This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\UseRawInputEncoding
thanks: Present address: Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, USA.thanks: Corresponding author

High-Level Coupled-Cluster Energetics by Merging Moment Expansions with Selected Configuration Interaction

Karthik Gururangan Department of Chemistry, Michigan State University, East Lansing, Michigan 48824, USA    J. Emiliano Deustua Department of Chemistry, Michigan State University, East Lansing, Michigan 48824, USA    Jun Shen Department of Chemistry, Michigan State University, East Lansing, Michigan 48824, USA    Piotr Piecuch [email protected]. Department of Chemistry, Michigan State University, East Lansing, Michigan 48824, USA Department of Physics and Astronomy, Michigan State University, East Lansing, Michigan 48824, USA
Abstract

Inspired by our earlier semi-stochastic work aimed at converging high-level coupled-cluster (CC) energetics [J. E. Deustua, J. Shen, and P. Piecuch, Phys. Rev. Lett. 119, 223003 (2017); J. Chem. Phys. 154, 124103 (2021)], we propose a novel form of the CC(PP;QQ) theory in which the stochastic Quantum Monte Carlo propagations, used to identify dominant higher–than–doubly excited determinants, are replaced by the selected configuration interaction (CI) approach using the CIPSI algorithm. The advantages of the resulting CIPSI-driven CC(PP;QQ) methodology are illustrated by a few molecular examples, including the dissociation of F2\mathrm{F_{2}} and the automerization of cyclobutadiene, where we recover the electronic energies corresponding to the CC calculations with a full treatment of singles, doubles, and triples based on the information extracted from compact CI wave functions originating from relatively inexpensive Hamiltonian diagonalizations.

I Introduction

One of the key objectives of quantum chemistry is to obtain accurate energetics of molecular systems in a computationally efficient manner. Among the various post-Hartree–Fock (post-HF) theories, the size extensive approaches derived from the exponential ansatz Hubbard (1957); Hugenholtz (1957) of coupled-cluster (CC) theoryCoester (1958); Coester and Kümmel (1960); Čížek (1966, 1969); Paldus, Čížek, and Shavitt (1972) are among the best techniques to accomplish this task.Paldus and Li (1999); Bartlett and Musiał (2007) We recall that the CC wave function for an NN-electron system is defined as

|Ψ=eT|Φ,|\Psi\rangle=e^{T}|\Phi\rangle, (1)

where |Φ\ket{\Phi} is the reference (usually, HF) determinant and

T=n=1NTnT=\sum_{n=1}^{N}T_{n} (2)

is the cluster operator, with TnT_{n} representing its nn-body (nn-particle–nn-hole) component. In practice, one truncates the cluster operator TT at a given many-body rank to define the standard CC hierarchy of approximations. The most basic and most practical one, obtained when TT is truncated at T2T_{2}, which has computational steps that scale as 𝒪(𝒩6)\mathcal{O}(\mathcal{N}^{6}) with the system size 𝒩\mathcal{N}, is the CC method with singles and doubles (CCSD).Purvis and Bartlett (1982); Cullen and Zerner (1982) The next two levels, namely, the CC approach with singles, doubles, and triples, abbreviated as CCSDT,Noga and Bartlett (1987); Scuseria and Schaefer (1988); Watts and Bartlett (1990) which interests us in this study most, obtained when TT is truncated at T3T_{3}, and the CC method with singles, doubles, triples, and quadruples, abbreviated as CCSDTQ,Oliphant and Adamowicz (1991); Kucharski and Bartlett (1992); Piecuch and Adamowicz (1994) in which TT is truncated at T4T_{4}, involve the 𝒪(𝒩8)\mathcal{O}(\mathcal{N}^{8}) and 𝒪(𝒩10)\mathcal{O}(\mathcal{N}^{10}) steps, respectively. It is well established that in the majority of chemical applications, including molecules near equilibrium geometries, bond dissociations involving smaller numbers of strongly correlated electrons, noncovalent interactions, and photochemistry, the conventional CCSD, CCSDT, CCSDTQ, etc. hierarchy and its equation-of-motion (EOM) Emrich (1981); Geertsen, Rittby, and Bartlett (1989); Stanton and Bartlett (1993); Kowalski and Piecuch (2001a, b); Kucharski et al. (2001); Kállay and Gauss (2004); Hirata (2004) and linear-response Monkhorst (1977); Dalgaard and Monkhorst (1983); Mukherjee and Mukherjee (1979); Sekino and Bartlett (1984); Takahashi and Paldus (1986); Koch and Jørgensen (1990); Koch et al. (1990); Kondo, Piecuch, and Paldus (1995a, b) extensions rapidly approach the exact, full configuration interaction (FCI) limit, so that by the time one reaches the CCSDT or CCSDTQ levels, the results are usually converged with respect to the relevant many-electron correlation effects,Bartlett and Musiał (2007) but the 𝒪(𝒩8)\mathcal{O}(\mathcal{N}^{8}) computational steps of CCSDT or the 𝒪(𝒩10)\mathcal{O}(\mathcal{N}^{10}) steps of CCSDTQ render the usage of such methods unfeasible for most problems of interest. Thus, one of the main activities in the CC development work has been the design of high-fidelity approximations to CCSDT and CCSDTQ, capable of reducing the above costs, while being more robust than perturbative approaches of the CCSD(T)Raghavachari et al. (1989) type, which fail in more multi-reference situations. Paldus and Li (1999); Bartlett and Musiał (2007); Piecuch et al. (2002, 2004); Piecuch (2010)

To that end, our group has recently developed the semi-stochastic CC(PP;QQ) formalism, Deustua, Shen, and Piecuch (2017, 2021); Yuwono et al. (2020) a novel methodology that can efficiently converge the energetics of high-level CC calculations, such as CCSDT, CCSDTQ, and EOMCCSDT,Kowalski and Piecuch (2001a, b); Kucharski et al. (2001) by combining the deterministic CC(PP;QQ) framework Shen and Piecuch (2012a, b, c); Bauman, Shen, and Piecuch (2017); Magoulas et al. (2018) with the stochastic Quantum Monte Carlo (QMC) wave function propagations in the many-electron Hilbert space defining the CIQMCBooth, Thom, and Alavi (2009); Cleland, Booth, and Alavi (2010); Dobrautz, Smart, and Alavi (2019); Ghanem, Lozovoi, and Alavi (2019); Ghanem, Guther, and Alavi (2020) and CC Monte Carlo (CCMC)Thom (2010); Franklin et al. (2016); Spencer and Thom (2016); Scott and Thom (2017) approaches. The semi-stochastic CC(PP;QQ) methodology of Refs. Deustua, Shen, and Piecuch, 2017, 2021; Yuwono et al., 2020 leverages the fact, recognized long time ago in the context of active-space CC considerations (cf. Ref. Piecuch, 2010 for a review), that higher-order cluster operators, such as T3T_{3} and T4T_{4}, and their counterparts utilized in EOMCC are usually relatively sparse. In the semi-stochastic CC(PP;QQ) approaches, the dominant higher–than–doubly excited cluster/excitation amplitudes relevant to the parent CC/EOMCC theory of interest are automatically selected using stochastic CIQMC or CCMC wave function propagations that provide lists of Slater determinants for the initial CC(PP) Deustua, Shen, and Piecuch (2017, 2021) or EOMCC(PP) Deustua et al. (2019); Yuwono et al. (2020) calculations, which are subsequently corrected using the biorthogonal moment expansions adopted in the CC(PP;QQ) formalism to capture the remaining correlations. The semi-stochastic CC(PP;QQ) methods have demonstrated their ability to rapidly converge the CCSDT,Deustua, Shen, and Piecuch (2017, 2021); Yuwono et al. (2020) CCSDTQ,Deustua, Shen, and Piecuch (2021) and EOMCCSDTYuwono et al. (2020) energetics out of the early stages of the underlying CIQMC or CCMC propagations, with minimal reliance on user- and system-dependent inputs.

Encouraged by the above findings, in this study we explore the use of selected CI as an alternative provider of the lists of the leading higher–than–doubly excited determinants needed to drive the CC(PP;QQ) computations. The selected CI schemes, which date back to the pioneering efforts in the late 1960s and early 1970s,Whitten and Hackmeyer (1969); Bender and Davidson (1969); Huron, Malrieu, and Rancurel (1973); Buenker and Peyerimhoff (1974) have recently regained significant interest, as their modern implementations have demonstrated the ability to capture the bulk of many-electron correlation effects in a computationally efficient manner using a conceptually straightforward linear wave function ansatz. Schriber and Evangelista (2016, 2017); Tubman et al. (2016, 2020); Liu and Hoffmann (2016); Zhang, Liu, and Hoffmann (2020); Holmes, Tubman, and Umrigar (2016); Sharma et al. (2017); Li et al. (2018); Garniron et al. (2017, 2019); Loos, Damour, and Scemama (2020); Eriksen et al. (2020) The selected CI model adopted in the CC(PP;QQ) considerations reported in this work is the CI method using perturbative selection made iteratively (CIPSI),Huron, Malrieu, and Rancurel (1973) as recently reformulated and further developed in Refs. Garniron et al., 2017, 2019.

II Theory and Algorithmic Details

We begin by reviewing the key elements of the ground-state CC(PP;QQ) formalism relevant to this work. Each CC(PP;QQ) calculation starts by identifying two disjoint subspaces of the NN-electron Hilbert space, the PP space designated as (P)\mathscr{H}^{(P)} and the QQ space denoted as (Q)\mathscr{H}^{(Q)}. The former space is spanned by the excited determinants |ΦK=EK|Φ|\Phi_{K}\rangle=E_{K}|\Phi\rangle, where EKE_{K} is the elementary particle–hole excitation operator generating |ΦK|\Phi_{K}\rangle from |Φ|\Phi\rangle, which together with |Φ|\Phi\rangle dominate the ground-state wave function, whereas the determinants in (Q)\mathscr{H}^{(Q)} are used to construct the correction due to the correlation effects the CC calculations in the PP space do not describe. Once the PP and QQ spaces are defined, we solve the CC amplitude equations

𝔐K(P)=0,|ΦK(P),\mathfrak{M}_{K}(P)=0,\;\;|\Phi_{K}\rangle\in\mathscr{H}^{(P)}, (3)

where

𝔐K(P)=ΦK|H¯(P)|Φ,\mathfrak{M}_{K}(P)=\langle\Phi_{K}|\overline{H}^{(P)}|\Phi\rangle, (4)

with

H¯(P)=eT(P)HeT(P),\overline{H}^{(P)}=e^{-T^{(P)}}He^{T^{(P)}}, (5)

are moments of the CC equations,Jankowski, Paldus, and Piecuch (1991); Piecuch and Kowalski (2000); Kowalski and Piecuch (2000) to obtain the approximate form of the cluster operator in the PP space,

T(P)=|ΦK(P)tKEK,T^{(P)}=\sum_{\ket{\Phi_{K}}\in\mathscr{H}^{(P)}}t_{K}E_{K}, (6)

and the corresponding ground-state energy

E(P)=Φ|H¯(P)|Φ,E^{(P)}=\langle\Phi|\overline{H}^{(P)}|\Phi\rangle, (7)

and calculate the noniterative correction δ(P;Q)\delta(P;Q) to determine the final CC(PP;QQ) energy as

E(P+Q)=E(P)+δ(P;Q).E^{(P+Q)}=E^{(P)}+\delta(P;Q). (8)

The correction δ(P;Q)\delta(P;Q) to the energy E(P)E^{(P)} obtained in the PP-space CC [CC(PP)] calculations is given byShen and Piecuch (2012a, b)

δ(P;Q)=|ΦK(Q)K(P)𝔐K(P),\delta(P;Q)=\sum_{\ket{\Phi_{K}}\in\mathscr{H}^{(Q)}}\ell_{K}(P)\>\mathfrak{M}_{K}(P), (9)

where 𝔐K(P)\mathfrak{M}_{K}(P) is defined by Eq. (4) and

K(P)=Φ|(1+Λ(P))H¯(P)|ΦK/DK(P),\ell_{K}(P)=\langle\Phi|(1+\Lambda^{(P)})\overline{H}^{(P)}|\Phi_{K}\rangle/D_{K}^{(P)}, (10)

with

DK(P)=E(P)ΦK|H¯(P)|ΦK.D_{K}^{(P)}=E^{(P)}-\langle\Phi_{K}|\overline{H}^{(P)}|\Phi_{K}\rangle. (11)

The Λ(P)\Lambda^{(P)} operator entering Eq. (10), given by

Λ(P)=|ΦK(P)λK(EK)\Lambda^{(P)}=\sum_{\ket{\Phi_{K}}\in\mathscr{H}^{(P)}}\lambda_{K}(E_{K})^{\dagger} (12)

and obtained by solving the linear system

Φ|(1+Λ(P))H¯(P)|ΦK=E(P)λK,|ΦK(P),\bra{\Phi}(1+\Lambda^{(P)})\overline{H}^{(P)}\ket{\Phi_{K}}=E^{(P)}\lambda_{K},\;\;\ket{\Phi_{K}}\in\mathscr{H}^{(P)}, (13)

is the hole–particle deexcitation operator defining the bra state Ψ~(P)|=Φ|(1+Λ(P))eT(P)\langle\tilde{\Psi}^{(P)}|=\langle\Phi|(1+\Lambda^{(P)})e^{-T^{(P)}} associated with the CC(PP) ket state |Ψ(P)=eT(P)|Φ|\Psi^{(P)}\rangle=e^{T^{(P)}}|\Phi\rangle.

The CC(PP;QQ) formalism includes the completely renormalized CC methods, such as CR-CC(2,3),Piecuch and Włoch (2005); Piecuch et al. (2006); Włoch et al. (2006); Włoch, Gour, and Piecuch (2007) which works better than CCSD(T) in bond breaking situations, but its main advantage is the freedom to make unconventional choices of the PP and QQ spaces, allowing one to relax the lower-order T1T_{1} and T2T_{2} clusters in the presence of their higher-order counterparts, such as the leading T3T_{3} contributions, which the CCSD(T), CR-CC(2,3), and other triples corrections to CCSD cannot do. One can use active orbitals to identify the leading higher–than–doubly excited determinants for the inclusion in the PP space used in the CC(PP) calculations, employing the δ(P;Q)\delta(P;Q) corrections to capture the remaining correlations of interest, as in the CC(t;3) and similar approaches, Shen and Piecuch (2012a, b, c); Bauman, Shen, and Piecuch (2017); Magoulas et al. (2018) or adopt a more black-box semi-stochastic CC(PP;QQ) framework, in which the selection of the dominant higher–than–doubly excited determinants entering the PP space is automated with the help of CIQMC or CCMC propagations. Deustua, Shen, and Piecuch (2017, 2021); Yuwono et al. (2020) In this article, we propose an alternative to the semi-stochastic CC(PP;QQ) methodology, in which we use the information extracted from the CIPSI runs to populate the PP spaces employed in the CC(PP) calculations preceding the determination of the δ(P;Q)\delta(P;Q) corrections.

We recall that the CIPSI approach, originally proposed in Ref. Huron, Malrieu, and Rancurel, 1973 and further developed in Refs. Garniron et al., 2017, 2019, seeks to construct an approximation to the FCI wave function by a series of Hamiltonian diagonalizations in increasingly large, iteratively defined, subspaces of the many-electron Hilbert space, designated as 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)}, where k=0,1,2,k=0,1,2,\ldots enumerates the consecutive CIPSI iterations. The initial subspace 𝒱int(0)\mathcal{V}_{\text{int}}^{(0)} can be one-dimensional, if the CIPSI calculations are started from a single determinant, such as the restricted HF (RHF) wave function used throughout this work as a reference, or multi-dimensional, if one prefers to start from a multi-determinantal state generated in some preliminary truncated CI computation, and the remaining subspaces are constructed via a recursive process in which 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)} is obtained by augmenting 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} with a subset of the leading singly and doubly excited determinants out of 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} identified with the help of the many-body perturbation theory (MBPT). Thus, if |Ψk(CIPSI)=|ΦI𝒱int(k)cI|ΦI|\Psi^{(\text{CIPSI})}_{k}\rangle=\sum_{|\Phi_{I}\rangle\in\mathcal{V}_{\text{int}}^{(k)}}c_{I}|\Phi_{I}\rangle is a CI wave function obtained by diagonalizing the Hamiltonian in the current subspace 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} and Evar,kE_{\text{var},k} is the corresponding energy, and if 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} is the space of all singly and doubly excited determinants out of |Ψk(CIPSI)|\Psi^{(\text{CIPSI})}_{k}\rangle, for each determinant |Φα𝒱ext(k)|\Phi_{\alpha}\rangle\in\mathcal{V}_{\text{ext}}^{(k)} we evaluate the second-order MBPT correction eα,k(2)=|Φα|H|Ψk(CIPSI)|2/(Evar,kΦα|H|Φα)e_{\alpha,k}^{(2)}=|\langle\Phi_{\alpha}|H|\Psi_{k}^{(\text{CIPSI})}\rangle|^{2}/(E_{\text{var},k}-\langle\Phi_{\alpha}|H|\Phi_{\alpha}\rangle) and use the resulting eα,k(2)e_{\alpha,k}^{(2)} values to decide which determinants from 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} should be added to the determinants |ΦI|\Phi_{I}\rangle already in 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} to construct the next diagonalization space 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)}. We can also use the eα,k(2)e_{\alpha,k}^{(2)} values to calculate the perturbatively corrected CIPSI energies Evar,k+ΔEk(2)E_{\text{var},k}+\Delta E^{(2)}_{k}, where ΔEk(2)=|Φα𝒱ext(k)eα,k(2)\Delta E^{(2)}_{k}=\sum_{|\Phi_{\alpha}\rangle\in\mathcal{V}_{\text{ext}}^{(k)}}e_{\alpha,k}^{(2)}, and, after further manipulations, their Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} counterparts, in which ΔEk(2)\Delta E^{(2)}_{k} is replaced by its renormalized ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} form introduced in Ref. Garniron et al., 2019.

In the modern implementation of CIPSI, formulated in Refs. Garniron et al., 2017, 2019 and available in the Quantum Package 2.0 software,Garniron et al. (2019) which we used in the present study, the process of enlarging 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} to generate 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)} is executed in the following manner. First, prior to examining the eα,k(2)e_{\alpha,k}^{(2)} corrections, one stochastically samples the most important singly and doubly excited determinants out of |Ψk(CIPSI)|\Psi^{(\text{CIPSI})}_{k}\rangle, so that not all singles and doubles from 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} are included in the accompanying 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} space, only the sampled ones. Next, one arranges the sampled determinants |Φα𝒱ext(k)|\Phi_{\alpha}\rangle\in\mathcal{V}_{\text{ext}}^{(k)} in descending order according to their |eα,k(2)||e^{(2)}_{\alpha,k}| values. The process of enlarging the current subspace 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} to construct the 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)} space for the subsequent Hamiltonian diagonalization, which starts from the determinants |Φα|\Phi_{\alpha}\rangle characterized by the largest |eα,k(2)||e^{(2)}_{\alpha,k}| contributions, moving toward those that have smaller |eα,k(2)||e^{(2)}_{\alpha,k}| values, continues until the number of determinants in 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)} exceeds the dimension of 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} multiplied by a user-defined factor f>1f>1. In this study, we used f=2f=2, which is the default in Quantum Package 2.0 (we will examine other choices of ff in the future). In practice, a typical dimension of 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)}, including each of the final diagonalization spaces used to generate lists of higher–than–doubly excited determinants for the CC(PP) calculations reported in this work, is slightly larger than ff times the dimension of 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)}, since the CIPSI algorithm adds extra determinants to 𝒱int(k+1)\mathcal{V}_{\text{int}}^{(k+1)} to ensure that the resulting |Ψk+1(CIPSI)|\Psi_{k+1}^{(\text{CIPSI})}\rangle wave function is an eigenstate of the total spin S2S^{2} and SzS_{z} operators. The final wave function |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle of a given CIPSI run and the associated variational (EvarE_{\text{var}}) and perturbatively corrected [Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)}] energies are obtained by terminating the above procedure in one of the following two ways: (i) stopping at the first iteration kk for which the second-order MBPT correction |ΔEk(2)||\Delta E^{(2)}_{k}| falls below a user-defined threshold η\eta, indicating that the CIPSI wave function is within a tolerable distance from FCI, or (ii) stopping at the first iteration kk for which the number of determinants in the corresponding 𝒱int(k)\mathcal{V}_{\text{int}}^{(k)} space exceeds a user-defined input parameter Ndet(in)N_{\text{det(in)}}. Since our main objective is to employ the CIPSI-driven CC(PP;QQ) algorithm to accurately reproduce the high-level CC rather than FCI energetics, without having to converge the underlying CIPSI sequence, we chose the latter option, which we enforced by using η=106\eta=10^{-6} hartree. As a result of setting the input parameter ff at 2, the sizes of the final wave functions |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle produced by our CIPSI runs, denoted as Ndet(out)N_{\text{det(out)}}, were always between Ndet(in)N_{\text{det(in)}} and 2Ndet(in)2N_{\text{det(in)}}.

Having discussed the key ingredients of the CC(PP;QQ) and CIPSI methodologies relevant to this work, we proceed to the description of the CIPSI-driven CC(PP;QQ) algorithm, which consists of the following steps:

  • 1.

    Given a reference state |Φ|\Phi\rangle, which in all of the calculations reported in this article was the RHF determinant, choose an input parameter Ndet(in)N_{\text{det(in)}}, used to terminate the CIPSI wave function growth, and execute a CIPSI run to obtain the final state |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle spanned by Ndet(out)N_{\text{det(out)}} determinants.

  • 2.

    Extract a list of higher–than–doubly excited determinants relevant to the desired CC theory level from |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle to define the PP space. If the goal is to converge the CCSDT energetics, the PP space consists of all singly and doubly excited determinants plus the triply excited determinants contained in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. To recover the CCSDTQ energetics, quadruply excited determinants contributing to |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle are included in the PP space as well.

  • 3.

    Solve the nonlinear CC(PP) system, Eq. (3), and the associated linear system given by Eq. (13), where E(P)E^{(P)} is defined by Eq. (7), in the PP space determined in Step 2 to obtain the cluster operator T(P)T^{(P)} and the deexcitation operator Λ(P){\Lambda}^{(P)}. If the target approach is CCSDT, define T(P)=T1+T2+T3(CIPSI)T^{(P)}=T_{1}+T_{2}+T_{3}^{(\text{CIPSI})} and Λ(P)=Λ1+Λ2+Λ3(CIPSI)\Lambda^{(P)}=\Lambda_{1}+\Lambda_{2}+\Lambda_{3}^{(\text{CIPSI})}, where the list of triples entering T3(CIPSI)T_{3}^{(\text{CIPSI})} and Λ3(CIPSI)\Lambda_{3}^{(\text{CIPSI})} is identical to that extracted from |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle in Step 2. If the goal is to converge the CCSDTQ energetics, define T(P)=T1+T2+T3(CIPSI)+T4(CIPSI)T^{(P)}=T_{1}+T_{2}+T_{3}^{(\text{CIPSI})}+T_{4}^{(\text{CIPSI})} and Λ(P)=Λ1+Λ2+Λ3(CIPSI)+Λ4(CIPSI)\Lambda^{(P)}=\Lambda_{1}+\Lambda_{2}+\Lambda_{3}^{(\text{CIPSI})}+\Lambda_{4}^{(\text{CIPSI})}, where the list of triples entering T3(CIPSI)T_{3}^{(\text{CIPSI})} and Λ3(CIPSI)\Lambda_{3}^{(\text{CIPSI})} and the list of quadruples entering T4(CIPSI)T_{4}^{(\text{CIPSI})} and Λ4(CIPSI)\Lambda_{4}^{(\text{CIPSI})} are again extracted from |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle.

  • 4.

    Use the information obtained in Step 3 to determine correction δ(P;Q)\delta(P;Q), Eq. (9), which describes the remaining correlations of interest that were not captured by the CC(PP) calculations. If the goal is to converge the CCSDT energetics, define the QQ space needed to calculate δ(P;Q)\delta(P;Q) as the remaining triply excited determinants that are not contained in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. If the target approach is CCSDTQ, define the QQ space as the triply and quadruply excited determinants absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. Add the resulting correction δ(P;Q)\delta(P;Q) to E(P)E^{(P)} to obtain the CC(PP;QQ) energy E(P+Q)E^{(P+Q)}, Eq. (8).

  • 5.

    To check convergence, repeat Steps 1–4 for a larger value of Ndet(in)N_{\text{det(in)}}. The CIPSI-driven CC(PP;QQ) calculations can be regarded as converged if the difference between consecutive E(P+Q)E^{(P+Q)} energies falls below a user-defined threshold. In analogy to the semi-stochastic CC(PP;QQ) framework of Refs. Deustua, Shen, and Piecuch, 2017, 2021; Yuwono et al., 2020, one can also stop if the fraction(s) of higher–than–doubly excited determinants contained in the final CIPSI state |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle is (are) sufficiently large to produce the desired accuracy level.

In this initial exploratory study, we implemented the CIPSI-driven CC(PP;QQ) approach that allows us to converge the CCSDT energetics. We did this by modifying our standalone CC(PP;QQ) codes, described in Refs. Deustua, Shen, and Piecuch, 2017, 2021; Yuwono et al., 2020; Shen and Piecuch, 2012a, b, c; Bauman, Shen, and Piecuch, 2017 and interfaced with the RHF and integral transformation routines available in GAMESS,Schmidt et al. (1993); Barca et al. (2020) such that they can use the lists of triply excited determinants extracted from the CIPSI wave functions |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle, generated with Quantum Package 2.0, to set up the relevant PP spaces (as already explained, the corresponding QQ spaces are automatically defined as the remaining triples absent in the |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle wave functions). By design, as the input parameter Ndet(in)N_{\text{det(in)}} used to terminate CIPSI runs increases, producing longer and longer CI expansions to represent wave functions |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle, the CC(PP;QQ) energies E(P+Q)E^{(P+Q)} approach their CCSDT parents. The underlying CC(PP) calculations converge the CCSDT energetics too, but, as further elaborated on in Section III, by ignoring the triples that were not captured by CIPSI, they do it at a much slower rate. In examining the convergence of the CIPSI-driven CC(PP) and CC(PP;QQ) energies toward CCSDT, we sampled the Ndet(in)N_{\text{det(in)}} values in a roughly semi-logarithmic manner, starting from Ndet(in)=1N_{\text{det(in)}}=1. Since all of the calculations reported in this work adopted RHF determinants as reference functions, the |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state becomes the RHF determinant and the resulting CC(PP) and CC(PP;QQ) energies become identical to those obtained in the RHF-based CCSD and CR-CC(2,3) calculations, respectively, when Ndet(in)=1N_{\text{det(in)}}=1. Thus, in analogy to the QMC propagation time τ\tau used in our semi-stochastic CC(PP)/EOMCC(PP) and CC(PP;QQ) studies,Deustua, Shen, and Piecuch (2017); Deustua et al. (2019); Yuwono et al. (2020); Deustua, Shen, and Piecuch (2021) we can regard the Ndet(in)N_{\text{det(in)}} input variable defining CIPSI computations as the parameter connecting CCSD [in the CC(PP) case] or CR-CC(2,3) [in the case of CC(PP;QQ) runs] with CCSDT. As a result, similarly to CCSD, CR-CC(2,3), and CCSDT, the CIPSI-driven CC(PP) and CC(PP;QQ) approaches considered in this work remain size extensive for all values of Ndet(in)N_{\text{det(in)}}. The CC(PP) calculations are size extensive, since they are nothing else than the usual CC computations in which we solve the connected amplitude equations, Eq. (3), for the cluster operator T(P)T^{(P)} defined by Eq. (6). In the case of the CIPSI-driven CC(PP) method implemented in this study, T(P)=T1+T2+T3(CIPSI)T^{(P)}=T_{1}+T_{2}+T_{3}^{(\text{CIPSI})}, where T3(CIPSI)=|Φijkabc|Ψ(CIPSI)tabcijkEijkabcT_{3}^{(\text{CIPSI})}=\sum_{|\Phi_{ijk}^{abc}\rangle\in|\Psi^{(\text{CIPSI})}\rangle}t_{abc}^{ijk}\>E_{ijk}^{abc} is the T3T_{3} operator defined using the list of triply excited determinants |Φijkabc|\Phi_{ijk}^{abc}\rangle contained in the final CIPSI state |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle (we use the usual notation in which i,j,ki,j,k and a,b,ca,b,c designate the occupied and unoccupied spin-orbitals in |Φ|\Phi\rangle, respectively, and EijkabcE_{ijk}^{abc} is the elementary triple excitation operator generating |Φijkabc|\Phi_{ijk}^{abc}\rangle from |Φ|\Phi\rangle). The noniterative correction δ(P;Q)\delta(P;Q), Eq. (9), which in the case of the CIPSI-driven CC(PP;QQ) approach developed in this work captures the T3T_{3} effects not described by T3(CIPSI)T_{3}^{(\text{CIPSI})} and which involves the summation over the remaining triply excited determinants that are not included in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle, i.e., δ(P;Q)=|Φijkabc|Ψ(CIPSI)ijkabc𝔐abcijk\delta(P;Q)=\sum_{|\Phi_{ijk}^{abc}\rangle\notin|\Psi^{(\text{CIPSI})}\rangle}\ell_{ijk}^{abc}\>\mathfrak{M}_{abc}^{ijk}, being the connected quantity similar to that used in the CR-CC(2,3) and CC(t;3) methods, is size extensive too (for the early numerical illustration of the size extensivity of CR-CC(2,3), see Ref. Piecuch et al., 2006).

The numerical demonstration of the size extensivity of the CIPSI-driven CC(PP) and CC(PP;QQ) methods implemented in this work is shown in Table 1. Our example is the noninteracting F2+Ne{\rm F}_{2}+{\rm Ne} system, described by the cc-pVDZ basis set,Dunning (1989) obtained by placing the Ne atom at 1,000 bohr from the stretched fluorine molecule in which the F–F bond length was set at twice its equilibrium value to increase the magnitude of T3T_{3} correlations. Along with the F2+Ne{\rm F}_{2}+{\rm Ne} system, we consider the isolated F2{\rm F}_{2} molecule having the same geometry as in F2+Ne{\rm F}_{2}+{\rm Ne} and the isolated neon atom, both described by the cc-pVDZ basis. The CIPSI diagonalization sequence used to provide the list of triply excited determinants for the inclusion in the PP space corresponding to the F2+Ne{\rm F}_{2}+{\rm Ne} system was initiated from the RHF reference determinant and defined by setting the wave function termination parameter Ndet(in)N_{\text{det(in)}}, the input parameter ff controlling the CIPSI wave function growth, and the MBPT-based stopping parameter η\eta at 5,000, 2, and 10610^{-6} hartree, respectively. Following the above description, the PP space used in the CIPSI-driven CC(PP) calculation for the noninteracting F2+Ne{\rm F}_{2}+{\rm Ne} dimer consisted of all singly and doubly excited determinants and a subset of triply excited determinants contained in the last |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state of the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run. The QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples not included in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. To ensure the consistency of the PP spaces used in the CC(PP) calculations for the F2+Ne{\rm F}_{2}+{\rm Ne} system and the F2{\rm F}_{2} and Ne fragments, we generated the PP space for F2{\rm F}_{2} by removing the triply excited determinants involving the orbitals of Ne from the list of triples obtained in the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for F2+Ne{\rm F}_{2}+{\rm Ne}. Similarly, the PP space used in the CC(PP) calculations for Ne was obtained by starting from the list of triples produced in the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI calculation for F2+Ne{\rm F}_{2}+{\rm Ne} and removing the triply excited determinants involving the orbitals of F2{\rm F}_{2}. As in all CC(PP;QQ) calculations considered in this work, the QQ spaces associated with the F2{\rm F}_{2} and Ne monomers were defined as the remaining triples missing in the respective PP spaces. We chose the Ndet(in)=5,000N_{\text{det(in)}}=5,000 value in the size extensivity test reported in Table 1, since it is sufficiently large to introduce the leading triply excited determinants into the relevant PP spaces, while being small enough to produce the CC(PP) and CC(PP;QQ) energies that are visibly different than their CCSDT counterparts.

It is clear from the results presented in Table 1 that, in analogy to CCSD, CR-CC(2,3), and CCSDT, the CIPSI-driven CC(PP) and CC(PP;QQ) methods are size extensive. Indeed, the CC(PP) and CC(PP;QQ) energies of the F2+Ne{\rm F}_{2}+{\rm Ne} dimer are numerically identical to the corresponding sums of the energies of the F2{\rm F}_{2} and Ne monomers. We observe the same behavior for other values of the CIPSI wave function termination parameter Ndet(in)N_{\text{det(in)}}. One may ask a question why the interfragment triply excited determinants |Φijkabc|\Phi_{ijk}^{abc}\rangle having spin-orbital indices located on different monomers, which are present in the final CIPSI state |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle of the noninteracting F2+Ne{\rm F}_{2}+{\rm Ne} system and thus end up in the corresponding PP space, do not result in the violation of size extensivity in the CC(PP) and CC(PP;QQ) calculations. The answer to this question is that the connected triply excited cluster amplitudes tabcijkt_{abc}^{ijk} carrying indices located on different noninteracting fragments vanish when obtained by solving the explicitly connected CC(PP) amplitude equations, Eq. (3). We did not remove such interfragment tabcijkt_{abc}^{ijk} amplitudes from our CC(PP) calculations for the F2+Ne{\rm F}_{2}+{\rm Ne} system and confirmed that they do indeed vanish. The use of CI diagonalizations in constructing the PP spaces for the CC(PP) and CC(PP;QQ) computations does not affect the size extensivity of the CIPSI-driven CC(PP) and CC(PP;QQ) approaches, since all we need from these diagonalizations are the lists of higher–than–doubly excited determinants relevant to the CC theory of interest (in our case, where we target the CCSDT energetics, the lists of triples), not the CI excitation amplitudes themselves. For example, as in all conventional CC calculations, the contributions from the interfragment triply excited determinants |Φijkabc|\Phi_{ijk}^{abc}\rangle to the ground-state wave function of the noninteracting F2+Ne{\rm F}_{2}+{\rm Ne} dimer are represented in the CC(PP) calculations by the disconnected T1T2T_{1}T_{2} and (1/6)T13(1/6)T_{1}^{3} clusters. The noniterative correction δ(P;Q)\delta(P;Q), which provides information about those T3T_{3} correlations that were not captured by the preceding CC(PP) run, becomes the sum of the δ(P;Q)\delta(P;Q) values for the isolated F2{\rm F}_{2} and Ne fragments.

Aside from size extensivity, as analyzed above, and high efficiency in converging the parent CCSDT energetics discussed in Section III, and in analogy to the active-orbital-based Shen and Piecuch (2012a, b, c); Bauman, Shen, and Piecuch (2017) and semi-stochastic Deustua, Shen, and Piecuch (2017, 2021); Yuwono et al. (2020) CC(PP;QQ) approaches, the CIPSI-driven CC(PP;QQ) methodology examined in this work offers significant savings in the computational effort compared to full CCSDT. This is largely related to the fact that, as shown in Section III, the convergence of the CIPSI-driven CC(PP;QQ) energies toward their CCSDT parents with the wave function termination parameter Ndet(in)N_{\text{det(in)}}, with the number of determinants used to generate the final CIPSI state Ndet(out)N_{\text{det(out)}}, and with the fractions of triples in the PP space captured by the CIPSI algorithm is very fast. Indeed, the CPU times associated with the CIPSI runs using smaller Ndet(in)N_{\text{det(in)}} values, resulting in smaller diagonalization spaces, are relatively short compared to the converged CIPSI computations. Next, as explained in detail in Refs. Deustua, Shen, and Piecuch, 2017, 2021; Yuwono et al., 2020, the CC(PP) calculations using small fractions of triples in the PP space, which is all one needs to converge the CCSDT-level energetics in the CIPSI-driven CC(PP;QQ) runs, are much faster than the corresponding CCSDT computations. Finally, as also explained in Refs. Deustua, Shen, and Piecuch, 2017, 2021; Yuwono et al., 2020, the computational cost of determining the CC(PP;QQ) correction δ(P;Q)\delta(P;Q) is less than the cost of a single iteration of CCSDT.

In examining the CIPSI-driven CC(PP) and CC(PP;QQ) energies in Section III, we are primarily interested in how fast they converge toward their parent CCSDT values as Ndet(in)N_{\text{det(in)}} and the fraction of triples in the PP space increase. In the case of the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies, we do what is often done in CIPSI calculations (see, e.g., Refs. Garniron et al., 2019; Loos, Damour, and Scemama, 2020) and compare them to their counterparts obtained by extrapolating the data obtained in the CIPSI runs defined by the largest Ndet(in)N_{\text{det(in)}} values to the FCI limit. Specifically, following the procedure used in Ref. Loos, Damour, and Scemama, 2020, we performed a linear fit of the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} energies leading to the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state obtained for the largest value of Ndet(in)N_{\text{det(in)}} in a given CIPSI sequence, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, and extrapolated the resulting line to the ΔEr,k(2)=0\Delta E_{\text{r},k}^{(2)}=0 limit.

III Numerical Examples

We illustrate potential benefits offered by the CIPSI-driven CC(PP;QQ) methodology, when applied to recovering the CCSDT energetics, using a few molecular examples. Our first example is the frequently studied dissociation of the fluorine molecule, as described by the cc-pVDZ basis set. We chose this example, since it is well established that the CCSDT approach provides an accurate description of bond breaking in F2{\rm F}_{2} (cf., e.g., Refs. Piecuch and Włoch, 2005; Piecuch et al., 2006; Kowalski and Piecuch, 2001c; Musiał and Bartlett, 2005) and since we previously used it to benchmark the CC(PP;QQ)-based CC(t;3) approachShen and Piecuch (2012a) and the semi-stochastic CC(PP;QQ) methods driven by CIQMCDeustua, Shen, and Piecuch (2017, 2021) and CCMCDeustua, Shen, and Piecuch (2017) propagations. The results of our calculations for the F2\mathrm{F_{2}}/cc-pVDZ system, in which the F–F bond length RR was stretched from its equilibrium, Re=2.66816R_{e}=2.66816 bohr, value, where electron correlation effects are largely dynamical in nature, to 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e}, where they gain an increasingly nondynamical character, are summarized in Table 2 and Fig. 1. The complexity of electron correlations in F2\mathrm{F_{2}} manifests itself in the rapidly growing magnitude of T3T_{3} contributions as the F–F distance increases, as exemplified by the unsigned differences between the CCSDT and CCSD energies, which are 9.485 millihartree at R=ReR=R_{e}, 32.424 millihartree at R=1.5ReR=1.5R_{e}, 45.638 millihartree at R=2ReR=2R_{e}, and 49.816 millihartree at R=5ReR=5R_{e}, when the cc-pVDZ basis set is employed. They grow with RR so fast that in the R=2Re5ReR=2R_{e}-5R_{e} region they become larger than the depth of the CCSDT potential (estimated at \sim44 millihartree when the CCSDT energy at R=ReR=R_{e} is subtracted from its R=5ReR=5R_{e} counterpart) and highly nonperturbative. The latter feature of T3T_{3} contributions in the stretched F2\mathrm{F_{2}} molecule can be seen by examining the errors relative to CCSDT obtained in the CCSD(T) calculations at R=1.5ReR=1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e}, which are 5.711-5.711, 23.596-23.596, and 39.348-39.348 millihartree, respectively, when the cc-pVDZ basis set is used. As shown in Table 2 [see the Ndet(in)=1N_{\text{det(in)}}=1 CC(PP;QQ) energies], the CR-CC(2,3) triples correction to CCSD helps, reducing the large errors characterizing CCSD(T) to 1.735 millihartree at R=1.5ReR=1.5R_{e}, 1.862 millihartree at R=2ReR=2R_{e}, and 1.613 millihartree at R=5ReR=5R_{e}, which are much more acceptable, but, as demonstrated in our earlier active-orbital-based and semi-stochastic CC(PP;QQ) studies, Shen and Piecuch (2012a); Deustua, Shen, and Piecuch (2017, 2021) further error reduction requires the relaxation of T1T_{1} and T2T_{2} clusters in the presence of the dominant T3T_{3} contributions. This is what the CIPSI-driven CC(PP;QQ) methodology, where we use CIPSI runs to identify the leading triple excitations for the inclusion in the PP space, allows us to do.

Indeed, with as little as 1,006–1,442 Sz=0S_{z}=0 determinants of the Ag(D2h)A_{g}(D_{2h}) symmetry in the final Hamiltonian diagonalization spaces (we used D2hD_{2h} group, which is the largest Abelian subgroup of the DhD_{\infty h} symmetry group of F2\mathrm{F_{2}}, in our calculations), generated by the inexpensive Ndet(in)=1,000N_{\text{det(in)}}=1,000 CIPSI runs at R=1.5ReR=1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e}, which capture very small fractions, on the order of 0.1–0.2 %, of all triples, the errors in the resulting CC(PP;QQ) energies relative to CCSDT are 0.202 millihartree at R=1.5ReR=1.5R_{e}, 0.132 millihartree at R=2ReR=2R_{e}, and 0.144 millihartree at R=5ReR=5R_{e}. This is an approximately tenfold error reduction compared to the CR-CC(2,3) calculations, in which T1T_{1} and T2T_{2} clusters, obtained with CCSD, are decoupled from T3T_{3}, and an improvement of the faulty CCSD(T) energetics by orders of magnitude. As explained in detail in our papers on the CIQMC/CCMC-driven CC(PP;QQ) approaches, Deustua, Shen, and Piecuch (2017, 2021); Yuwono et al. (2020) with the fractions of triples in the relevant PP spaces being so small, the post-CIPSI steps of the CC(PP;QQ) calculations are not much more expensive than the CCSD-based CR-CC(2,3) computations and a lot faster than the corresponding CCSDT computations. The CC(PP;QQ) calculations using Ndet(in)=1,000N_{\text{det(in)}}=1,000 do not offer any improvements over CR-CC(2,3) at the equilibrium geometry, since the final diagonalization space of the underlying CIPSI run does not yet contain any triply excited determinants, and the CR-CC(2,3) energy at R=ReR=R_{e} is already very accurate anyway, but with the relatively small additional effort corresponding to Ndet(in)=10,000N_{\text{det(in)}}=10,000, which results in 10,150 Sz=0S_{z}=0 determinants of the Ag(D2h)A_{g}(D_{2h}) symmetry in the final CIPSI diagonalization space and only 1.2 % of all triples in the PP space, the unsigned error in the CC(PP;QQ) energy relative to its CCSDT parent substantially decreases, from 0.240 millihartree, when Ndet(in)1,000N_{\text{det(in)}}\leq 1,000, to 67 microhartree, when Ndet(in)N_{\text{det(in)}} is set at 10,000. The use of Ndet(in)=10,000N_{\text{det(in)}}=10,000 for the remaining three geometries considered in Table 2 and Fig. 1 produces similarly compact |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle wave functions, spanned by 11,578–19,957 determinants, similarly small fractions of triples in the corresponding PP spaces, ranging from 1.5 % at R=1.5ReR=1.5R_{e} to 2.2 % at R=5ReR=5R_{e}, and even smaller errors in the CC(PP;QQ) energies relative to CCSDT.

It is clear from Table 2 and Fig. 1 that the convergence of the CIPSI-driven CC(PP;QQ) energies toward CCSDT with the wave function termination parameter Ndet(in)N_{\text{det(in)}}, with the number of determinants used to generate the final CIPSI state |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle [Ndet(out)N_{\text{det(out)}}], and with the fraction of triples in the PP space captured by the CIPSI procedure is very fast. The uncorrected CC(PP) energies converge to CCSDT too, but they do it at a considerably slower rate than their CC(PP;QQ) counterparts. For example, the CIPSI-driven CC(PP) calculations reduce the 9.485, 32.424, 45.638, and 49.816 millihartree errors relative to CCSDT obtained with CCSD to 1.419, 0.991, 0.922, and 0.764 millihartree, respectively, when Ndet(in)=50,000N_{\text{det(in)}}=50,000, which translates in the Ndet(out)N_{\text{det(out)}} values ranging between 65,172 and 92,682 and about 5–9 % of all triples included in the underlying PP spaces, but the errors characterizing the corresponding CC(PP;QQ) energies are already at the level of 20–30 microhartree at this stage, which is obviously a substantial improvement over the CC(PP) results. It is also worth noticing that the convergence of the CIPSI-driven CC(PP) and CC(PP;QQ) energies toward their CCSDT parents with Ndet(in)N_{\text{det(in)}} [or Ndet(out)N_{\text{det(out)}}] is considerably faster than the convergence of the corresponding variational and perturbatively corrected CIPSI energies toward the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} values. This is in line with the above observations that indicate that the CIPSI-driven CC(PP;QQ) calculations are capable of recovering the parent CCSDT energetics, even when electronic quasi-degeneracies and T3T_{3} clusters become significant, out of the unconverged CIPSI runs that rely on relatively small diagonalization spaces. We observed similar patterns when comparing the semi-stochastic, CIQMC- and CCMC-driven, CC(PP)/EOMCC(PP) and CC(PP;QQ) calculations with the underlying CIQMC/CCMC propagations. Deustua, Shen, and Piecuch (2017); Deustua et al. (2019); Yuwono et al. (2020); Deustua, Shen, and Piecuch (2021)

In analogy to the previously considered deterministic, active-orbital-based Shen and Piecuch (2012a, b); Bauman, Shen, and Piecuch (2017); Magoulas et al. (2018); Yuwono et al. (2019) and semi-stochastic, CIQMC/CCMC-based Deustua, Shen, and Piecuch (2017, 2021) CC(PP;QQ) studies, the convergence of the CIPSI-driven CC(PP;QQ) computations toward the parent CCSDT energetics remains equally rapid when we use basis sets larger than cc-pVDZ. This is illustrated in Table 3, where we show the results of the CIPSI-driven CC(PP;QQ) calculations for the F2\mathrm{F_{2}} molecule at R=2ReR=2R_{e} using the cc-pVTZ basis set.Dunning (1989) As pointed out in Ref. Deustua, Shen, and Piecuch, 2021, and in analogy to the cc-pVDZ basis, the T3T_{3} contribution characterizing the stretched F2\mathrm{F_{2}}/cc-pVTZ system in which the internuclear separation is set at twice the equilibrium bond length, estimated by forming the difference between the CCSDT and CCSD energies at 62.819-62.819 millihartree, is not only very large, but also larger, in absolute value, than the corresponding CCSDT dissociation energy, which is about 57 millihartree, when the CCSDT energy at ReR_{e} is subtracted from its 5Re5R_{e} counterpart. It is also highly nonperturbative at the same time, as demonstrated by the 26.354-26.354 millihartree error relative to CCSDT obtained with CCSD(T). Again, the CR-CC(2,3) triples correction to CCSD, equivalent to the Ndet(in)=1N_{\text{det(in)}}=1 CC(PP;QQ) calculation in Table 3, works a lot better than CCSD(T), but the 4.254 millihartree error relative to CCSDT remains. With as little as 5,118 Sz=0S_{z}=0 determinants of the Ag(D2h)A_{g}(D_{2h}) symmetry in the final diagonalization space obtained by the nearly effortless Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run, which captures 0.03 % of all triples, the difference between the CC(PP;QQ) and CCSDT energies decreases to 0.345 millihartree, and with the help of the Ndet(in)=50,000N_{\text{det(in)}}=50,000 CIPSI calculation, which is still relatively inexpensive, resulting in 82,001 Sz=0S_{z}=0 Ag(D2h)A_{g}(D_{2h})-symmetric determinants in the final diagonalization space and less than 1 % of the triples in the PP space, the error in the CC(PP;QQ) energy relative to its CCSDT parent reduces to less than 0.1 millihartree. Similarly to the cc-pVDZ basis, the convergence of the CIPSI-driven CC(PP;QQ) energies toward CCSDT with Ndet(in)N_{\text{det(in)}}, Ndet(out)N_{\text{det(out)}}, and the fraction of triples in the PP space captured by the CIPSI algorithm is not only fast, when the larger cc-pVTZ basis set is employed, but also much faster than in the case of the uncorrected CC(PP) calculations. Once again, as Ndet(in)N_{\text{det(in)}} increases, the rate of convergence of the CIPSI-driven CC(PP) and CC(PP;QQ) energies toward their CCSDT parent is higher than those characterizing the corresponding variational and perturbatively corrected CIPSI energies in their attempt to recover the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy.

Our final test, shown in Table 4, is the frequently examined Shen and Piecuch (2012b); Balková and Bartlett (1994); Lyakh, Lotrich, and Bartlett (2011); Whitman and Carpenter (1982); Carpenter (1983); Hess, Čarsky, and Schaad (1983); Voter and Goddard III (1986); Čarsky et al. (1988); Demel and Pittner (2006); Eckert-Maksić et al. (2006); Bhaskaran-Nair, Demel, and Pittner (2008); Karadakov (2008); Demel et al. (2008); Shen et al. (2008); Li and Paldus (2009); Deustua, Shen, and Piecuch (2017); Zhang, Li, and Evangelista (2019); Aroeira et al. (2021); Deustua, Shen, and Piecuch (2021) automerization of cyclobutadiene. In this case, one of the key challenges is an accurate determination of the activation barrier, which requires a well-balanced description of the nondegenerate, rectangle-shaped, closed-shell reactant (or the equivalent product) species, in which electron correlation effects are largely dynamical in nature, and the quasi-degenerate, square-shaped, transition state characterized by substantial nondynamical correlations associated with its strongly diradical character. Experimental estimates of the activation barrier for the automerization of cyclobutadiene, which range from 1.6 to 10 kcal/mol,Whitman and Carpenter (1982); Carpenter (1983) are not very precise, but the most accurate single- and multi-reference ab initio computations, compiled, for example, in Refs. Shen and Piecuch, 2012b; Lyakh, Lotrich, and Bartlett, 2011; Zhang, Li, and Evangelista, 2019, place it in the 6–10 kcal/mol range. This, in particular, applies to the CCSDT approach,Balková and Bartlett (1994); Shen and Piecuch (2012b) which is of the primary interest in the present study. Indeed, if we, for example, use the reactant and transition-state geometries obtained with the multi-reference average-quadratic CC (MR-AQCC) approachSzalay and Bartlett (1993, 1995) in Ref. Eckert-Maksić et al., 2006 and the cc-pVDZ basis set, the CCSDT value of the activation energy characterizing the automerization of cyclobutadiene becomes 7.627 kcal/mol,Shen and Piecuch (2012b) in good agreement with the best ab initio calculations carried out to date. By adopting the same geometries and basis set in this initial benchmark study of the CIPSI-driven CC(PP;QQ) methodology, we can examine if the CC(PP;QQ) calculations using the PP spaces constructed with the help of CIPSI runs are capable of converging this result. The main challenge here is that the T3T_{3} effects, estimated as the difference between the CCSDT and CCSD energies, are not only very large, but also hard to balance. When the cc-pVDZ basis set used in this study is employed, they are 26.827-26.827 millihartree for the reactant and 47.979-47.979 millihartree for the transition state. Furthermore, in the case of the transition state, the coupling of the lower-rank T1T_{1} and T2T_{2} clusters with their higher-rank T3T_{3} counterpart is so large that none of the noniterative triples corrections to CCSD provide a reasonable description of the activation barrier.Shen and Piecuch (2012b); Balková and Bartlett (1994); Lyakh, Lotrich, and Bartlett (2011) This, in particular, applies to the CR-CC(2,3) approach, equivalent to the Ndet(in)=1N_{\text{det(in)}}=1 CC(PP;QQ) calculation, which produces an activation barrier exceeding 16 kcal/mol, when the cc-pVDZ basis set is employed, instead of less than 8 kcal/mol obtained with CCSDT (cf. Table 4). The failure of CR-CC(2,3) to provide an accurate description of the activation energy is a consequence of its inability to accurately describe the transition state. Indeed, the difference between the CR-CC(2,3) and CCSDT energies at the transition-state geometry is 14.636 millihartree, when the cc-pVDZ basis set is employed, as opposed to only 0.848 millihartree obtained for the reactant. As discussed in detail in Refs. Shen and Piecuch, 2012b; Lyakh, Lotrich, and Bartlett, 2011, other triples corrections to CCSD, including the widely used CCSD(T) approach, face similar problems. We demonstrated in Refs. Shen and Piecuch, 2012b; Deustua, Shen, and Piecuch, 2017, 2021 that the deterministic CC(PP;QQ)-based CC(t;3) approach and the semi-stochastic CC(PP;QQ) calculations using CIQMC and CCMC are capable of accurately approximating the CCSDT energies of the reactant and transition-state species and the CCSDT activation barrier, so it is interesting to explore if the CIPSI-driven CC(PP;QQ) methodology can do the same.

As shown in Table 4, the CC(PP;QQ) calculations using CIPSI to identify the dominant triply excited determinants for the inclusion in the PP space are very efficient in converging the CCSDT energetics. With the final diagonalization spaces spanned by a little over 110,000 Sz=0S_{z}=0 determinants of the Ag(D2h)A_{g}(D_{2h}) symmetry (we used D2hD_{2h} group for both the D2hD_{2h}-symmetric reactant and the D4hD_{4h}-symmetric transition state in our calculations), generated in the relatively inexpensive CIPSI runs defined by Ndet(in)=100,000N_{\text{det(in)}}=100,000 that capture 0.1 % of all triples, the 0.848 millihartree, 14.636 millihartree, and 8.653 kcal/mol errors in the reactant, transition-state, and activation energies relative to CCSDT obtained with CR-CC(2,3) are reduced by factors of 2–4, to 0.382 millihartree, 3.507 millihartree, and 1.961 kcal/mol, respectively, when the CC(PP;QQ) approach is employed. When Ndet(in)N_{\text{det(in)}} is increased to 500,000, resulting in about 890,000–900,000 Sz=0S_{z}=0 determinants of the Ag(D2h)A_{g}(D_{2h}) symmetry in the final diagonalization spaces used by CIPSI and 1.0–1.2 % of the triples in the resulting PP spaces, the errors in the CC(PP;QQ) reactant, transition-state, and activation energies relative to CCSDT become 0.267 millihartree, 0.432 millihartree, and 0.104 kcal/mol. Clearly, these are great improvements compared to the initial Ndet(in)=1N_{\text{det(in)}}=1, i.e., CR-CC(2,3), values, especially if we realize that with the fractions of triples being so small, the post-CIPSI steps of the CC(PP;QQ) computations are not only a lot faster than the parent CCSDT runs, but also not much more expensive than the corresponding CR-CC(2,3) calculations, as elaborated on in Refs. Deustua, Shen, and Piecuch, 2017, 2021; Yuwono et al., 2020.

In analogy to the previously discussed case of bond breaking in F2\mathrm{F_{2}}, the convergence of the CIPSI-driven CC(PP;QQ) energies toward CCSDT for the reactant and transition-state species defining the automerization of cyclobutadiene with Ndet(in)N_{\text{det(in)}}, Ndet(out)N_{\text{det(out)}}, and the fractions of triples in the relevant PP spaces captured by the underlying CIPSI runs is not only very fast, but also significantly faster than that characterizing the uncorrected CC(PP) calculations. For each of the two species, the CC(PP) energies converge toward their CCSDT parent in a steady fashion, but, as shown in Table 4, their convergence is rather slow, emphasizing the importance of correcting the results of the CC(PP) calculations for the missing triple excitations not captured by the CIPSI runs using smaller diagonalization spaces. Similarly to the previously examined active-orbital-based Shen and Piecuch (2012a, b); Bauman, Shen, and Piecuch (2017); Magoulas et al. (2018); Yuwono et al. (2019) and CIQMC/CCMC-based Deustua, Shen, and Piecuch (2017, 2021) CC(PP;QQ) approaches, our moment correction δ(P;Q)\delta(P;Q), defined by Eq. (9), is very effective in this regard. Another interesting observation, which can be made based on the results presented in Table 4, is that while the CC(PP) energies for the individual reactant and transition-state species converge toward their CCSDT parent values in a steady fashion, the corresponding activation barrier values behave in a less systematic manner, oscillating between about 1-1 and 1 kcal/mol when Ndet(in)=500,00015,000,000N_{\text{det(in)}}=500,000\mbox{--}15,000,000. One might try to eliminate this behavior, which is a consequence of a different character of the many-electron correlation effects in the reactant and transition-state species, by merging the PP spaces used to perform the CC(PP) calculations for the two structures, but, as shown in Table 4, the CC(PP;QQ) correction δ(P;Q)\delta(P;Q), which is highly effective in capturing the missing T3T_{3} correlations, takes care of this problem too. As Ndet(in)N_{\text{det(in)}}, Ndet(out)N_{\text{det(out)}}, and the fractions of triples in the PP spaces used in the CC(PP) calculations for the reactant and transition state increase, the CC(PP;QQ) values of the activation barrier converge toward its CCSDT parent rapidly and in a smooth manner, eliminating, at least to a large extent, the need to equalize the PP spaces used in the underlying CC(PP) steps. As in the case of bond breaking in the fluorine molecule, the convergence of the CIPSI-driven CC(PP) and CC(PP;QQ) energies toward their CCSDT parents with Ndet(in)N_{\text{det(in)}}/Ndet(out)N_{\text{det(out)}} is considerably faster than the convergence of the variational and perturbatively corrected CIPSI energies toward the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} values, but we must keep in mind that the calculated CCSDT and extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies, while representing the respective parent limits for the CC(PP;QQ) and CIPSI calculations, are fundamentally different quantities, especially when higher–than–triply excited cluster components, which are not considered in this work, become significant. As one might anticipate, the Ndet(in)N_{\text{det(in)}} values needed to accurately represent the CCSDT energies of the reactant and transition-state species of cyclobutadiene by the CIPSI-driven CC(PP;QQ) approach are considerably larger than those used in the analogous CC(PP;QQ) calculations for the smaller F2{\rm F}_{2} system, but they are orders of magnitude smaller than the values of Ndet(in)N_{\text{det(in)}} required to obtain the similarly well converged Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energetics in the underlying CIPSI runs.

IV Conclusions

Inspired by our recent studiesDeustua, Shen, and Piecuch (2017); Yuwono et al. (2020); Deustua, Shen, and Piecuch (2021) aimed at determining accurate electronic energies equivalent to the results of high-level CC calculations, in which we combined the deterministic CC(PP;QQ) framework Shen and Piecuch (2012a, b, c); Bauman, Shen, and Piecuch (2017); Magoulas et al. (2018) with the stochastic CIQMCBooth, Thom, and Alavi (2009); Cleland, Booth, and Alavi (2010); Dobrautz, Smart, and Alavi (2019); Ghanem, Lozovoi, and Alavi (2019); Ghanem, Guther, and Alavi (2020) and CCMCThom (2010); Franklin et al. (2016); Spencer and Thom (2016); Scott and Thom (2017) propagations, and the successes of modern formulationsSchriber and Evangelista (2016, 2017); Tubman et al. (2016, 2020); Liu and Hoffmann (2016); Zhang, Liu, and Hoffmann (2020); Holmes, Tubman, and Umrigar (2016); Sharma et al. (2017); Li et al. (2018); Garniron et al. (2017, 2019) of selected CI techniques,Whitten and Hackmeyer (1969); Bender and Davidson (1969); Huron, Malrieu, and Rancurel (1973); Buenker and Peyerimhoff (1974) we have proposed a new form of the CC(PP;QQ) approach, in which we identify the dominant higher–than–doubly excited determinants for the inclusion in the underlying PP spaces with the help of the selected CI algorithm abbreviated as CIPSI.Huron, Malrieu, and Rancurel (1973); Garniron et al. (2017, 2019) To illustrate potential benefits offered by the proposed merger of the CC(PP;QQ) and CIPSI methodologies, we have implemented the CIPSI-driven CC(PP;QQ) method designed to converge CCSDT energetics. The advantages of the CIPSI-driven CC(PP;QQ) methodology have been illustrated by a few numerical examples, including bond breaking in F2{\rm F}_{2} and the automerization of cyclobutadiene, which are accurately described by CCSDT.

The reported benchmark calculations demonstrate that the convergence of the CIPSI-driven CC(PP;QQ) energies toward CCSDT with the wave function termination parameter Ndet(in)N_{\text{det(in)}} adopted by CIPSI, with the number of determinants used to generate the final CIPSI state [Ndet(out)N_{\text{det(out)}}], and with the fractions of triples in the PP space captured by the CIPSI procedure is very fast. As a result, one can obtain CCSDT-level energetics, even when electronic quasi-degeneracies and T3T_{3} clusters become substantial, based on the information extracted from the relatively inexpensive CIPSI runs. This can be attributed to two key factors. The first one is a tempered wave function growth through iterative Hamiltonian diagonalizations in the modern implementation of CIPSI available in Quantum Package 2.0,Garniron et al. (2017, 2019) which we utilized in this work, resulting in an economical selection of the dominant triply excited determinants (in general, the dominant higher–than–doubly excited determinants) for the inclusion in the PP spaces driving the CC(PP;QQ) computations. The second one is the efficiency of the moment corrections δ(P;Q)\delta(P;Q) defining the CC(PP;QQ) formalism, which provide an accurate and robust description of the missing T3T_{3} contributions that cannot be captured by the underlying CC(PP) calculations using small fractions of triples identified by the CIPSI runs employing smaller diagonalization spaces. We have also shown that the uncorrected CC(PP) energies converge with Ndet(in)N_{\text{det(in)}}, Ndet(out)N_{\text{det(out)}}, and the fractions of triples in the PP spaces constructed with the help of CIPSI to their CCSDT parent values too, but they do it at a much slower rate, so that we do not recommend the uncorrected CC(PP) approach in calculations aimed at recovering high-level CC energetics.

Clearly, the present study is only our initial exploration of the CIPSI-driven (or, in general, selected-CI-driven) CC(PP;QQ) methodology, which needs more work. In addition to code optimization and more numerical tests, especially including larger molecules and basis sets, we would like to extend the proposed CIPSI-driven CC(PP;QQ) framework to higher levels of the CC theory, beyond CCSDT, as we already did in the context of the active-orbital-basedBauman, Shen, and Piecuch (2017); Magoulas et al. (2018) and CIQMC-basedDeustua, Shen, and Piecuch (2021) CC(PP;QQ) considerations, and examine if other selected CI methods, such as heat-bath CIHolmes, Tubman, and Umrigar (2016); Sharma et al. (2017); Li et al. (2018) or adaptive-CI,Schriber and Evangelista (2016, 2017) to mention a couple of examples, are as useful in the context of CC(PP;QQ) considerations as the CIPSI approach adopted in this work. Following our recent semi-stochastic EOMCC(PP) and CC(PP;QQ) work, Deustua et al. (2019); Yuwono et al. (2020) we are also planning to extend the CIPSI-driven CC(PP;QQ) methodology to excited electronic states. One of the main advantages of CIPSI and other selected-CI methods, which are based on Hamiltonian diagonalizations, is that they can be easily applied to excited states (see, e.g., Refs. Schriber and Evangelista, 2017; Chien et al., 2018; Loos et al., 2018, 2019, 2020a; Loos, Scemama, and Jacquemin, 2020; Loos et al., 2020b). This would allow us to construct the state-specific PP spaces, adjusted to the individual electronic states of interest, which is more difficult to accomplish within the CIQMC framework (see Refs. Deustua et al., 2019; Yuwono et al., 2020 for additional comments). Encouraged by our recent work on the semi-stochastic CC(PP;QQ) models using truncated CIQMC rather than FCIQMC propagations to determine the underlying PP spaces,Deustua, Shen, and Piecuch (2021) we would like to examine if one can replace the unconstrained CIPSI algorithm used in this study, which explores the entire many-electron Hilbert space in the iterative wave function build-up, by its less expensive truncated analogs compatible with the determinantal spaces needed in the CC calculations of interest (e.g., the CISDT or CISDTQ analogs of CIPSI if one is interested in converging the CCSDT or CCSDTQ energetics through CIPSI-driven CC(PP;QQ) computations). This may help us to achieve the desired high accuracy levels in the CIPSI-driven CC(PP;QQ) calculations with the relatively short CI wave function expansions, even when larger systems are examined, since the diagonalization spaces generated by the truncated CIPSI models will be significantly smaller than those produced when CIPSI is allowed to explore the entire many-electron Hilbert space. Last, but not least, inspired by our recent work on the CIPSI-driven externally corrected CC models,Magoulas et al. (2021) we would like to investigate the effect of the CIPSI input parameter ff that controls the wave function growth in successive Hamiltonian diagonalizations, which was set in this study at the default value of 2, on the rate of convergence of the CIPSI-driven CC(PP;QQ) energies toward their high-level CC parents, such as those obtained with CCSDT.

Acknowledgements.
This work has been supported by the Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Science, U.S. Department of Energy (Grant No. DE-FG02-01ER15228 to P.P), and Phase I and II Software Fellowships awarded to J.E.D. by the Molecular Sciences Software Institute funded by the National Science Foundation grant ACI-1547580. We thank Drs. Pierre-François Loos and Anthony Scemama for useful discussions about Quantum Package 2.0 employed in our CIPSI computations.

Data Availability

The data that support the findings of this study are available within the article.

References

References

  • Hubbard (1957) J. Hubbard, Proc. R. Soc. Lond., Ser. A 240, 539 (1957).
  • Hugenholtz (1957) N. M. Hugenholtz, Physica 23, 481 (1957).
  • Coester (1958) F. Coester, Nucl. Phys. 7, 421 (1958).
  • Coester and Kümmel (1960) F. Coester and H. Kümmel, Nucl. Phys. 17, 477 (1960).
  • Čížek (1966) J. Čížek, J. Chem. Phys. 45, 4256 (1966).
  • Čížek (1969) J. Čížek, Adv. Chem. Phys. 14, 35 (1969).
  • Paldus, Čížek, and Shavitt (1972) J. Paldus, J. Čížek,  and I. Shavitt, Phys. Rev. A 5, 50 (1972).
  • Paldus and Li (1999) J. Paldus and X. Li, Adv. Chem. Phys. 110, 1 (1999).
  • Bartlett and Musiał (2007) R. J. Bartlett and M. Musiał, Rev. Mod. Phys. 79, 291 (2007).
  • Purvis and Bartlett (1982) G. D. Purvis, III and R. J. Bartlett, J. Chem. Phys. 76, 1910 (1982).
  • Cullen and Zerner (1982) J. M. Cullen and M. C. Zerner, J. Chem. Phys. 77, 4088 (1982).
  • Noga and Bartlett (1987) J. Noga and R. J. Bartlett, J. Chem. Phys. 86, 7041 (1987), 89, 3401 (1988) [Erratum].
  • Scuseria and Schaefer (1988) G. E. Scuseria and H. F. Schaefer, III, Chem. Phys. Lett. 152, 382 (1988).
  • Watts and Bartlett (1990) J. D. Watts and R. J. Bartlett, J. Chem. Phys. 93, 6104 (1990).
  • Oliphant and Adamowicz (1991) N. Oliphant and L. Adamowicz, J. Chem. Phys. 95, 6645 (1991).
  • Kucharski and Bartlett (1992) S. A. Kucharski and R. J. Bartlett, J. Chem. Phys. 97, 4282 (1992).
  • Piecuch and Adamowicz (1994) P. Piecuch and L. Adamowicz, J. Chem. Phys. 100, 5792 (1994).
  • Emrich (1981) K. Emrich, Nucl. Phys. A 351, 379 (1981).
  • Geertsen, Rittby, and Bartlett (1989) J. Geertsen, M. Rittby,  and R. J. Bartlett, Chem. Phys. Lett. 164, 57 (1989).
  • Stanton and Bartlett (1993) J. F. Stanton and R. J. Bartlett, J. Chem. Phys. 98, 7029 (1993).
  • Kowalski and Piecuch (2001a) K. Kowalski and P. Piecuch, J. Chem. Phys. 115, 643 (2001a).
  • Kowalski and Piecuch (2001b) K. Kowalski and P. Piecuch, Chem. Phys. Lett. 347, 237 (2001b).
  • Kucharski et al. (2001) S. A. Kucharski, M. Włoch, M. Musiał,  and R. J. Bartlett, J. Chem. Phys. 115, 8263 (2001).
  • Kállay and Gauss (2004) M. Kállay and J. Gauss, J. Chem. Phys. 121, 9257 (2004).
  • Hirata (2004) S. Hirata, J. Chem. Phys. 121, 51 (2004).
  • Monkhorst (1977) H. Monkhorst, Int. J. Quantum Chem. Symp. 11, 421 (1977).
  • Dalgaard and Monkhorst (1983) E. Dalgaard and H. Monkhorst, Phys. Rev. A 28, 1217 (1983).
  • Mukherjee and Mukherjee (1979) D. Mukherjee and P. K. Mukherjee, Chem. Phys. 39, 325 (1979).
  • Sekino and Bartlett (1984) H. Sekino and R. J. Bartlett, Int. J. Quantum Chem. Symp. 18, 255 (1984).
  • Takahashi and Paldus (1986) M. Takahashi and J. Paldus, J. Chem. Phys. 85, 1486 (1986).
  • Koch and Jørgensen (1990) H. Koch and P. Jørgensen, J. Chem. Phys. 93, 3333 (1990).
  • Koch et al. (1990) H. Koch, H. J. A. Jensen, P. Jørgensen,  and T. Helgaker, J. Chem. Phys. 93, 3345 (1990).
  • Kondo, Piecuch, and Paldus (1995a) A. E. Kondo, P. Piecuch,  and J. Paldus, J. Chem. Phys. 102, 6511 (1995a).
  • Kondo, Piecuch, and Paldus (1995b) A. E. Kondo, P. Piecuch,  and J. Paldus, J. Chem. Phys. 104, 8566 (1995b).
  • Raghavachari et al. (1989) K. Raghavachari, G. W. Trucks, J. A. Pople,  and M. Head-Gordon, Chem. Phys. Lett. 157, 479 (1989).
  • Piecuch et al. (2002) P. Piecuch, K. Kowalski, I. S. O. Pimienta,  and M. J. McGuire, Int. Rev. Phys. Chem. 21, 527 (2002).
  • Piecuch et al. (2004) P. Piecuch, K. Kowalski, I. S. O. Pimienta, P.-D. Fan, M. Lodriguito, M. J. McGuire, S. A. Kucharski, T. Kuś,  and M. Musiał, Theor. Chem. Acc. 112, 349 (2004).
  • Piecuch (2010) P. Piecuch, Mol. Phys. 108, 2987 (2010).
  • Deustua, Shen, and Piecuch (2017) J. E. Deustua, J. Shen,  and P. Piecuch, Phys. Rev. Lett. 119, 223003 (2017).
  • Deustua, Shen, and Piecuch (2021) J. E. Deustua, J. Shen,  and P. Piecuch, J. Chem. Phys. 154, 124103 (2021).
  • Yuwono et al. (2020) S. H. Yuwono, A. Chakraborty, J. E. Deustua, J. Shen,  and P. Piecuch, Mol. Phys. 118, e1817592 (2020).
  • Shen and Piecuch (2012a) J. Shen and P. Piecuch, Chem. Phys. 401, 180 (2012a).
  • Shen and Piecuch (2012b) J. Shen and P. Piecuch, J. Chem. Phys. 136, 144104 (2012b).
  • Shen and Piecuch (2012c) J. Shen and P. Piecuch, J. Chem. Theory Comput. 8, 4968 (2012c).
  • Bauman, Shen, and Piecuch (2017) N. P. Bauman, J. Shen,  and P. Piecuch, Mol. Phys. 115, 2860 (2017).
  • Magoulas et al. (2018) I. Magoulas, N. P. Bauman, J. Shen,  and P. Piecuch, J. Phys. Chem. A 122, 1350 (2018).
  • Booth, Thom, and Alavi (2009) G. H. Booth, A. J. W. Thom,  and A. Alavi, J. Chem. Phys. 131, 054106 (2009).
  • Cleland, Booth, and Alavi (2010) D. Cleland, G. H. Booth,  and A. Alavi, J. Chem. Phys. 132, 041103 (2010).
  • Dobrautz, Smart, and Alavi (2019) W. Dobrautz, S. D. Smart,  and A. Alavi, J. Chem. Phys. 151, 094104 (2019).
  • Ghanem, Lozovoi, and Alavi (2019) K. Ghanem, A. Y. Lozovoi,  and A. Alavi, J. Chem. Phys. 151, 224108 (2019).
  • Ghanem, Guther, and Alavi (2020) K. Ghanem, K. Guther,  and A. Alavi, J. Chem. Phys. 153, 224115 (2020).
  • Thom (2010) A. J. W. Thom, Phys. Rev. Lett. 105, 263004 (2010).
  • Franklin et al. (2016) R. S. T. Franklin, J. S. Spencer, A. Zoccante,  and A. J. W. Thom, J. Chem. Phys. 144, 044111 (2016).
  • Spencer and Thom (2016) J. S. Spencer and A. J. W. Thom, J. Chem. Phys. 144, 084108 (2016).
  • Scott and Thom (2017) C. J. C. Scott and A. J. W. Thom, J. Chem. Phys. 147, 124105 (2017).
  • Deustua et al. (2019) J. E. Deustua, S. H. Yuwono, J. Shen,  and P. Piecuch, J. Chem. Phys. 150, 111101 (2019).
  • Whitten and Hackmeyer (1969) J. Whitten and M. Hackmeyer, J. Chem. Phys. 51, 5584 (1969).
  • Bender and Davidson (1969) C. Bender and E. Davidson, Phys. Rev. 183, 23 (1969).
  • Huron, Malrieu, and Rancurel (1973) B. Huron, J. P. Malrieu,  and P. Rancurel, J. Chem. Phys. 58, 5745 (1973).
  • Buenker and Peyerimhoff (1974) R. Buenker and S. Peyerimhoff, Theor. Chim. Acta. 35, 33 (1974).
  • Schriber and Evangelista (2016) J. Schriber and F. Evangelista, J. Chem. Phys. 144, 161106 (2016).
  • Schriber and Evangelista (2017) J. Schriber and F. Evangelista, J. Chem. Theory Comput. 13, 5354 (2017).
  • Tubman et al. (2016) N. M. Tubman, J. Lee, T. Takeshita, M. Head-Gordon,  and K. Whaley, J. Chem. Phys. 145, 044112 (2016).
  • Tubman et al. (2020) N. M. Tubman, C. Freeman, D. Levine, D. Hait, M. Head-Gordon,  and K. Whaley, J. Chem. Theory Comput. 16, 2139 (2020).
  • Liu and Hoffmann (2016) W. Liu and M. Hoffmann, J. Chem. Theory Comput. 12, 1169 (2016), 12, 3000 (2016) [Erratum].
  • Zhang, Liu, and Hoffmann (2020) N. Zhang, W. Liu,  and M. Hoffmann, J. Chem. Theory Comput. 16, 2296 (2020).
  • Holmes, Tubman, and Umrigar (2016) A. A. Holmes, N. M. Tubman,  and C. J. Umrigar, J. Chem. Theory Comput. 12, 3674 (2016).
  • Sharma et al. (2017) S. Sharma, A. A. Holmes, G. Jeanmairet, A. Alavi,  and C. J. Umrigar, J. Chem. Theory Comput. 13, 1595 (2017).
  • Li et al. (2018) J. Li, M. Otten, A. A. Holmes, S. Sharma,  and C. J. Umrigar, J. Chem. Phys. 149, 214110 (2018).
  • Garniron et al. (2017) Y. Garniron, A. Scemama, P.-F. Loos,  and M. Caffarel, J. Chem. Phys. 147, 034101 (2017).
  • Garniron et al. (2019) 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).
  • Loos, Damour, and Scemama (2020) P.-F. Loos, Y. Damour,  and A. Scemama, J. Chem. Phys. 153, 176101 (2020).
  • Eriksen et al. (2020) J. J. Eriksen, T. A. Anderson, J. E. Deustua, K. Ghanem, D. Hait, M. R. Hoffmann, 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).
  • Jankowski, Paldus, and Piecuch (1991) K. Jankowski, J. Paldus,  and P. Piecuch, Theor. Chim. Acta 80, 223 (1991).
  • Piecuch and Kowalski (2000) P. Piecuch and K. Kowalski, in Computational Chemistry: Reviews of Current Trends, Vol. 5, edited by J. Leszczyński (World Scientific, Singapore, 2000) pp. 1–104.
  • Kowalski and Piecuch (2000) K. Kowalski and P. Piecuch, J. Chem. Phys. 113, 18 (2000).
  • Piecuch and Włoch (2005) P. Piecuch and M. Włoch, J. Chem. Phys. 123, 224105 (2005).
  • Piecuch et al. (2006) P. Piecuch, M. Włoch, J. R. Gour,  and A. Kinal, Chem. Phys. Lett. 418, 467 (2006).
  • Włoch et al. (2006) M. Włoch, M. D. Lodriguito, P. Piecuch,  and J. R. Gour, Mol. Phys. 104, 2149 (2006), 104, 2991 (2006) [Erratum].
  • Włoch, Gour, and Piecuch (2007) M. Włoch, J. R. Gour,  and P. Piecuch, J. Phys. Chem. A 111, 11359 (2007).
  • Schmidt et al. (1993) 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).
  • Barca et al. (2020) 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. 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. Włoch, P. Xu, F. Zahariev,  and M. S. Gordon, J. Chem. Phys. 152, 154102 (2020).
  • Dunning (1989) T. H. Dunning, Jr., J. Chem. Phys. 90, 1007 (1989).
  • Kowalski and Piecuch (2001c) K. Kowalski and P. Piecuch, Chem. Phys. Lett. 344, 165 (2001c).
  • Musiał and Bartlett (2005) M. Musiał and R. J. Bartlett, J. Chem. Phys. 122, 224102 (2005).
  • Yuwono et al. (2019) S. H. Yuwono, I. Magoulas, J. Shen,  and P. Piecuch, Mol. Phys. 117, 1486 (2019).
  • Balková and Bartlett (1994) A. Balková and R. J. Bartlett, J. Chem. Phys. 101, 8972 (1994).
  • Lyakh, Lotrich, and Bartlett (2011) D. I. Lyakh, V. F. Lotrich,  and R. J. Bartlett, Chem. Phys. Lett. 501, 166 (2011).
  • Whitman and Carpenter (1982) D. W. Whitman and B. K. Carpenter, J. Am. Chem. Soc. 104, 6473 (1982).
  • Carpenter (1983) B. K. Carpenter, J. Am. Chem. Soc. 105, 1700 (1983).
  • Hess, Čarsky, and Schaad (1983) B. A. Hess, P. Čarsky,  and L. J. Schaad, J. Am. Chem. Soc. 105, 695 (1983).
  • Voter and Goddard III (1986) A. F. Voter and W. A. Goddard III, J. Am. Chem. Soc. 108, 2830 (1986).
  • Čarsky et al. (1988) P. Čarsky, R. J. Bartlett, G. Fitzgerald, J. Nova,  and V. Špirko, J. Chem. Phys. 89, 3008 (1988).
  • Demel and Pittner (2006) O. Demel and J. Pittner, J. Chem. Phys. 124, 144112 (2006).
  • Eckert-Maksić et al. (2006) M. Eckert-Maksić, M. Vazdar, M. Barbatti, H. Lischka,  and Z. B. Maksić, J. Chem. Phys. 125, 064310 (2006).
  • Bhaskaran-Nair, Demel, and Pittner (2008) K. Bhaskaran-Nair, O. Demel,  and J. Pittner, J. Chem. Phys. 129, 184105 (2008).
  • Karadakov (2008) P. B. Karadakov, J. Phys. Chem. A 112, 7303 (2008).
  • Demel et al. (2008) O. Demel, K. R. Shamasundar, L. Kong,  and M. Nooijen, J. Phys. Chem. A 112, 11895 (2008).
  • Shen et al. (2008) J. Shen, T. Fang, S. Li,  and Y. Jiang, J. Phys. Chem. A 112, 12518 (2008).
  • Li and Paldus (2009) X. Li and J. Paldus, J. Chem. Phys. 131, 114103 (2009).
  • Zhang, Li, and Evangelista (2019) T. Zhang, C. Li,  and F. A. Evangelista, J. Chem. Theory Comput. 15, 4399 (2019).
  • Aroeira et al. (2021) G. J. R. Aroeira, M. M. Davis, J. M. Turney,  and H. F. Schaefer, J. Chem. Theory Comput. 17, 182 (2021).
  • Szalay and Bartlett (1993) P. G. Szalay and R. J. Bartlett, Chem. Phys. Lett. 214, 481 (1993).
  • Szalay and Bartlett (1995) P. G. Szalay and R. J. Bartlett, J. Chem. Phys. 103, 3600 (1995).
  • Chien et al. (2018) A. D. Chien, A. A. Holmes, M. Otten, C. J. Umrigar, S. Sharma,  and P. M. Zimmerman, J. Phys. Chem. A 122, 2714 (2018).
  • Loos et al. (2018) P.-F. Loos, A. Scemama, A. Blondel, Y. Garniron, M. Caffarel,  and D. Jacquemin, J. Chem. Theory Comput. 14, 4360 (2018).
  • Loos et al. (2019) P.-F. Loos, M. Boggio-Pasqua, A. Scemama, M. Caffarel,  and D. Jacquemin, J. Chem. Theory Comput. 15, 1939 (2019).
  • Loos et al. (2020a) P.-F. Loos, F. Lipparini, M. Boggio-Pasqua, A. Scemama,  and D. Jacquemin, J. Chem. Theory Comput. 16, 1711 (2020a).
  • Loos, Scemama, and Jacquemin (2020) P.-F. Loos, A. Scemama,  and D. Jacquemin, J. Phys. Chem. Lett. 11, 2374 (2020).
  • Loos et al. (2020b) P.-F. Loos, A. Scemama, M. Boggio-Pasqua,  and D. Jacquemin, J. Chem. Theory Comput. 16, 3720 (2020b).
  • Magoulas et al. (2021) I. Magoulas, K. Gururangan, P. Piecuch, J. E. Deustua,  and J. Shen, J. Chem. Theory Comput. 17, 4006 (2021).
Table 1: Numerical demonstration of the size extensivity of the CIPSI-driven CC(PP) and CC(PP;QQ) approaches, alongside the analogous CCSD, CR-CC(2,3), and CCSDT calculations, using the noninteracting F2+Ne{\rm F}_{2}+{\rm Ne} system, as described by the cc-pVDZ basis set, in which the F–F bond length RR was fixed at twice its equilibrium value. In all post-RHF calculations, the core orbitals correlating with the 1s1s shells of the fluorine and neon atoms were frozen and the Cartesian components of dd orbitals were employed throughout. All energy values are total electronic energies in hartree.
Method E(F2+Ne)E(\mathrm{F_{2}}+\mathrm{Ne})111 The noninteracting F2+Ne{\rm F}_{2}+{\rm Ne} system was obtained by placing the Ne atom along the axis of the F–F bond at 1,000 bohr from the center of mass of the stretched fluorine molecule in which the internuclear separation RR was set at 2Re2R_{e}, where Re=2.66816R_{e}=2.66816 bohr is the equilibrium geometry of F2{\rm F}_{2}. E(F2)E(\mathrm{F_{2}})222 The stretched F2{\rm F}_{2} molecule in which the F–F bond length RR was set at 2Re2R_{e}. E(Ne)E(\mathrm{Ne}) E(F2+Ne)[E(F2)+E(Ne)]E(\mathrm{F_{2}}+\mathrm{Ne})-[E(\mathrm{F_{2}})+E(\mathrm{Ne})]
CCSD333 Equivalent to the CC(PP) calculations with Ndet(in)=1N_{\text{det(in)}}=1. 327.692849962-327.692849962 199.012562571-199.012562571 128.680287394-128.680287394 0.000000003
CR-CC(2,3)444 Equivalent to the CC(PP;QQ) calculations with Ndet(in)=1N_{\text{det(in)}}=1. 327.737915219-327.737915219 199.056339293-199.056339293 128.681575920-128.681575920 0.000000006-0.000000006
CC(PP)/Ndet(in)=5,000N_{\text{det(in)}}=5,000 327.736961010-327.736961010555 The PP space used in the CC(PP) calculation for the F2+Ne{\rm F}_{2}+{\rm Ne} system consisted of all singles and doubles and a subset of triples contained in the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state of the underlying Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run. The QQ space needed to compute the CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. The Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for F2+Ne{\rm F}_{2}+{\rm Ne}, which was initiated from the RHF reference determinant, used f=2f=2 and η=106\eta=10^{-6} hartree. 199.056233029-199.056233029666 The PP space used in the CC(PP) calculation for F2\mathrm{F_{2}} was obtained by removing the triply excited determinants involving Ne orbitals from the list of triples provided by the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for the F2+Ne\mathrm{F_{2}}+\mathrm{Ne} system. The QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples missing in the PP space. 128.680728060-128.680728060777 The PP space used in the CC(PP) calculation for Ne was obtained by removing the triply excited determinants involving F2\mathrm{F_{2}} orbitals from the list of triples provided by the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for the F2+Ne\mathrm{F_{2}}+\mathrm{Ne} system. The QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples missing in the PP space. 0.000000080
CC(PP;QQ)/Ndet(in)=5,000N_{\text{det(in)}}=5,000 327.739651938-327.739651938555 The PP space used in the CC(PP) calculation for the F2+Ne{\rm F}_{2}+{\rm Ne} system consisted of all singles and doubles and a subset of triples contained in the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state of the underlying Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run. The QQ space needed to compute the CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. The Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for F2+Ne{\rm F}_{2}+{\rm Ne}, which was initiated from the RHF reference determinant, used f=2f=2 and η=106\eta=10^{-6} hartree. 199.058190353-199.058190353666 The PP space used in the CC(PP) calculation for F2\mathrm{F_{2}} was obtained by removing the triply excited determinants involving Ne orbitals from the list of triples provided by the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for the F2+Ne\mathrm{F_{2}}+\mathrm{Ne} system. The QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples missing in the PP space. 128.681461627-128.681461627777 The PP space used in the CC(PP) calculation for Ne was obtained by removing the triply excited determinants involving F2\mathrm{F_{2}} orbitals from the list of triples provided by the Ndet(in)=5,000N_{\text{det(in)}}=5,000 CIPSI run for the F2+Ne\mathrm{F_{2}}+\mathrm{Ne} system. The QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples missing in the PP space. 0.000000040
CCSDT 327.739605196-327.739605196 199.058201287-199.058201287 128.681403900-128.681403900 0.000000009-0.000000009
Table 2: Convergence of the CC(PP) and CC(PP;QQ) energies toward CCSDT, alongside the variational and perturbatively corrected CIPSI energies, for the F2\mathrm{F_{2}}/cc-pVDZ molecule in which the F–F bond length RR was set at ReR_{e}, 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e}, where Re=2.66816R_{e}=2.66816 bohr is the equilibrium geometry. For each value of the wave function termination parameter Ndet(in)N_{\text{det(in)}}, the PP space used in the CC(PP) calculations consisted of all singles and doubles and a subset of triples contained in the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state of the underlying CIPSI run, whereas the QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. In all post-RHF calculations, the two lowest-lying core orbitals were frozen and the Cartesian components of dd orbitals were employed throughout. Each CIPSI run was initiated from the RHF reference determinant and the MBPT-based stopping parameter η\eta was set at 10610^{-6} hartree. The input parameter ff controlling the CIPSI wave function growth was set at the default value of 2.
R/ReR/R_{e} Ndet(in)N_{\text{det(in)}} / Ndet(out)N_{\text{det(out)}} % of triples EvarE_{\text{var}}111 For each internuclear separation RR, the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=5,000,000N_{\text{det(in)}}=5,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. These extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies at R=ReR=R_{e}, 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e} are 199.104422(6)-199.104422(6), 199.069043(1)-199.069043(1), 199.060152(8)-199.060152(8), and 199.059647(11)-199.059647(11) hartree, respectively, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained in the above extrapolation procedure. Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}111 For each internuclear separation RR, the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=5,000,000N_{\text{det(in)}}=5,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. These extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies at R=ReR=R_{e}, 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e} are 199.104422(6)-199.104422(6), 199.069043(1)-199.069043(1), 199.060152(8)-199.060152(8), and 199.059647(11)-199.059647(11) hartree, respectively, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained in the above extrapolation procedure. Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)}111 For each internuclear separation RR, the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=5,000,000N_{\text{det(in)}}=5,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. These extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies at R=ReR=R_{e}, 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e} are 199.104422(6)-199.104422(6), 199.069043(1)-199.069043(1), 199.060152(8)-199.060152(8), and 199.059647(11)-199.059647(11) hartree, respectively, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained in the above extrapolation procedure. CC(P)(P)222 The CC(PP) and CC(PP;QQ) energies are reported as errors relative to CCSDT, in millihartree. The total CCSDT energies at R=ReR=R_{e}, 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e} are 199.102796-199.102796, 199.065882-199.065882, 199.058201-199.058201, and 199.058586-199.058586 hartree, respectively. CC(P;Q)(P;Q)222 The CC(PP) and CC(PP;QQ) energies are reported as errors relative to CCSDT, in millihartree. The total CCSDT energies at R=ReR=R_{e}, 1.5Re1.5R_{e}, 2Re2R_{e}, and 5Re5R_{e} are 199.102796-199.102796, 199.065882-199.065882, 199.058201-199.058201, and 199.058586-199.058586 hartree, respectively.
1.0 1 / 1 0 418.057333 Equivalent to RHF. 94.150-94.150444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 12.651-12.651 9.485555 Equivalent to CCSD. 0.240-0.240666 Equivalent to CR-CC(2,3).
10 / 17 0 330.754 32.707-32.707 -4.877 9.485 0.240-0.240
100 / 154 0 232.186 7.963-7.963 2.338 9.485 0.240-0.240
1,000 / 1,266 0 65.926 1.480 2.079 9.485 0.240-0.240
5,000 / 5,072 0.4 23.596 0.133(0)-0.133(0) 0.069(0)-0.069(0) 4.031 0.129-0.129
10,000 / 10,150 1.2 19.197 0.045(2) 0.084(2) 3.010 0.067-0.067
50,000 / 81,288 7.9 11.282 0.133(1) 0.145(1) 1.419 0.031-0.031
100,000 / 162,430 14.5 9.222 0.138(1) 0.146(1) 0.983 0.020-0.020
500,000 / 649,849 34.3 5.630 0.092(1) 0.095(1) 0.519 0.009-0.009
1,000,000 / 1,300,305 42.2 4.816 0.072(0) 0.074(0) 0.464 0.008-0.008
5,000,000 / 5,187,150 85.1 1.161 0.015(2) 0.016(2) 0.023 0.001-0.001
1.5 1 / 1 0 541.109333 Equivalent to RHF. 130.718-130.718444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 137.819 32.424555 Equivalent to CCSD. 1.735666 Equivalent to CR-CC(2,3).
10 / 18 0 319.363 11.279-11.279 10.126 32.424 1.735
100 / 177 0 235.819 2.527 12.175 32.424 1.735
1,000 / 1,442 0.1 77.306 5.218 5.948 16.835 0.202
5,000 / 5,773 0.7 21.091 0.811(2) 0.856(2) 2.490 0.009
10,000 / 11,578 1.5 17.333 0.811(2) 0.839(2) 1.892 0.028
50,000 / 92,682 8.8 10.879 0.762(1) 0.771(1) 0.991 0.033
100,000 / 185,350 13.9 9.243 0.632(1) 0.639(1) 0.727 0.023
500,000 / 742,754 30.8 5.586 0.391(1) 0.393(1) 0.390 0.005
1,000,000 / 1,484,218 37.1 4.795 0.330(0) 0.332(0) 0.362 0.004
5,000,000 / 5,907,228 74.3 1.165 0.079(2) 0.079(2) 0.028 0.000-0.000
2.0 1 / 1 0 640.056333 Equivalent to RHF. 159.482-159.482444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 289.080 45.638555 Equivalent to CCSD. 1.862666 Equivalent to CR-CC(2,3).
10 / 10 0 337.263 3.392-3.392 19.484 45.638 1.862
100 / 122 0.0 250.492 6.090 16.021 38.309 1.411
1,000 / 1,006 0.1 105.265 5.589 7.036 21.727 0.132
5,000 / 8,118 1.1 17.355 0.787(1) 0.815(1) 1.725 0.003-0.003
10,000 / 16,291 2.1 14.555 0.860(1) 0.878(1) 1.338 0.012
50,000 / 65,172 5.2 11.064 0.800(1) 0.810(1) 0.922 0.015
100,000 / 130,448 8.4 9.410 0.655(1) 0.662(1) 0.695 0.009
500,000 / 521,578 19.8 5.929 0.375(1) 0.378(1) 0.400 0.005
1,000,000 / 1,043,539 28.0 4.820 0.306(0) 0.308(0) 0.314 0.002
5,000,000 / 8,190,854 72.8 0.764 0.047(1) 0.047(1) 0.009 0.000-0.000
5.0 1 / 1 0 730.244333 Equivalent to RHF. 183.276-183.276444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 430.051 49.816555 Equivalent to CCSD. 1.613666 Equivalent to CR-CC(2,3).
10 / 15 0 310.757 4.700 21.059 49.816 1.613
100 / 151 0.0 236.876 13.785 21.508 37.524 1.418
1,000 / 1,241 0.2 70.879 6.966 7.491 5.154 0.144
5,000 / 9,977 1.2 14.531 1.033(0) 1.050(0) 1.489 0.029
10,000 / 19,957 2.2 12.550 1.039(0) 1.050(0) 1.156 0.029
50,000 / 79,866 4.6 9.025 0.764(1) 0.770(1) 0.764 0.022
100,000 / 159,668 7.6 7.495 0.580(1) 0.584(1) 0.584 0.013
500,000 / 639,593 18.0 4.391 0.276(0) 0.277(0) 0.294 0.003
1,000,000 / 1,278,976 22.0 3.682 0.238(0) 0.239(0) 0.259 0.003
5,000,000 / 5,099,863 46.1 0.675 0.041(1) 0.041(1) 0.009 0.000-0.000
Table 3: Convergence of the CC(PP) and CC(PP;QQ) energies toward CCSDT, alongside the variational and perturbatively corrected CIPSI energies, for the F2\mathrm{F_{2}}/cc-pVTZ molecule in which the F–F bond length RR was fixed at 2Re2R_{e}, where Re=2.66816R_{e}=2.66816 bohr is the equilibrium geometry. For each value of the wave function termination parameter Ndet(in)N_{\text{det(in)}}, the PP space used in the CC(PP) calculations consisted of all singles and doubles and a subset of triples contained in the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state of the underlying CIPSI run, whereas the QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. In all post-RHF calculations, the two lowest-lying core orbitals were frozen and the spherical components of dd and ff orbitals were employed throughout. Each CIPSI run was initiated from the RHF reference determinant and the MBPT-based stopping parameter η\eta was set at 10610^{-6} hartree. The input parameter ff controlling the CIPSI wave function growth was set at the default value of 2.
Ndet(in)N_{\text{det(in)}} / Ndet(out)N_{\text{det(out)}} % of triples EvarE_{\text{var}}111 The EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=5,000,000N_{\text{det(in)}}=5,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. The extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy is 199.242119(59)-199.242119(59) hartree, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy obtained in the above extrapolation procedure. Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}111 The EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=5,000,000N_{\text{det(in)}}=5,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. The extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy is 199.242119(59)-199.242119(59) hartree, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy obtained in the above extrapolation procedure. Evar+ΔEr(2)E_{\text{var}}+\Delta E^{(2)}_{\text{r}}111 The EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=5,000,000N_{\text{det(in)}}=5,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. The extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy is 199.242119(59)-199.242119(59) hartree, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy obtained in the above extrapolation procedure. CC(P)(P)222 The CC(PP) and CC(PP;QQ) energies are reported as errors relative to CCSDT, in millihartree. The total CCSDT energy is 199.238344-199.238344 hartree. CC(P;Q)(P;Q)222 The CC(PP) and CC(PP;QQ) energies are reported as errors relative to CCSDT, in millihartree. The total CCSDT energy is 199.238344-199.238344 hartree.
1 / 1 0 758.849333 Equivalent to RHF. 165.740-165.740444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 340.460 62.819555 Equivalent to CCSD. 4.254666 Equivalent to CR-CC(2,3).
10 / 18 0 441.567 0.554-0.554 31.337 62.819 4.254
100 / 156 0.00 393.749 6.420 28.790 58.891 3.683
1,000 / 1,277 0.01 253.172 13.595(0) 20.323 42.564 1.579
5,000 / 5,118 0.03 123.591 10.874(1) 12.149(1) 18.036 0.345
10,000 / 10,239 0.06 73.122 7.202(5) 7.636(5) 11.439 0.198
50,000 / 82,001 0.84 29.674 3.371(2) 3.428(2) 4.898 0.061
100,000 / 163,866 1.58 27.002 3.068(2) 3.113(2) 4.157 0.049
500,000 / 655,859 3.75 22.301 2.517(1) 2.547(1) 3.111 0.014
1,000,000 / 1,311,633 5.58 20.244 2.292(1) 2.316(1) 2.739 0.009
5,000,000 / 5,253,775 13.3 14.499 1.645(1) 1.657(1) 1.866 0.015-0.015
Table 4: Convergence of the CC(PP) and CC(PP;QQ) energies toward CCSDT, alongside the variational and perturbatively corrected CIPSI energies, for the reactant (R) and transition-state (TS) species involved in the automerization of cyclobutadiene, as described by the cc-pVDZ basis set, and for the corresponding barrier height. The R and TS geometries, optimized using the MR-AQCC approach, were taken from Ref. Eckert-Maksić et al., 2006. For each value of the wave function termination parameter Ndet(in)N_{\text{det(in)}}, the PP space used in the CC(PP) calculations consisted of all singles and doubles and a subset of triples contained in the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle state of the underlying CIPSI run, whereas the QQ space needed to compute the corresponding CC(PP;QQ) correction δ(P;Q)\delta(P;Q) was defined as the remaining triples absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. In all post-RHF calculations, the four lowest-lying core orbitals were frozen and the spherical components of dd orbitals were employed throughout. Each CIPSI run was initiated from the RHF reference determinant and the MBPT-based stopping parameter η\eta was set at 10610^{-6} hartree. The input parameter ff controlling the CIPSI wave function growth was set at the default value of 2.
Species Ndet(in)N_{\text{det(in)}} / Ndet(out)N_{\text{det(out)}} % of triples EvarE_{\text{var}}111 For each of the two species, the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=15,000,000N_{\text{det(in)}}=15,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. These extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies for the R and TS species are 154.249292(314)-154.249292(314) and 154.235342(321)-154.235342(321) hartree, respectively, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained in the above extrapolation procedure. The EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} barrier heights are reported as errors, in kcal/mol, relative to the reference value of 8.753(0) kcal/mol obtained using the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies of the R and TS species. Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}111 For each of the two species, the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=15,000,000N_{\text{det(in)}}=15,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. These extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies for the R and TS species are 154.249292(314)-154.249292(314) and 154.235342(321)-154.235342(321) hartree, respectively, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained in the above extrapolation procedure. The EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} barrier heights are reported as errors, in kcal/mol, relative to the reference value of 8.753(0) kcal/mol obtained using the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies of the R and TS species. Evar+ΔEr(2)E_{\text{var}}+\Delta E^{(2)}_{\text{r}}111 For each of the two species, the EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies are reported as errors, in millihartree, relative to the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energy found using a linear fit based on the last four Evar,k+ΔEr,k(2)E_{\text{var},k}+\Delta E_{\text{r},k}^{(2)} values leading to the largest CIPSI wave function obtained with Ndet(in)=15,000,000N_{\text{det(in)}}=15,000,000, plotted against the corresponding ΔEr,k(2)\Delta E_{\text{r},k}^{(2)} corrections, following the procedure used in Ref. Loos, Damour, and Scemama, 2020. These extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies for the R and TS species are 154.249292(314)-154.249292(314) and 154.235342(321)-154.235342(321) hartree, respectively, where the error bounds in parentheses correspond to the uncertainty associated with the linear fit. The error bounds for the Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)} and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained at the various values of Ndet(in)N_{\text{det(in)}} reflect on the semi-stochastic design of the 𝒱ext(k)\mathcal{V}_{\text{ext}}^{(k)} spaces discussed in the main text, but they ignore the uncertainties characterizing the reference Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies obtained in the above extrapolation procedure. The EvarE_{\text{var}}, Evar+ΔE(2)E_{\text{var}}+\Delta E^{(2)}, and Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} barrier heights are reported as errors, in kcal/mol, relative to the reference value of 8.753(0) kcal/mol obtained using the extrapolated Evar+ΔEr(2)E_{\text{var}}+\Delta E_{\text{r}}^{(2)} energies of the R and TS species. CC(P)(P)222 The CC(PP) and CC(PP;QQ) energies of the R and TS species are reported as errors relative to CCSDT, in millihartree. The total CCSDT energies of the R and TS species are 154.244157-154.244157 and 154.232002-154.232002 hartree, respectively. The CC(PP) and CC(PP;QQ) barrier heights are reported in kcal/mol relative to the CCSDT value of 7.627 kcal/mol. CC(P;Q)(P;Q)222 The CC(PP) and CC(PP;QQ) energies of the R and TS species are reported as errors relative to CCSDT, in millihartree. The total CCSDT energies of the R and TS species are 154.244157-154.244157 and 154.232002-154.232002 hartree, respectively. The CC(PP) and CC(PP;QQ) barrier heights are reported in kcal/mol relative to the CCSDT value of 7.627 kcal/mol.
R 1 / 1 0 598.120333 Equivalent to RHF. 83.736-83.736444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 120.809 26.827555 Equivalent to CCSD. 0.848666 Equivalent to CR-CC(2,3).
50,000 / 55,653 0.0 121.880 26.065(182) 28.096(178) 25.468 0.678
100,000 / 111,321 0.1 109.688 23.819(163) 25.397(160) 22.132 0.382
500,000 / 890,582 1.2 93.413 19.049(141) 20.167(139) 16.260 0.267
1,000,000 / 1,781,910 2.0 89.989 18.322(137) 19.348(135) 15.359 0.251
5,000,000 / 7,125,208 7.9 78.122 16.311(123) 17.045(122) 10.794 0.150
10,000,000 / 14,253,131 11.8 73.250 15.514(115) 16.146(114) 9.632 0.127
15,000,000 / 28,493,873 25.8 60.872 12.842(96) 13.260(95) 4.817 0.046
TS 1 / 1 0 632.707333 Equivalent to RHF. 102.816-102.816444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 282.246 47.979555 Equivalent to CCSD. 14.636666 Equivalent to CR-CC(2,3).
50,000 / 56,225 0.0 146.895 45.357(180) 47.696(176) 42.132 9.563
100,000 / 112,481 0.1 130.832 36.716(183) 38.673(179) 31.723 3.507
500,000 / 899,770 1.0 93.288 18.106(139) 19.251(137) 14.742 0.432
1,000,000 / 1,800,183 1.6 89.049 17.458(142) 18.482(140) 13.645 0.412
5,000,000 / 7,195,780 5.4 78.472 15.587(124) 16.346(123) 10.720 0.260
10,000,000 / 14,400,744 9.7 71.784 14.397(114) 15.016(113) 8.358 0.155
15,000,000 / 28,793,512 15.2 63.375 12.587(102) 13.058(101) 7.080 0.108
Barrier 1 / 1 ; 1 0 ; 0 21.703333 Equivalent to RHF. 11.973-11.973444 Equivalent to the second-order MBPT energy using the Epstein–Nesbet denominator. 101.303 13.274555 Equivalent to CCSD. 8.653666 Equivalent to CR-CC(2,3).
50,000 / 55,653 ; 56,225 0.0 ; 0.0 15.697 12.106(161) 12.299(157) 10.457 5.576
100,000 / 111,321 ; 112,481 0.1 ; 0.1 13.268 8.093(154) 8.331(151) 6.018 1.961
500,000 / 890,582 ; 899,770 1.2 ; 1.0 0.079-0.079 0.592(124)-0.592(124) 0.574(122)-0.574(122) 0.953-0.953 0.104
1,000,000 / 1,781,910 ; 1,800,183 2.0 ; 1.6 0.590-0.590 0.542(124)-0.542(124) 0.544(122)-0.544(122) 1.075-1.075 0.101
5,000,000 / 7,125,208 ; 7,195,780 7.9 ; 5.4 0.220 0.454(110)-0.454(110) 0.439(109)-0.439(109) 0.047-0.047 0.069
10,000,000 / 14,253,131 ; 14,400,744 11.8 ; 9.7 0.920-0.920 0.701(102)-0.701(102) 0.710(100)-0.710(100) 0.800-0.800 0.017
15,000,000 / 28,493,873 ; 28,793,512 25.8 ; 15.2 1.571 0.159(88)-0.159(88) 0.127(87)-0.127(87) 1.420 0.039
Refer to caption
Figure 1: Convergence of the CC(PP) (red lines and circles) and CC(PP;QQ) (black lines and squares) energies toward their CCSDT parents as functions of the actual numbers of determinants, Ndet(out)N_{\mathrm{det(out)}}, defining the sizes of the final wave functions |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle generated in the underlying CIPSI runs, for the F2\mathrm{F_{2}}/cc-pVDZ molecule in which the F–F bond length RR was set at (a) ReR_{\mathrm{e}}, (b) 1.5ReR_{\mathrm{e}}, (c) 2ReR_{\mathrm{e}}, and (d) 5ReR_{\mathrm{e}}, where Re=2.66816R_{e}=2.66816 bohr is the equilibrium geometry. The PP spaces used in the CC(PP) calculations consisted of all singles and doubles and subsets of triples contained in the final |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle states of the underlying CIPSI runs, whereas the QQ spaces needed to compute the corresponding CC(PP;QQ) corrections δ(P;Q)\delta(P;Q) were defined as the remaining triples absent in |Ψ(CIPSI)|\Psi^{(\text{CIPSI})}\rangle. The insets show the percentages of triples captured by the CIPSI runs as functions of Ndet(out)N_{\text{det(out)}}.