 0 −4641d 588 330 24 −8 −14 −17 −14 −3 −1 0 38 14 1 −6 −7 0 −2 −1 0 0 a Unless otherwise stated, all singlet–triplet gaps are reported as errors relative to CCSDT in cm−1. b Equivalent to CCSD. c Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR- CC(2,3),D or CR-CC(2,3)D]. d The CCSDT singlet–triplet gap in cm−1. 82 Table 3.10 The total numbers of walkers, reported as percentages of the total walker popula- tions at 200000 MC iterations, characterizing the i-FCIQMC propagations with δτ = 0.0001 g and A 3Σ+ a.u. that were needed to generate the CC(P) and CC(P;Q) results for the X 1Σ+ states of (HFH)− reported in Tables 3.7 and 3.8. Adapted from Ref. [102]. u MC Iters. X 1Σ+ 0.02a 0 0.11 2000 0.20 4000 0.27 6000 0.32 8000 0.37 10000 0.55 20000 1.42 50000 100000 6.21 150000 25.44 100b 200000 RH-F = 1.50 ˚A RH-F = 1.75 ˚A RH-F = 2.00 ˚A RH-F = 2.50 ˚A RH-F = 4.00 ˚A g A 3Σ+ 0.13a 0.73 1.23 1.59 1.77 1.91 2.46 4.81 13.61 37.71 100g g A 3Σ+ u X 1Σ+ u 0.59a 0.00a 1.74 0.01 2.45 0.02 2.97 0.02 3.20 0.03 3.29 0.03 4.09 0.06 6.83 0.21 1.61 16.64 12.45 41.09 100k 100j u X 1Σ+ 0.01a 0.03 0.05 0.07 0.08 0.09 0.15 0.47 2.88 17.11 100h u X 1Σ+ 0.01a 0.08 0.15 0.19 0.23 0.26 0.40 1.10 5.16 23.37 100d g A 3Σ+ 0.10a 0.64 1.14 1.46 1.66 1.79 2.38 4.59 12.87 36.28 100e g A 3Σ+ 0.24a 0.95 1.54 1.86 2.12 2.29 3.03 5.58 15.13 38.88 100i g A 3Σ+ 0.09a 0.59 1.08 1.36 1.55 1.68 2.19 4.19 12.24 35.16 100c u X 1Σ+ 0.01a 0.06 0.10 0.13 0.16 0.18 0.28 0.81 4.25 21.01 100f g state) and ROHF (A 3Σ+ u state) a The initial walker population, meaning 1500 walkers on the RHF (X 1Σ+ reference determinants. b The total number of walkers at 200000 MC iterations is 9865967. c The total number of walkers at 200000 MC iterations is 1749699. d The total number of walkers at 200000 MC iterations is 12468454. e The total number of walkers at 200000 MC iterations is 1431689. f The total number of walkers at 200000 MC iterations is 15510033. g The total number of walkers at 200000 MC iterations is 1123676. h The total number of walkers at 200000 MC iterations is 24265207. i The total number of walkers at 200000 MC iterations is 632102. j The total number of walkers at 200000 MC iterations is 50189301. k The total number of walkers at 200000 MC iterations is 254390. 83 Figure 3.8 Total electronic energies of the X 1Σ+ g (open circles and solid line) and A 3Σ+ u (filled circles and dotted line) states of (HFH)− with increase in the H–F distance, from 1.5 ˚A to 4.0 ˚A, obtained from the FCI (red circles), CCSD (blue circles), and CCSDT (green circles) methods. Recreated from the data reported in Refs. [66, 84, 88, 197]. 84 H-F distance increases Figure 3.9 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1Σ+ g [panels (a) and (b)] and A 3Σ+ u [panels (c) and (d)] states of (HFH)−, as described by the 6-31G(d,p) basis set, and of the corresponding singlet–triplet gaps [panels (e) and (f)] toward their parent CCSDT values. The H–F distances RH-F used are 1.50 ˚A, 1.75 ˚A, 2.00 ˚A, 2.50 ˚A, and 4.00 ˚A. The P spaces consisted of all singles and doubles and subsets of triples identified during i-FCIQMC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [102]. 85 05010015020001234Error rel. to CCSDT (mEh)(a)X 1Σ+g, CC(P)1.50 Å1.75 Å2.00 Å2.50 Å4.00 Å05010015020001234(b)X 1(+g, CC(P;Q)1.50 Å1.75 Å2.00 Å2.50 Å4.00 Å05010015020001234Error rel. to CCSDT (mEh)(c)A 3Σ+u, CC(P)1.50 Å1.75 Å2.00 Å2.50 Å4.00 Å05010015020001234(d)A 3(+u, CC(P;Q)1.50 Å1.75 Å2.00 Å2.50 Å4.00 Å050100150200MC Iterations (103)010203040Error rel. to CCSDT (c −1)(e)ΔES–T,ΔCC(P)1.50ΔÅ1.75ΔÅ2.00ΔÅ2.50ΔÅ4.00ΔÅ050100150200MCΔIterationsΔ(103)010203040(f)ΔES–T,ΔCC(P;Q)1.50ΔÅ1.75ΔÅ2.00ΔÅ2.50ΔÅ4.00ΔÅ 3.3.3 Cyclobutadiene and cyclopentadienyl cation We now proceed to the examination of the performance of the semi-stochastic CC(P ;Q) algorithm in calculations involving medium-sized organic biradicals, starting from two pro- totypical anti-aromatic systems, cyclobutadiene and cyclopentadienyl cation, both described using the cc-pVDZ basis set. As in the rest of this chapter, we are mainly interested in how efficient the CIQMC-driven CC(P ;Q) methodology is in recovering the CCSDT energies of the lowest singlet and triplet states and gaps between them. In the case of cyclobutadiene and cyclopentadienyl cation discussed in this subsection, we focus on examining vertical singlet–triplet gaps. We begin with the FCIQMC-driven CC(P ;Q) calculations for cyclobutadiene, in which we adopted the D4h-symmetric geometry that represents the transition state for the au- tomerization of cyclobutadiene proceeding on the lowest singlet potential, optimized with the MR average-quadratic CC (MR-AQCC) approach [200, 201] using the cc-pVDZ basis in Ref. [202]. We employed this geometry for two reasons. One of them is the fact that we used the same geometry in our earlier CIQMC- and CCMC-based [99, 101], CIPSI-driven [120], and active-orbital-based [87] CC(P ;Q) calculations for cyclobutadiene, when examining its automerization. Because of this, we could verify the correctness of our FCIQMC-driven CC(P ;Q) calculations for the lowest-energy singlet state, which is also the ground state of the system. Another is the observation that the D4h-symmetric transition-state structure char- acterizing the automerization of cyclobutadiene is practically identical to the D4h-symmetric minimum on the lowest triplet surface. Indeed, the MR-AQCC/cc-pVDZ C–C and C–H bond lengths defining the transition state on the ground-state singlet potential differ from those characterizing the triplet minimum optimized using unrestricted CCSD (UCCSD) in Ref. [203] by less than 0.009 and 0.001 ˚A, respectively. At the D4h-symmetric geometry used in our calculations, cyclobutadiene is characterized by the delocalization of four π electrons over four π MOs, which gives rise to the close-lying singlet and triplet states that require a highly accurate treatment of electron correlation 86 effects if we are to obtain a well-balanced description of the two states and the small energy separation between them. One can understand this by examining the valence π network of the D4h-symmetric cyclobutadiene species, which consists of the doubly occupied nondegenerate a2u orbital, the doubly degenerate eg level, in which each component MO is occupied by a single electron, and the nondegenerate b1u orbital, which in the zeroth-order description of the lowest singlet and triplet states remains empty. The two valence electrons in the degenerate eg shell can couple to a singlet or a triplet, resulting in the open-shell singlet ground state, X 1B1g, which has a substantial MR character, and the first excited triplet state, A 3A2g, which is predominantly SR in nature (see Fig. 3.10 for an illustration of the π MO network of cyclobutadiene along with the triplet electronic configuration). In order to balance the substantial nondynamical correlation effects, needed for an accurate description of the low- spin X 1B1g state, with the dynamical correlations dominating its high-spin triplet A 3A2g companion within a conventional, particle-conserving, SRCC framework and produce reliable ∆ES-T values for cyclobutadiene, which could compete with the high-accuracy ab initio data reported in Refs. [154, 155, 202–208], one has to consider robust treatments of the connected triply excited clusters, such as that offered by CCSDT [204, 208]. Indeed, full CCSDT, which is the target of this investigation, produces high-quality results for the lowest singlet and triplet states of the D4h-symmetric cyclobutadiene system and the energy separation between them. For example, the ∆ES-T value obtained in the CCSDT/cc-pVDZ calculations at the transition-state geometry used in the present study, of −4.8 kcal/mol, is practically identical to the results of the state-of-the-art DEA-EOMCC computations including the high-rank 4p-2h correlations on top of CCSD, reported in Refs. [154, 155], which give −5.0 kcal/mol when the cc-pVDZ basis set is employed (for similar recent observations regarding the reliability of full CCSDT in generating virtually exact singlet–triplet gap values for cyclobutadiene, see Ref. [208]). It is, therefore, interesting to explore if the semi-stochastic CC(P ;Q) methodology investigated in this work is capable of converging the CCSDT results for the X 1B1g and A 3A2g states of cyclobutadiene and vertical gap between them out of the 87 early stages of CIQMC propagations. The results of our FCIQMC-driven CC(P ) and CC(P ;Q) computations for cyclobutadi- ene are summarized in Table 3.11 and Fig. 3.11. In all of our calculations, starting with the stochastic i-FCIQMC steps and ending with the deterministic CC(P ;Q) and CCSDT runs, we used the D2h Abelian subgroup of the D4h point group characterizing the cyclobutadiene’s geometry adopted in this work. Consequently, the i-FCIQMC propagations for the X 1B1g and A 3A2g states were set up to converge the lowest states of the 1Ag (D2h) and 3B1g (D2h) symmetries. Consistent with the CC(P ) and CC(P ;Q) runs that follow the i-FCIQMC steps and the accompanying CCSD, CR-CC(2,3), and CCSDT computations, the reference de- terminants used to initiate our i-FCIQMC propagations were the closed-shell, D2h-adapted, RHF function obtained by placing two electrons on one of the eg valence orbitals for the lowest-energy 1Ag (D2h) state and the high-spin ROHF determinant, adapted to D2h as well, for the lowest 3B1g (D2h) state. As a result, the lists of triply excited determinants extracted from the i-FCIQMC runs at the various propagation times τ > 0, needed to define the P spaces for the CC(P ) and CC(P ;Q) computations, consisted of the Sz = 0 triples of the Ag (D2h) symmetry for the X 1B1g state and the Sz = 1 triples of the B1g (D2h) symmetry in the case of the A 3A2g state. Given our interest in converging the CCSDT energetics, the Q spaces used to construct the δ(P ; Q) corrections consisted of the remaining triply excited determinants, absent in the i-FCIQMC wave functions of the X 1B1g and A 3A2g states at a given τ . The results shown in Table 3.11 and Fig. 3.11 display several similarities with the previ- ously discussed methylene and (HFH)− cases. One cannot, for example, obtain an accurate description of the more MR singlet ground state and the energy separation between the X 1B1g and A 3A2g states without incorporating the leading triply excited determinants in the P space. Indeed, when the P space consists of only singly and doubly excited determi- nants, as in the τ = 0 CC(P ) (i.e., CCSD) and CC(P ;Q) [i.e., CR-CC(2,3)] calculations, one ends up with the enormous errors in the energies of the X 1B1g state relative to their 88 CCSDT parent, which are 47.979 millihartree in the former case and 14.636 millihartree when the latter computation is considered. The τ = 0 CC(P ;Q) energy of the A 3A2g state is a lot more accurate, reducing the large, 23.884 millihartree, error relative to CCSDT ob- tained in the underlying CC(P ) calculation to −60 microhartree, but this does not help too much. The corresponding CR-CC(2,3) triples correction to CCSD, which neglects the coupling of the low-order T1 and T2 clusters with their higher-order T3 counterpart, is inca- pable of offering a balanced description of the X 1B1g and A 3A2g states, so that the resulting singlet–triplet gap is very poor. The 9.2 kcal/mol difference between the ∆ES-T values ob- tained in the τ = 0 CC(P ;Q) or CR-CC(2,3) and CCSDT calculations is so large that the X 1B1g − A 3A2g separation predicted by CR-CC(2,3) has a wrong sign compared to its −4.8 kcal/mol CCSDT counterpart, while being nearly identical in magnitude. This difference becomes even larger when the uncorrected τ = 0 CC(P ), meaning CCSD, calculations are considered (15.1 kcal/mol). As shown in Table 3.11 and Fig. 3.11, the situation dramatically changes when the P spaces used in the CC(P ) and CC(P ;Q) calculations are enriched with the subsets of triply excited determinants captured by the i-FCIQMC propagations. The convergence of the CC(P ;Q) energies of the X 1B1g and A 3A2g states, especially the former ones, and the vertical separations between them is particularly impressive. For example, after as few as 6000 δτ = 0.0001 a.u. MC time steps and i-FCIQMC capturing less than 30% of all triples in the P space, where, as demonstrated in Table 3.12, the walker population characterizing the i-FCIQMC run for the X 1B1g state is only 0.02% of the total number of walkers at τ = 8.0 a.u. (the termination time for our i-FCIQMC propagations for cyclobutadiene), the CC(P ;Q) approach reduces the 14.636 millihartree difference between the CR-CC(2,3) and CCSDT energies of the strongly correlated singlet ground state to 2.223 millihartree. While the CR-CC(2,3) description of the A 3A2g state, which has a largely SR character, is already excellent, the CC(P ;Q) calculation performed after 6000 MC iterations, which uses only 26% of triples in the P space and a tiny walker population that amounts to 0.04% of all walkers at 89 τ = 8.0 a.u. in the underlying i-FCIQMC propagation, improves it too, reducing the small, 60 microhartree, unsigned difference between the CR-CC(2,3) and CCSDT energies to an even smaller 51 microhartree. As a consequence of the above improvements, especially for the X 1B1g state, the error relative to CCSDT characterizing the ∆ES-T value obtained in the FCIQMC-driven CC(P ;Q) calculations after 6000 δτ = 0.0001 a.u. MC time steps, where the underlying i-FCIQMC propagations are still in their early stages, is only 1.4 kcal/mol, as opposed to 9.2 kcal/mol obtained at τ = 0 with CR-CC(2,3). The resulting X 1B1g − A 3A2g energy separation, of −3.4 kcal/mol, has not only the correct sign, but is also very close to the −4.8 kcal/mol value obtained with CCSDT. If we wait a little longer, by executing the extra 2000 MC iterations, so that the i-FCIQMC propagations can capture 34%–39% of all triply excited determinants, we can reduce the already small 2.223 millihartree, 51 microhartree, and 1.4 kcal/mol errors in the CC(P ;Q) energies of the X 1B1g and A 3A2g states and separation between them relative to CCSDT, obtained after 6000 MC time steps, to 0.835 millihartree, 31 microhartree, and 0.5 kcal/mol, respectively. It is clear from Table 3.11 and Fig. 3.11 that the convergence of the semi-stochastic CC(P ;Q) results for the lowest- energy singlet and triplet states of cyclobutadiene, especially the X 1B1g energies and the X 1B1g − A 3A2g gap values, which the τ = 0 CC(P ;Q) or CR-CC(2,3) calculations describe poorly, toward CCSDT is very fast, even when the underlying i-FCIQMC propagations are far from convergence. It is also apparent from our calculations that the noniterative corrections δ(P ; Q) play a significant role in accelerating convergence of the corresponding CC(P ) energetics toward CCSDT. As shown, for example, in Table 3.11, the relatively large differences between the uncorrected CC(P ) energies of the X 1B1g and A 3A2g states and vertical gap between them obtained at τ = 0.8 a.u., i.e., after 8000 δτ = 0.0001 a.u. MC iterations, and the corresponding CCSDT data, which exceed 11 and 7 millihartree and 3 kcal/mol, respectively, are reduced to 0.835 millihartree, 31 microhartree, and 0.5 kcal/mol, when the CC(P ;Q) approach is employed. We can see similar improvements in the CC(P ) energies at other τ values. 90 Table 3.11 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1B1g and A 3A2g states of cyclobutadiene, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet–triplet gaps toward their parent CCSDT values. All calculations were performed at the D4h-symmetric transition-state geometry of the X 1B1g state optimized in the MR- AQCC calculations reported in Ref. [202]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited deter- minants not captured by the corresponding i-FCIQMC runs. The i-FCIQMC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walkers on the RHF (X 1B1g state) and ROHF (A 3A2g state) reference determinants and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the four lowest core orbitals were kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 80000 ∞ P a 47.979d 40.663 27.235 17.188 11.207 8.299 2.030 0.049 0.001 X 1B1g (P ; Q)a %Tb 14.636e 11.059 5.921 2.223 0.835 0.429 0.013 0.000 0.000 0 3.5 16.6 29.5 39.2 46.6 70.0 96.9 99.9 P a 23.884d 21.004 14.317 10.016 7.463 5.865 2.461 0.166 0.009 A 3A2g (P ; Q)a %Tb -0.060e 0.031 0.068 0.051 0.031 0.020 0.005 0.000 0.000 0 3.0 14.2 25.5 34.3 41.0 62.8 94.2 99.6 −154.232002f −154.224380f X 1B1g − A 3A2g (P ; Q)c P c 9.2e 15.1d 6.9 12.3 3.7 8.1 1.4 4.6 0.5 3.3 0.3 1.5 0.0 -0.3 0.0 -0.1 0.0 0.0 −4.8g a Unless otherwise stated, all energies are reported as errors relative to CCSDT in millihartree. b The %T values are the percentages of triples captured during the i-FCIQMC propagations [the Sz = 0 triply excited determinants of the Ag(D2h) symmetry in the case of the X 1B1g state and the Sz = 1 triply excited determinants of the B1g (D2h) symmetry in the case of the A 3A2g state]. c Unless otherwise specified, the values of the singlet–triplet gaps are reported as errors relative to CCSDT in kcal/mol. d Equivalent to CCSD. e Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR- CC(2,3),D or CR-CC(2,3)D]. f Total CCSDT energy in hartree. g The CCSDT singlet–triplet gap in kcal/mol. 91 Table 3.12 The total numbers of walkers, reported as percentages of the total walker popu- lations at 80000 MC iterations, characterizing the i-FCIQMC propagations with δτ = 0.0001 a.u. that were needed to generate the CC(P ) and CC(P ;Q) results for cyclobutadiene re- ported in Table 3.11. Adapted from Ref. [102]. MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 80000 X 1B1g 0.00a 0.00 0.01 0.02 0.03 0.05 0.16 3.81 100b A 3A2g 0.00a 0.00 0.02 0.04 0.07 0.09 0.28 4.93 100c a The initial walker population, meaning 1500 walkers on the RHF (X 1B1g state) and ROHF (A 3A2g state) reference determinants. b The total number of walkers at 80000 MC iterations is 8457504823. c The total number of walkers at 80000 MC iterations is 4067481034. Figure 3.10 π molecular orbital network of the cyclobutadiene molecule, obtained at the HF/cc-pVDZ level, at the D4h-symmetric transition-state geometry of the X 1B1g state opti- mized in the MR-AQCC calculations in Ref. [202]. The orbital irreducible representations in the D4h symmetry are shown in black and the corresponding labels in the C2v symmetry are shown in the parenthesis in orange. This electronic configuration refers to the triplet state. 92 Figure 3.11 Convergence of the CC(P ) and CC(P ;Q) energies of the X 1B1g [panel (a)] and A 3A2g [panel (b)] states of cyclobutadiene, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet–triplet gaps [panel (c)] toward their parent CCSDT values. All calculations were performed at the D4h-symmetric transition-state geometry of the X 1B1g state optimized in the MR-AQCC calculations in Ref. [202]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i-FCIQMC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-FCIQMC. Adapted from Ref. [102]. Most of the observations regarding the performance of the semi-stochastic CC(P ;Q) methodology and its CC(P ) counterpart remain valid when the larger cyclopentadienyl cation, which is also the largest molecular system considered in our CC(P )/CC(P ;Q) work to date, is examined. Following the previous DEA-EOMCC studies of cyclopentadienyl cation from the Piecuch group [154, 155], where the effect of high-order 4p-2h correlations on the singlet–triplet gap was investigated, we used the D5h-symmetric geometry corresponding to a minimum on the lowest triplet surface obtained in the UCCSD/cc-pVDZ optimization in Ref. [203]. At this geometry, cyclopentadienyl cation is characterized by the delocalization of four π electrons over five π MOs, resulting in the doubly occupied nondegenerate a′′ 2 or- bital, the doubly degenerate e′′ 1 shell, in which each component MO is occupied by a single electron, and the doubly degenerate e′′ 2 shell, which in the zeroth-order description of the lowest-energy singlet and triplet states remains empty. In analogy to the previously discussed cyclobutadiene system, the two electrons in the degenerate e′′ 1 MOs can couple to a singlet 93 02040608001020304050(a)X 1B1gCC(P)CC(P;Q)020406080MC Iterations (103)01020304050Error re . to CCSDT (mEh)(b)A 3A2gCC(P)CC(P;Q)020406080MC Iterations (103)03691215Error re . to CCSDT (kca /mo )(c)ΔES–TCC(P)CC(P;Q) or triplet, but compared to cyclobutadiene, where the lowest-energy singlet state is also a ground state, the state ordering in cyclopentadienyl cation is reversed, so that the lowest triplet, designated as X 3A′ 2, is the ground state and the lowest-energy singlet, denoted as A 1E′ 2, is the first excited state (see Fig. 3.12 for an illustration of the π MOs). Similar to all other examples considered in this work, in order to obtain a well-balanced description of the X 3A′ 2 state, which has a SR character dominated by dynamical correlations, and its A 1E′ 2 companion, which is an open-shell singlet characterized by significant MR correlations, and obtain an accurate value of ∆ES-T within a conventional SRCC framework, one must turn to higher-level theories that can offer a robust treatment of Tn clusters with n > 2. Otherwise, as shown in Ref. [203], and as confirmed in our calculations, the results can be very poor. For example, the A 1E′ 2 − X 3A′ 2 separation in cyclopentadienyl cation result- ing from the restricted CCSD calculations using the cc-pVDZ basis, which are equivalent to our τ = 0 CC(P ) computations, is about 23 kcal/mol. This is in large disagreement with the most accurate ab initio calculations of the singlet–triplet gap in cyclopentadienyl cation performed to date using the DEA-EOMCC formalism including 3p-1h as well as 4p-2h correlations on top of the CCSD treatment of the underlying closed-shell core, which give about 14 kcal/mol when the cc-pVDZ basis set is employed [154, 155] (for the examples of other high-level SRCC and MRCC calculations of the singlet–triplet gap in cyclopentadienyl cation, see Ref. [203]; Ref. [155] also provides the well-converged MR perturbation theory data, which agree with the state-of-the-art DEA-EOMCC computations reported in Refs. [154, 155]). The restricted CCSDT approach, which is the target SRCC method in this study, provides a much better description, reducing the approximately 9 kcal/mol error relative to the most accurate DEA-EOMCC calculations with up to 4p-2h excitations reported in Refs. [154, 155] obtained with restricted CCSD to less than 3 kcal/mol, when the cc-pVDZ basis set is employed. It would certainly be interesting to examine if the inclusion of higher– than–triply excited clusters, such as T4, could further improve the CCSDT description of the singlet–triplet gap in cyclopentadienyl cation, but in this work we focus on the ability of the 94 semi-stochastic, CIQMC-based, CC(P ;Q) methodology to improve the CR-CC(2,3) ∆ES-T values and converge the results of CCSDT computations. We hope to return to the topic of the role of T4 clusters in describing the singlet–triplet gap in cyclopentadienyl cation in one of our future studies. It may be worth pointing out that the A 1E′ 2 − X 3A′ 2 gap obtained in the restricted CCSDT/cc-pVDZ calculations, which give ∆ES-T = 16.7 kcal/mol, is in very good agreement with the 16.1 kcal/mol resulting from the DEA-EOMCC/cc-pVDZ computations truncated at 3p-1h excitations [154, 155]. The results of our CIQMC-driven CC(P ) and CC(P ;Q) computations for cyclopentadi- enyl cation are reported in Table 3.13 and Fig. 3.13. As already alluded to above, to reduce the computational costs of the CIQMC propagations preceding the CC(P ) and CC(P ;Q) steps, especially in the later stages of the CIQMC runs that are included in Table 3.13 and Fig. 3.13 for the completeness of our presentation, we replaced the i-FCIQMC algorithm, which we exploited in our calculations for methylene, (HFH)−, and cyclobutadiene, by its truncated i-CISDTQ-MC counterpart. It has been established in Ref. [101] that the replace- ment of i-FCIQMC by i-CISDTQ-MC, when identifying the leading higher–than–doubly excited determinants for the inclusion in the P spaces used in the semi-stochastic CC(P ) and CC(P ;Q) runs, has virtually no effect on the rate at which these runs converge the parent SRCC energetics. In analogy to cyclobutadiene, all of our i-CISDTQ-MC, semi-stochastic CC(P ) and CC(P ;Q), and deterministic CCSD, CR-CC(2,3), and CCSDT computations utilized the largest Abelian subgroup of the D5h point group characterizing the cyclopenta- dienyl cation’s structure examined in the present study, which is C2v. This means that in setting up our calculations for the X 3A′ 2 state, we treated it as the lowest state of the 3B2 (C2v) symmetry, whereas the doubly degenerate A 1E′ 2 state was represented by its 1A1 (C2v) component. Similar to cyclobutadiene, and to remain consistent with the CC(P ), CC(P ;Q), and other SRCC runs for cyclopentadienyl cation carried out in this study, the reference determinant used to initiate the i-CISDTQ-MC propagation for the lowest-energy 3B2 (C2v) state was the triplet ROHF determinant. In the case of the 1A1 (C2v) component of the 95 A 1E′ 2 state, we used the RHF determinant obtained by pairing the two valence electrons in one of the e′′ 1 MOs to initiate the corresponding i-CISDTQ-MC run. Consistent with the above description, the subsets of triply excited determinants used to construct the P spaces for the semi-stochastic CC(P ) and CC(P ;Q) computations for the X 3A′ 2 state were the Sz = 1 triples of the B2 (C2v) symmetry captured by i-CISDTQ-MC. In the case of the A 1E′ 2 state, represented, as explained above, by its 1A1 (C2v) component, we used the Sz = 0 triples of the A1 (C2v) symmetry identified by the i-CISDTQ-MC propagation set up to converge the lowest A1 (C2v) state. As usual, the corresponding Q spaces were spanned by the remaining triply excited determinants that were not captured by the i-CISDTQ-MC runs when the lists of P -space triples were created. Our calculations for cyclopentadienyl cation, summarized in Table 3.13 and Fig. 3.13, demonstrate that the CC(P ;Q) energies of the X 3A′ 2 and A 1E′ 2 states and vertical gaps between them display fast convergence toward the respective CCSDT values with the propa- gation time τ . This is particularly apparent in the case of the CC(P ;Q) energies of the more MR A 1E′ 2 state and the A 1E′ 2 − X 3A′ 2 separation, which cannot be accurately described if the underlying P spaces contain only singly and doubly excited determinants. Indeed, the CR-CC(2,3) energy of the A 1E′ 2 state, which is equivalent to the τ = 0 CC(P ;Q) value, is much more accurate than the result of the associated CC(P ) or CCSD calculation, which produces the enormous error relative to CCSDT exceeding 38 millihartree, but the substan- tial, > 6 millihartree, difference with the CCSDT energy remains. The situation for the SR X 3A′ 2 state, where the CR-CC(2,3) approach reduces the nearly 29 millihartree error relative to CCSDT obtained in the CCSD calculations to ∼0.2 millihartree, is a lot better, but this does not help the resulting ∆ES-T value, which differs from its CCSDT counterpart by almost 4 kcal/mol (almost a quarter of the CCSDT value of ∆ES-T). The discrepancy between the errors in the CR-CC(2,3) energies of the X 3A′ 2 and A 1E′ 2 states is simply too large. Clearly, one needs to incorporate some triples in the corresponding P spaces, especially in the case of the more challenging A 1E′ 2 state. 96 Once the τ = 0 P spaces are augmented with the leading triply excited determinants identified by the i-CISDTQ-MC propagations and the noniterative corrections δ(P ; Q) are added to the CC(P ) energies to estimate the effects of the remaining T3 correlations, we observe smooth convergence of the resulting CC(P ;Q) energetics toward their respective CCSDT limits. This includes significant improvements in the poor description of the A 1E′ 2 state and the A 1E′ 2 − X 3A′ 2 separation by CR-CC(2,3). As shown in Table 3.13, already after 10000 δτ = 0.0001 a.u. MC time steps, where the i-CISDTQ-MC propagations are still in their infancy, capturing only 25–30% of all triples and using tiny walker populations, on the order of 0.1–0.2% of the total numbers of walkers at τ = 8.0 a.u. (see Table 3.14), the 6.245 millihartree and 3.8 kcal/mol errors in the energy of the A 1E′ 2 state and the ∆ES-T value relative to CCSDT obtained with CR-CC(2,3) reduce in the CC(P ;Q) calculations to 2.248 millihartree and 1.3 kcal/mol, respectively. By running i-CISDTQ-MC a little longer and capturing about 50–60% of all triples in the relevant P spaces, as is the case after 20000 MC iterations, where the walker populations compared to τ = 8.0 a.u. are still tiny, the errors in the CC(P ;Q) values of the A 1E′ 2 energy and ∆ES-T relative to their CCSDT parents drop down by an order of magnitude compared to 10000 MC iterations, to 0.217 millihartree and 0.1 kcal/mol, respectively. Although the excellent description of the predominantly SR X 3A′ 2 state by the CR-CC(2,3) approach hardly needs any improvement, the i-CISDTQ-MC-driven CC(P ;Q) calculations are helping here too, reducing the 0.245 millihartree difference between the CR-CC(2,3) and CCSDT energies to 0.108 millihartree after 10000 MC iterations (26 microhartree when the number of MC iterations is increased to 20000). As anticipated, the uncorrected CC(P ) energies of the X 3A′ 2 and A 1E′ 2 states converge to the respective CCSDT limits too, but they do it at a much slower pace than their CC(P ;Q) counterparts. A comparison of the results of the CC(P ) and CC(P ;Q) calculations for the A 1E′ 2 − X 3A′ 2 gap shown in Table 3.13 and Fig. 3.13 (c) may create an impression as if the noniterative corrections δ(P ; Q) offer very little, but this would be misleading. The relatively fast convergence of the CC(P ) values of ∆ES-T toward their CCSDT parent in 97 the early stages of the underlying i-CISDTQ-MC propagations, which compares well with that observed in the corresponding CC(P ;Q) computations, is a result of the fortuitous cancellation of large errors characterizing the CC(P ) energies of the X 3A′ 2 and A 1E′ 2 states. Since no other system examined in this study displays similar error cancellations, and since costs of computing corrections δ(P ; Q), which offer major error reductions in the individual CC(P ) energies, while accelerating their convergence toward the SRCC target, are low, we recommend using the δ(P ; Q)-corrected CC(P ;Q) energetics. 98 2 and A 1E′ Table 3.13 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 states of cyclopentadienyl cation, as described by the cc-pVDZ basis set, and of the correspond- ing vertical singlet–triplet gaps toward their parent CCSDT values. All calculations were performed at the D5h-symmetric geometry of the X 3A′ 2 state optimized using the unre- stricted CCSD/cc-pVDZ approach reported in Ref. [203]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-CISDTQ-MC runs. The i-CISDTQ-MC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by plac- ing 1500 walkers on the ROHF (X 3A′ 2 state) reference determinants and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock cal- culations, the five lowest core orbitals were kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. 2 state) and RHF (A 1E′ MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 80000 ∞ P a 28.840d 27.396 22.253 17.394 13.743 11.027 4.250 0.155 0.007 X 3A′ 2 (P ; Q)a %Tb 0.245e 0.272 0.267 0.212 0.152 0.108 0.026 0.001 0.000 0 0.8 5.1 11.6 18.3 24.8 52.1 95.3 99.8 −192.615924f P a 38.572d 35.598 27.946 21.124 16.042 12.947 3.964 0.060 0.001 A 1E′ 2 (P ; Q)a %Tb 6.245e 5.948 5.078 3.971 2.756 2.248 0.217 0.001 0.000 −192.589235f 0 1.0 6.5 14.7 23.0 30.9 61.4 98.3 100.0 A 1E′ P c 6.1d 5.1 3.6 2.3 1.4 1.2 -0.2 -0.1 0.0 2 − X 3A′ 2 (P ; Q)c 3.8e 3.6 3.0 2.4 1.6 1.3 0.1 0.0 0.0 16.7g a Unless otherwise stated, all energies are reported as errors relative to CCSDT in millihartree. b The %T values are the percentages of triples captured during the i-CISDTQ-MC propagations [the Sz = 1 triply excited determinants of the B2 (C2v) symmetry in the case of the X 3A′ 2 state and the Sz = 0 triply excited determinants of the A1 (C2v) symmetry in the case of the A 1E′ 2 state]. c Unless otherwise specified, the values of the singlet–triplet gaps are reported as errors relative to CCSDT in kcal/mol. d Equivalent to CCSD. e Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR- CC(2,3),D or CR-CC(2,3)D]. f Total CCSDT energy in hartree. g The CCSDT singlet–triplet gap in kcal/mol. 99 Table 3.14 The total numbers of walkers, reported as percentages of the total walker populations at 80000 MC iterations, characterizing the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. that were needed to generate the CC(P ) and CC(P ;Q) results for the cyclopentadienyl cation reported in Table 3.13. Adapted from Ref. [102]. MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 80000 X 3A′ 2 0.00a 0.01 0.03 0.06 0.11 0.16 0.62 13.20 100b A 1E′ 2 0.00a 0.00 0.02 0.05 0.09 0.13 0.54 15.73 100c a The initial walker population, meaning 1500 walkers on the ROHF (X 3A′ reference determinants. b The total number of walkers at 80000 MC iterations is 7867091953. c The total number of walkers at 80000 MC iterations is 11371381724. 2 state) and RHF (A 1E′ 2 state) Figure 3.12 π molecular orbital network of the cyclopentadienyl cation molecule, obtained at the HF/cc-pVDZ level, at the D5h-symmetric geometry of the X 3A′ 2 state optimized using the unrestricted CCSD/cc-pVDZ approach in Ref. [203]. The orbital irreducible representations in the D5h symmetry are shown in black and the corresponding labels in the C2v symmetry are shown in the parenthesis in orange. This electronic configuration refers to the triplet state. 100 Figure 3.13 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 [panel (a)] and A 1E′ 2 [panel (b)] states of cyclopentadienyl cation, as described by the cc-pVDZ basis set, and of the corresponding vertical singlet–triplet gaps [panel (c)] toward their parent CCSDT values. All calculations were performed at the D5h-symmetric geometry of the X 3A′ 2 state optimized using the unrestricted CCSD/cc-pVDZ approach in Ref. [203]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i-CISDTQ- MC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-CISDTQ-MC. Adapted from Ref. [102]. 3.3.4 Trimethylenemethane Our final example is trimethylenemethane, a fascinating non-Kekul´e hydrocarbon exam- ined as early as in 1948 [209] and 1950 [210], in which four valence π electrons are delocalized over four closely spaced π-type orbitals. Assuming the D3h symmetry, which is the symmetry of the minimum-energy structure on the ground-state triplet surface of trimethylenemethane, the four MOs of this system’s valence π network consist of the nondegenerate 1a′′ 2 orbital, the doubly degenerate 1e′′ shell, and the nondegenerate 2a′′ 2 orbital. If one adopts the C2v symmetry, relevant to the low-lying singlet states, which is also the largest Abelian sub- group of D3h exploited in our CCSD, CR-CC(2,3), CCSDT, and CIQMC-driven CC(P ) and CC(P ;Q) computations, the nondegenerate 1a′′ 2 and 2a′′ 2 orbitals in a D3h description become the 1b1 and 3b1 MOs, respectively, whereas the degenerate 1e′′ shell splits into the 1a2 and 2b1 components (see Fig. 3.14 for an illustration of the π MOs). The first experimental identification of trimethylenemethane dates back to 1966 [211], 101 0204060800122436(a)X 3A′2CC(P)CC(P;Q)020406080MC Iteratio s (103)0122436Error rel. to CCSDT (mEh)(b)A 1E′2CC(P)CC(P;Q)020406080MC Iteratio s (103)0.01.53.04.56.0Error rel. to CCSDT (kcal/mol)(c)ΔES–TCC(P)CC(P;Q) a definitive experimental verification, using electron paramagnetic resonance, of its triplet ground state was accomplished already in 1976 [212], and the electronic structure of trimethylen- emethane has been well understood for decades (cf., e.g., Ref. [213] and references therein), but an accurate characterization of its triplet ground state and low-lying singlet states and energy separations between them continues to present a significant challenge to quantum chemistry approaches [80, 88, 153, 154, 214–239]. The D3h-symmetric triplet ground state, designated as X 3A′ 2 (in a C2v description adopted in this study, X 3B2), which is dominated by the |{core}(1a′′ 2)2(1e′′ 1)1(1e′′ 2)1| configuration (in C2v, |{core}(1b1)2(1a2)1(2b1)1|), is rela- tively easy to describe, but the next two states, which are the nearly degenerate singlets sta- bilized by the Jahn–Teller distortion that lifts their exact degeneracy in a D3h description, are not. The lower of the two singlets, which is characterized by a Cs-symmetric minimum that can be approximated by a twisted C2v structure and which is, therefore, usually designated as the A 1B1 state, is an open-shell singlet that emerges from the |{core}(1b1)2(1a2)1(2b1)1| configuration. The second singlet, labeled as the B 1A1 state, is a C2v-symmetric multi- configurational state dominated by the |{core}(1b1)2(1a2)2| and |{core}(1b1)2(2b1)2| closed- shell determinants. The A 1B1 state, although lower in energy compared to its B 1A1 coun- terpart, has not been observed experimentally due to unfavorable Franck–Condon factors [229, 240], so we do not consider it in this work. However, the second singlet, B 1A1, has been detected in photoelectron spectroscopy experiments reported in Refs. [240, 241], which located it at 16.1 ± 0.1 kcal/mol above the X 3A′ 2 ground state. Thus, following our pre- vious deterministic, active-orbital-based, CC(P ;Q) work [88] and the state-of-the-art DEA- and DIP-EOMCC computations with up to 4p-2h and 4h-2p excitations reported in Refs. [80, 153, 154], in carrying out the CIQMC-driven CC(P ) and CC(P ;Q) calculations dis- cussed in this subsection and executing the accompanying CCSD, CR-CC(2,3), and CCSDT runs, we focused on the D3h-symmetric triplet ground state, X 3A′ 2, the C2v-symmetric B 1A1 singlet, and the adiabatic gap between them, adopting the geometries of the two states op- timized using the spin-flip density functional theory (SF-DFT) and the 6-31G(d) basis in 102 Ref. [231]. In analogy to other organic biradicals discussed in this chapter, we employed the cc-pVDZ basis set, so that the parent CCSDT computations, needed to judge the perfor- mance of our semi-stochastic CC(P ) and CC(P ;Q) methods, and the more expensive CC(P ) and CC(P ;Q) calculations employing large, near-100%, fractions of triples in the relevant P spaces (captured in the later stages of the underlying CIQMC propagations) were not too difficult to execute on the computers available to us. As shown in our earlier deter- ministic CC(P ;Q) work [88], in which we tested the active-orbital-based CC(t;3) method, which recovers the CCSDT energetics to within small fractions of kilocalorie per mole, and as confirmed by the authors of Ref. [237], who managed to perform the CCSDT/cc-pVTZ calcu- lations, the use of a larger cc-pVTZ basis changes the adiabatic B 1A1 − X 3A′ 2 gap by about 0.5–1 kcal/mol, i.e., the use the cc-pVDZ basis is sufficient to draw meaningful conclusions regarding the performance of the semi-stochastic CC(P ) and CC(P ;Q) approaches. While the main goal of this study is to examine the efficiency of the CIQMC-driven CC(P ;Q) approaches in converging the CCSDT energetics, it is worth pointing out that the parent CCSDT calculations using the ROHF reference determinant for the X 3A′ 2 state and the RHF reference for the more strongly correlated B 1A1 state, in spite of their SR character, are capable of producing a reasonably accurate description of the adiabatic B 1A1 − X 3A′ separation in trimethylenemethane. Indeed, the purely electronic B 1A1 − X 3A′ 2 gap, desig- 2 nated, in analogy to other singlet–triplet gaps considered in this work, as ∆ES-T, resulting from the ROHF/RHF-based CCSDT/cc-pVDZ computations using the SF-DFT/6-31G(d) geometries of the X 3A′ 2 and B 1A1 states optimized in Ref. [231] is 21.7 kcal/mol [88] (cf. Table 3.15). The corresponding experimentally derived result, obtained by subtracting the zero-point vibrational energy correction ∆ZPVE resulting from the SF-DFT/6-31G(d) cal- culations reported in Ref. [231] from the experimental B 1A1 − X 3A′ 2 gap determined in Refs. [240, 241], is 18.1 kcal/mol. The CCSDT/cc-pVDZ value of ∆ES-T is not as accurate as the electronic B 1A1 − X 3A′ 2 gaps generated in the high-level DEA- and DIP-EOMCC calculations with the explicit inclusion of 4p-2h and 4h-2p correlations on top of CCSD, 103 which produce 18–19 kcal/mol [80, 153, 154], but it is certainly much better than 46.1, 24.4, and 29.8 kcal/mol obtained with the ROHF/RHF-based CCSD, CCSD(T), and CR-CC(2,3) methods, respectively, when the cc-pVDZ basis set is employed [88] [as demonstrated in Ref. [88], the use of a larger cc-pVTZ basis makes the CCSD, CCSD(T), and CR-CC(2,3) results even worse; the CCSD/cc-pVDZ and CR-CC(2,3)/cc-pVDZ values of ∆ES-T are in- cluded in Table 3.15 as the τ = 0 CC(P ) and CC(P ;Q) data, respectively]. While much of the 3.6 kcal/mol difference between the electronic B 1A1 − X 3A′ 2 separation obtained in the ROHF/RHF-based CCSDT/cc-pVDZ calculations and its experimentally derived estimate of 18.1 kcal/mol determined in Ref. [231] is, most likely, a consequence of the neglect of T4 clusters in the CCSDT approach, we should keep in mind that the latter estimate de- pends on the source of the information about the ∆ZPVE correction. For example, if one replaces the ∆ZPVE value obtained in the SF-DFT/6-31G(d) calculations reported in Ref. [231] by its CCSD(T)/6-311++G(2d,2p) estimate and accounts for the core polarization effects determined with the help of the CCSD(T)/cc-pCVQZ computations, combining the resulting information with the experimental B 1A1 − X 3A′ 2 separation determined in Refs. [240, 241], the purely electronic, experimentally derived, adiabatic ∆ES-T gap increases to 19.4 kcal/mol [237], which differs from our CCSDT/cc-pVDZ result by 2.3 kcal/mol. On the other hand, as shown in Ref. [237], the CCSDT value of the adiabatic B 1A1 − X 3A′ 2 gap increases with the basis set too, to 23.1 kcal/mol when the cc-pVTZ basis is employed, which reinforces our view that without accounting for T4 correlations one cannot bring the results of conventional SRCC computations to a close agreement with the experimentally derived data. While the examination of the role of T4 clusters, basis set, geometries of the X 3A′ 2 and B 1A1 states employed in the calculations, ∆ZPVE corrections, etc., would certainly be interesting, it would also be outside the scope of the present study. Thus, in the remainder of this subsection, we return to the analysis of the performance of the CIQMC-driven CC(P ;Q) approach and its CC(P ) counterpart, especially their ability to converge the parent CCSDT energetics when the cc-pVDZ basis is employed. 104 The results of our semi-stochastic CC(P )/cc-pVDZ and CC(P ;Q)/cc-pVDZ computa- tions for the X 3A′ 2 and B 1A1 states of trimethylenemethane and the adiabatic gap be- tween them, along with the associated CCSD, CR-CC(2,3), and CCSDT data, are sum- marized in Table 3.15 and Fig. 3.15. As in the case of cyclopentadienyl cation, to reduce the computational costs of the underlying CIQMC propagations, especially in their later stages, we resorted to the truncated i-CISDTQ-MC approach. In analogy to cyclobutadiene and cyclopentadienyl cation, we terminated our i-CISDTQ-MC propagations after 80000 δτ = 0.0001 a.u. MC time steps, where the differences between the CC(P ;Q) and CCSDT energies of the X 3A′ 2 and B 1A1 states fall below 1 microhartree. Consistent with the CC(P ), CC(P ;Q), and other SRCC calculations for trimethylenemethane reported in Table 3.15 and Fig. 3.15, we used the ROHF determinant to initiate the i-CISDTQ-MC propagation for the D3h-symmetric X 3A′ 2 (in C2v, X 3B2) state and the RHF determinant to initiate the i- CISDTQ-MC run for the C2v-symmetric B 1A1 state. The lists of triply excited determinants captured by the i-CISDTQ-MC runs at the various times τ > 0, needed to construct the P spaces for the CC(P ) and CC(P ;Q) computations, were the Sz = 1 triples of the B2 (C2v) symmetry in the case of the X 3A′ 2 state and the Sz = 0 triples of the A1 (C2v) symmetry when considering the B 1A1 state. The remaining triples not captured by i-CISDTQ-MC defined the corresponding Q spaces. It is clear from the results presented in Table 3.15 and Fig. 3.15 that the semi-stochastic CC(P ;Q) approach is very effective in converging the parent CCSDT energetics character- izing the X 3A′ 2 and B 1A1 states of trimethylenemethane and the adiabatic gap between them. It offers substantial improvements in the results of the CR-CC(2,3) calculations in the early stages of the underlying i-CISDTQ-MC propagations, especially when the multi- configurational B 1A1 state and the adiabatic B 1A1 − X 3A′ 2 separation ∆ES-T, which are poorly described by CR-CC(2,3), are examined, while greatly accelerating the convergence of the CC(P ) energies toward CCSDT. Indeed, after 6000 δτ = 0.0001 a.u. MC iterations, which is a very short propagation time engaging only ∼0.1% of the total walker popula- 105 tions at τ = 8.0 a.u., where we terminated our i-CISDTQ-MC runs (cf. Table 3.16), and i-CISDTQ-MC capturing as little as 14–17% of all triply excited determinants, the semi- stochastic CC(P ;Q) methodology reduces the 13.370 millihartree difference between the CR-CC(2,3) and CCSDT energies of the B 1A1 state and the 8.1 kcal/mol error in the CR- CC(2,3) value of the B 1A1 − X 3A′ 2 gap relative to CCSDT to 1.260 millihartree and 0.6 kcal/mol, respectively, which is a chemical accuracy regime. Interestingly, the i-CISDTQ- MC-based CC(P ;Q) value of ∆ES-T obtained after 6000 MC iterations matches the quality of the B 1A1 − X 3A′ 2 gap resulting from the fully deterministic CC(P ;Q) calculations using the CC(t;3) approach, which give a 0.5 kcal/mol error relative to CCSDT when the cc-pVDZ ba- sis set is employed [88]. After the additional 4000 MC time steps, where the i-CISDTQ-MC propagations for the X 3A′ 2 and B 1A1 states are still very far from convergence and where the fractions of triples captured by i-CISDTQ-MC increase to about 30%, the small errors relative to CCSDT characterizing the i-CISDTQ-MC-based CC(P ;Q) values of the energy of the B 1A1 state and ∆ES-T at τ = 0.6 a.u. drop down by factors of 4–6, to 0.314 millihartree and 0.1 kcal/mol, respectively, illustrating how rapid the convergence of the CIQMC-driven CC(P ;Q) calculations toward the parent SRCC data can be. While the CR-CC(2,3) de- scription of the X 3A′ 2 state, which has a SR character, is much better than in the case of its strongly correlated B 1A1 counterpart, the semi-stochastic CC(P ;Q) computations offer great improvements in this case too. They are, for example, capable of reducing the ∼0.4 millihartree difference between the CR-CC(2,3) and CCSDT energies to a 0.1 millihartree level after 10000 MC iterations and i-CISDTQ-MC capturing less than 30% of all triples. In analogy to all other molecular examples considered in this chapter, the uncorrected CC(P ) values of the energies of the X 3A′ 2 and B 1A1 states and separation between them converge to their CCSDT limits too, but it is clear from Table 3.15 and Fig. 3.15 that they do it at a much slower rate than their CC(P ;Q) counterparts. This can be illustrated by comparing the errors relative to CCSDT characterizing the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 and B 1A1 states and separation between them obtained after 6000 MC iterations. They 106 are more than 11 millihartree, about 21 millihartree, and almost 6 kcal/mol, respectively, in the former case and only 0.253 millihartree, 1.260 millihartree, and 0.6 kcal/mol, when the CC(P ) energies are corrected for the remaining T3 correlations using the CC(P ;Q) approach. As explained in Section 3.1, the CC(P ) energies converge to CCSDT more slowly than their δ(P ; Q)-corrected CC(P ;Q) counterparts, since the initial, τ = 0, CC(P ) calculation for a given electronic state is equivalent to CCSD, where T3 = 0. The CIQMC-driven CC(P ;Q) calculations start from CR-CC(2,3), which provides information about T3 clusters via non- iterative corrections to CCSD. This once again emphasizes the benefits of using corrections δ(P ; Q) in the context of the semi-stochastic CC(P ;Q) work. 107 Table 3.15 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 and B 1A1 states of trimethylenemethane, as described by the cc-pVDZ basis set, and of the corresponding adi- abatic singlet–triplet gaps toward their parent CCSDT values. The D3h- and C2v-symmetric 2 and B 1A1 states, respectively, optimized in the SF-DFT/6-31G(d) geometries of the X 3A′ calculations, were taken from Ref. [231]. The P spaces used in the CC(P ) and CC(P ;Q) calculations were defined as all singly and doubly excited determinants and subsets of triply excited determinants extracted from the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. The Q spaces used to determine the CC(P ;Q) corrections consisted of the triply excited determinants not captured by the corresponding i-CISDTQ-MC runs. The i-CISDTQ-MC calculations preceding the CC(P ) and CC(P ;Q) steps were initiated by placing 1500 walk- 2 state) and RHF (B 1A1 state) reference determinants and the na ers on the ROHF (X 3A′ parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the four lowest core orbitals were kept frozen and the spherical components of d orbitals were employed throughout. Adapted from Ref. [102]. MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 80000 ∞ P a 19.202d 17.975 14.462 11.319 9.066 7.429 3.294 0.213 0.012 X 3A′ 2 (P ; Q)a %Tb 0.418e 0.422 0.357 0.253 0.173 0.123 0.031 0.001 0.000 0 1.1 6.6 14.1 21.3 27.9 52.3 92.8 99.5 P a 58.051d 50.012 32.925 20.628 14.601 10.680 2.675 0.061 0.002 B 1A1 (P ; Q)a %Tb 13.370e 9.362 3.236 1.260 0.649 0.314 0.028 0.000 0.000 0 1.2 7.7 16.8 25.5 33.1 61.1 97.1 99.9 −155.466242f −155.431596f B 1A1 − X 3A′ 2 (P ; Q)c P c 8.1e 24.4d 5.6 20.1 1.8 11.6 0.6 5.8 0.3 3.5 0.1 2.0 0.0 -0.4 0.0 -0.1 0.0 0.0 21.7g a Unless otherwise stated, all energies are reported as errors relative to CCSDT in millihartree. b The %T values are the percentages of triples captured during the i-CISDTQ-MC propagations [the Sz = 1 triply excited determinants of the B2 (C2v) symmetry in the case of the X 3A′ 2 state and the Sz = 0 triply excited determinants of the A1 symmetry in the case of the B 1A1 state]. c Unless otherwise specified, the values of the singlet–triplet gaps are reported as errors relative to CCSDT in kcal/mol. d Equivalent to CCSD. e Equivalent to CR-CC(2,3) [the most complete variant of CR-CC(2,3) abbreviated sometimes as CR- CC(2,3),D or CR-CC(2,3)D]. f Total CCSDT energy in hartree. g The CCSDT singlet–triplet gap in kcal/mol. 108 Table 3.16 The total numbers of walkers, reported as percentages of the total walker populations at 80000 MC iterations, characterizing the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. that were needed to generate the CC(P ) and CC(P ;Q) results for trimethylenemethane reported in Table 3.15. Adapted from Ref. [102]. MC Iterations 0 2000 4000 6000 8000 10000 20000 50000 80000 X 3A′ 2 0.00a 0.01 0.06 0.14 0.24 0.34 1.09 14.60 100b B 1A1 0.00a 0.01 0.05 0.11 0.18 0.26 0.93 15.05 100c a The initial walker population, meaning 1500 walkers on the ROHF (X 3A′ reference determinants. b The total number of walkers at 80000 MC iterations is 2363904677. c The total number of walkers at 80000 MC iterations is 3543757954. 2 state) and RHF (B 1A1 state) Figure 3.14 π molecular orbital network of the trimethylenemethane molecule, obtained at the HF/cc-pVDZ level, at the D3h-symmetric geometry of the X 3A′ 2 state optimized in the SF-DFT/6-31G(d) calculations and taken from Ref. [231] The orbital irreducible representations in the D3h symmetry are shown in black and the corresponding labels in the C2v symmetry are shown in the parenthesis in orange. This electronic configuration refers to the triplet state. 109 Figure 3.15 Convergence of the CC(P ) and CC(P ;Q) energies of the X 3A′ 2 [panel (a)] and B 1A1 [panel (b)] states of trimethylenemethane, as described by the cc-pVDZ basis set, and of the corresponding adiabatic singlet–triplet gaps [panel (c)] toward their parent CCSDT 2 and B 1A1 states, optimized in the SF-DFT/6-31G(d) values. The geometries of the X 3A′ calculations, were taken from Ref. [231]. The P spaces consisted of all singles and doubles and subsets of triples identified during the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. and the Q spaces consisted of the triples not captured by i-CISDTQ-MC. Adapted from Ref. [102]. 110 02040608001224364860(a)X 3A′2CC(P)CC(P;Q)020406080MC Iterations (103)01224364860Error re . to CCSDT (mEh)(b)B 1A1CC(P)CC(P;Q)020406080MC Iterations (103)0510152025Error re . to CCSDT (kca /mo )(c)ΔES–TCC(P)CC(P;Q) CHAPTER 4 THE SEMI-STOCHASTIC EXTENSIONS OF PARTICLE NONCONSERVING EQUATION-OF-MOTION COUPLED-CLUSTER THEORIES This chapter is based on the method development and programming work, followed by benchmark computations, aimed at extending the semi-stochastic approaches to particle nonconserving EOMCC schemes, which is subject to the manuscript in preparation [242], in which I have played a lead role in every aspect of the work other than the proposal for developing such ideas by my advisor Professor Piotr Piecuch and oversight from him. As described in the Introduction, despite remarkable strides in computer hardware and efficient software advancements, providing an accurate and reliable description of MR prob- lems and open-shell systems remain a challenge. This includes excited states dominated by two-electron transitions and electronic spectra of radical and biradical species, to name a few. In the previous chapter, we have already discussed how we can use the semi-stochastic CC(P;Q) approach to tackle these challenging situations, and in this chapter, we describe the how the semi-stochastic ideas can be combined with the particle nonconserving EOMCC formalisms to achieve highly accurate results for open-shell systems. As previously discussed, the particle nonconserving EOMCC approaches offer a simple and elegant way of describing the electronic spectra of open-shell systems by formally adding electrons to (electron attachment) or by removing electrons from (electron ionization) the nearest closed-shell reference core. Within this family of EOMCC methods, we find the EA and and IP EOMCC approaches [44, 78, 79, 138–146, 243–248], where the reference and the target systems differ by one electron (|Ntarget − Nreference| = 1), the DEA and DIP EOMCC methodologies [80, 147–154, 157, 249, 250], where |Ntarget − Nreference| = 2, and their higher order extensions. These approaches prove particularly advantageous in the study of electronic spectra of open-shell systems because the operation of adding electrons to or removing electrons from a closed-shell reference to attain a target system automatically creates an appropriate multi-reference model space specific to the system of interest while 111 relaxing the remaining electrons. In contrast to employing genuine MRCC approaches, these methods, being formally single reference in nature, are considerably much simpler to use. Furthermore, compared to traditional spin-integrated, spin-orbital implementations of particle-conserving CC/EOMCC treatments employing unrestricted or restricted open-shell reference determinants they offer distinct advantages, such as rigorous spin and symmetry adaptation of the computed states and the capability to balance both high-spin and low-spin states in an accurate and equally balanced footing. Within this area, the EA-EOMCC(3p-2h) and IP-EOMCC(3h-2p) methods [78, 140– 142, 162, 163, 251–253], have demonstrated notable success in accurately describing open- shell systems featuring one unpaired electron outside a closed-shell core i.e., they can effec- tively treat the (1,0) and (0,1) sectors of the Fock space. Additionally, the DEA-EOMCC(4p- 2h) and DIP-EOMCC(4h-2p) methods [80, 153–156, 254–256] enable quantitative descrip- tions of the (2,0) and (0,2) sectors of the Fock space, thereby facilitating the precise deter- mination of singlet–triplet gaps in biradical systems. To facilitate the application of these approaches without compromising on the accuracy, the Piecuch group has previously lever- aged the active space ideas based on active orbitals to select the leading components in the respective electron attaching or removing operators, which resulted in methods [78, 80, 140– 142, 153–156, 162, 163, 251–256] designated as EA-EOMCCSDt, IP-EOMCCSDt, DEA- EOMCC(3p-1h,4p-2h){Nu}, DEA-EOMCC(4p-2h){Nu}, and DIP-EOMCC(4h-2p){No} (No and Nu indicate the numbers of active occupied and unoccupied orbitals used to define the active-space). While the aforementioned methodologies have proven to be very reliable in describing the electronic spectra of radicals and determining singlet–triplet gaps in biradical systems, the selection of an appropriate active-space relies heavily on the specific system un- der study and the chemical intuition of the investigator. Notably, the active-space selection can pose a significant challenge in large and complex molecular systems. In this work, we propose an alternative to using active orbitals to select the dominant higher-order excita- tions. Drawing inspiration from our group’s prior semi-stochastic CC(P )/EOMCC(P ) and 112 CC(P ;Q) work [99–102, 137], where the CIQMC wave function propagations [91–94, 257] were fused with deterministic CC/EOMCC computations to provide high-level CC/EOMCC energetics at a reduced computational cost and in a black-box manner. To be more spe- cific, we extend the semi-stochastic particle-conserving EOMCC(P ) approach [100, 137] to the particle nonconserving regime, resulting in the EA-EOMCC(P ), IP-EOMCC(P ), DEA- EOMCC(P ), and DIP-EOMCC(P ) methodologies [242]. These methods utilize CIQMC wave function propagations to stochastically select lists of dominant 3p-2h/3h-2p/4p-2h/4h- 2p components in the EOMCC electron-attaching and electron-removing operators in an au- tomated fashion and subsequently solve the particle-nonconserving EOMCC equations based on those stochastically determined lists. To validate the efficiency of these semi-stochastic approaches, we examine the C2N, CNC, N3, and NCO radicals and methylene and TMM biradicals. 4.1 Theory Consistent with the previously developed semi-stochastic CC(P ) and EOMCC(P ) schemes [99–102, 137], the CIQMC-driven EA/IP/DEA/DIP-EOMCC methods, abbreviated as EA- EOMCC(P ), IP-EOMCC(P ), DEA-EOMCC(P ), and DIP-EOMCC(P ), respectively, utilize stochastic QMC wave function propagations to select the dominant high-order correlations in an automated black-box manner. The key algorithmic details of the semi-stochastic EA/IP/DEA/DIP-EOMCC calculations are as follows [242]: 1. First, perform a CCSD calculation, for the (N ∓ 1)- or (N ∓ 2)-electron closed-shell system in the case of the EA/IP-EOMCC, or DEA/DIP-EOMCC, respectively, to obtain the cluster amplitudes and to construct the similarity transformed Hamiltonian. 2. Next, start a CIQMC propagation by placing a certain number of walkers on the reference determinant pertaining to the N -electron target system, while using the one- and two-electron integrals from the closed-shell reference. 3. Afterwards extract a list of the most important determinants relevant to the EA/IP 113 and DEA/DIP EOMCC theory of interest [e.g., 3p-2h excited determinants relative to |Φ(N −1)⟩ for EA-EOMCC(3p-2h), 3h-2p excited determinants relative to |Φ(N +1)⟩ for IP-EOMCC(3h-2p), 4p-2h excited determinants relative to |Φ(N −2)⟩ for DEA-EOMCC(4p- 2h), and 4h-2p excited determinants relative to |Φ(N +2)⟩ for DIP-EOMCC(4h-2p)] from the CIQMC wave function at a given time τ to define the P spaces for the EA/IP- EOMCC(P ) and DEA/DIP-EOMCC(P ) calculations. If the target approach is EA- EOMCC(3p-2h), the P space (H (P )) is defined as all 1p (G1p), all 2p-1h (G2p-1h), and a subset of 3p-2h (g3p-2h) excited determinants having at least nP (e.g., one) positive Figure 4.1 A schematic illustration depicting the construction of P -spaces in the EA/IP/DEA/DIP-EOMCC(P ) computations. Panel (a) showcases the stabilization of cor- relation energy (green line) and the corresponding increase in the total number of walkers is shown in panel (b) (red line). On the right four snapshots from a QMC calculation are presented, featuring the lists of determinants picked up by the QMC algorithm at various time steps (light blue for 2000, dark blue for 20000, violet for 40000, and magenta for 100000 QMC iterations). It is evident that QMC deems some determinants more important than others by placing more walkers on them. or negative walker, such that H (P ) = G1p ⊕ G2p-1h ⊕ g3p-2h. Similarly, define H (P ) = G1h ⊕ G2h-1p ⊕ g3h-2p for IP-EOMCC(3h-2p), H (P ) = G2p ⊕ G3p-1h ⊕ g4p-2h for DEA- EOMCC(4p-2h), and in case of DIP-EOMCC(4h-2p) use H (P ) = G2h ⊕ G3h-1p ⊕ g4h-2p (see Fig. 4.1 for an illustration). 4. Then solve the semi-stochastic EA/IP-EOMCC and DEA/DIP-EOMCC equations in 114 utf-8[unix]QMC-20001 102 1 2 3 4 5 6 7 8 9 10 11 12 132 -4 1 2 3 5 6 7 8 10 11 12 13 22 473 3 1 2 3 4 6 7 9 10 11 12 13 20 494 -3 1 2 3 4 6 7 8 9 11 12 13 18 455 3 1 2 3 4 5 6 7 8 9 10 11 13 146 -3 1 2 3 4 5 6 7 8 9 12 13 18 417 3 1 2 3 4 5 6 7 9 10 11 12 50 538 -3 1 3 4 5 6 7 8 9 11 12 13 42 689 3 1 2 3 4 5 7 8 9 10 11 13 14 4610 -3 1 2 3 4 5 6 7 8 9 10 12 13 1911 -3 1 3 4 5 6 7 8 9 10 11 13 42 5212 -2 1 2 3 4 5 6 7 8 11 12 13 15 1613 2 1 2 3 4 5 6 7 8 9 11 13 18 2014 2 1 2 3 4 5 6 7 8 10 11 13 55 7215 2 1 2 3 4 5 6 7 8 10 11 12 53 5516 -2 1 2 3 4 7 8 9 10 11 12 13 15 1617 2 1 2 3 4 6 7 8 10 11 12 13 17 5318 2 1 2 3 4 5 7 8 9 10 11 13 16 6819 -2 1 3 4 5 6 7 8 9 10 12 13 14 5320 2 1 2 3 5 6 7 8 9 10 11 12 13 4421 -2 2 3 4 5 6 7 8 9 11 12 13 28 3722 -2 1 2 4 5 6 7 8 10 11 12 13 49 5523 -2 1 2 3 4 6 7 8 9 10 11 13 20 3724 2 1 2 3 4 5 7 8 10 11 12 13 31 3825 -2 1 2 3 4 6 7 8 9 10 11 13 20 23utf-8[unix]QMC-200001 193 1 2 3 4 5 6 7 8 9 10 11 12 132 -19 1 2 3 4 7 8 9 10 11 12 13 15 163 -16 1 2 3 4 5 6 7 8 11 12 13 17 184 -15 1 2 3 4 5 6 9 10 11 12 13 15 165 15 1 2 3 4 5 6 7 8 9 10 11 13 146 -14 1 2 3 4 5 6 7 8 9 11 13 14 167 -14 1 2 3 4 5 6 7 8 9 12 13 18 198 -14 1 2 3 4 5 6 7 8 10 11 12 15 199 -13 1 2 3 4 5 6 7 8 11 12 13 15 1610-13 1 2 3 4 5 6 8 9 10 11 12 13 2111 12 1 2 3 4 5 6 7 8 10 11 13 14 1512-11 1 2 3 4 5 6 7 8 11 12 13 15 1813 11 1 2 3 4 5 6 7 9 10 11 12 13 2414-11 1 2 3 4 5 8 9 10 11 12 13 15 1615 10 1 2 3 4 5 6 7 8 9 11 13 16 2016 10 1 2 3 5 6 7 8 9 10 11 12 19 2217 -9 1 2 3 4 5 6 7 8 9 10 12 13 1918 -9 1 2 3 4 5 6 7 8 11 12 13 47 4819 -9 1 2 3 4 5 6 7 9 10 11 12 13 3820 8 1 2 3 4 5 6 9 10 11 12 13 35 4621 -8 2 3 4 5 6 7 8 9 10 11 13 20 3722 -8 1 2 3 4 5 6 7 8 9 12 13 18 4123 -8 1 2 3 4 5 6 7 9 10 11 12 13 3624 8 1 2 4 5 6 7 8 10 11 12 13 15 2125 8 1 2 3 4 5 6 7 9 10 11 12 13 22utf-8[unix]QMC-400001 419 1 2 3 4 5 6 7 8 9 10 11 12 132 -44 1 2 3 4 5 6 7 8 10 11 12 15 193 -38 1 2 3 4 5 8 9 10 11 12 13 15 164 -37 1 2 3 4 5 6 7 8 11 12 13 15 165 -37 1 2 3 4 5 6 9 10 11 12 13 15 166 -34 1 2 3 4 5 6 7 8 9 11 13 14 167 33 1 2 3 4 5 6 7 9 10 11 12 13 248 -29 1 2 3 4 5 6 8 9 10 11 12 13 219 -28 1 2 3 4 5 6 7 8 9 12 13 18 1910 28 1 2 3 4 5 6 7 8 10 11 13 14 1511 27 1 2 3 4 5 6 7 8 9 11 12 13 1612 -26 1 2 3 4 5 6 7 8 9 10 13 19 2013 -24 1 2 3 4 7 8 9 10 11 12 13 15 1614 -24 1 2 3 4 5 6 7 8 9 10 12 13 1915 -23 1 2 3 4 5 6 7 8 9 10 11 12 1916 -23 1 2 4 5 6 7 8 9 11 12 13 16 2117 -23 1 2 3 4 5 6 7 9 10 11 12 13 3818 21 1 2 3 4 5 6 7 8 9 11 13 16 2019 20 1 2 3 4 5 7 8 9 10 11 12 13 3620 -20 1 2 3 4 6 7 9 10 11 12 13 15 1821 19 1 2 3 4 5 6 7 8 9 10 11 13 1422 19 1 2 3 4 5 6 7 8 9 10 11 13 2623 -19 1 2 3 4 5 6 7 8 10 11 13 15 2024 19 1 2 3 4 5 6 7 8 10 11 13 14 1725 -18 1 2 3 4 7 8 9 10 11 12 13 17 18utf-8[unix]QMC-1000001 4398 1 2 3 4 5 6 7 8 9 10 11 12 132 -404 1 2 3 4 5 6 7 8 11 12 13 15 163 -400 1 2 3 4 5 6 7 8 9 10 12 13 194 360 1 2 3 4 5 6 7 8 9 10 11 14 195 -352 1 2 3 4 5 6 9 10 11 12 13 15 166 -331 1 2 3 4 5 6 7 8 9 10 13 19 207 -316 1 2 3 4 5 6 7 8 11 12 13 17 188 -295 1 2 3 4 5 6 7 8 10 11 12 15 199 293 1 2 3 4 5 6 7 9 10 11 12 13 2410 -279 1 2 3 4 5 6 7 8 11 12 13 15 1811 259 1 2 3 4 5 6 7 8 10 11 13 14 1512 255 1 2 3 4 5 6 7 8 11 12 13 16 1713 -250 1 2 3 4 5 6 7 8 9 11 13 14 1614 228 1 2 3 4 5 6 7 8 9 11 12 13 1615 -220 1 2 3 4 5 6 8 9 10 11 12 13 2316 215 1 2 3 4 5 7 8 9 10 11 12 13 3617 -200 1 2 3 4 5 6 7 8 10 11 13 17 2018 191 1 2 3 4 5 6 8 9 10 11 12 13 3719 -189 1 2 3 4 5 6 7 8 10 11 13 15 2020 -183 1 2 5 6 7 8 9 10 11 12 13 15 1621 -179 1 2 3 4 5 6 7 8 9 12 13 18 1922 -162 1 2 3 4 5 6 8 9 10 11 12 13 2123 -160 1 2 3 4 5 6 7 9 10 11 12 13 3824 159 1 2 3 4 5 6 7 8 9 10 11 13 1425 159 1 2 3 4 5 6 9 10 11 12 13 15 18(a)(b) the stochastically determined P spaces, i.e., use R(+1) µ = Rµ,1p + Rµ,2p-1h + R(MC) µ,3p-2h for EA-EOMCC(P ), R(−1) µ = Rµ,1h + Rµ,2h-1p + R(MC) Rµ,2p + Rµ,3p-1h + R(MC) µ,4p-2h for DEA-EOMCC(P ), and R(−2) µ,3h-2p for IP-EOMCC(P ), R(+2) µ = Rµ,2h + Rµ,3h-1p + R(MC) µ,4h-2p µ = for DIP-EOMCC(P ). 5. Finally, check the convergence of the resulting energies by repeating the steps 3 and 4 at some later CIQMC propagation time τ ′ > τ . If the energies do not change within a given convergence threshold, we can stop the calculation. One can also stop the calculation if τ in steps 3 and 4 is chosen such that the stochastically determined P spaces contain sufficiently large fraction of higher order excitations relevant to the target EOMCC level. Before going to the next section, we must discuss an interesting aspect of the semi- stochastic particle nonconserving EOMCC(P ) approaches. At τ = 0, the EA-EOMCC(P ), IP-EOMCC(P ), DEA-EOMCC(P ), and DIP-EOMCC(P ) energies are identical to the en- ergies obtained in the EA-EOMCC(2p-1h), IP-EOMCC(2h-1p), DEA-EOMCC(3p-1h), and DIP-EOMCC(3h-1p) calculations. This is because at τ = 0 the P spaces in the respective cal- culations do not contain any 3p-2h, 3h-2p, 4p-2h, or 4h-2p determinants. On the other hand, when τ = ∞, the P spaces in the EA-EOMCC(P ), IP-EOMCC(P ), DEA-EOMCC(P ), and DIP-EOMCC(P ) calculations contain all the 3p-2h, 3h-2p, 4p-2h, and 4h-2p determi- nants, respectively, and, as a result, the semi-stochastic EA/IP/DEA/DIP-EOMCC calcu- lations provide energetics identical to that of the EA-EOMCC(3p-2h), IP-EOMCC(3h-2p), DEA-EOMCC(4p-2h), and DIP-EOMCC(4h-2p) approaches. This relationship between the EA-EOMCC(P ), IP-EOMCC(P ), DEA-EOMCC(P ), and DIP-EOMCC(P ) methods and the fully deterministic EA-EOMCC(3p-2h), IP-EOMCC(3h-2p), DEA-EOMCC(4p-2h), and DIP-EOMCC(4h-2p) methodologies were helpful in examining the correctness of our codes. They also point to the ability of these semi-stochastic particle nonconserving EOMCC ap- proaches driven by the information extracted from CIQMC to offer a systematically improv- 115 able description as τ goes from 0 to ∞. In the next section, we examine the performance of the semi-stochastic EA-EOMCC(P ) and IP-EOMCC(P ) approaches in converging the high-level EA-EOMCC(3p-2h) and IP-EOMCC(3h-2p) methods and the section after that illustrates the capability of their double electron attachment [DEA-EOMCC(P )] and double ionization potential [DIP-EOMCC(P )] extensions to converge the DEA-EOMCC(4p-2h) and DIP-EOMCC(4h-2p) methodologies. 4.2 Adiabatic Excitations in C2N, CNC, N3, and NCO In order to demonstrate the ability of the semi-stochastic EA-EOMCC(P ) and IP- EOMCC(P ) approaches in converging their high-level, fully-deterministic parents, EA-EOMCC(3p- 2h) and IP-EOMCC(3h-2p), we carried out benchmark calculations for the ground and a few lowest-lying doublet states along with the corresponding adiabatic excitation energies of the C2N [Table 4.1 and Fig. 4.2 ], CNC [Table 4.2 and Fig. 4.3], N3 [Table 4.3 and Fig. 4.4], and NCO [Table 4.4 and Fig. 4.5] radicals. All the systems utilized the DZP[4s2p1d] basis set [167, 168] and employed geometries optimized with the SAC-CI SDT-R/PS method combined with the DZP[4s2p1d] basis set, as reported in Ref. [162]. Following our previ- ous semi-stochastic work [99–101, 137], we used the HANDE software package [170, 171] to execute all our i-FCIQMC calculations needed to generate the lists of 3p-2h and 3h-2p de- terminants entering the P spaces for the EA-EOMCC(P) and IP-EOMCC(P) calculations. Our standalone CC/EOMCC codes, interfaced with the RHF, ROHF, and integral transfor- mation routines available in the GAMESS software package [172–174], were used to carry out the EA-EOMCC(P ), IP-EOMCC(P ), and the fully deterministic CCSD, EA-EOMCC(2p- 1h), EA-EOMCC(3p-2h), IP-EOMCC(2h-1p), and IP-EOMCC(3h-2p) computations. Each i-FCIQMC run was initiated by placing 500 walkers on the relevant reference function (see Tables 4.1–4.4 for the details), the initiator parameter na was set at 3, and all of the i- FCIQMC propagations used a time step of τ = 0.0001 a.u. As the true point group of the system of interest was not Abelian for all four molecules, namely C2N, CNC, N3, and NCO, we utilized their largest Abelian subgroups in the calculations. This choice was necessary, 116 since all of our CC/EOMCC codes interfaced with GAMESS and the CIQMC routines in HANDE can only handle Abelian symmetries. In all the post-HF computations, the core electrons corresponding to the 1s shells of the carbon, nitrogen, and oxygen atoms were kept frozen. 4.2.1 Application of the EA-EOMCC(P ) approach to C2N and CNC radicals In this subsection, we investigate the performance of the semi-stochastic EA-EOMCC(P ) approach in converging the EA-EOMCC(3p-2h) total electronic and adiabatic excitation energies of a few low-lying doublet states of C2N and CNC radicals. The reference wave functions are obtained by performing CCSD calculations on the nearest closed-shell C2N+ and CNC+ cations using the DZP[4s2p1d] basis set [167, 168] and, as mentioned above, the nuclear geometries of the ground and excited states of C2N and CNC, optimized using EA SAC-CI-SDT-R/PS and DZP[4s2p1d] basis set, are taken from Ref. [162]. 4.2.1.1 C2N We begin our discussion by using the C2N radical, where we study the X 2Π ground state and three low-lying valence doublet excited states, A 2∆, B 2Σ−, and C 2Σ+. Following Ref. [162], we used the EA SAC-CI-SDT-R/PS with DZP[4s2p1d] optimized geometries, which are RCC = 1.400 ˚A and RCN = 1.185 ˚A for the X 2Π state, RCC = 1.315 ˚A and RCN = 1.207 ˚A for the A 2∆ state, RCC = 1.302 ˚A and RCN = 1.223 ˚A for the B 2Σ− state, and RCC = 1.311 ˚A and RCN = 1.214 ˚A for the C 2Σ+ state. At these geometries, all the three valence excited states studied here, are characterized by two-electron transitions as well as significant 2p-1h components (the B 2Σ− also has non-negligible 3p-2h contributions) in the wave functions relative to the reference C2N+ ion. Therefore, to properly describe these states, an accurate treatment of 3p-2h excitations is required and this is evident from Table 1 of Ref. [162] or Table IV of Ref. [68] (cf., also, Table 4.1 of this work), where the basic EA-EOMCCSD or EA-EOMCC(2p-1h) approach predicts an incorrect state ordering, i.e., suggesting the B 2Σ− state to be higher in energy than the C 2Σ+ state. Furthermore, the A 2∆ − X 2Π, B 2Σ− − X 2Π, and C 2Σ+ − X 2Π adiabatic excitation energies obtained 117 from EA-EOMCC(2p-1h) (6.190 eV, 7.860 eV, and 6.722 eV) are ∼3–5 eV away from the experimental results (2.636 eV, 2.779 eV, and 3.306 eV) reported in Ref. [258]. The higher order EA-EOMCC(3p-2h) method that accounts for a complete treatment of 3p-2h terms in the electron attaching operator R(+1) µ , reduces these errors to ∼1 eV or lower (3.055 eV, 3.678 eV, and 3.809 eV), necessitating the need to use at least the EA-EOMCC(3p-2h) level of theory, if a realistic description is wanted. However, as mentioned before, EA-EOMCC(3p- 2h) has a scaling of N 7 and, hence, could become very expensive very quickly. It would, therefore, be interesting to see if the semi-stochastic EA-EOMCC(P) approach is capable of recovering the EA-EOMCC(3p-2h)-quality results in this challenging case of C2N with a fraction of 3p-2h terms selected in a automated manner using the i-FCIQMC wave function propagations. Following the semi-stochastic EA-EOMCC(P ) algorithm, as described above, and our interest in converging the EA-EOMCC(3p-2h) energetics of the ground and three lowest lying valence excited states of C2N, the cluster operator and the electron attaching operator were approximated as T = T1 + T2 and R(+1) µ = Rµ,1p + Rµ,2p-1h + R(MC) µ,3p-2h, where the lists of 3p-2h determinants defining the R(MC) µ,3p-2h components at a given time τ were obtained from the corresponding i-FCIQMC propagations at the same value of τ , meaning the B1 (C2v) component of 2Π for the X 2Π state, the A1 (C2v) component of 2∆ for the A 2∆ state, A2 (C2v) for the B 2Σ− state, and the A1 (C2v) for the C 2Σ+ state (C2v is the largest Abelian subgroup of the true point group of C2N, C∞v). The results of our EA-EOMCC(P ) computations are reported in Table 4.1 and Fig. 4.2. As can be seen from Table 4.1, at τ = 0.0, the the EA-EOMCC(P ) results are far from the EA-EOMCC(3p-2h) energetics, which is expected since, at this stage, the P spaces lack any 3p-2h determinants, rendering these results equivalent to those obtained from the basic EA-EOMCC(2p-1h) method. The ground state of the C2N radical, X 2Π, is dominated by 1p excitations relative to the C2N+ reference and, as a result, it is described reasonably well by EA-EOMCCSD [EA-EOMCC(P |τ =0)]. Still errors of ∼5 millihartree remain relative to 118 EA-EOMCC(3p-2h). This error rapidly decreases to the 1–3 millihartree range as soon as we start incorporating ∼20–30% 3p-2h determinants in the P space. Indeed, at 20000 MC iterations and with about 30% of 3p-2h determinants the error in EA-EOMCC(P ) energy is only 1.421 millihartree. The next state denoted as A 2∆, which is also the first excited state, has significant 2p-1h contributions relative to the reference determinant and, therefore, the effect of adding 3p-2h determinants in the P space is more prominent. For this state, the EA-EOMCC(P |τ =0) approach (EA-EOMCCSD) has, massive, 120 millihartree errors relative to the target EA-EOMCC(3p-2h) method. This huge error is brought down to less than 20 millihartree with 40% of 3p-2h components and at 20000 QMC iterations. If we let QMC run a little longer until 50000 QMC iterations, it recovers ∼70% of 3p-2h determinants and the errors in the total energies are reduced to less than 1 millihartree relative to EA-EOMCC(3p- 2h). If we look at the adiabatic excitation energies, they are converged to 0.2 eV or better at a much earlier QMC time step. For example, already at 20000 Monte Carlo time steps the A 2∆ − X 2Π excitation energy is less than 0.5 eV in error and after 50000 iterations the error is only 0.013 eV. The next state of interest is B 2Σ− and this is also the most difficult to describe state in this system. It is characterized by significant 3p-2h contributions in addition to dominant 2p-1h correlations, and as a result, the EA-EOMCC(2p-1h) approach fails miserably to describe this state, producing ∼160 millihartree errors relative to EA- EOMCC(3p-2h). As we start including 3p-2h determinants in the P space with the help of CIQMC wave function propagations, the errors decrease at an unbelievable rate (see Fig. 4.2). For example, just after 4000 MC iterations and with only 4% of 3p-2h determinants in the P space, the huge > 150 millihartree errors decrease to 12.217 millihartree. After 30000 MC iterations, the CIQMC wave function samplings select 22.8% of 3p-2h determinants in the P space and the EA-EOMCC(P ) energy at this stage is only ∼1 millihartree from EA-EOMCC(3p-2h). This extremely fast convergence in the total electronic energy is also reflected in the rapid convergence of the B 2Σ− − X 2Π gap. At 4000 MC time steps this gap is already within 0.2 eV from the EA-EOMCC(3p-2h) gap and as soon as we reach 30000 119 MC iterations this B 2Σ− − X 2Π gap is < 0.01 eV from the target. For the final C 2Σ+ state for this system in our study, the errors in the EA-EOMCC(2p-1h) energy is > 110 millihartree relative to EA-EOMCC(3p-2h). For this state also, with incorporation of 3p-2h determinants in the P space, the errors decrease rapidly. At 10000 MC iterations and with 18% of 3p-2h determinants in the P space the errors in total electronic energies are reduced to half the errors in EA-EOMCCSD. As we incorporate more and more 3p-2h terms in the P space, the errors keep decreasing further. In terms of the C 2Σ+ − X 2Π gap, the results are even more favorable. For example, the 2.913 eV error at τ = 0 reduces to 0.432 eV at 30000 MC iterations and this error decreases even further to less than 0.1 eV if we let QMC run an additional 10000 iterations. 120 Table 4.1 Convergence of the EA-EOMCC(P ) energies [abbreviated as EA(P )] of the X 2Π, A 2∆, B 2Σ−, and C 2Σ+ states of C2N, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and of the corresponding adiabatic excitation energies toward their parent EA-EOMCC(3p-2h) values. The geometries of the X 2Π, A 2∆, B 2Σ−, and C 2Σ+ states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the EA-EOMCC(P ) calculations were defined as all 1p and 2p-1h determinants and subsets of 3p-2h determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the EA-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determinants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the carbon and nitrogen atoms were kept frozen. MC Iters. 0 4000 10000 20000 30000 40000 50000 60000 100000 ∞ X 2Π EA(P )a %(3p-2h)b 4.696d 5.634 2.686 1.421 0.883 0.679 0.428 0.253 0.071 −130.404919e 0 8.1 18.8 30.2 39.8 48.1 57.1 64.8 82.4 A 2∆ EA(P )a %(3p-2h)b 119.913d 104.423 56.413 18.608 6.901 4.193 0.907 0.582 0.011 0 8.1 21.6 38.3 50.1 60.5 68.0 74.1 90.1 −130.292647e B 2Σ− EA(P )a %(3p-2h)b 158.397d 12.217 5.812 2.407 1.212 0.757 0.306 0.405 0.045 −130.269764e 0 4.3 9.3 15.8 22.8 30.4 38.2 47.2 76.3 c C 2Σ+ c ∆E2 EA(P )a %(3p-2h)b ∆E1 111.733d 0 7.1 98.803 17.7 66.416 31.0 33.336 42.7 16.752 52.6 9.944 60.9 4.400 67.7 3.028 85.1 0.180 Ad. Excit. Energy c ∆E3 3.135d 4.182d 2.913d 2.535 0.179 2.688 1.734 0.085 1.462 0.868 0.027 0.468 0.164 0.432 0.009 0.252 0.002 0.096 0.013 −0.003 0.108 0.076 0.004 0.009 −0.002 −0.001 0.003 3.678f 3.809f 3.055f −130.264924e a Unless otherwise stated, all energies are reported as errors relative to EA-EOMCC(3p-2h) in millihartree. b The %(3p-2h) values are the percentages of 3p-2h determinants captured during the i-FCIQMC propagations (the Sz = 1/2 3p-2h determinants of the B1 symmetry in the case of the X 2Π state, the A1 symmetry in the case of the A 2∆ state, the A2 symmetry in the case of the B 2Σ− state, and the A1 symmetry in the case of the C 2Σ+ state). c Unless otherwise specified, the adiabatic excitation energies are reported as errors relative to EA-EOMCC(3p-2h) in eV; ∆E1 = A 2∆ − X 2Π, ∆E2 = B 2Σ− − X 2Π, and ∆E3 = C 2Σ+ − X 2Π. d Equivalent to EA-EOMCC(2p-1h). e Total EA-EOMCC(3p-2h) energy in hartree. f The EA-EOMCC(3p-2h) adiabatic excitation energy in eV. 121 Figure 4.2 Convergence of (a) the EA-EOMCC(P ) energies of the X 2Π, A 2∆, B 2Σ−, and C 2Σ+ states of C2N, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energies toward their parent EA-EOMCC(3p-2h) values. 122 020406080100MC Iterations (103)04080120160Error rel. to EA-EOMCC(3p-2h) ( Eh)(a)X 2ΠA 2ΔB 2Σ−C 2Σ+020406080100MC Iterations (103)012345Error rel. to EA-EOMCC(3p-2h) (eV)(b)A 2Δ− X 2ΠB 2Σ−− X 2ΠC 2Σ+− X 2Π 4.2.1.2 CNC The next example studied using the semi-stochastic EA-EOMCC(P ) approach is the linear, D∞h symmetric, CNC radical, where we considered the ground X 2Πg and the two low-lying A 2∆u and B 2Σ+ u doublet excited electronic states and the corresponding adiabatic excita- tion energies, A 2∆u − X 2Πg and B 2Σ+ u − X 2Πg. Following Ref. [162], we used the EA SAC-CI-SDT-R/PS with DZP[4s2p1d] optimized geometries, which are RCN = 1.253 ˚A for the X 2Πg state, RCN = 1.256 ˚A for the A 2∆u, and RCN = 1.259 ˚A for the B 2Σ+ u state. At these geometries, the X 2Πg state is dominated by 1p excitations out of the closed-shell reference state CNC+, whereas the A 2∆u and B 2Σ+ u states are characterized by significant two-electron excitations, resulting in large 2p-1h contributions in the corresponding wave functions. Therefore, it is of no surprise that the effect of incorporating 3p-2h correlations is more important in the for A 2∆u and B 2Σ+ u states, but not so much in case of the X 2Πg state. So, in order to accurately describe the total electronic energies of the A 2∆u and B 2Σ+ u states, we need at least the EA-EOMCC(3p-2h) level of theory and this is reflected in the A 2∆u − X 2Πg and B 2Σ+ u − X 2Πg adiabatic excitation energies. It can be seen from Table 1 of Ref. [162] or Table IV of Ref. [68] (cf., also, Table 4.2 of this work) that the EA- EOMCCSD predicted A 2∆u − X 2Πg and B 2Σ+ u − X 2Πg adiabatic excitation energies are 7.206 eV and 7.639 eV, respectively, which are as much as 3.3–3.5 eV away from the experi- mentally obtained values of 3.761 eV and 4.315 eV [258]. When the EA-EOMCC calculations include all possible 3p-2h correlations, the resulting EA-EOMCC(3p-2h) approach predicts these excitation energies to be 4.105 eV and 4.718 eV, respectively, which deviate from the experimental values by only 0.3–0.4 eV. The active-space EA-EOMCCSDt method also per- form very well, not only improving the EA-EOMCCSD results, but also yielding results as good as the target EA-EOMCC(3p-2h) methodology. Our target in this work is, therefore, to reproduce the EA-EOMCC(3p-2h) energetics using the semi-stochastic, i-FCIQMC-driven, EA-EOMCC(P ) approach. In all of our i-FCIQMC calculations for CNC, we used the D2h Abelian subgroup of its true point group D∞h. In particular, following the computational 123 protocol described earlier, the underlying i-FCIQMC calculations required to select the lists of 3p-2h determinants defining the R(MC) 3p-2h components in the EA-EOMCC(P ) calculations for CNC were set to converge the lowest energy states of B3g, Au, and B1u symmetry for the X 2Πg, A 2∆u, and B 2Σ+ u states at their respective equilibrium geometries. The results of our semi-stochastic EA-EOMCC(P ) approach in recovering EA-EOMCC(3p-2h)-quality energetics for the X 2Πg, A 2∆u, and B 2Σ+ u states of CNC and the corresponding A 2∆u − X 2Πg and B 2Σ+ u − X 2Πg excitation energies are reported in Table 4.2 and Fig. 4.3 of this subsection. As previously mentioned, the first row of Table 4.2 refers to EA-EOMCC(P ) results for τ = 0, which are equivalent to EA-EOMCCSD and as already discussed above, it pro- vides a very poor description of the lowest-lying doublet states of CNC. As the QMC wave function evolves in time, which in this context refers to accumulating more and more 3p-2h determinants that can be used to enrich the P spaces in the semi-stochastic calculations, EA- EOMCC(P ) systematically converges to our target EA-EOMCC(3p-2h). The X 2Πg state, that is 4.881 millihartree away from EA-EOMCC(3p-2h) at τ = 0, narrows down to within ∼1 millihartree of EA-EOMCC(3p-2h) with only 16% of 3p-2h determinants in the P space and this happens only at 20000 MC iterations. If the i-FCIQMC propagation is allowed to run an additional 10000 MC iteration, the EA-EOMCC(P) energies are less than 1 milli- hartree away from target and the percentage of 3p-2h determinants in this case is only 21%. For the remaining two states, A 2∆u and B 2Σ+ u , the gradual increase in the 3p-2h determi- nants in the P space, as dictated by the underlying i-FCIQMC wave function propagation, is more pronounced. For the A 2∆u state, only at 4000 MC iterations and with only 5.5% of 3p-2h determinants in the P space, the ∼119 millihartree error in the EA-EOMCC(P |τ =0) [EA-EOMCC(2p-1h] energy, relative to EA-EOMCC(3p-2h) steeply decreases to 12 milli- hartree and this happens only at 4000 MC iterations. If we let the i-FCIQMC propagations to run longer, the EA-EOMCC(P) energies converge to the target EA-EOMCC(3p-2h) in a very rapid pace. For example, at 10000 MC iteration and with 12% of 3p-2h determinants 124 in the P space, the errors reduce to less than 5 millihartree and by the time 20%–30% 3p-2h determinants are captured, EA-EOMCC(P) is only 1–2 millihartree away from the target. In case of the A 2∆u − X 2Πg adiabatic excitation energy the situation is much more favorable. Already at 4000 MC iteration, the errors are less than 0.2 eV relative to EA-EOMCC(3p-2h) and after 10000 MC iterations the errors are always much lower than 0.1 eV. For the final, and most difficult to describe, B 2Σ+ u state, at 10000 MC iterations and with 18% of 3p-2h determinants, the errors in the EA-EOMCC(P) energies are reduced by almost a factor of 3 compared to EA-EOMCC(P |τ =0), which is ∼112 millihartree away from EA-EOMCC(3p- 2h). This error is further reduced to ∼5 millihartree or better when we include about 40% or more 3p-2h determinants in the P space. Interestingly, the B 2Σ+ u − X 2Πg adiabatic excitation energies are much more well behaved, and only at 20000 MC iterations and with 30% of 3p-2h determinants in the P space corresponding to the EA-EOMCC(P ) calculation pertaining to the B 2Σ+ u state, this energy difference is converged to within 0.3 eV relative to EA-EOMCC(3p-2h). The two above examples demonstrate that the semi-stochastic EA-EOMCC(P ) method is capable of rapidly converging the EA-EOMCC(3p-2h) energetics for the individual electronic states and the corresponding adiabatic excitation energies of radicals with a very small percentage of 3p-2h determinants in the respective P spaces. 125 Table 4.2 Convergence of the EA-EOMCC(P ) energies [abbreviated as EA(P )] of the X 2Πg, A 2∆u, and B 2Σ+ u states of CNC, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and of the corresponding adiabatic excitation energies toward their parent EA-EOMCC(3p-2h) values. The geometries of the X 2Πg, A 2∆u, and B 2Σ+ u states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the EA-EOMCC(P ) calculations were defined as all 1p and 2p-1h determinants and subsets of 3p-2h determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the EA-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determinants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the carbon and nitrogen atoms were kept frozen. MC Iters. 0 4000 10000 20000 30000 40000 50000 60000 100000 ∞ X 2Πg EA(P )a %(3p-2h)b 4.881d 4.591 2.419 1.370 0.817 0.572 0.369 0.264 0.040 0 5.1 11.1 16.0 20.9 24.5 28.8 32.0 40.9 −130.411530e A 2∆u EA(P )a %(3p-2h)b 118.827d 11.668 4.816 1.851 0.865 0.407 0.194 0.096 0.015 0 5.5 12.0 20.6 27.5 35.4 43.4 52.6 79.7 −130.260673e B 2Σ+ u Ad. Excit. Energy c c EA(P )a %(3p-2h)b ∆E1 ∆E2 2.921d 3.101d 112.241d 0 2.419 0.193 8.1 93.500 1.164 0.065 18.0 45.183 0.299 0.013 31.3 12.348 0.124 42.2 5.369 0.001 0.045 −0.004 51.1 2.221 0.036 −0.005 59.1 1.707 −0.005 −0.001 67.0 0.228 −0.001 −0.001 86.9 0.006 4.718f 4.105f −130.238150e a Unless otherwise stated, all energies are reported as errors relative to EA-EOMCC(3p-2h) in millihartree. b The %(3p-2h) values are the percentages of 3p-2h determinants captured during the i-FCIQMC propagations (the Sz = 1/2 3p-2h determinants of the B3g symmetry in the case of the X 2Πg state, the Au symmetry in the case of the A 2∆u state, the B1u symmetry in the case of the B 2Σ+ u state). c Unless otherwise specified, the adiabatic excitation energies are reported as errors relative to EA-EOMCC(3p-2h) in eV; ∆E1 = A 2∆u − X 2Πg and ∆E2 = B 2Σ+ u − X 2Πg. d Equivalent to EA-EOMCC(2p-1h). e Total EA-EOMCC(3p-2h) energy in hartree. f The EA-EOMCC(3p-2h) adiabatic excitation energy in eV. 126 Figure 4.3 Convergence of (a) the EA-EOMCC(P ) energies of the X 2Πg, A 2∆u, and B 2Σ+ u states of CNC, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energies toward their parent EA-EOMCC(3p-2h) values. 127 020406080100MC Iterations (103)04080120160Error rel. to EA-EOMCC(3p-2h) ( Eh)(a)X 2ΠgA 2ΔuB 2Σ+u020406080100MC Iterations (103)012345Error rel. to EA-EOMCC(3p-2h) (eV)(b)A 2Δu− X 2ΠgB 2Σ+u− X 2Πg 4.2.2 Application of the IP-EOMCC(P ) approach to N3 and NCO radicals After examining the performance of the EA-EOMCC(P) methodology in converging the high-level EA-EOMCC(3p-2h) energetics, we proceed to assess the capability of the IP- EOMCC(P) approach in recovering the IP-EOMCC(3h-2h) energetics by applying the IP- EOMCC(P) approach in computing the total electronic energies of the ground and a few excited doublet states of N3 and NCO radicals and the corresponding adiabatic excitation energies. The reference wave functions were obtained by performing CCSD calculations on the N− 3 and NCO− anions using the DZP basis set [167, 168] and geometries obtained from IP SAC-CI-SDT-R/PS / DZP optimizations as reported in Ref. [162]. 4.2.2.1 N3 Our first system under investigation is the N3 radical, where we studied the X 2Πg and B 2Σ+ u states, along with the corresponding B 2Σ+ u − X 2Πg adiabatic excitation energy. For this, we utilized the IP SAC-CI-SDT-R/PS (in combination with the DZP[4s2p1d] basis set) optimized geometries obtained from Ref. [162]. These geometries are RN−N = 1.188 ˚A for the X 2Πg state and RN−N = 1.185 ˚A for the B 2Σ+ u state. At these geometries both the states are characterized by predominant 1h excitations from the reference wave function N3 −, with the B 2Σ+ u state having some 2h-1p contributions. Consequently, the effect of including higher-level correlations, such as 3h-2p, is less pronounced than in C2N and CNC. Nevertheless, there are still noticeable improvements when going from IP-EOMCC(2h-1p) to IP-EOMCC(3h-2p). As evidenced in Table 1 of Ref. [162] and Table V of Ref. [68] (cf., also Table 4.3 of this work), the lower-order IP-EOMCCSD approach estimates the B 2Σ+ u − X 2Πg energy to be 4.640 eV, which exhibits a discrepancy of about 0.1 eV compared to the experimental value of 4.555 eV reported in Ref. [258]. The IP-EOMCC(3h-2p) method, that includes a complete treatment of 3h-2p components of R(+1) µ , reduces this error to 0.04 eV, marking a significant improvement. Therefore, it would be interesting to see how effective the i-FCIQMC driven IP-EOMCC(P) approaches are in recovering IP-EOMCC(3h- 2p) energetics. In all the i-FCIQMC calculations the D2h Abelian subgroup of the true point 128 group of the molecule D∞h was utilized. The underlying i-FCIQMC propagations were set to converge the lowest energy states of B3g (D2h) symmetry for the X 2Πg and the B1u (D2h) symmetry in case of the B 2Σ+ u state at their respective equilibrium geometries. The results of our semi-stochastic IP-EOMCC(P ) approach in recovering total electronic energies of the X 2Πg and B 2Σ+ u states of N3 and the corresponding X 2Πg − B 2Σ+ u adiabatic excitation energy is reported in Table 4.3 and Fig. 4.4. From Table 4.3, it can be observed that at τ = 0, the IP-EOMCC(P ) calculated to- tal electronic energies of the X 2Πg and B 2Σ+ u states are approximately 13-15 millihartree away from the IP-EOMCC(3h-2p) energies. As soon as the accumulation of 3h-2p deter- minants defining the R(MC) µ,3h-2p component of the electron ionizing operator R(−1) µ begins, the IP-EOMCC(P) energies quickly start converging towards the IP-EOMCC(3h-2p) energetics. For the X 2Πg state, at just 4000 MC iterations and with 28.7% of 3h-2p determinants in the P space, the IP-EOMCC(P) approach reduces the 13 millihartree error in the IP-EOMCCSD method to 2.627 millihartree. By 10000 MC iterations, IP-EOMCC(P) is already within ∼1 millihartree relative to IP-EOMCC(3h-2p), with the underlying P space containing about 40% of 3h-2p terms. The situation is similar in case of the B 2Σ+ u state, where the ∼15 millihartree error in the IP-EOMCC(2h-1p) is reduced to 5 millihartree after only 4000 MC iterations and with 32% of 3h-2p determinants. At 10000 MC iterations, this error is further decreased to 1.340 millihartree, and at this state the underlying P space is populated by 48.1% 3h-2p determinants. In case of the X 2Πg − B 2Σ+ u gap, convergence is much easier due to favorable cancellation of errors. Here, IP-EOMCC(2h-1p) already has a small er- ror of 0.042 eV relative to IP-EOMCC(3h-2p), but the benefit of using the IP-EOMCC(P) approach is still evident. At only 10000 MC iterations, this already small energy gap of 0.042 eV is further decreased to 0.006 eV, demonstrating the ability of the semi-stochastic IP-EOMCC(P ) approach to improve the IP-EOMCCSD energetics, even when the scope of improvement is not very large. 129 Table 4.3 Convergence of the IP-EOMCC(P ) energies [abbreviated as IP(P )] of the X 2Πg and B 2Σ+ u states of N3, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and of the corresponding adiabatic excitation energy toward their parent IP-EOMCC(3h-2p) values. The geometries of the X 2Πg and B 2Σ+ u states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the IP-EOMCC(P ) calculations were defined as all 1h and 2h-1p determinants and subsets of 3h-2p determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the IP-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determinants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the nitrogen atoms were kept frozen. MC Iters. 0 4000 10000 20000 30000 40000 50000 60000 100000 ∞ X 2Πg IP(P )a %(3h-2p)b 13.078d 2.627 1.127 0.488 0.271 0.189 0.104 0.082 0.052 0 28.7 41.4 53.8 61.3 68.4 72.5 75.9 84.3 −163.729333e B 2Σ+ u IP(P )a %(3h-2p)b 14.623d 5.035 1.340 0.343 0.134 0.036 0.012 0.004 0.000 0 31.9 48.1 61.1 68.2 73.9 77.2 80.1 85.7 −163.560374e Ad. Excit. Energy ∆Ec 0.042d 0.066 0.006 −0.004 −0.004 −0.004 −0.003 −0.002 −0.001 4.598f a Unless otherwise stated, all energies are reported as errors relative to IP-EOMCC(3h-2p) in millihartree. b The %(3h-2p) values are the percentages of 3h-2p determinants captured during the i-FCIQMC propaga- tions (the Sz = 1/2 3h-2p determinants of the B3g symmetry in the case of the X 2Πg state and the B1u symmetry in the case of the B 2Σ+ c Unless otherwise specified, the adiabatic excitation energies are reported as errors relative to IP- EOMCC(3h-2p) in eV; ∆E = B 2Σ+ d Equivalent to IP-EOMCC(2h-1p). e Total IP-EOMCC(3h-2p) energy in hartree. f The IP-EOMCC(3h-2p) adiabatic excitation energy in eV. u − X 2Πg. u state). 130 Figure 4.4 Convergence of (a) the IP-EOMCC(P ) energies of the X 2Πg and B 2Σ+ u states of N3, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energy toward their parent IP-EOMCC(3h-2p) values. 131 020406080100MC Iterations (103)0481216Error rel. to IP-EOMCC(3h-2p) (mEh)(a)X 2ΠgB 2Σ+u020406080100MC Iterati ns (103)0.000.020.040.060.080.10Err r rel. t IP-EOMCC(3h-2p) (eV)(b)B 2Σ+u− X 2Πg 4.2.2.2 NCO Our final example is the NCO radical, where we investigated the ground X 2Π state and two low-lying valence excited doublet states A 2∆ and B 2Π and the corresponding A 2∆ − X 2Π and B 2Π − X 2Π excitation energies. Following Ref. [162], we employed the IP SAC-CI-SDT-R/PS / DZP[4s2p1d] optimized geometries, which are RNC = 1.230 ˚A and RCO = 1.193 ˚A for the X 2Π state, RNC = 1.191 ˚A and RCO = 1.190 ˚A for the A 2∆ state, and RNC = 1.220 ˚A and RCO = 1.309 ˚A for the B 2Π state. At these geometries, all the low-lying states mentioned above are dominated by 1h excitations from the closed-shell reference ion NCO− with some 2h-1p contributions being important for the B 2Π state. So, the basic IP-EOMCCSD approach is fairly reasonable in this case, however, similar to the case of N3, the effect of incorporating 3h-2p correlations using the Rµ,3h-2p component of the electron ionizing operator R(−1) µ offers non-negligible improvements. This is evident from Table 1 of Ref. [162] and Table V of Ref. [68] (see also, Table 4.4 of this work), where the IP-EOMCCSD method predicts the A 2∆ − X 2Π and B 2Π − X 2Π excitation energies to be 2.900 eV and 4.199 eV, respectively, which are ∼0.1–0.2 eV from the experimental results of 2.821 eV and 3.937 eV reported in Ref. [258]. The IP-EOMCC(3h-2p) approach further improves these results bringing them to within 0.04 eV or better relative to the experimental values. Now, it would be interesting to see how the semi-stochastic, i-FCIQMC-driven, IP- EOMCC(P) method performs in this case. All the results pertaining to the IP-EOMCC(P) approach are reported in Table 4.4 and Fig. 4.5. In all the i-FCIQMC calculations we used the C2v Abelian subgroup of the true point group of NCO, C∞v. In particular the underlying i-FCIQMC calculations were set to converge the lowest energy states of B2, A1, and B2 symmetry for the X 2Π, A 2∆, and B 2Π states at the respective equilibrium geometries reported above. The first row of Table 4.4 contains IP-EOMCC(P) results for τ = 0, which is the start of i-FCIQMC wave function propagations. At this stage, the P spaces utilized in the IP- EOMCC(P) calculations do not contain any 3h-2p determinants, and, as described earlier, 132 this is equivalent to the IP-EOMCCSD method and for the X 2Π state, this energy is about 10 millihartree away from the high-level IP-EOMCC(3h-2p) energy. This error is readily decreased to 3.503 millihartree as soon as the i-FCIQMC propagations complete 4000 steps and recover 23% of 3h-2p determinants. When the i-FCIQMC propagations reach 10000 MC iterations, this error is further reduced to 1.862 millihartree and with an additional 10000 MC iterations the errors are in the sub-millihartree regime. In case of the A 2∆ state, IP-EOMCC(2h-1p) has an error of 11 millihartree relative to IP-EOMCC(3h-2p). At 4000 MC iterations and with ∼24% of 3h-2p determinants in the P space, the IP-EOMCC(P) approach brings this error down to 3 millihartree and after 10000 MC iterations the IP- EOMCC(P) energy is in the sub-millihartree region relative to IP-EOMCC(3h-2p). At this stage, the P space contains 37.4% of 3h-2p determinants pertaining to the A 2∆ state. In case of the A 2∆ − X 2Π energy gap, the convergence is even faster. At the start of the i-FCIQMC propagation, when the P spaces do not contain any 3h-2p determinants, the error in the adiabatic excitation energy relative to IP-EOMCC(3h-2p) is only 0.036 eV. However, as soon as we start including 3h-2p determinants in the P spaces pertaining to the IP-EOMCC(P) computations, this error rapidly decreases as can be seen from Table 4.4 and Fig. 4.5. In case of the B 2Π state, which is the most difficult to describe between the three states of NCO investigated here, the IP-EOMCCSD method produces an error of > 20 millihartree relative to the IP-EOMCC(3h-2p) method. At 4000 MC iterations and with 26% of 3h-2p determinants in the P space, the IP-EOMCC(P) energies reach within ∼5 millihartree to the IP-EOMCC(3h-2p) energetics. With an additional 6000 MC iterations, this is further reduced to 1.964 millihartree and at this stage the P spaces contain ∼42% of 3h-2p determinants. The B 2Π − X 2Π adiabatic excitation energies are much easier to converge to the IP-EOMCC(3h-2p) results. The IP-EOMCCSD predicted energy gap, which has an error of 0.288 eV is already within 0.05 eV at 4000 MC iterations and at 1000 MC iterations this error is a mere 0.003 eV. The two above examples demonstrate that the semi-stochastic IP-EOMCC(P ) method is 133 capable of rapidly converging the IP-EOMCC(3h-2p) energetics for the individual electronic states as well as the corresponding adiabatic excitation energies of radicals with very small percentages of 3h-2p determinants in the respective P spaces. 134 Table 4.4 Convergence of the IP-EOMCC(P ) energies [abbreviated as IP(P )] of the X 2Π, A 2∆, and B 2Π states of NCO, as described by the DZP[4s2p1d] basis set of Refs. [167, 168], and the corresponding adiabatic excitation energies toward their parent IP-EOMCC(3h-2p) values. The geometries of the X 2Π, A 2∆u, and B 2Π states, optimized in the SAC-CI SDT-R/PS calculations using the same basis set, were taken from Ref. [162]. The P spaces used in the IP-EOMCC(P ) calculations were defined as all 1h and 2h-1p determinants and subsets of 3h-2p determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations preceding the IP-EOMCC(P ) steps were initiated by placing 500 walkers on the ROHF reference determinants of the corresponding states and the na parameter of the initiator algorithm was set at 3. In all post-Hartree–Fock calculations, the lowest core orbitals of the carbon, nitrogen, and oxygen atoms were kept frozen. MC Iters. 0 4000 10000 20000 30000 40000 50000 60000 100000 ∞ X 2Π IP(P )a %(3h-2p)b 9.587d 3.503 1.862 0.973 0.428 0.329 0.164 0.097 0.011 0 23.2 34.9 46.6 55.0 62.5 68.1 72.4 81.8 −167.591596e A 2∆ IP(P )a %(3h-2p)b 10.921d 3.198 0.991 0.359 0.153 0.102 0.032 0.015 0.000 0 23.6 37.4 52.5 61.4 68.0 72.6 76.1 84.1 −167.486358e B 2Π Ad. Excit. Energy c c IP(P )a %(3h-2p)b ∆E1 ∆E2 0.288d 0.036d 20.154d 0 0.049 −0.008 25.6 5.302 −0.024 41.5 1.964 0.003 −0.017 −0.001 54.9 0.943 −0.008 63.9 0.428 0.000 −0.006 −0.012 70.2 -0.106 −0.004 −0.004 74.8 0.003 0.001 −0.002 77.9 0.133 0.000 0.000 85.2 0.002 3.911f 2.864f −167.447865e a Unless otherwise stated, all energies are reported as errors relative to IP-EOMCC(3h-2p) in millihartree. b The %(3h-2p) values are the percentages of 3h-2p determinants captured during the i-FCIQMC propagations (the Sz = 1/2 3h-2p determinants of the B2 symmetry in the case of the X 2Π state, the A1 symmetry in the case of the A 2∆ state, and the B2 symmetry in the case of the B 2Π state). c Unless otherwise specified, the adiabatic excitation energies are reported as errors relative to IP-EOMCC(3h-2p) in eV; ∆E1 = A 2∆ − X 2Π and ∆E2 = B 2Π − X 2Π. d Equivalent to IP-EOMCC(2h-1p). e Total IP-EOMCC(3h-2p) energy in hartree. f The IP-EOMCC(3h-2p) adiabatic excitation energy in eV. 135 Figure 4.5 Convergence of (a) the IP-EOMCC(P ) energies of the X 2Π, A 2∆, and B 2Π states of NCO, as described by the DZP[4s2p1d] basis set, and (b) the corresponding adiabatic excitation energies toward their parent IP-EOMCC(3h-2p) values. 136 020406080100MC Iterations (103)06121824Error rel. to IP-EOMCC(3h-2p) (mEh)(a)X 2ΠA 2ΔB 2Π020406080100MC Iterati ns (103)Δ0.10.00.10.20.30.4Err r rel. t IP-EOMCC(3h-2p) (eV)(b)A 2Δ− X 2ΠB 2Π− X 2Π 4.3 Singlet–Triplet Gaps in Methylene and Trimethylenemethane After exploring the performance of the semi-stochastic EA-EOMCC(P) and IP-EOMCC(P) methods, which describe the (1,0) and (0,1) sectors of the Fock space, we move on to their extensions to the (2,0) and (0,2) sectors of the Fock space, which we refer to as the semi-stochastic DEA- [DEA-EOMCC(P)] and DIP-EOMCC [DIP-EOMCC(P)] approaches. These methods are particularly suitable for studying singlet and triplet states of biradical systems and the corresponding singlet–triplet gaps. To explore the capability of the semi- stochastic DEA-EOMCC(P) and DIP-EOMCC(P) approaches in converging their high-level fully-deterministic parent DEA-EOMCC(4p-2h) and DIP-EOMCC(4h-2p) methods, we car- ried out benchmark calculations for the lowest singlet and triplet states along with the corresponding singlet–triplet gaps (∆ES-T = Esinglet − Etriplet) in the CH2 [Tables 4.5 and 4.6 and Figs. 4.6 and 4.7] and trimethylenemethane [Table 4.7 and Figs. 4.9 and 4.10]. We utilized the TZ2P basis set [180] and FCI/TZ2P geometries as reported in Ref. [181] for the CH2 molecule and for trimethylenemethane, we employed the 6-31G(d) basis set [178, 179] with the SF-DFT/6-31G(d) geometries reported in Ref. [231]. For the smaller CH2 molecule we used i-FCIQMC, while for the larger trimethylenemethane system we exploited the i- CISDTQ-MC propagations to generate the lists of 4p-2h and 4h-2p determinants entering the P spaces for the DEA-EOMCC(P) and DIP-EOMCC(P) calculations. Again, following our previous semi-stochastic work [99–101, 137], we used the HANDE software package [170, 171] to execute all our QMC calculations. Our standalone CC/EOMCC codes, interfaced with the RHF, ROHF, and integral transformation routines available in the GAMESS software package [172–174], were used to carry out the DEA-EOMCC(P ), DIP-EOMCC(P ), and the fully deterministic CCSD, DEA-EOMCC(3p-1h), DEA-EOMCC(4p-2h), DIP-EOMCC(3h- 1p), and DIP-EOMCC(4h-2p) computations. Each i-FCIQMC and i-CISDTQ-MC run was initiated by placing 500 walkers on the relevant reference function (see Tables 4.5–4.7 for the details), the initiator parameter na was set at 3, and all of the i-FCIQMC and i-CISDTQMC propagations used a time step of τ = 0.0001 a.u. If the true point group of the system 137 of interest was not Abelian, which was the case for trimethylenemethane, we utilized the largest Abelian subgroups in the calculations. This choice was necessary, since all of our CC/EOMCC codes interfaced with GAMESS and the CIQMC routines in HANDE can only handle Abelian symmetries. In all the post-HF computations, the core electrons correspond- ing to the 1s shells of the carbon atom was kept frozen. 4.3.1 Methylene The discussions for the semi-stochastic DEA- and DIP-EOMCC methods, DEA-EOMCC(P) and DIP-EOMCC(P), begin with the results for total electronic energies of the X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene and the corresponding singlet–triplet gaps. In the DEA-EOMCC calculations we used the CH2+ 2 dication as the reference and for the DIP- EOMCC approach the CH2− 2 dianion was used as the reference. The geometries used were the FCI/TZ2P geometries as reported in Ref. [181] for the CH2 molecule and throughout this work we utilized the TZ2P basis set [180]. All the results of the DEA-EOMCC(P) and DIP-EOMCC(P) calculations are reported in Tables 4.5 and 4.6 and Figs. 4.6 and 4.7. The ground X 3B1 and the second excited state B 1B1 are characterized as having a SR char- acter that can be well represented by high-spin triplet and open-shell singlet configurations of (1a1)2(2a1)2(1b2)2(3a1)1(1b1)1 type. The first and the third excited states, A 1A1 and C 1A1, have a significant MR character originating due to the mixing of the two configura- tions (1a1)2(2a1)2(1b2)2(3a1)2 and (1a1)2(2a1)2(1b2)2(1b1)2, which makes it very challenging for many electronic structure methods to provide an accurate description. As a result, in order to obtain accurate results for the A 1A1 − X 3B1, B 1B1 − X 3B1, and C 1A1 − X 3B1 singlet–triplet gaps, one needs to use methods that can offer well balanced treatments of both dynamical and nondynamical correlations. This makes CH2 a very popular example to test the performance of newly developed electronic structure methods. The most basic DEA-EOMCCSD [DEA-EOMCC(3p-1h)] and DIP-EOMCCSD [DIP-EOMCC(3h-1p)] fail to provide an accurate description of the triplet ground and the low-lying singlet states of methylene and the corresponding singlet–triplet gaps, and one needs to incorporate 4p-2h 138 and 4h-2p correlations for an accurate description. From Table 1 of Ref. [153] one can com- pare the FCI, DEA-EOMCC(3p-1h), DEA-EOMCC(4p-2h), DIP-EOMCC(3h-1p), and DIP- EOMCC(4h-2p) calculated singlet–triplet gaps A 1A1 − X 3B1, B 1B1 − X 3B1, and C 1A1 − X 3B1. It is evident that the DEA-EOMCC(3p-1h) predicted gaps are −0.11, −1.89, and −3.64 kcal/mol in error compared to the FCI singlet–triplet gaps of 11.14, 35.59, and 61.67 kcal/mol, respectively. The complete incorporation of 4p-2h correlations via the DEA- EOMCC(4p-2h) methodology improves these results, especially for the second and third gaps, bringing down the errors to 0.38, −0.02, and 0.21 kcal/mol compared to FCI. On the other hand, the DIP-EOMCC(3h-1p) method predicts these singlet–triplet gaps to be −4.53, −3.22, and −4.63 kcal/mol away from FCI. We see a significant improvement by using the DIP-EOMCC(4h-2p) approach, which reduces these errors to only −0.44, −0.51, and −0.48 kcal/mol relative to FCI. So, it would be interesting to see how our DEA-EOMCC(P) and DIP-EOMCC(P) approaches perform in this interesting case of methylene. We used the i-FCIQMC method to extract the lists of 4p-2h determinants, in case of DEA-EOMCC(P), and the 4h-2p determinants, in case of the DIP-EOMCC(P) approach. The i-FCIQMC runs were initiated on the ROHF determinant of B1 symmetry for the X 3B1 state and the RHF determinants of A1 symmetries for the A 1A1 and C 1A1 states, and the A2 sym- metry B 1B1, and with the one- and two-body integrals obtained from the reference ionic systems (CH2+ 2 in case of DEA-EOMCC(P) and CH2− 2 for the DIP-EOMCC(P) case). In the next paragraph, we explore the performance of the semi-stochastic, i-FCIQMC-driven, DEA-EOMCC(P) approach and in the following paragraph we investigate the performance of the DIP-EOMCC(P) approach. From Table 4.5, one can notice that at τ = 0, the DEA-EOMCC(P) [equivalent to the DEA-EOMCC(3p-1h)] energies are 14 millihartree in error compared to DEA-EOMCC(4p- 2h) for the X 3B1 state. As soon as we start incorporating 4p-2h determinants in the P spaces defining the R(MC) µ,4p-2h components of R(+2) µ operator in the DEA-EOMCC(P) calculations, the energies rapidly converge to DEA-EOMCC(4p-2h). At 4000 MC iterations and with just 139 1.8% of 4p-2h determinants in the P space, the DEA-EOMCC(P) errors are already within 5 millihartree and as soon as the i-FCIQMC propagations select 7% of 4p-2h determinants, the error in the DEA-EOMCC(P) energy becomes only ∼1 millihartree. This happens at just 10000 MC iterations, and if we allow i-FCIQMC to complete 50000 MC steps, the P spaces contain 15.3% of 4p-2h determinants and the error in energy of the X 3B1 state reaches the sub-millihartree region. This fast convergence to DEA-EOMCC(4p-2h) remains true even when we go to the first excited singlet state A 1A1, where the > 13 millihartree error in the DEA-EOMCC(3p-1h) energies are quickly reduced to < 2 millihartree at only 10000 MC iterations and with just 4.3% of 4p-2h determinants in the P space. The situation improves even more when the i-FCIQMC propagations reach 20000 MC time steps, where with only 7.6% of 4p-2h determinants the total electronic energy is already in the sub- millihartree regime. If we look at the A 1A1 − X 3B1 singlet triplet gap, it can be seen that the 0.487 kcal/mol error at τ = 0 is quickly reduced to 0.2 kcal/mol at already 10000 MC iterations and by the time we reach 50000 MC iterations, the error is less than 0.1 kcal/mol relative to DEA-EOMCC(4p-2h). For the next state B 1B1, which is mostly single reference in nature, the error at τ = 0 is 11.129 millihartree. However, with only ∼5% and ∼8% of 4p-2h determinants in the P spaces and at just 10000 and 20000 MC iterations, respectively, the ∼2 millihartree and ∼1 millihartree marks are reached by the errors in the total electronic energy predicted by DEA-EOMCC(P). The situation in the corresponding B 1B1 − X 3B1 singlet–triplet gap is more favorable. For example, the 1.875 kcal/mol errors in the DEA-EOMCC(P |τ =0) method is quickly brought down to 0.1 kcal/mol relative to DEA- EOMCC(4p-2h) at only 10000 MC time steps, with additional iterations further improving the results as seen in Table 4.5 and in Fig. 4.6. In case of the most challenging C 1A1 state, the errors in the DEA-EOMCC(3p-1h) energies relative to DEA-EOMCC(4p-2h) is 7.987 millihartree. As the underlying i-FCIQMC wave function propagation gradually select more and more 4p-2h determinants, this error quickly gets down to the sub-millihartree regime. At 20000 MC iteration with 5% of 4p-2h determinants in the P space the DEA- 140 EOMCC(P) energies are about 3 millihartree away from the DEA-EOMCC(4p-2h) method and after 50000 MC iterations and with 13% of 4p-2h determinants in the P space, these errors are less than 1 millihartree. This is also reflected in the C 1A1 − X 3B1 singlet– triplet gap. The 3.847 kcal/mol error at τ = 0 rapidly becomes ∼1 kcal/mol in absolute value at 20000 MC iterations and by the time the i-FCIQMC propagations complete 50000 time steps, the absolute value of the error is only about 0.2 kcal/mol from the high-level DEA-EOMCC(4p-2h) result. It is interesting to note that, due to the different rates of convergence of the low-lying states, the singlet–triplet gaps fluctuate before converging to the DEA-EOMCC(4p-2h) results. 141 Table 4.5 Convergence of the DEA-EOMCC(P ) energies [abbreviated as DEA(P )] of the X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set of Ref. [180], and of the corresponding adiabatic singlet–triplet gaps toward their parent DEA-EOMCC(4p-2h) values. The geometries of the X 3B1, A 1A1, B 1B1, and C 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces used in the DEA-EOMCC(P ) calculations were defined as all 2p and 3p-1h determinants and subsets of 4p-2h determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations were initiated by placing 500 walkers on the corresponding ROHF reference determinant for the X 3B1 state and the corresponding RHF reference determinants for the remaining states and the na parameter of the initiator algorithm was set at 3. In all the post-Hartree–Fock calculations, the lowest core orbital of the carbon atom was kept frozen. MC Iters. DEA(P )a %(4p-2h)b DEA(P )a %(4p-2h)b DEA(P )a %(4p-2h)b DEA(P )a %(4p-2h)b ∆E1 A 1A1 C 1A1 X 3B1 B 1B1 0 4000 10000 20000 50000 100000 200000 ∞ 14.118d 4.863 2.252 1.323 0.448 0.079 0.001 0.0 1.8 4.4 7.0 15.3 38.2 85.2 −39.066449e 13.342d 4.611 1.953 0.996 0.316 0.028 0.000 0.0 1.5 4.3 7.6 20.1 51.9 88.9 −39.048089e 11.129d 4.477 2.085 1.163 0.394 0.050 0.000 0.0 2.0 4.8 7.9 18.3 48.5 89.3 −39.009764e 7.987d 6.344 4.794 3.161 0.701 0.078 0.000 0.0 1.3 3.3 5.1 13.2 40.7 88.0 −38.967833e c ∆E2 Singlet–Triplet Gap c c ∆E3 0.487d 1.875d 3.847d 0.242 −0.929 0.158 0.105 −1.595 0.188 0.101 −1.154 0.205 0.034 −0.158 0.083 0.001 0.018 0.032 0.000 0.000 0.000 11.521f 35.570f 61.881f a Unless otherwise stated, all energies are reported as errors relative to DEA-EOMCC(4p-2h) in millihartree. b The %(4p-2h) values are the percentages of 4p-2h determinants captured during the i-FCIQMC propagations [the Sz = 1 4p-2h determinants of the B1 symmetry in the case of the X 3B1 state, the Sz = 0 4p-2h determinants of the B1 symmetry in case of the B 1B1 state , and the Sz = 0 4p-2h determinants of the A1 symmetry in the case of the A 1A1, and the A1 symmetry in case of the C 1A1 state]. c Unless otherwise specified, the singlet–triplet gaps are reported as errors relative to DEA-EOMCC(4p-2h) in kcal/mol; ∆E1 = A 1A1 − X 3B1, ∆E2 = B 1B1 − X 3B1, and ∆E3 = C 1A1 − X 3B1. d Equivalent to DEA-EOMCC(3p-1h). e Total DEA-EOMCC(4p-2h) energy in hartree. f The DEA-EOMCC(4p-2h) singlet–triplet gap in kcal/mol. 142 Figure 4.6 Convergence of (a) DEA-EOMCC(P ) energies of X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set, and (b) of the corresponding adiabatic singlet–triplet gaps towards their parent DEA-EOMCC(4p-2h) values. 143 04080120160200MC Iterations (103)0246810Error rel. to DEA-EOMCC(4p-2h) (mEh)(a)X 3B1A 1A1B 1B1C 1A104080120160200MC Iterations (103)−101234Error rel. to DEA-EOMCC(4p-2h) (kcal/mol)(b)A 1A1− X 3B1B 1B1− X 3B1C 1A1− X 3B1 Table 4.6 contains the results of our semi-stochastic, i-FCIQMC-driven, DIP-EOMCC(P) computations. At the start of the i-FCIQMC propagations, the total electronic energy of the X 3B1 state calculated by the DIP-EOMCC(P) approach has an error of 26.147 millihartree relative to DIP-EOMCC(4h-2p). This error is readily decreased to about 3 millihartree at only 4000 MC iterations, where the P spaces contain about 35% of 4h-2p determinants in the P space. Allowing the i-FCIQMC propagations to complete a total of 10000 and 20000 MC iterations results in errors of 1.357 and 0.766 millihartree, respectively. The percentage of 4h-2p determinants in the P spaces at this stage are 43.4% and 50.3%, respectively. For the first excited singlet state A 1A1, the 19.624 millihartree error reported at τ = 0 is sharply decreased to 3 millihartree at just 4000 MC iteration and with ∼19% of 4h-2p determinants in the P space. By the time the i-FCIQMC propagations reach 10000 MC time steps, this error is already below 1 millihartree and this happens with 30% of 4h-2p determinants in the P space. Similarly, the −4.093 kcal/mol error in the A 1A1 − X 3B1 singlet–triplet gap narrows down to just −0.152 kcal/mol at 4000 MC iterations and by the time 50000 MC time steps are completed, this error is only −0.028 millihartree away from the DIP-EOMCC(4h- 2p) value. The next state, which is an open-shell singlet state, is referred to as B 1B1, and in this case the ∼22 millihartree errors in the DIP-EOMCC(3h-1p) energetics improves to only 3.681 millihartree at just 4000 MC iterations and with ∼33% of 4h-2p determinants in the P space. Then with i-FCIQMC propagation it steadily converges to DIP-EOMCC(4h-2p). By the time i-FCIQMC completes 20000 MC iterations, the errors in the total electronic energies in this state falls in the sub-millihartree regime. At this stage, the B 1B1 − X 3B1 singlet–triplet gap is already converged to 0.032 kcal/mol compared to the target DIP- EOMCC(4h-2p) values, and this is a significant improvement from the DIP-EOMCC(P |τ =0) result of −2.714. The final state considered here is the C 1A1 state, which has a strong multireference character as described above. In this case, the DIP-EOMCC(P) calculations start with an error of 19.355 millihartree at τ = 0. After 4000 MC time steps and with 17.3% of 4h-2p determinants in the P space, this error is down to 8.869 millihartree. It gradually 144 decreases to 3.480 millihartree at 20000 MC iterations before reaching the sub-millihartree mark at 50000 MC iterations. At these stages, the percentages of the 4h-2p determinants are 36.6% and 51.7%, respectively. The convergence of the C 1A1 − X 3B1 singlet–triplet gap mirrors this convergence pattern. The −4.149 kcal/mol error in this gap shrinks to 1.703 kcal/mol at 20000 MC iterations and by the time one reaches 50000 MC iterations the error is only 0.149 kcal/mol (cf., Fig. 4.7 for a graphical representation). 145 Table 4.6 Convergence of the DIP-EOMCC(P ) energies [abbreviated as DIP(P )] of the X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set of Ref. [180], and of the corresponding adiabatic singlet–triplet gaps toward their parent DIP-EOMCC(4h-2p) values. The geometries of the X 3B1, A 1A1, B 1B1, and C 1A1 states, optimized in the FCI calculations using the TZ2P basis set, were taken from Ref. [181]. The P spaces used in the DIP-EOMCC(P ) calculations were defined as all 2h and 3h-1p determinants and subsets of 4h-2p determinants extracted from the i-FCIQMC propagations with δτ = 0.0001 a.u. The i-FCIQMC calculations were initiated by placing 500 walkers on the corresponding ROHF reference determinant for the X 3B1 state and the corresponding RHF reference determinants for the remaining states and the na parameter of the initiator algorithm was set at 3. In all the post-Hartree–Fock calculations, the lowest core orbital of the carbon atom was kept frozen. MC Iters. DIP(P )a %(4h-2p)b DIP(P )a %(4h-2p)b DIP(P )a %(4h-2p)b DIP(P )a %(4h-2p)b ∆E1 Singlet–Triplet Gap c ∆E3 c ∆E2 c A 1A1 C 1A1 X 3B1 B 1B1 0 4000 10000 20000 50000 100000 200000 ∞ 26.147d 3.289 1.357 0.766 0.131 0.009 0.000 0.0 34.6 43.4 50.3 59.5 68.1 70.6 −39.066449e 19.624d 3.046 0.946 0.524 0.085 0.005 0.000 0.0 19.4 29.8 39.8 55.5 63.8 65.2 −39.048089e 21.821d 3.681 1.853 0.818 0.168 0.009 0.000 0.0 32.9 42.0 48.6 59.0 68.1 70.6 −39.009764e 19.355d 8.869 5.741 3.480 0.369 0.057 0.000 0.0 17.3 28.2 36.6 51.7 62.8 65.2 −38.967833e −4.093d −2.714d −4.149d 3.502 −0.152 0.246 2.751 −0.257 0.311 1.703 −0.152 0.032 0.149 −0.028 0.023 0.030 −0.003 0.000 0.000 0.000 0.000 10.701f 35.083f 61.191f a Unless otherwise stated, all energies are reported as errors relative to DIP-EOMCC(4h-2p) in millihartree. b The %(4h-2p) values are the percentages of 4h-2p determinants captured during the i-FCIQMC propagations [the Sz = 1 4h-2p determinants of the B1 symmetry in the case of the X 3B1 state, the Sz = 0 4h-2p determinants of the B1 symmetry in case of the B 1B1 state , and the Sz = 0 4h-2p determinants of the A1 symmetry in the case of the A 1A1, and the A1 symmetry in case of the C 1A1 state]. c Unless otherwise specified, the singlet–triplet gaps are reported as errors relative to DIP-EOMCC(4h-2p) in kcal/mol; ∆E1 = A 1A1 − X 3B1, ∆E2 = B 1B1 − X 3B1, and ∆E3 = C 1A1 − X 3B1. d Equivalent to DIP-EOMCC(3h-1p). e Total DIP-EOMCC(4h-2p) energy in hartree. f The DIP-EOMCC(4h-2p) singlet–triplet gap in kcal/mol. 146 Figure 4.7 Convergence of (a) DIP-EOMCC(P ) energies of X 3B1, A 1A1, B 1B1, and C 1A1 states of methylene, as described by the TZ2P basis set, and (b) of the corresponding adiabatic singlet–triplet gaps towards their parent DIP-EOMCC(4h-2p) values. 147 04080120160200MC Iterations (103)03691215Error rel. to DIP-EOMCC(4h-2p) (mEh)(a)X 3B1A 1A1B 1B1C 1A104080120160200MC Iterations (103)0246Error rel. to DIP-EOMCC(4h-2p) (kcal/mol)(b)A 1A1− X 3B1B 1B1− X 3B1C 1A1− X 3B1 4.3.2 Trimethylenemethane Our final example is trimethylenemethane (TMM), a fascinating non-Kekul´e hydrocar- bon, in which four valence π electrons are de-localized over four closely spaced π-type or- bitals. Assuming D3h symmetry, which is the true point group symmetry of the minimum energy structure of the ground-state triplet surface of trimethylenemethane, the four MOs involved in the π-orbital network of trimethylenemethane consists of the nondegenerate 1a′′ 2 orbital, the doubly degenerate 1e′′ orbitals, and the nondegenerate 2a′′ 2 orbital. If, on the other hand, the symmetry relevant to the low-lying singlet states, C2v is adopted, the 1a′′ 2 and 2a′′ 2 orbitals in the D3h description becomes 1b1 and 3b1, respectively, and the doubly degenerate 1e′′ shell splits into the 1a2 and 2b1 components (see Fig. 3.14 for a pictorial representation of the orbitals). Although the electronic structure of trimethylenemethane has been well understood for decades (cf., e.g., Ref. [213] and references therein), an ac- curate characterization of its triplet ground state and low-lying singlet states and the en- ergy separation between them continues to represent a significant challenge to quantum chemistry approaches [80, 88, 153, 154, 214–239]. The D3h-symmetric triplet ground state, designated as X 3A′ 2 (in a C2v description adopted in this study, X 3B2), which is domi- nated by the |{core}(1a′′ 2)2(1e′′ 1)1(1e′′ 2)1| configuration (in C2v, |{core}(1b1)2(1a2)1(2b1)1|), is relatively easy to describe, but the next two states, which are the nearly degenerate sin- glets, are not. These states undergo Jahn–Teller distortion that lifts their exact degeneracy in a D3h description, splitting the two states. The lower of the two singlets, character- ized by a Cs-symmetric minimum that can be approximated by a twisted C2v structure and is usually designated as the open-shell singlet A 1B1 state and the second state is a C2v-symmetric multi configurational state referred to as B 1A1. The first state emerges from the |{core}(1b1)2(1a2)1(2b1)1| configuration, while the second one is dominated by the |{core}(1b1)2(1a2)2| and |{core}(1b1)2(2b1)2| closed-shell determinants. The open-shell sin- glet state A 1B1, after the Jahn–Teller distortion is lower in energy compared to the B 1A1 state, but it has not been observed experimentally due to unfavorable Franck–Condon fac- 148 tors [229, 240], so we do not consider it in this work. The second singlet state B 1A1, on the other hand, has been detected in photoelectron spectroscopy experiments reported in Refs. [240, 241], which located it at 16.1 ± 0.1 kcal/mol above the X 3A′ 2 ground state. Thus, following our group’s previous deterministic, active-orbital-based, CC(P ;Q) work [88], the state-of-the-art DEA- and DIP-EOMCC computations with up to 4p-2h and 4h-2p excita- tions reported in Refs. [80, 153, 154], and the semi-stochastic CC(P;Q) calculations reported in Ref. [102], we focused on the D3h-symmetric X 3A′ 2 ground state and the C2v-symmetric B 1A1 singlet state and the adiabatic singlet–triplet gap between them (see, Fig. 4.8 for a schematic illustration of the Jahn–Teller splitting in trimethylenemethane and the singlet– triplet gap targeted). Similar to our previous work, we utilized the geometries optimized using the the spin-flip density functional theory (SF-DFT) and the 6-31G(d) basis reported in Ref. [231]. The purely electronic singlet–triplet gap derived from the experiments, ob- tained by subtracting the zero-point vibrational energy correction ∆ZPVE resulting from the SF-DFT/6-31G(d) calculations reported in Ref. [231] from the experimental B 1A1 − X 3A′ 2 gap determined in Refs. [240, 241], is 18.1 kcal/mol. However, this estimate depends on the source of the information about the ∆ZPVE correction. For example, if one replaces the ∆ZPVE value obtained in the SF-DFT/6-31G(d) calculations reported in Ref. [231] by its CCSD(T)/6-311++G(2d,2p) estimate and accounts for the core polarization effects de- termined with the help of the CCSD(T)/cc-pCVQZ computations, combining the resulting information with the experimental B 1A1 − X 3A′ 2 separation determined in Refs. [240, 241], the purely electronic, experimentally derived, adiabatic ∆ES-T gap increases to 19.4 kcal/mol [237]. The DEA-EOMCC(4p-2h)/6-31G(d) and DIP-EOMCC(4h-2p)/6-31G(d) approaches using the SF-DFT/6-31G(d) geometries predict this gap to be 19.856 kcal/mol and 19.907 kcal/mol, respectively. The remainder of this section describes how the DEA-EOMCC(P) and DIP-EOMCC(P) employing the 6-31G(d) basis set perform in this challenging situation. Following our previous semi-stochastic CC(P;Q) approach [102], we used the i-CISDTQ-MC wave function propagation for constructing the P spaces. Based on our previous experience 149 in semi-stochastic methods, replacing FCIQMC propagations with CISDTQ-MC will not affect the rate of convergence of our DEA-EOMCC(P) and DIP-EOMCC(P) approaches, while offering additional computational savings in the QMC part. In case of the DEA- EOMCC calculations, we used the TMM2+ dication as the closed-shell reference and for the DIP-EOMCC calculations, we used the TMM2− dianion as the reference determinant. The leading 4p-2h (4h-2p) determinants captured during the i-CISDTQ-MC propagations at various time steps τ are the determinants of B1 symmetry with Sz = 1 in the case of the X 3A′ 2 state, and the Sz = 0 A1 symmetric determinants in case of the B 1A1 state. For the CIQMC propagations we used the ROHF reference for the X 3A′ 2 state and the RHF reference for B 1A1 state. The one- and two-body integrals used in the DEA-EOMCC (DIP- EOMCC) calculations are extracted from the TMM2+ (TMM2−) system. All the results of the semi-stochastic calculations are reported in Table 4.7 and Figs. 4.9 and 4.10. We first discuss the results of our DEA-EOMCC(P) calculations in the next paragraph and then the following paragraph contains the results of DIP-EOMCC(P) computations. The results reported in Table 4.7 and Figure 4.9 demonstrate that, for both the X 3A′ 2 and B 1A1 states, the large errors which range from 19 to 27 millihartree at τ = 0 get reduced to half its value with just 1% of all 4p-2h determinants in the stochastically determined P spaces captured by the i-CISDTQ-MC runs which happens at 10000 MC iteration. After the additional 20000 MC time steps, which results in capturing 7–8% of all 4p-2h determinants, the errors go down to ∼3 millihartree for both states. This error in the energies of the two states reaches the sub-millihartree regime when the underlying P spaces contain 18–26% of the 4p-2h determinants and this happens at 50000 MC iteration. For the B 1A1 − X 3A′ 2 singlet–triplet gap, ∆ES-T, the ∼5 kcal/mol error relative to DEA-EOMCC(4p-2h) at τ = 0 quickly gets reduced to 1 kcal/mol at 10000 MC iterations. By the time we reach 20000 QMC iterations the singlet–triplet gap is only 0.349 kcal/mol away from the target 19.856 kcal/mol value obtained with DEA-EOMCC(4p-2h). From Table 4.7 and Figure 4.10 we can see that the DIP-EOMCC(P |τ =0) calculations, 150 where the Rµ,4h-2p components are completely neglected, produce very large, 26–30 milli- hartree, errors for both the states relative to the parent DIP-EOMCC(4h-2p) method. As soon as we propagate a little in imaginary time i.e., after only 10000 QMC iterations the DIP-EOMCC(P ) calculated energies for the X 3A′ 2 and B 1A1 states are 4.259 and 9.124 millihartree away from their target values. This is a significant improvement, especially considering the fact that the P spaces at this time step contain only 7.3% of 4h-2p determi- nants for the X 3A′ 2 state and 7.7% of 4h-2p determinants in case of the B 1A1 state. The B 1A1 − X 3A′ 2 gap at this point is ∼3 kcal/mol away from its DIP-EOMCC(4h-2p) predicted target of 19.907 kcal/mol. After 30000 QMC iterations, where the P spaces for the singlet and the triplet states contain about 30% and 16% of 4h-2p determinants respectively. At this point the error in the singlet–triplet gap is less than 1 kcal/mol and the underlying electronic states are only 1–3 millihartree in error. 151 Table 4.7 Convergence of the DEA-EOMCC(P ) and DIP-EOMCC(P ) energies [abbreviated as DEA(P ) and DIP(P ), respec- 2 and B 1A1 states of TMM, as described by the 6-31G(d) basis set of Ref. [178], and of the corresponding tively] of the X 3A′ adiabatic singlet–triplet (S–T) gaps toward their parent DEA-EOMCC(4p-2h) and DIP-EOMCC(4h-2p) values. The geometries 2 and B 1A1 states, optimized using the SF-DFT/6-31G(d) calculations, were taken from Ref. [231]. The P spaces of the X 3A′ used in the DEA-EOMCC(P ) calculations were defined as all 2p and 3p-1h determinants and subsets of 4p-2h determinants extracted from the i-CISDTQ-MC propagations with δτ = 0.0001 a.u. The P spaces used in the DIP-EOMCC(P ) calculations were defined as all 2h and 3h-1p determinants and subsets of 4h-2p determinants extracted from the i-CISDTQ-MC propaga- tions with δτ = 0.0001 a.u. In all the post-SCF calculations, the lowest core orbitals of the carbon atoms were kept frozen and the spherical components of the carbon d orbitals were employed throughout. MC Iters. DEA(P )a %(4p-2h)b DEA(P )a %(4p-2h)b X 3A′ 2 DEA-EOMCC(P ) B 1A1 0 4000 10000 20000 30000 40000 50000 80000 ∞ 19.933g 16.852 10.967 6.092 3.274 1.583 0.727 0.048 0.0 0.4 1.4 3.6 6.9 11.8 18.3 53.9 −155.399202i 27.755g 19.554 12.653 6.648 3.738 1.901 0.843 0.092 0.0 0.3 1.4 4.1 8.4 15.4 26.0 74.3 −155.367559i S–T Gap ∆Ec 4.909g 1.695 1.058 0.349 0.291 0.200 0.073 0.028 19.856j X 3A′ 2 DIP-EOMCC(P ) B 1A1 DIP(P )d %(4h-2p)e DIP(P )d %(4h-2p)e 26.513h 8.053 4.259 1.747 2.649 1.526 1.110 0.007 0.0 3.8 7.3 13.9 15.9 22.3 28.8 54.7 −155.399528k 0.0 2.1 7.7 18.5 30.4 43.0 54.0 61.9 −155.367804k 29.721h 16.608 9.124 3.720 1.080 0.229 0.044 0.001 S–T Gap ∆Ef 2.013h 5.368 3.053 1.238 −0.985 −0.814 −0.669 −0.004 19.907l 2 state and the Sz = 0 4p-2h determinants of the A1 symmetry in case of the B 1A1 state]. a Unless otherwise stated, all energies are reported as errors relative to DEA-EOMCC(4p-2h) in millihartree. b The %(4p-2h) values are the percentages of 4p-2h determinants captured during the i-CISDTQ-MC propagations [the Sz = 1 4p-2h determinants of the B1 symmetry in case of the X 3A′ c Unless otherwise specified, the singlet–triplet gaps ∆E = B 1A1 − X 3A′ d Unless otherwise stated, all energies are reported as errors relative to DIP-EOMCC(4h-2p) in millihartree. e The %(4h-2p) values are the percentages of 4h-2p determinants captured during the i-CISDTQ-MC propagations [the Sz = 1 4h-2p determinants of the B1 symmetry in case of the X 3A′ f Unless otherwise specified, the singlet–triplet gaps ∆E = B 1A1 − X 3A′ g Equivalent to DEA-EOMCC(3p-1h). h Equivalent to DIP-EOMCC(3h-1p). i Total DEA-EOMCC(3p-1h) energy in hartree. j The DEA-EOMCC(4p-2h) singlet–triplet gap in kcal/mol. k Total DIP-EOMCC(3h-1p) energy in hartree. l The DIP-EOMCC(4h-2p) singlet–triplet gap in kcal/mol. 2 state and the Sz = 0 4h-2p determinants of the A1 symmetry in case of the B 1A1 state]. 2 are reported as errors relative to DEA-EOMCC(4p-2h) in kcal/mol. 2 are reported as errors relative to DIP-EOMCC(4h-2p) in kcal/mol. 152 Figure 4.8 Jahn-Teller distortion in the trimethylenemethane molecule. At the geometry of the D3h-symmetric triplet ground state (shown in blue), trimethylenemethane has a doubly degenerate singlet excited state. Due to Jahn–Teller distortion these states split into an open-shell singlet state A 1B1 (shown in green) and a multi-configurational singlet state B 1A1 (shown in red). Although the A 1B1 state becomes the first excited state, it is not observed experimentally due to unfavorable Franck–Condon factors. Thus, we calculate the singlet–triplet gap between the ground triplet and the second excited singlet state B 1A1. 153 2 and (b) B 1A1 states Figure 4.9 Convergence of DEA-EOMCC(P ) energies of (a) X 3A′ of TMM, as described by the 6-31G(d) basis set, and (c) of the corresponding adiabatic singlet–triplet gap towards their parent DEA-EOMCC(4p-2h) values. Figure 4.10 Convergence of DIP-EOMCC(P ) energies of (a) X 3A′ 2 and (b) B 1A1 states of TMM, as described by the 6-31G(d) basis set, and (c) of the corresponding adiabatic singlet–triplet gap towards their parent DIP-EOMCC(4h-2p) values. 154 01224(a)X 3A′2020406080MC Iterations (103)01224Error rel. to DEA-EOMCC(4p-2h) (mEh)(b)B 1A1020406080MC Iterations (103)−20246Error rel. to DEA-EOMCC(4p-2h) (kcal/mol)(c)B 1A1−X 3A′201224(a)X 3A′2020406080MC Iterations (103)01224Error rel. to DIP-EOMCC(4h-2p) (mEh)(b)B 1A1020406080MC Iterations (103)−20246Error rel. to DIP-EOMCC(4h-2p) (kcal/mol)(c)B 1A1−X 3A′2 CHAPTER 5 CONCLUDING REMARKS AND FUTURE OUTLOOK In this dissertation, we have discussed some of the recent advances in the CC and EOMCC theories to which the author of this thesis has had the opportunity to contribute during his doctoral work in Professor Piotr Piecuch’s group. In particular, we have explored the semi-stochastic, CIQMC-driven, CC(P;Q) framework and and its extension to ground and excited states of open-shell systems. We have also discussed the semi-stochastic formulation of the particle nonconserving EA, IP, DEA, and DIP EOMCC frameworks, again taking advantage of CIQMC. In the first part of this dissertation, we have discussed the CC theory as one of the best ways of approaching the many-electron correlation problem in molecular systems in a computationally tractable manner. After highlighting the advantages of the CC theory and some of its key challenges, and discussing its extensions to excited and open-shell states via the EOMCC formalism, we have focused on the CC(P;Q) methodology, where the flexibility in defining the P and Q excitation spaces allows one to obtain highly accurate energet- ics equivalent or very close to those obtained with the high-level CCSDT, CCSDTQ, etc. methods and their EOMCC extensions at small fractions of the computational costs, even when higher–than–two-body components of the cluster and EOM excitation operators be- come large, nonperturbative, and strongly coupled to their lower-rank components. We have also discussed the particle nonconserving EOMCC formalisms of the EA, IP, DEA, and DIP EOMCC types that offer an elegant and orthogonally spin-adapted approach to open-shell species within the SR framework. This includes problems involving electronic excitation spectra of radicals and singlet–triplet gaps in biradicals. We have also briefly reviewed the stochastic FCIQMC wave function propagation and sampling approach and its truncated CIQMC counterparts. In the second part of this dissertation, we have focused on the semi-stochastic CC(P;Q) methodology that combines the flexible deterministic CC(P;Q) framework with the stochas- 155 tic CIQMC wave function propagations to automatically identify the P and Q spaces needed in the CC(P;Q) calculations without any reference to the previously exploited user- and system-dependent active-orbital concepts. By examining the excitation spectra of CH+, CH, and CNC, we have demonstrated the ability of the semi-stochastic CC(P;Q) methodol- ogy to converge ground and excited states, including challenging non-singlet excited states of open-shell systems, out of the early stages of CIQMC propagations. Many of the excited states examined in this dissertation are dominated by two-electron transitions, making them difficult to describe by the majority of the existing quantum chemistry approaches. We have also applied the semi-stochastic CC(P;Q) methodology to converge the CCSDT energet- ics of the lowest singlet and triplet states of several biradical species, including methylene, (HFH)−, cyclobutadiene, cyclopentadienyl cation, and trimethylenemethane, and the cor- responding singlet–triplet gaps, which is another challenging problem for many quantum chemistry methods because one has to balance the predominantly weakly correlated triplet states with the multiconfigurational, often strongly correlated, singlet states. In the third part of this dissertation, we have extended the semi-stochastic ideas to the particle nonconserving EOMCC formalisms of the EA, IP, DEA, and DIP types. We have demonstrated that by combining CIQMC wave function propagations with the determinis- tic EA- and IP-EOMCC frameworks, one can converge the high-level EA-EOMCC(3p-2h) and IP-EOMCC(3h-2p) energetics at small fractions of the computational costs out of the early stages of CIQMC propagations, even in the presence of strong 3p-2h and 3h-2p cor- relations. We have done this by studying the C2N, CNC, N3, and NCO radicals, where we have calculated the low-lying doublet states and the associated adiabatic excitation energies. We have then extended the semi-stochastic CIQMC-driven ideas to the DEA-EOMCC and DIP-EOMCC formalisms, which has allowed us to converge the high-level DEA-EOMCC(4p- 2h) and DIP-EOMCC(4h-2p) energetics in an automated and efficient manner using small fractions of 4p-2h and 4h-2p amplitudes, again identified in the early stages of the under- lying CIQMC runs. To illustrate the efficiency of our semi-stochastic DEA-EOMCC(4p-2h) 156 and DIP-EOMCC(4h-2p) approaches, we have investigated the ground and three low-lying singlet excited states of methylene, along with the corresponding singlet–triplet gaps, and the trimethylenemethane biradical, where we have determined the ground triplet and the low-lying and multiconfigurational singlet states and the corresponding singlet–triplet gap. While this dissertation has explored the semi-stochastic CC(P;Q) approach aimed at CCSDT/EOMCCSDT and the EA/IP/DEA/DIP EOMCC methodologies aimed at a highly accurate description of 3p-2h/3h-2p/4p-2h/4h-2p correlations, showing a lot of promise, there are still several areas that need to be examined. For example, recent work on the active space [84] and QMC-driven [101] CC(P;Q) frameworks targeting the ground-state CCSDTQ energetics have shown encouraging results, and hence it would be useful to extend the semi- stochastic CC(P;Q) methodology investigated in this dissertation to target the EOMCCS- DTQ energetics of excited states. The selected-CI-driven [120] and adaptive [121] CC(P;Q) approaches have shown promising results in converging CCSDT energies as well, so it would be interesting to extend them to target EOMCCSDT or even CCSDTQ/EOMCCSDTQ. In the case of the particle nonconserving EOMCC approaches, the CC(P;Q)-type moment corrections have not been implemented yet. It would, thus, be very beneficial to extend the noniterative CC(P;Q) corrections to the particle nonconserving EOMCC schemes. Based on the significant acceleration toward the desired high-level CC/EOMCC energetics these corrections offer in the particle conserving cases, one may anticipate that they will be very effective in the EA-EOMCC, IP-EOMCC, DEA-EOMCC, and DIP-EOMCC approaches as well. Combining the particle nonconserving EOMCC formalisms with the selected-CI-based or adaptive selections of higher-order correlations, examined in this work using CIQMC- driven ideas, is another interesting aspect worth investigation. In analogy to other post-SCF ab initio quantum chemistry approaches, in the longer- term, one can also think of extending the applicability of the methods developed in this dissertation to larger and more complex molecular systems containing hundreds of electrons and dozens of non-hydrogen atoms using techniques such as the fragment molecular orbital 157 (FMO) approach [259–261], the effective fragment potential (EFP) embedding scheme [262] and its merger with FMO abbreviated as EFMO [263–265], the cluster-in-molecule frame- work [266–269], and, ultimately, the quantum mechanics/molecular mechanics [270–274] and polarizable continuum [275, 276] models, to name a few examples. By doing so, especially when combined with code parallelization across multiple multi-core nodes, at least some of our semi-stochastic CC(P;Q) and EOMCC algorithms developed in this dissertation can be made applicable to studies of molecular electronic excitation spectra and singlet–triplet gaps in solution and condensed phases. Clearly, additional molecular applications of the semi- stochastic CC(P;Q) and EA/IP/DEA/DIP EOMCC approaches investigated in this work would be very useful too. 158 [1] M. S. Gordon and D. G. Truhlar, Theor. Chim. Acta 71, 1 (1987). REFERENCES [2] [3] [4] [5] [6] [7] [8] [9] F. Coester, Nucl. Phys. 7, 421 (1958). F. Coester and H. K¨ummel, Nucl. Phys. 17, 477 (1960). J. ˇC´ıˇzek, J. Chem. Phys. 45, 4256 (1966). J. ˇC´ıˇzek, Adv. Chem. Phys. 14, 35 (1969). J. ˇC´ıˇzek and J. Paldus, Int. J. Quantum Chem. 5, 359 (1971). J. Paldus, J. ˇC´ıˇzek, and I. Shavitt, Phys. Rev. A 5, 50 (1972). J. Paldus and X. Li, Adv. Chem. Phys. 110, 1 (1999). R. J. Bartlett and M. Musia l, Rev. Mod. Phys. 79, 291 (2007). [10] K. A. Brueckner, Phys. Rev. 100, 36 (1955). [11] J. Goldstone, Proc. R. Soc. London A Math. Phys. Eng. Sci. 239, 267 (1957). [12] J. Hubbard, Proc. R. Soc. London A Math. Phys. Eng. Sci. 240, 539 (1957). [13] N. M. Hugenholtz, Physica 23, 481 (1957). [14] J. Paldus, P. Piecuch, L. Pylypow, and B. Jeziorski, Phys. Rev. A 47, 2738 (1993). [15] K. Kowalski and P. Piecuch, Phys. Rev. A 61, 052506 (2000). [16] P. Piecuch and K. Kowalski, Int. J. Mol. Sci. 3, 676 (2002). [17] D. I. Lyakh, M. Musia l, V. F. Lotrich, and R. J. Bartlett, Chem. Rev. 112, 182 (2012). [18] F. A. Evangelista, J. Chem. Phys. 149, 030901 (2018). [19] B. O. Roos, Adv. Chem. Phys. 69, 399 (1987). [20] M. W. Schmidt and M. S. Gordon, Annu. Rev. Phys. Chem. 49, 233 (1998). [21] P. G. Szalay, T. M¨uller, G. Gidofalvi, H. Lischka, and R. Shepard, Chem. Rev. 112, 108 (2012). 159 [22] D. Roca-Sanju´an, F. Aquilante, and R. Lindh, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2, 585 (2012). [23] S. Chattopadhyay, R. K. Chaudhuri, U. S. Mahapatra, A. Ghosh, and S. S. Ray, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 6, 266 (2016). [24] K. Kowalski and P. Piecuch, Int. J. Quantum Chem. 80, 757 (2000). [25] P. Piecuch, R. Tobo la, and J. Paldus, Chem. Phys. Lett. 210, 243 (1993). [26] P. Piecuch and J. Paldus, Phys. Rev. A 49, 3479 (1994). [27] K. Jankowski, J. Paldus, I. Grabowski, and K. Kowalski, J. Chem. Phys. 97, 7600 (1992), 101, 1759 (1994) [Erratum]. [28] K. Jankowski, J. Paldus, I. Grabowski, and K. Kowalski, J. Chem. Phys. 101, 3085 (1994). [29] G. D. Purvis, III and R. J. Bartlett, J. Chem. Phys. 76, 1910 (1982). [30] J. M. Cullen and M. C. Zerner, J. Chem. Phys. 77, 4088 (1982). [31] G. E. Scuseria, A. C. Scheiner, T. J. Lee, J. E. Rice, and H. F. Schaefer, III, J. Chem. Phys. 86, 2881 (1987). [32] P. Piecuch and J. Paldus, Int. J. Quantum Chem. 36, 429 (1989). [33] J. Noga and R. J. Bartlett, J. Chem. Phys. 86, 7041 (1987), 89, 3401 (1988) [Erratum]. [34] G. E. Scuseria and H. F. Schaefer, III, Chem. Phys. Lett. 152, 382 (1988). [35] M. R. Hoffman and H. F. Schaefer, III, Adv. Quant. Chem. 18, 207 (1986). [36] N. Oliphant and L. Adamowicz, J. Chem. Phys. 95, 6645 (1991). [37] S. A. Kucharski and R. J. Bartlett, Theor. Chim. Acta 80, 387 (1991). [38] P. Piecuch and L. Adamowicz, J. Chem. Phys. 100, 5792 (1994). [39] S. A. Kucharski and R. J. Bartlett, J. Chem. Phys. 97, 4282 (1992). [40] K. Emrich, Nucl. Phys. A 351, 379 (1981). [41] J. Geertsen, M. Rittby, and R. J. Bartlett, Chem. Phys. Lett. 164, 57 (1989). 160 [42] D. C. Comeau and R. J. Bartlett, Chem. Phys. Lett. 207, 414 (1993). [43] J. F. Stanton and R. J. Bartlett, J. Chem. Phys. 98, 7029 (1993). [44] P. Piecuch and R. J. Bartlett, Adv. Quantum Chem. 34, 295 (1999). [45] H. Monkhorst, Int. J. Quantum Chem. Symp. 11, 421 (1977). [46] E. Dalgaard and H. Monkhorst, Phys. Rev. A 28, 1217 (1983). [47] M. Takahashi and J. Paldus, J. Chem. Phys. 85, 1486 (1986). [48] H. Koch and P. Jørgensen, J. Chem. Phys. 93, 3333 (1990). [49] H. Koch, H. J. A. Jensen, P. Jørgensen, and T. Helgaker, J. Chem. Phys. 93, 3345 (1990). [50] H. Nakatsuji, Chem. Phys. Lett. 59, 362 (1978). [51] K. Kowalski and P. Piecuch, J. Chem. Phys. 115, 643 (2001). [52] K. Kowalski and P. Piecuch, Chem. Phys. Lett. 347, 237 (2001). [53] S. A. Kucharski, M. W loch, M. Musia l, and R. J. Bartlett, J. Chem. Phys. 115, 8263 (2001). [54] S. Hirata, J. Chem. Phys. 121, 51 (2004). [55] M. K´allay and J. Gauss, J. Chem. Phys. 121, 9257 (2004). [56] K. Raghavachari, G. W. Trucks, J. A. Pople, and M. Head-Gordon, Chem. Phys. Lett. 157, 479 (1989). [57] P. Piecuch and K. Kowalski, in Computational Chemistry: Reviews of Current Trends, Vol. 5, edited by J. Leszczy´nski (World Scientific, Singapore, 2000) pp. 1–104. [58] K. Kowalski and P. Piecuch, J. Chem. Phys. 113, 18 (2000). [59] P. Piecuch, K. Kowalski, I. S. O. Pimienta, and M. J. McGuire, Int. Rev. Phys. Chem. 21, 527 (2002). [60] P. Piecuch, K. Kowalski, I. S. O. Pimienta, P.-D. Fan, M. Lodriguito, M. J. McGuire, S. A. Kucharski, T. Ku´s, and M. Musia l, Theor. Chem. Acc. 112, 349 (2004). [61] K. Kowalski and P. Piecuch, J. Chem. Phys. 120, 1715 (2004). 161 [62] M. W loch, J. R. Gour, K. Kowalski, and P. Piecuch, J. Chem. Phys. 122, 214107 (2005). [63] P. Piecuch and M. W loch, J. Chem. Phys. 123, 224105 (2005). [64] P. Piecuch, M. W loch, J. R. Gour, and A. Kinal, Chem. Phys. Lett. 418, 467 (2006). [65] M. W loch, M. D. Lodriguito, P. Piecuch, and J. R. Gour, Mol. Phys. 104, 2149 (2006), 104, 2991 (2006) [Erratum]. [66] M. W loch, J. R. Gour, and P. Piecuch, J. Phys. Chem. A 111, 11359 (2007). [67] P. Piecuch, M. W loch, M. Lodriguito, and J. R. Gour, in Recent Advances in the Theory of Chemical and Physical Systems, Progress in Theoretical Chemistry and Physics, Vol. 15, edited by S. Wilson, J.-P. Julien, J. Maruani, E. Br¨andas, and G. Delgado- Barrio (Springer, Dordrecht, 2006) pp. 45–106. [68] P. Piecuch, J. R. Gour, and M. W loch, Int. J. Quantum Chem. 109, 3268 (2009). [69] G. Fradelos, J. J. Lutz, T. A. Weso lowski, P. Piecuch, and M. W loch, J. Chem. Theory Comput. 7, 1647 (2011). [70] J. Shen and P. Piecuch, Chem. Phys. 401, 180 (2012). [71] N. Oliphant and L. Adamowicz, J. Chem. Phys. 94, 1229 (1991). [72] N. Oliphant and L. Adamowicz, J. Chem. Phys. 96, 3739 (1992). [73] P. Piecuch, N. Oliphant, and L. Adamowicz, J. Chem. Phys. 99, 1875 (1993). [74] P. Piecuch and L. Adamowicz, J. Chem. Phys. 102, 898 (1995). [75] L. Adamowicz, P. Piecuch, and K. B. Ghose, Mol. Phys. 94, 225 (1998). [76] P. Piecuch, S. A. Kucharski, and R. J. Bartlett, J. Chem. Phys. 110, 6103 (1999). [77] K. Kowalski and P. Piecuch, J. Chem. Phys. 113, 8490 (2000). [78] J. R. Gour, P. Piecuch, and M. W loch, J. Chem. Phys. 123, 134113 (2005). [79] P. Piecuch, Mol. Phys. 108, 2987 (2010). [80] J. Shen and P. Piecuch, J. Chem. Phys. 138, 194102 (2013). [81] K. Kowalski and P. Piecuch, J. Chem. Phys. 122, 074107 (2005). 162 [82] P.-D. Fan, K. Kowalski, and P. Piecuch, Mol. Phys. 103, 2191 (2005). [83] M. Horoi, J. R. Gour, M. W loch, M. D. Lodriguito, B. A. Brown, and P. Piecuch, Phys. Rev. Lett. 98, 112501 (2007). [84] N. P. Bauman, J. Shen, and P. Piecuch, Mol. Phys. 115, 2860 (2017). [85] J. D. Watts and R. J. Bartlett, Chem. Phys. Lett. 233, 81 (1995). [86] J. D. Watts and R. J. Bartlett, Chem. Phys. Lett. 258, 581 (1996). [87] J. Shen and P. Piecuch, J. Chem. Phys. 136, 144104 (2012). [88] J. Shen and P. Piecuch, J. Chem. Theory Comput. 8, 4968 (2012). [89] S. H. Yuwono, I. Magoulas, J. Shen, and P. Piecuch, Mol. Phys. 117, 1486 (2019). [90] I. Magoulas, N. P. Bauman, J. Shen, and P. Piecuch, J. Phys. Chem. A 122, 1350 (2018). [91] G. H. Booth, A. J. W. Thom, and A. Alavi, J. Chem. Phys. 131, 054106 (2009). [92] D. Cleland, G. H. Booth, and A. Alavi, J. Chem. Phys. 132, 041103 (2010). [93] K. Ghanem, A. Y. Lozovoi, and A. Alavi, J. Chem. Phys. 151, 224108 (2019). [94] K. Ghanem, K. Guther, and A. Alavi, J. Chem. Phys. 153, 224115 (2020). [95] A. J. W. Thom, Phys. Rev. Lett. 105, 263004 (2010). [96] R. S. T. Franklin, J. S. Spencer, A. Zoccante, and A. J. W. Thom, J. Chem. Phys. 144, 044111 (2016). [97] J. S. Spencer and A. J. W. Thom, J. Chem. Phys. 144, 084108 (2016). [98] C. J. C. Scott and A. J. W. Thom, J. Chem. Phys. 147, 124105 (2017). [99] J. E. Deustua, J. Shen, and P. Piecuch, Phys. Rev. Lett. 119, 223003 (2017). [100] S. H. Yuwono, A. Chakraborty, J. E. Deustua, J. Shen, and P. Piecuch, Mol. Phys. 118, e1817592 (2020). [101] J. E. Deustua, J. Shen, and P. Piecuch, J. Chem. Phys., 154, 124103 (2021). [102] A. Chakraborty, S. H. Yuwono, J. E. Deustua, J. Shen, and P. Piecuch, J. Chem. Phys. 163 157, 134101 (2022). [103] J. Whitten and M. Hackmeyer, J. Chem. Phys. 51, 5584 (1969). [104] C. Bender and E. Davidson, Phys. Rev. 183, 23 (1969). [105] B. Huron, J. P. Malrieu, and P. Rancurel, J. Chem. Phys. 58, 5745 (1973). [106] R. Buenker and S. Peyerimhoff, Theor. Chim. Acta. 35, 33 (1974). [107] J. Schriber and F. Evangelista, J. Chem. Phys. 144, 161106 (2016). [108] J. Schriber and F. Evangelista, J. Chem. Theory Comput. 13, 5354 (2017). [109] N. M. Tubman, J. Lee, T. Takeshita, M. Head-Gordon, and K. Whaley, J. Chem. Phys. 145, 044112 (2016). [110] N. M. Tubman, C. Freeman, D. Levine, D. Hait, M. Head-Gordon, and K. Whaley, J. Chem. Theory Comput. 16, 2139 (2020). [111] W. Liu and M. Hoffmann, J. Chem. Theory Comput. 12, 1169 (2016), 12, 3000 (2016) [Erratum]. [112] N. Zhang, W. Liu, and M. Hoffmann, J. Chem. Theory Comput. 16, 2296 (2020). [113] A. A. Holmes, N. M. Tubman, and C. J. Umrigar, J. Chem. Theory Comput. 12, 3674 (2016). [114] S. Sharma, A. A. Holmes, G. Jeanmairet, A. Alavi, and C. J. Umrigar, J. Chem. Theory Comput. 13, 1595 (2017). [115] J. Li, M. Otten, A. A. Holmes, S. Sharma, and C. J. Umrigar, J. Chem. Phys. 149, 214110 (2018). [116] Y. Garniron, A. Scemama, P.-F. Loos, and M. Caffarel, J. Chem. Phys. 147, 034101 (2017). [117] Y. Garniron, T. Applencourt, K. Gasperich, A. Benali, A. Ferte, J. Paquier, B. Pradines, R. Assaraf, P. Reinhardt, J. Toulouse, P. Barbaresco, N. Renon, G. David, J.-P. Malrieu, M. Veril, M. Caffarel, P.-F. Loos, E. Giner, and A. Scemama, J. Chem. Theory Comput. 15, 3591 (2019). [118] P.-F. Loos, Y. Damour, and A. Scemama, J. Chem. Phys. 153, 176101 (2020). [119] J. J. Eriksen, T. A. Anderson, J. E. Deustua, K. Ghanem, D. Hait, M. R. Hoffmann, 164 S. Lee, D. S. Levine, I. Magoulas, J. Shen, N. M. Tubman, K. B. Whaley, E. Xu, Y. Yao, N. Zhang, A. Alavi, G. K.-L. Chan, M. Head-Gordon, W. Liu, P. Piecuch, S. Sharma, S. L. Ten-no, C. J. Umrigar, and J. Gauss, J. Phys. Chem. Lett. 11, 8922 (2020). [120] K. Gururangan, J. E. Deustua, J. Shen, and P. Piecuch, J. Chem. Phys. 155, 174114 (2021). [121] K. Gururangan and P. Piecuch, J. Chem. Phys. 159, 084108 (2023). [122] N. Metropolis and S. Ulam, J. Am. Stat. Assoc. 44, 335 (1949). [123] N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, and E. Teller, J. Chem. Phys. 21, 1087 (1953). [124] W. K. Hastings, Biometrika 57, 97 (1970). [125] W. L. McMillan, Phys. Rev. 138, A442 (1965). [126] D. Schiff and L. Verlet, Phys. Rev. 160, 208 (1967). [127] D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980). [128] P. J. Reynolds, D. M. Ceperley, B. J. Alder, and W. A. Lester, Jr., J. Chem. Phys. 77, 5593 (1982). [129] W. M. C. Foulkes, L. Mitas, R. J. Needs, and G. Rajagopal, Rev. Mod. Phys. 73, 33 (2001). [130] D. Ceperley, G. V. Chester, and M. H. Kalos, Phys. Rev. B 16, 3081 (1977). [131] D. M. Ceperley, Rev. Mineral. Geochem. 71, 129 (2010). [132] J. Toulouse, R. Assaraf, and C. J. Umrigar, Adv. Quantum Chem. 73, 285 (2016). [133] J. B. Anderson, J. Chem. Phys. 63, 1499 (1975). [134] D. J. Klein and H. M. Pickett, J. Chem. Phys. 64, 4811 (1976). [135] J. B. Anderson, J. Chem. Phys. 65, 4121 (1976). [136] J. B. Anderson, Int. J. Quantum Chem. 15, 109 (1979). [137] J. E. Deustua, S. H. Yuwono, J. Shen, and P. Piecuch, J. Chem. Phys. 150, 111101 (2019). 165 [138] M. Nooijen and R. J. Bartlett, J. Chem. Phys. 102, 3629 (1995). [139] M. Nooijen and R. J. Bartlett, J. Chem. Phys. 102, 6735 (1995). [140] J. R. Gour, P. Piecuch, and M. W loch, Int. J. Quantum Chem. 106, 2854 (2006). [141] J. R. Gour and P. Piecuch, J. Chem. Phys. 125, 234107 (2006). [142] Y. Ohtsuka, P. Piecuch, J. R. Gour, M. Ehara, and H. Nakatsuji, J. Chem. Phys. 126, 164111 (2007). [143] M. Nooijen and J. G. Snijders, Int. J. Quantum Chem. Symp. 26, 55 (1992). [144] M. Nooijen and J. G. Snijders, Int. J. Quantum Chem. 48, 15 (1993). [145] J. F. Stanton and J. Gauss, J. Chem. Phys. 101, 8938 (1994). [146] R. J. Bartlett and J. F. Stanton, in Reviews in Computational Chemistry, Vol. 5, edited by K. B. Lipkowitz and D. B. Boyd (VCH Publishers, New York, 1994) pp. 65–169. [147] M. Nooijen and R. J. Bartlett, J. Chem. Phys. 106, 6441 (1997). [148] M. Nooijen, Int. J. Mol. Sci. 3, 656 (2002). [149] M. Musia l, A. Perera, and R. J. Bartlett, J. Chem. Phys. 134, 114108 (2011). [150] M. Musia l, S. A. Kucharski, and R. J. Bartlett, J. Chem. Theory Comput. 7, 3088 (2011). [151] T. Ku´s and A. I. Krylov, J. Chem. Phys. 135, 084109 (2011). [152] T. Ku´s and A. I. Krylov, J. Chem. Phys. 136, 244109 (2012). [153] J. Shen and P. Piecuch, Mol. Phys. 112, 868 (2014). [154] A. O. Ajala, J. Shen, and P. Piecuch, J. Phys. Chem. A 121, 3469 (2017). [155] S. J. Stoneburner, J. Shen, A. O. Ajala, P. Piecuch, D. G. Truhlar, and L. Gagliardi, J. Chem. Phys. 147, 164120 (2017). [156] J. Shen and P. Piecuch, Mol. Phys. 119, e1966534 (2021). [157] S. Gulania, E. F. Kjønstad, J. F. Stanton, H. Koch, and A. I. Krylov, J. Chem. Phys. 154, 114115 (2021). 166 [158] K. E. Schmidt and M. H. Kalos, in Applications of the Monte Carlo method in Statistical Physics, Vol. 36 (Springer-Verlag, Berlin, 1984). [159] N. S. Blunt, S. D. Smart, G. H. Booth, and A. Alavi, J. Chem. Phys. 143, 134117 (2015). [160] N. S. Blunt, G. H. Booth, and A. Alavi, J. Chem. Phys. 146, 244105 (2017). [161] S. Hirata, J. Chem. Phys. 121, 51 (2004). [162] M. Ehara, J. R. Gour, and P. Piecuch, Mol. Phys. 107, 871 (2009). [163] J. A. Hansen, P. Piecuch, J. J. Lutz, and J. R. Gour, Phys. Scr. 84, 028110 (2011). [164] J. Olsen, A. M. S. de Mer´as, H. Jensen, and P. Jørgensen, Chem. Phys. Lett. 154, 380 (1989). [165] T. H. Dunning, Jr., J. Chem. Phys. 90, 1007 (1989). [166] R. A. Kendall, T. H. Dunning, Jr., and R. J. Harrison, J. Chem. Phys. 96, 6769 (1992). [167] T. H. Dunning, Jr., J. Chem. Phys. 53, 2823 (1970). [168] T. H. Dunning, Jr. and P. J. Hay, in Methods of Electronic Structure Theory, Vol. 2, edited by H. F. Schaefer, III (Plenum, New York, 1977) pp. 1–28. [169] J. E. Deustua, I. Magoulas, J. Shen, and P. Piecuch, J. Chem. Phys. 149, 151101 (2018). [170] J. S. Spencer, N. S. Blunt, W. A. Vigor, F. D. Malone, W. M. C. Foulkes, J. J. Shepherd, and A. J. W. Thom, J. Open Res. Softw. 3, e9 (2015). [171] J. S. Spencer, N. S. Blunt, S. Choi, J. Etrych, M.-A. Filip, W. M. C. Foulkes, R. S. T. Franklin, W. J. Handley, F. D. Malone, V. A. Neufeld, R. Di Remigio, T. W. Rogers, C. J. C. Scott, J. J. Shepherd, W. A. Vigor, J. Weston, R. Xu, and A. J. W. Thom, J. Chem. Theory Comput. 15, 1728 (2019). [172] M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. J. Su, T. L. Windus, M. Dupuis, and J. A. Montgomery, J. Comput. Chem. 14, 1347 (1993). [173] G. M. J. Barca, C. Bertoni, L. Carrington, D. Datta, N. De Silva, J. E. Deustua, D. G. Fedorov, J. R. Gour, A. O. Gunina, E. Guidez, T. Harville, S. Irle, J. Ivanic, K. Kowalski, S. S. Leang, H. Li, W. Li, J. J. Lutz, I. Magoulas, J. Mato, V. Mironov, H. Nakata, B. Q. Pham, P. Piecuch, D. Poole, S. R. Pruitt, A. P. Rendell, L. B. 167 Roskop, K. Ruedenberg, T. Sattasathuchana, M. W. Schmidt, J. Shen, L. Slipchenko, M. Sosonkina, V. Sundriyal, A. Tiwari, J. L. G. Vallejo, B. Westheimer, M. W loch, P. Xu, F. Zahariev, and M. S. Gordon, J. Chem. Phys. 152, 154102 (2020). [174] F. Zahariev, P. Xu, B. M. Westheimer, S. Webb, J. Galvez Vallejo, A. Tiwari, V. Sun- driyal, M. Sosonkina, J. Shen, G. Schoendorff, M. Schlinsog, T. Sattasathuchana, K. Ruedenberg, L. B. Roskop, A. P. Rendell, D. Poole, P. Piecuch, B. Q. Pham, V. Mironov, J. Mato, S. Leonard, S. S. Leang, J. Ivanic, J. Hayes, T. Harville, K. Gu- rurangan, E. Guidez, I. S. Gerasimov, C. Friedl, K. N. Ferreras, G. Elliott, D. Datta, D. D. A. Cruz, L. Carrington, C. Bertoni, G. M. J. Barca, M. Alkan, and M. S. Gordon, J. Chem. Theory Comput. 19, 7031 (2023). [175] M. Zachwieja, J. Mol. Spectrosc. 170, 285 (1995). [176] R. K¸epa, A. Para, M. Rytel, and M. Zachwieja, J. Mol. Spectrosc. 178, 189 (1996). [177] K. P. Huber and G. Herzberg, Molecular Spectra and Molecular Structure: Constants of Diatomic Molecules (Van Nostrand Reinhold, New York, 1979). [178] W. J. Hehre, R. Ditchfield, and J. A. Pople, J. Chem. Phys. 56, 2257 (1972). [179] P. C. Hariharan and J. A. Pople, Theor. Chim. Acta 28, 213 (1973). [180] T. H. Dunning, Jr., J. Chem. Phys. 55, 716 (1971). [181] C. D. Sherrill, M. L. Leininger, T. J. V. Huis, and H. F. Schaefer, III, J. Chem. Phys. 108, 1040 (1998). [182] C. W. Bauschlicher and P. R. Taylor, J. Chem. Phys. 85, 6510 (1986). [183] C. W. Bauschlicher, S. R. Langhoff, and P. R. Taylor, J. Chem. Phys. 87, 387 (1987). [184] A. D. McLean, P. R. Buenker, R. M. Escribano, and P. Jensen, J. Chem. Phys. 87, 2166 (1987). [185] D. C. Comeau, I. Shavitt, P. Jensen, and P. R. Bunker, J. Chem. Phys. 90, 6491 (1989). [186] D. B. Knowles, J. Ramon, A. Collado, G. Hirsch, and R. J. Buekner, J. Chem. Phys. 92, 585 (1990). [187] X. Li, P. Piecuch, and J. Paldus, Chem. Phys. Lett. 224, 267 (1994). [188] P. Piecuch, X. Li, and J. Paldus, Chem. Phys. Lett. 230, 377 (1994). 168 [189] A. Balkov´a and R. J. Bartlett, J. Chem. Phys. 102, 7116 (1995). [190] X. Li and J. Paldus, J. Chem. Phys. 124, 174101 (2006). [191] J. D. Watts and R. J. Bartlett, J. Chem. Phys. 93, 6104 (1990). [192] J. R. Gour, P. Piecuch, and M. W loch, Mol. Phys. 108, 2633 (2010). [193] P. Jensen and P. R. Bunker, J. Chem. Phys. 89, 1327 (1988). [194] E. R. Davidson, D. Feller, and P. Phillips, Chem. Phys. Lett. 76, 416 (1980). [195] N. C. Handy, Y. Yamaguchi, and H. F. Schaefer, III, J. Chem. Phys. 84, 4481 (1986). [196] J. R. Hart, A. K. Rappe, S. M. Gorun, and T. H. Upton, J. Phys. Chem. 96, 6264 (1992). [197] P. Piecuch, J. R. Gour, and M. W loch, Int. J. Quantum Chem. 108, 2128 (2008). [198] J. Shen, Z. Kou, E. Xu, and S. Li, J. Chem. Phys. 134, 044134 (2011). [199] H. Schurkus, D.-T. Chen, H.-P. Cheng, G. Chan, and J. Stanton, J. Chem. Phys. 152, 234115 (2020). [200] P. G. Szalay and R. J. Bartlett, Chem. Phys. Lett. 214, 481 (1993). [201] P. G. Szalay and R. J. Bartlett, J. Chem. Phys. 103, 3600 (1995). [202] M. Eckert-Maksi´c, M. Vazdar, M. Barbatti, H. Lischka, and Z. B. Maksi´c, J. Chem. Phys. 125, 064310 (2006). [203] T. Saito, S. Nishihara, S. Yamanaka, Y. Kitagawa, T. Kawakami, S. Yamada, H. Isobe, M. Okumura, and K. Yamaguchi, Theor. Chem. Acc. 130, 749 (2011). [204] A. Balkov´a and R. J. Bartlett, J. Chem. Phys. 101, 8972 (1994). [205] S. V. Levchenko and A. I. Krylov, J. Chem. Phys. 120, 175 (2004). [206] P. M. Zimmerman, J. Phys. Chem. A 121, 4712 (2017). [207] J.-N. Boyn and D. A. Mazziotti, J. Chem. Phys. 154, 134103 (2021). [208] E. Monino, M. Boggio-Pasqua, A. Scemama, D. Jacquemin, and P.-F. Loos, J. Phys. Chem. A 126, 4664 (2022). 169 [209] C. A. Coulson, J. Chim. Phys. Phys.-Chim. Biol. 45, 243 (1948). [210] H. C. Longuet-Higgins, J. Chem. Phys. 18, 265 (1950). [211] P. Dowd, J. Am. Chem. Soc. 88, 2587 (1966). [212] R. J. Baseman, D. W. Pratt, M. Chow, and P. Dowd, J. Am. Chem. Soc. 98, 5726 (1976). [213] W. T. Borden and E. R. Davidson, Acc. Chem. Res. 14, 69 (1981). [214] D. R. Yarkony and H. F. Schaefer, III, J. Am. Chem. Soc. 96, 3754 (1974). [215] W. J. Hehre, L. Salem, and M. R. Willcott, J. Am. Chem. Soc. 96, 4328 (1974). [216] J. H. Davis and W. A. Goddard, III, J. Am. Chem. Soc. 99, 4242 (1977). [217] D. M. Hood, R. M. Pitzer, and H. F. Schaefer, III, J. Am. Chem. Soc. 100, 2227 (1978). [218] D. M. Hood, H. F. Schaefer, III, and R. M. Pitzer, J. Am. Chem. Soc. 100, 8009 (1978). [219] D. A. Dixon, R. Foster, T. A. Halgren, and W. N. Lipscomb, J. Am. Chem. Soc. 100, 1359 (1978). [220] D. Feller, W. T. Borden, and E. R. Davidson, J. Chem. Phys. 74, 2256 (1981). [221] S. B. Auster, R. M. Pitzer, and M. S. Platz, J. Am. Chem. Soc. 104, 3812 (1982). [222] W. T. Borden, E. R. Davidson, and D. Feller, Tetrahedron 38, 737 (1982). [223] P. M. Lahti, A. R. Rossi, and J. A. Berson, J. Am. Chem. Soc. 107, 2273 (1985). [224] A. Skanche, L. J. Schaad, and B. A. Hess, Jr., J. Am. Chem. Soc. 110, 5315 (1988). [225] S. Olivella and J. Salvador, Int. J. Quantum Chem. 37, 713 (1990). [226] T. P. Radhakrishnan, Tetrahedron Lett. 32, 4601 (1991). [227] A. S. Ichimura, N. Koga, and H. Iwamura, J. Phys. Org. Chem. 7, 207 (1994). [228] W. T. Borden, Mol. Cryst. Liq. Cryst. 232, 195 (1993). [229] C. J. Cramer and B. A. Smith, J. Phys. Chem. 100, 9664 (1996). 170 [230] B. Ma and H. F. Schaefer, III, Chem. Phys. 207, 31 (1996). [231] L. V. Slipchenko and A. I. Krylov, J. Chem. Phys. 117, 4694 (2002). [232] L. V. Slipchenko and A. I. Krylov, J. Chem. Phys. 118, 6874 (2003). [233] L. V. Slipchenko and A. I. Krylov, J. Chem. Phys. 123, 084107 (2005). [234] J. Brabec and J. Pittner, J. Phys. A 110, 11765 (2006). [235] J. Shen, T. Fang, S. Li, and Y. Jiang, J. Phys. Chem. A 112, 12518 (2008). [236] X. Li and J. Paldus, J. Chem. Phys. 129, 174101 (2008). [237] A. Perera, R. W. Molt, Jr., V. F. Lotrich, and R. J. Bartlett, Theor. Chem. Acc. 133, 1514 (2014). [238] S. Sinha Ray, S. Manna, A. Ghosh, R. K. Chaudhuri, and S. Chattopadhyay, Int. J. Quantum Chem. 119, e25776 (2019). [239] S. Chattopadhyay, ACS Omega 6, 1668 (2021). [240] P. G. Wenthold, J. Hu, R. R. Squires, and W. C. Lineberger, J. Am. Chem. Soc. 118, 475 (1996). [241] P. G. Wenthold, J. Hu, R. R. Squires, and W. C. Lineberger, J. Am. Soc. Mass Spec- trom. 10, 800 (1999). [242] A. Chakraborty, S. Basumallick, J. Shen, and P. Piecuch, “Extension of the Semi- Stochastic Equation-of-Motion Coupled-Cluster Formalism to the Single and Double Electron Attachment and Ionization Potential Schemes”, in preparation for J. Phys. Chem. A. [243] M. Musia l and R. J. Bartlett, J. Chem. Phys. 119, 1901 (2003). [244] M. Kamiya and S. Hirata, J. Chem. Phys. 125, 074111 (2006). [245] P.-D. Fan, M. Kamiya, and S. Hirata, J. Chem. Theor. Comput. 3, 1036 (2007). [246] M. Musia l, S. A. Kucharski, and R. J. Bartlett, J. Chem. Phys. 118, 1128 (2003). [247] M. Musia l and R. J. Bartlett, Chem. Phys. Lett. 384, 210 (2004). [248] Y. J. Bomble, J. C. Saeh, J. F. Stanton, P. G. Szalay, and M. K´allay, J. Chem. Phys. 122, 154107 (2005). 171 [249] M. Wladyslawski and M. Nooijen, in Low-Lying Potential Energy Surfaces, ACS Sym- posium Series, Vol. 828, edited by M. R. Hoffmann and K. G. Dyall (American Chem- ical Society, Washington, D.C., 2002) pp. 65–92. [250] K. W. Sattelmeyer, H. F. Schaefer, III, and J. F. Stanton, Chem. Phys. Lett. 378, 42 (2003). [251] M. Ehara, P. Piecuch, J. J. Lutz, and J. R. Gour, Chem. Phys. 399, 94 (2012). [252] N. P. Bauman, J. A. Hansen, M. Ehara, and P. Piecuch, J. Chem. Phys. 141, 101102 (2014). [253] N. P. Bauman, J. A. Hansen, and P. Piecuch, J. Chem. Phys. 145, 084306 (2016). [254] J. R. Reimets, J. Shen, M. Kianinia, C. Bradac, I. Aharonovich, M. J. Ford, and P. Piecuch, Phys. Rev. B 102, 144105 (2020). [255] W. Park, J. Shen, S. Lee, P. Piecuch, M. Filatov, and C. H. Choi, J. Phys. Chem. Lett. 12, 9720 (2021). [256] W. Park, J. Shen, S. Lee, P. Piecuch, T. Joo, M. Filatov(Gulak), and C. H. Choi, J. Phys. Chem. C 126, 14976 (2022). [257] W. Dobrautz, S. D. Smart, and A. Alavi, J. Chem. Phys. 151, 094104 (2019). [258] G. Herzberg, Molecular Spectra and Molecular Structure 1: Spectra of Diatomic Molecules (Van Nostrand, London, 1967). [259] K. Kitaura, E. Ikeo, T. Asada, T. Nakano, and M. Uebayasi, Chem. Phys. Lett. 313, 701 (1999). [260] D. G. Fedorov and K. Kitaura, J. Chem. Phys. 120, 6832 (2004). [261] D. G. Fedorov, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 7, e1322 (2017). [262] M. S. Gordon, L. Slipchenko, H. Li, and J. H. Jensen, in Annual Reports in Compu- tational Chemistry, edited by D. C. Spellmeyer and R. Wheeler (Elsevier, 2007) pp. 177–193. [263] C. Steinmann, D. G. Fedorov, and J. H. Jensen, J. Phys. Chem. A 114, 8705 (2010). [264] C. Steinmann, D. G. Fedorov, and J. H. Jensen, PLoS One 7, e41117 (2012). [265] C. Steinmann, D. G. Fedorov, and J. H. Jensen, PLoS One 8, e60602 (2013). 172 [266] S. Li, J. Shen, W. Li, and Y. Jiang, J. Chem. Phys. 125, 074109 (2006). [267] W. Li, J. R. Gour, P. Piecuch, and S. Li, J. Chem. Phys. 131, 114109 (2009). [268] W. Li and P. Piecuch, J. Phys. Chem. A 114, 8644 (2010). [269] W. Li and P. Piecuch, J. Phys. Chem. A 114, 6721 (2010). [270] A. Warshel and M. Levitt, J. Mol. Biol. 103, 227 (1976). [271] Y. Zhang, T.-S. Lee, and W. Yang, J. Chem. Phys. 110, 46 (1999). [272] H. Lin and D. G. Truhlar, J. Phys. Chem. A 109, 3991 (2005). [273] G. Giudetti, I. Polyakov, B. L. Grigorenko, S. Faraji, A. V. Nemukhin, and A. I. Krylov, J. Chem. Theoy Comput. 18, 5056 (2022). [274] A. Zlobin, J. Belyaeva, and A. Golovin, J. Chem. Inf. Model. 63, 546 (2023). [275] J. Tomasi, B. Mennucci, and R. Cammi, Chem. Rev. 105, 2999 (2005). [276] A. V. Marenich, C. J. Cramer, and D. G. Truhlar, J. Phys. Chem. B 113, 6378 (2009). 173