In this paper, we study the nuclear gradients of heat bath configuration interaction self-consistent field (HCISCF) wave functions and use them to optimize molecular geometries for various molecules. We show that HCISCF nuclear gradients are fairly insensitive to the size of the “selected” variational space, which allows us to reduce the computational cost without introducing significant errors. The ability of the HCISCF to treat larger active spaces combined with the flexibility for users to control the computational cost makes the method very attractive for studying strongly correlated systems, which require a larger active space than possible with a complete active space self-consistent field. Finally, we study the realistic catalyst, Fe(PDI), and highlight some of the challenges this system poses for density functional theory (DFT). We demonstrate how HCISCF can clarify the energetic stability of geometries obtained from DFT when the results are strongly dependent on the functional. We also use the HCISCF gradients to optimize geometries for this species and study the adiabatic singlet–triplet gap. During geometry optimization, we find that multiple near-degenerate local minima exist on the triplet potential energy surface.

Systems that contain transition metals, covalent bond breaking, and electronically excited states are often challenging for theoretical methods because their electronic structure can be dominated by more than one electronic configuration. To study such systems, we require multireference approaches such as the complete active space self-consistent field (CASSCF) method.1–3 In the CASSCF, we perform the full configuration interaction (FCI) procedure on a subset of the molecular orbitals termed the active space. The FCI expansion enumerates all possible configurations for a given set of orbitals and scales combinatorially with the size of the active space. This steep cost limits the size of active spaces that can be treated to roughly 22 electrons in 22 orbitals, which we abbreviate as (22e,22o).4 A substantial amount of research in quantum chemistry is devoted to reducing the cost of configuration interaction (CI) methods through a variety of approximations: restricted active space,3,5 generalized active space,2,6,7 density matrix renormalization,8–29 selected configuration interaction (SCI),30–49 and FCI quantum Monte Carlo.50–53 The heat bath configuration interaction (HCI) is a particularly efficient implementation of the SCI method.54–56 These methods can correlate more than 40 electrons in 40 orbitals in routine calculations, making the study of larger and more complex systems accessible.

With the ability to treat larger active spaces, extending these algorithms to calculate molecular properties other than single-point energy is vital for connecting theoretical and experimental research. One such property of interest is the nuclear gradients of the electronic energy, which are crucial for chemical applications because they are used to obtain minimum energy geometries, transition states, and reaction pathways, and are used in ab initio molecular dynamics.57 Many of the approximate CI/CASSCF schemes have already been extended to calculate the nuclear gradients,58–68 and in this work, we extend the family of HCI methods.

The recently developed HCI algorithm has proved to be an efficient approximation of FCI where the size of the CI expansion is controlled with a single user parameter ϵ1. This parameter allows one to continuously grow the wave function from a single determinant, e.g., Hartree–Fock (HF), to all configurations in the FCI expansion. The energy of this wave function can be corrected with a perturbative correction, which is often necessary to reduce the error, relative to FCI, to an acceptable level. In recent work, two of the authors extended the HCI algorithm to be used in CASSCF-like calculations and implemented this method in the open source quantum chemistry package PySCF.69–72 This work showed that the multireference orbitals can be obtained from HCISCF with relatively loose thresholds and as long as a final step in which the optimized orbitals are used to perform a single tight threshold HCI calculation, the final results are in good agreement with the full CASSCF calculation.56 In practice, this insensitivity of the final results to the optimized orbitals is advantageous because it means that accurate multireference orbitals can be obtained from relatively cheap HCI calculations. In general, other properties may also be insensitive to the accuracy of the HCI wave function, which motivates this work and the study of nuclear gradients.

Previous work on approximate CI/CASSCF schemes has shown that for most systems, properties such as gradients and relaxed geometries are often quite accurate even when the energy is not.58–68 In particular, Park has studied the gradients of a different selected CI scheme, adaptive sampling CI (ASCI), and demonstrated that CASSCF gradients and geometries were well approximated by this method.67,68 Park showed that geometries converged faster than energies with respect to the number of determinants in the wave function and that the convergence of both energy and relaxed geometries could be improved by a second order perturbation theory correction (PT2) and by extrapolating these corrections.68 As a result, important properties of these molecules such as singlet–triplet (S-T) gaps were also improvable with these corrections, but we note that for most systems, the difference after the correction was typically around 1 kcal/mol. However, Park’s work focused largely on aromatic systems, which are not particularly strongly correlated, while in this work, we focus on the transition metal complex Fe(PDI), which is highly multireference.73 

The rest of this paper is organized as follows: First, we briefly review HCI theory and its self-consistent variant along with analytical gradients of CI wave functions. Next, we discuss the gradients of CASSCF-like wave functions where HCI is used as the CI solver and verify that they converge smoothly with the HCI parameter ϵ1. Then, we highlight several strategies to improve the quality of the gradients when increasing the size of the CI expansion is not feasible, e.g., when memory is limited. Finally, we demonstrate the use of these gradients to find equilibrium geometries and adiabatic singlet–triplet gaps for the model catalytic system (PDI)Fe–N2, where PDI is 2,6-bis[1-(2,6-dimethylphenyl-imino)ethyl]pyridine. For simplicity, we refer to this complex as Fe(PDI) and show it in Fig. 1. For this challenging system, we also examine the important practical consideration of selecting the active space in a robust numerical manner.

FIG. 1.

The model Fe(PDI) complex. Ar = 2,6-dimethylphenyl. The numbering convention is the same as Stieber et al.74 for consistency.

FIG. 1.

The model Fe(PDI) complex. Ar = 2,6-dimethylphenyl. The numbering convention is the same as Stieber et al.74 for consistency.

Close modal

HCI is a two-step procedure consisting of a variational calculation and a perturbative correction. In the first stage of the HCI algorithm, a set of important determinants is iteratively built and used to expand the variational wave function, then a second-order correction to the energy is computed using multireference Epstein–Nesbet perturbation theory (PT).75,76 Like other selected CI (SCI) + PT methods, it derives from FCI and linearly parametrizes the wave function |Ψ⟩ = ici|Di⟩. Unlike FCI, where the expansion is a set of all possible determinants, SCI methods express the wave function as a set of important determinants. This allows SCI+PT methods to correlate much larger systems than possible with FCI.

1. Variational stage

In the first step, the HCI algorithm identifies a set of important determinants that approximate the full many-body wave function and we refer to this set as the variational space (V). At the beginning of the variational step, the user specifies a small number of determinants to add to V, and usually, this includes only the Hartree–Fock determinant. Then, the algorithm adds all connected determinants |Da⟩, which satisfy the following importance criteria:

(1)

where |Di⟩ is a determinant already in the variational space, Hai=Da|Ĥ|Di, ci is the amplitude of the |Di⟩, and ϵ1 is a user-specified parameter that controls the size of the variational space, i.e., the accuracy of the variational wave function. A smaller ϵ1 means that V is closer to the FCI determinant space and in the limit that ϵ1 goes to 0, we recover our FCI wave function and energy. We repeat this selection of important determinants and grow the variational space until the HCI energy converges.

The CIPSI selection criteria, iHkiciE0Hkk>ϵCIPSI, is more complicated and uses first order perturbation theory to determine whether a determinant should be added to the variational space. Despite its similarity to the CIPSI selection criteria, the HCI importance function offers several key advantages over the original CIPSI algorithm. Since the matrix elements between determinants connected by double-excitations are just the two-electron integral elements (and a parity factor of +1 or −1), evaluating the HCI selection criteria requires no calculations. While avoiding calculations is advantageous, the real gain in performance comes from the fact that the two electron integrals are stored and sorted once before the variational stage. Selecting important determinants only requires iterating through the sorted two-electrons integrals and the algorithm stops once those values fall below a certain threshold. As a result, this means that the algorithm never even “touches” the unimportant connected determinants, which often outnumber the important ones by many orders of magnitude. We note that HCI was inspired by challenges in the original formulation of CIPSI and more modern implementations have shown substantial improvements compared to their predecessors.77 

2. Perturbative stage

We then correct the variational energy using a second order correction from Epstein–Nesbet perturbation theory, evaluated as

(2)

where the inner summation is “screened” in a similar manner to the variational stage. Analogously, the perturbative expansion grows as ϵ2 becomes smaller, and in the limit ϵ2 → 0, the perturbative correction becomes exact.

While screening the inner sum in Eq. (2) mitigates the large memory requirement for calculating a perturbative correction, it does not eliminate it entirely. Calculating an accurate correction requires a small ϵ2 value and for large systems, evaluating Eq. (2) deterministically becomes prohibitively expensive. In these cases, we can reduce the memory requirement further by calculating E2(ϵ2) semi-stochastically. By that, we mean that we calculate as much of the correction deterministically as computational resources allow and then calculate the remainder stochastically. This procedure is advantageous compared to a purely stochastic one because the portion of E2(ϵ2) calculated stochastically is smaller, and it takes fewer iterations to converge the stochastic error. We refer to calculations that use this procedure as semi-stochastic HCI (SHCI) to distinguish them from HCI calculations that only use the variational step. There is another even more efficient scheme, which was recently introduced by one of the authors of this paper, but we do not use that scheme in this work.78 

We briefly review analytical energy gradients of variational wave functions. For a detailed discussion on the subject, we refer to the reader to the review by Yamaguchi and Schaefer.79 We start with the electronic Hamiltonian in the second quantization,

(3)

We define the one- and two-body integrals in the usual way,

(4)

where ϕ are orbitals, ZI are the nuclear charges, rI the electron–nuclear separations, and r12 the electron electron separation. We use I, j, k, and l as orthogonal orbital indices and σ, σ′ as spin indices. In most cases, these are molecular orbitals (MOs), which are linear combinations of atomic orbitals (AOs) |i=μAOCμi|μ. For the remainder of this work, we use the following convention for orbital indices: Greek letters, e.g., μ and ν refer to AOs, while i, j refer to MOs.

The active space Hamiltonian is a function of the basis set, the MO coefficients C, and the nuclear coordinates {ai}, while the electronic wave function |Ψ⟩ is a function of the parameters {ci} making the electronic energy a function of all four. Expanding the nuclear gradients in terms of these variables, we get

(5)

The first term is the contribution from the basis set, i.e., the atomic orbitals, the second term is from the MO coefficients, and the final term is from the wave function parameters. Since the CI coefficients are always determined variationally so Eelecci=0 the term always vanishes. Similarly, for wave functions where the molecular orbital parameters are variational, EelecCμi=0 and the second term on the right-hand side of Eq. (5) also vanishes. To derive an explicit form for the gradients, we start by expressing the electronic energy in the second-quantized form:

(6)

where γij=σΨ|aiσajσ|Ψ is the one-body reduced density matrix (RDM) and Γijkl=12σσΨ|aiσajσakσalσ|Ψ is the two-body RDM. Taking derivative of Eq. (6) with respect to a general nuclear coordinate a generates terms, including γija and Γijkla. Since the RDMs are just functions of the CI coefficients, we know that we can ignore their derivatives because the final term in Eq. (5) vanishes. As a result, we arrive at a concise expression for then the nuclear gradients of the energy,

(7)

Expanding this equation further, we can separate the terms relating to the MO response,

(8)

where hija, gijkla, and Sija are the “skeleton” derivatives of the one-electron, two-electron, and overlap integrals defined as

(9)

Uija is called the orbital response and is defined by

(10)

It can be determined by solving the z-vector equations.80 Finally, Xij is the Lagrangian from the Generalized Brillouin Theorem, defined as

(11)

For the CASSCF, Xij is symmetric, i.e., Xij = Xji, and we can simplify Eq. (8) to

(12)

and avoid solving the z-vector equations. We direct the reader to the derivation of Eq. (133) in the excellent review by Yamaguchi and Schaefer.79 All three terms in Eq. (12) arise from the response of the basis set to changes in nuclear coordinates, and no terms from the response of MO or CI coefficients are present.

This derivation hinges on the fact that Xij is symmetric and for approximate CASSCF schemes such as HCISCF and DMRGSCF, Xij is generally not symmetric and since we ignore the third term in Eq. (12), it is not exact. In practice, one can solve include rotation among the active space orbitals, which we refer to as active–active (AA) rotations during the orbital optimization to eliminate the error of using Eq. (12) with approximate schemes such as DMRGSCF and HCISCF. Despite the formal lack of exactness, we highlight that if these approximate CI methods are sufficiently converged, the expression for the gradients is a justified approximation. If these methods are not converged to the FCI limit, Eq. (12) must include a response term due to the AA rotations to be formally exact.

Finally, we note that screening in Eqs. (1) and (2) is dependent on the geometry, meaning that the potential energy surface can be discontinuous for a fixed value of ϵ1. During the finite-difference calculations used to verify the analytic gradients, we only observed discontinuities for extremely inaccurate HCI wave functions, where the HCI parameter ϵ1 is greater than 10 mhartree and the wave functions consisted of less than 100 determinants. Since typical HCI calculations use much smaller values of ϵ1 and many more determinants, we do not explore these discontinuities further in this work.

For all DFT geometry optimizations, we use Gaussian16 with the def2-TZVP basis set and density fitting to reduce the computational cost.81 We use the default convergence thresholds in the Gaussian16 package. In addition to finding optimal geometries, we used harmonic frequency analysis to confirm that the reported geometries are all genuine local minima, i.e., contain no imaginary frequencies. We perform all other quantum chemical calculations using PySCF69–72 and checked several of the mean-field calculations using Q-Chem.82 For HCISCF calculations we use our own implementation of the HCI algorithm, called Dice,56 in tandem with PySCF. The nuclear gradients are calculated in PySCF using the RDMs obtained from Dice and geomeTRIC83 is used to perform the geometry updates. Due to the challenging nature of the Fe(PDI) problem, we relaxed the geometry convergence thresholds for all optimizations of the Fe(PDI) species, which are each double the default thresholds used by both Gaussian16 and geomeTRIC. Our convergence thresholds are 2 · 10−6 hartree for the energy (τE), 6  ·  10−4 hartree/bohr for the root mean squared (rms) of the nuclear gradient (τgrms), 9  ·  10−4 hartree/bohr for the maximum of the nuclear gradient (τgmax), 2.4  ·  10−3 Å for the rms of the atomic displacement (τdrms), and 3.6  ·  10−3 Å for the maximum atomic displacement (τdmax).

For the Fe(PDI) species, we use a numerical procedure to select the active space to reduce the heuristics which typically plague active space calculations. All multireference calculations use the cc-pVDZ basis set.84 We outline the procedure for selecting the active space orbitals below:

  1. Run unrestricted DFT calculation using the PBE0 functional85,86 to generate an initial guess for Step 2.

  2. Run an unrestricted Hartree–Fock (UHF) calculation to generate spin orbitals without correlation from an exchange-correlation functional.

  3. Use the procedure outlined in Ref. 87 to generate unrestricted natural orbitals (UNOs).88 

  4. Use the UHF natural orbital occupation numbers (NOONs) along with the spin density information from the UHF calculation to ensure that the targeted state shows a strong correlation and qualitatively correct spin density.

  5. Use the UNOs in a “loose” HCI calculation where the ϵ1 = 10−4 hartree and the active space is (100e,100o). We select the orbitals based on their NOON so the 50 orbitals below the highest occupied natural orbital (HONO) and 50 above the lowest unoccupied natural orbital (LUNO) are part of the active space.

  6. Use the 1-RDM from the previous step to calculate new NOs and sort based on NOONs.

  7. Use these orbitals in all subsequent HCISCF calculations, selecting the orbitals solely based on their NOONs analogous to the way used in Step 5. We select the orbitals in pairs (i.e., one occupied and one virtual) around the HONO-LUNO gap. For all Fe(PDI) calculations, we use an active space of (40e,40o), which includes more than those recommended by the criteria UNO methods, i.e., 0.02 < NOON < 1.98.88 

We emphasize that Steps 1 and 2 are sensitive to the optimization strategy used for the SCF procedure, e.g., DIIS,89 ADIIS,90 or Newton’s Method. We surveyed a range of choices for these strategies in both steps and select the best pair of strategies for each species. By “best” here, we mean that the SCF solution is the lowest in energy, stable with respect to its variational parameters, and has a non-negligible spin-density on the Fe atom. All results from this survey are shown and described in  Appendix C. While we have tried to search thoroughly for the lowest energy UHF solution, we note that it is possible that even lower energy solutions exist in this complicated energy landscape. From our experience, these different SCF solutions could yield different multireference orbitals, i.e., distinct multiconfigurationalself-consistent field (MCSCF) wave functions, and as a result different optimized molecular structures.

For all geometry comparisons, we use Kabsch’s algorithm91 to align geometries and report the difference in geometry as Δ2na where Δ is the difference in aligned nuclear coordinates and na is the number of atoms. For brevity, we refer to this quantity as RMSD for the remainder of the paper. To align the geometries we use the open source package rmsd.92 

The input scripts necessary to reproduce all of the results from this work are publicly available on GitHub at https://github.com/jamesETsmith/hci_nuc_gradients.

In this subsection, we discuss the errors that arise when using Eq. (12) for approximate CASSCF wave functions. In Fig. 2, we compare different variants of HCISCF gradients to CASSCF gradients, all using the cc-pVDZ basis set.84 As mentioned earlier, the HCISCF gradients approximate CASSCF gradients because HCI is an approximation of FCI and as a result, the energy is not invariant for AA orbital rotation. In our discussions of gradients, we note that all gradients discussed suffer from basis set incompleteness errors, but since that is constant throughout the calculations, we do not discuss it further.

FIG. 2.

The molecular systems used Sec. IV A to study the gradients of HCI-type wave functions. (a) Stilbene and (b) bis-(ethylene-1,2-dithiolato)nickel [Ni(edt)2]. In these structures, gray corresponds to carbon, yellow to sulfur, green to nickel, and white to hydrogen.

FIG. 2.

The molecular systems used Sec. IV A to study the gradients of HCI-type wave functions. (a) Stilbene and (b) bis-(ethylene-1,2-dithiolato)nickel [Ni(edt)2]. In these structures, gray corresponds to carbon, yellow to sulfur, green to nickel, and white to hydrogen.

Close modal

Without solving the response term, there are two ways that we can eliminate the error of this approximation: we can include AA rotations in the MCSCF optimization or we can converge the HCI wave function to the FCI limit. Including the AA rotations in the MCSCF optimization removes the need for the z-vector equation completely and ensures that Eq. (12) is exact. However, it can significantly increase the number of MCSCF iterations required to reach convergence. AA rotations make the MCSCF procedure more challenging because the corresponding parameters are strongly coupled to the CI coefficients. Due to these challenges, optimizing AA rotations is often the most expensive way to improve the gradients. In the second approach, we can either increase the size of the variational space or include a perturbative correction and increase the size of the perturbative space to better converge an HCI wave function. By better converging to the FCI limit, we eliminate any effect of AA rotations on the energy, and as a result, any need for the z-vector equation, again ensuring that Eq. (12) is exact. It is more cost-efficient to increase the number of configurations treated perturbatively, and if we are memory limited, then this may be the only option. When performing CASSCF calculations using only the variational step of HCI, we refer to these as vHCISCF wave functions and when we use both the variational and the perturbative, we refer to them as HCISCF. In Fig. 2, we compare different flavors of HCI-based gradients for four systems: N2, Sc2, trans-stilbene, and bis-(ethylene-1,2-dithiolato)nickel [Ni(edt)2] using actives spaces of (10e,8o), (6e,18o), (14e,14o), and (18e,13o), respectively. The latter two systesms are shown in Fig. 3. For N2 and Sc2, these active space sizes correspond to the full valence space. These active spaces correspond to the full valence space for N2 and Sc2, the conjugated 2pz orbitals on carbon for trans-stilbene, and the conjugated 2pz orbitals of carbon and sulfur as well as the 3d orbitals of the nickel atom for Ni(edt)2. Schlimgen and Mazziotti have shown that Ni(edt)2 exhibits a strong multireference character making it an ideal candidate to study gradients of approximate CASSCF wave functions.63,93 The geometries for all systems studied are in the publicly available Github repository at https://github.com/jamesETsmith/hci_nuc_gradients. We start with an intentionally inaccurate vHCISCF wave function, to leave considerable room for improvement, and show that adding AA rotations or PT correction can decrease the gradient error. As expected, we find that AA rotations and the PT correction improve the quality of gradients and with the exception of N2 provide the largest reduction of error when used together.

FIG. 3.

Errors in the gradient relative to the CASSCF for various flavors of HCI-based wave functions for (a) N2 CAS(10e,8o)/cc-pVDZ at a bond length of 1 Å and (b) Sc2 CAS(6e,18o)/cc-pVDZ at 2.38 Å, (c) trans-stilbene CAS(14e,14o)/cc-pVDZ, and (d) Ni(edt)2 CAS(18e,13o)/cc-pVDZ. Calculations for (a) and (b) used values of ϵ1 = 5  ·  10−3 and ϵ2 = 10−10 hartree, and calculations for (c) and (d) used values of ϵ1 = 1  ·  10−3 and ϵ2 = 10−12 hartree. The error in the vHCISCF here is high due to the deliberately large ϵ1 parameter. This was done to highlight the ability of AA rotations and the perturbative correction to effectively reduce the overall error in the gradients.

FIG. 3.

Errors in the gradient relative to the CASSCF for various flavors of HCI-based wave functions for (a) N2 CAS(10e,8o)/cc-pVDZ at a bond length of 1 Å and (b) Sc2 CAS(6e,18o)/cc-pVDZ at 2.38 Å, (c) trans-stilbene CAS(14e,14o)/cc-pVDZ, and (d) Ni(edt)2 CAS(18e,13o)/cc-pVDZ. Calculations for (a) and (b) used values of ϵ1 = 5  ·  10−3 and ϵ2 = 10−10 hartree, and calculations for (c) and (d) used values of ϵ1 = 1  ·  10−3 and ϵ2 = 10−12 hartree. The error in the vHCISCF here is high due to the deliberately large ϵ1 parameter. This was done to highlight the ability of AA rotations and the perturbative correction to effectively reduce the overall error in the gradients.

Close modal

Although AA rotations and the PT correction in our MCSCF optimization can reduce the error in our nuclear gradients, they do so at a cost that is often difficult to predict. Both options increase the number of MCSCF iterations required for convergence and with AA rotations that the calculations can require more than 50 iterations and may not converge at all. This behavior is consistent with the previous DRMG-SCF94 and ASCI-SCF67,68 work but could likely be improved by using more sophisticated optimization like the ones described by Yao and Umrigar or Kreplin et al.94,95

Another strategy is to use a smaller ϵ1 value, which increases the number of determinants in the HCI wave function. Using ϵ1, we can control the accuracy of our gradients, compared to the CASSCF, and we find that we can use a relatively large value of ϵ1, i.e., a smaller number of determinants in our wave function, and still produce accurate gradients. For the same two systems as above, we examine the convergence of the analytic gradients with respect to ϵ1, i.e., the size of the variational space. Again, we use the cc-pVDZ basis set, and Fig. 4 shows the results for both systems. To contextualize the results, we show the loose, default, and tight gradient convergence tolerances used by the Gaussian16 package81 during geometry optimizations. In almost all cases, we find that AA rotations decrease the error relative to the CASSCF gradients, but, in several instances, such as ϵ1 = 5  ·  10−4 hartree in trans-stilbene, it does not. Since AA rotations are not guaranteed to reduce the vHCISCF or HCISCF gradient error relative to the CASSCF, it is not surprising that some cases exist where it increases the error slightly. For all wave functions, we observe a steady convergence of the gradient to the exact (CASSCF) value. The rate of convergence varies but shows similar behavior for all four systems, which is encouraging given the multireference nature of Sc2 and Ni(edt)2. For the typical range of ϵ1 values used in research applications (ϵ1 = 10−4–10−5 hartree), the HCISCF gradients agree well with their CASSCF counterparts. These encouraging results allow us to reduce the computational time required to calculate nuclear gradients of vHCISCF without significantly reducing their accuracy. For the remainder of this paper, all gradients are calculated with vHCISCF and no AA rotations, unless otherwise specified.

FIG. 4.

The error in the gradient as a function of ϵ1 for several HCI-based wave functions for (a) N2 CAS(10e,8o)/cc-pVDZ at a bond length of 1 Å and (b) Sc2 CAS(6e,18o)/cc-pVDZ at 2.38 Å, (c) trans-stilbene CAS(14e,14o)/cc-pVDZ, and (d) Ni(edt)2 CAS(18e,13o)/cc-pVDZ. The values of ϵ2 are the same as in Fig. 2. As the wave function converges to the CASSCF result, the improvements from AA rotations and PT contribution make a negligible difference in the gradient. We compare the errors in the gradient to the tolerances for the root mean squared (rms) gradient used by the Gaussian package during geometry optimization. In (c), vHCISCF+AA and HCISCF+AA values are omitted for ϵ1 = 10−6 because of MCSCF convergence problems. In (d), we note that the HCISCF provides such a small improvement over vHCISCF that they are indistinguishable in this figure.

FIG. 4.

The error in the gradient as a function of ϵ1 for several HCI-based wave functions for (a) N2 CAS(10e,8o)/cc-pVDZ at a bond length of 1 Å and (b) Sc2 CAS(6e,18o)/cc-pVDZ at 2.38 Å, (c) trans-stilbene CAS(14e,14o)/cc-pVDZ, and (d) Ni(edt)2 CAS(18e,13o)/cc-pVDZ. The values of ϵ2 are the same as in Fig. 2. As the wave function converges to the CASSCF result, the improvements from AA rotations and PT contribution make a negligible difference in the gradient. We compare the errors in the gradient to the tolerances for the root mean squared (rms) gradient used by the Gaussian package during geometry optimization. In (c), vHCISCF+AA and HCISCF+AA values are omitted for ϵ1 = 10−6 because of MCSCF convergence problems. In (d), we note that the HCISCF provides such a small improvement over vHCISCF that they are indistinguishable in this figure.

Close modal

We use two geometries reported by Ortuño and Cramer, which they obtained using DFT: the first, “1BS(1,1),” which we call A, optimized for the singlet spin symmetry, and the second, “3BS(3,1),” which we call C, optimized for the triplet.73 Both structures were calculated using unrestricted M06-L with the def2-TZVP basis and density fitting. In this work, we use the shorthand of (M,X) to describe a state with multiplicity M at geometry X.

Ortuño and Cramer note in their paper that “although we report DFT energies, we note that they are likely quantitatively unreliable owing to significant spin contamination and unpredictable sensitivity to [multireference] character.”73 Given this warning, we first studied Fe(PDI) with several DFT functionals to assess how sensitive properties such as energy and spin density (on the Fe atom) were to these choices. We chose several of the Minnesota functionals (M06-2X, M06-L, and MN15)97–98 because of their success with challenging transition metal complexes. We also optimized geometries to test how sensitive the relaxed structures were to the choice of these functionals.

For all optimizations, we used A as the starting geometry for singlets and C as the starting geometry for triplets. We found that for each multiplicity, the three functionals led to distinct geometries, i.e., six geometries in all. We compare the geometries in detail in  Appendixes A–E, see Table VI. In Fig. 5, we illustrate the difference in relative energy and Fe spin density at the relaxed geometries. The left panel demonstrates that the spin symmetry of the ground state depends on the functional, and the right panel highlights that spin density on the iron atom ranges from two to four unpaired electrons. We also found that the NOONs at each of these relaxed structures in Table I vary noticeably for different functionals. This qualitative difference makes it challenging to determine whether singlet and triplet geometries should start from A or C in subsequent multireference calculations. Throughout the study of this species, we have found that geometry optimization use of DFT and vHCISCF wave functions can depend on the initial geometry, which prompted us to select our initial geometries with care.

FIG. 5.

The effect of the functional on energy and spin density on the Fe atom for all systems studied with DFT. The energies and Fe spin density reported are from the relaxed geometries. The initial coordinates for calculations are from Ortuño and Cramer.

FIG. 5.

The effect of the functional on energy and spin density on the Fe atom for all systems studied with DFT. The energies and Fe spin density reported are from the relaxed geometries. The initial coordinates for calculations are from Ortuño and Cramer.

Close modal
TABLE I.

The natural orbital occupation numbers (NOONs) for the systems studied with DFT as a function of functional. All values reported are from the relaxed geometries. Here, HONO is the highest occupied NO and LUNO is the lowest unoccupied NO.

FunctionalMult.Geom.HONO-1HONOLUNOLUNO+1
M06-2X A 1.325 1.000 1.000 0.675 
M06-2X C 1.188 1.000 1.000 0.812 
M06-L A 1.921 1.000 1.000 0.079 
M06-L C 1.566 1.000 1.000 0.434 
MN15 A 1.617 1.000 1.000 0.383 
MN15 C 1.686 1.000 1.000 0.314 
FunctionalMult.Geom.HONO-1HONOLUNOLUNO+1
M06-2X A 1.325 1.000 1.000 0.675 
M06-2X C 1.188 1.000 1.000 0.812 
M06-L A 1.921 1.000 1.000 0.079 
M06-L C 1.566 1.000 1.000 0.434 
MN15 A 1.617 1.000 1.000 0.383 
MN15 C 1.686 1.000 1.000 0.314 

1. Selecting initial geometries

Since we hope to calculate accurate S-T gaps, which we can qualitatively compare to experiments, it is challenging to “decide,” which geometries and functionals that we should use for multireference calculations. To address this, we can use SHCI energies to determine which DFT geometry is the most stable in an unbiased manner and use these geometries in the subsequent study of the S-T gaps. For each geometry (A and C), we used the procedure outlined above to select the active space of size (40e,40o) and then ran single point vHCISCF calculations followed by a single tight SHCI calculation. We deliberately correlate a larger number of orbitals than would be selected by the NOON numerical thresholds mentioned earlier in order to confidently include all strongly correlated orbitals in the active space. For the vHCISCF calculations, we used a value of ϵ1 = 7.5·10−5 hartree. Given the sensitivity of the MCSCF solutions to the input orbitals, we tested several values of ϵ1 and found that the MCSCF orbitals converge faster than the energy with respect to this parameter agreeing with the previous work by several of the authors.56 We report this analysis in  Appendix D. To ensure that our choice of ϵ1 was appropriate, we tested the use of a more accurate ϵ1 value of 5 · 10−5 hartree and found that it produced geometries that are quite similar to those produced when using ϵ1 = 7.5·10−5. The RMSD value between the two optimized geometries is 5.6·10−3 Å, which is less than the loose convergence criteria used by Gaussian (6.7·10−3 Å). vHCISCF wave functions using ϵ1 = 7.5·10−5 contained roughly 3.5·106 determinants and those using ϵ1 = 5·10−5 contained roughly 7.7·106. The ability to use a larger value of epsilon reduced the time necessary to solve the CI problem from more than 1600 s to roughly 700 s. We point out that the savings would increase when compared to smaller values of ϵ1 and this speed-up is a rough lower bound. All timings reported here were recorded on a single AMD Rome node with 128 cores. For the final SHCI calculation, we used several values of ϵ1 to extrapolate to the FCI limit, the tightest one being 10−5, and ϵ2 = 5·10−8 hartree. We detail this extrapolation procedure in  Appendix A. Extrapolations for singlet and triplet states at all three geometries are shown in Figs. 6 and 7). The numerical data are shown in Table II.

FIG. 6.

The extrapolated SHCI energies at the initial geometry for both the singlet and triplet spin symmetries. For the extrapolated values, see Table II.

FIG. 6.

The extrapolated SHCI energies at the initial geometry for both the singlet and triplet spin symmetries. For the extrapolated values, see Table II.

Close modal
FIG. 7.

The extrapolated SHCI energies at the final (i.e., “relaxed”) geometries for both singlet and triplet spin symmetries. For the extrapolated values, see Table II.

FIG. 7.

The extrapolated SHCI energies at the final (i.e., “relaxed”) geometries for both singlet and triplet spin symmetries. For the extrapolated values, see Table II.

Close modal
TABLE II.

The energy relaxation for A, B, and C after vHCISCF geometry optimization. The initial and final energies are the extrapolated values, and all have uncertainties of roughly 1.5  ·  10−5 hartree.

Mult.Geom.Initial (hartree)Final (hartree)Relaxation (kcal/mol)
A −2497.697 79 −2497.735 52 −23.67 
B −2497.696 78 −2497.735 50 −24.29 
C −2497.671 40 −2497.719 16 −29.97 
A −2497.644 96 −2497.676 56 −19.83 
B −2497.680 06 −2497.709 22 −18.30 
C −2497.683 59 −2497.710 49 −16.88 
Mult.Geom.Initial (hartree)Final (hartree)Relaxation (kcal/mol)
A −2497.697 79 −2497.735 52 −23.67 
B −2497.696 78 −2497.735 50 −24.29 
C −2497.671 40 −2497.719 16 −29.97 
A −2497.644 96 −2497.676 56 −19.83 
B −2497.680 06 −2497.709 22 −18.30 
C −2497.683 59 −2497.710 49 −16.88 

Figure 6 indicates that A is the best structure to start geometry optimization for the singlet spin symmetry and C for the triplet, agreeing with the DFT results using M06-L and MN15. For completeness, we optimized the Fe(PDI) structure starting from both multiplicities at all initial geometries for a total of four vHCISCF geometry optimizations. For all vHCISCF gradients used in geometry optimizations, we use Eq. (12) without including any AA rotations or response terms mentioned in Sec. II making the gradients approximate. This approach is consistent with several previous studies where gradients of approximate CASSCF methods were calculated.59–61,67,68,99,100 In recent work, Park found little difference between ASCI-SCF geometries compared to the CASSCF, so we believe that this approximation is justified.68 Even though the systems studied by Park are not strongly correlated, our results for Ni(edt)2 suggest that even strongly correlated species follow the same trend.

2. Comparing vHCISCF optimized geometries

In the following discussion of optimized geometries, we consider differences in two geometries to be small when they are roughly equal to the tolerance that we use for convergence, τdrms (2.4  ·  10−3 Å), and large when they are more than one order of magnitude larger. The RMSD between initial and final geometries for all vHCISCF optimization is between 0.1 and 0.4 Å, indicating that the structures changed in a substantial manner, and the detailed values are shown in Table III. Since this difference is orders of magnitude larger than our threshold for convergence (3.6  ·  10−3 Å), the final geometries are all distinct from their starting structures. When comparing optimized geometries starting from different initial coordinates, the structural difference shrink for all singlet states after vHCISCF geometry optimization, while triplet states all diverge from each other. We show a numerical comparison of the geometries in Table IV.

TABLE III.

Comparing the change in geometry after optimization for the singlet and triplet species. All RMSD values are in Å and are calculated using the using the rmsd92 package.

GeometriesSinglet RMSDTriplet RMSD
A 1.7·10−1 1.3·10−1 
C 3.9·10−1 3.4·10−1 
GeometriesSinglet RMSDTriplet RMSD
A 1.7·10−1 1.3·10−1 
C 3.9·10−1 3.4·10−1 
TABLE IV.

Comparing pairs of geometries at the initial, singlet optimized, and triplet optimized structures. Optimizing the singlet decreases the differences in all structures while optimizing the triplet state increases them. All RMSD values are in Å and are calculated using the using the rmsd92 package.

GeometriesInitial RMSDOptimized singlet RMSDOptimized triplet RMSD
A,C 7.8·10−1 5.2·10−1 7.9·10−1 
GeometriesInitial RMSDOptimized singlet RMSDOptimized triplet RMSD
A,C 7.8·10−1 5.2·10−1 7.9·10−1 

Next, we examine the differences between the experimental and theoretical structures of Fe(PDI). For consistency, we use the same numbering scheme as Stieber et al.74Figure 8 shows errors in the initial and final geometries of state (1,A). To re-iterate, all initial geometry was optimized with symmetry-broken DFT, while the final geometries were obtained with the vHCISCF. We compare selected bond lengths from both geometries to experimental ones reported in Ref. 74, which were obtained from crystallographic experiments. We only report a subset of the bond lengths because (1,A) is only a model system for the experimental complex and they differ slightly in composition. Despite using multireference methods, we found that the optimized geometries did not outperform the M06-L, MN15, and M06-2X. However, the DFT geometry optimizations use def-TZVP while the vHCISCF calculations use cc-pVDZ. The def-TZVP basis is considerably larger than cc-pVDZ and should a priori provide a better description of geometry. Due to computational limitations, we were not able to run vHCISCF calculations using triple zeta basis sets. We found that the time to calculate the nuclear gradients increased by nearly 50 fold while the time per MCSCF iteration increased by less than a factor of two, making the gradients the clear bottleneck to using triple zeta basis sets. We note that agreement between the DFT-optimized and experimental geometries may be fortuitous since the Fe(PDI) complex contains more alkyl groups and the experimental evidence suggests that the electronic structure is sensitive to these changes.74 However, a further investigation is necessary to test this claim. The other geometry optimizations starting from other geometries and/or spin symmetry showed similar behavior, and we report them in Figs. 1012 in  Appendix E.

FIG. 8.

The errors in initial and final geometry compared to experimentally obtained bond lengths for the state (1,A).74 See Fig. 1 for a numerically labeled diagram Fe(PDI). We report the analogous figures for other spin states and starting geometries in  Appendix E.

FIG. 8.

The errors in initial and final geometry compared to experimentally obtained bond lengths for the state (1,A).74 See Fig. 1 for a numerically labeled diagram Fe(PDI). We report the analogous figures for other spin states and starting geometries in  Appendix E.

Close modal

3. Singlet-triplet gaps

Reference 74 argued that the ground state is a singlet with a thermally accessible triplet state based on the strong temperature dependence of NMR data for species similar to Fe(PDI). However, the set of compounds that they report in Table II of their paper differ by the alkyl (methyl, ethyl, or isopropyl) groups attached to C(2) and C(8) compared to Fe(PDI). The type of functional group noticeably impacts the extent of the temperature dependence, and it is not clear if those same arguments apply to our model species. In this work, we define thermally accessible as roughly 1 kcal/mol, i.e., “chemical accuracy.” Reference 73 also suggested that the S-T gap of Fe(PDI) small and reported vertical excitation values of 2.2 kcal/mol using RASSCF and 3.4 kcal/mol using RASPT2. Their RAS calculations correlated 22 electrons in 22 orbitals and use a mixed basis where a triple-zeta basis is used for Fe, a double-zeta basis is used for C and N, and a minimal basis is used for H.

Our calculations agree with previous experimental and theoretical work that the ground state has singlet spin symmetry but do not indicate that there is a low-lying triplet state. Our vertical excitation energies at A are roughly an order of magnitude larger than those reported by Ortuño and Cramer, while the initial adiabatic gap is more comparable. During optimization, the singlet species relax between 24 and 30 kcal/mol, while the triplet species relax by only 17–20 kcal/mol, leading to larger S-T gaps in contrast to Ortuño and Cramer’s expectations. We report the S-T gaps in Table V from extrapolated SHCI calculations along with DFT and RAS calculations from Ortuño and Cramer73 for reference.

TABLE V.

The singlet-triplet gaps in kcal/mol at the initial (DFT-optimized) reported by Ortuño and Cramer73 and final (vHCISCF-optimized) geometries. We show the states used to calculate each gap and point out that RASSCF and RASPT2 geometries are vertical excitations at geometry A, while all other initial gaps are adiabatic, i.e., at two distinct relaxed geometries.

MethodStatesInitialFinal
Ortuño and Cramer73  
M06-L (3,C)–(1,A9.5 ⋯ 
RASSCF (22e,22o) (3,A)–(1,A2.2 ⋯ 
RASPT2 (22e,22o) (3,A)–(1,A3.4 ⋯ 
This work 
SHCI (40e,40o) (3,A)–(1,A33.2 37.0 
SHCI (40e,40o) (3,C)–(1,A8.9 15.7 
MethodStatesInitialFinal
Ortuño and Cramer73  
M06-L (3,C)–(1,A9.5 ⋯ 
RASSCF (22e,22o) (3,A)–(1,A2.2 ⋯ 
RASPT2 (22e,22o) (3,A)–(1,A3.4 ⋯ 
This work 
SHCI (40e,40o) (3,A)–(1,A33.2 37.0 
SHCI (40e,40o) (3,C)–(1,A8.9 15.7 

4. Multiple local minima

During initial DFT geometries optimizations, we found another triplet geometry optimization with M06-L, which we call B. We found that this geometry was more similar to A than C and when compared to A had an RMSD value of 2.8·10−2 Å and differed in energy by roughly 1 kcal/mol. After running both singlet and triplet vHCISCF geometry optimizations starting from B, we found that the singlet geometry became much closer to A (8.0·10−3 Å) and had an extrapolated energy that agreed within 1.72·10−5 hartree, which is nearly the standard error (1.56·10−5) hartree, leading us to believe that these are the same structure. On the other hand, after geometry optimization for triplet spin symmetry, we found that the geometry differed from the relaxed C geometry by 7.9·10−1 Å but had an extrapolated energy that differs by less than 1 kcal/mol implying that there are at least two distinct stable triplet structures and that initial geometry choices are important for this species.

We have benchmarked nuclear gradients of HCI-type wave functions and demonstrated that like orbitals, nuclear gradients are relatively insensitive to the empirical parameter in HCI. While the nuclear gradients are formally inexact, this error tends to lead to only minimal differences in the final optimized geometries. If more accurate gradients or optimized structures are required, we note that the gradients are systematically improvable if more accurate gradients are required, they are systematically improvable. As a result, accurate molecular structures can be obtained with less computational cost and if accurate energies are desired, one can run a final more accurate SHCI calculation at this geometry.

We studied the challenging catalytic model, Fe(PDI), and used HCI-type wave functions to assess geometries obtained from DFT and to optimize new structures. Given the sensitivity of geometry optimization to the starting coordinates, HCI-based methods can unambiguously compare different geometries and determine the lowest energy structure when DFT cannot. For multireference systems, where single-reference methods, such as DFT and coupled-cluster, will give ambiguous results or even fail, HCI-type wave functions are a useful alternative. Our results agree with previous experimental and theoretical work that the ground state of the Fe(PDI) complex is a singlet. However, we found that the S-T was larger than previously predicted and was not thermally accessible. As alluded to earlier, there are several factors that could explain this difference such as the missing functional groups in our model system as well as different correlated methods and basis sets. When compared to the experimentally obtained structure, the vHCISCF performs similarly, albeit slightly worse, than the three DFT functionals studied. Fe(PDI) is a challenging system even for accurate multireference methods, and we encourage others in the community to use it as a benchmark system going forward.

This work utilized resources from the University of Colorado Boulder Research Computing Group, which was supported by the National Science Foundation (Award Nos. ACI-1532235 and ACI-1532236), the University of Colorado Boulder, and Colorado State University. S.S. acknowledges funding from NSF Award No. CHE1800584 and the Sloan research fellowship. J.E.T.S. gratefully acknowledges support from a fellowship through The Molecular Sciences Software Institute under NSF Grant No. ACI-1547580. J.E.T.S. was also supported by the Flatiron Institute, which is a division of the Simons Foundation. J.L. thanks David Reichman for encouragement and support.

The authors have no conflicts to disclose.

James E. T. Smith: Investigation (lead); Software (lead); Writing – original draft (lead); Writing – review & editing (lead). Joonho Lee: Investigation (supporting); Supervision (supporting). Sandeep Sharma: Supervision (lead); Writing – original draft (supporting); Writing – review & editing (supporting).

The data that support the findings of this study are openly available at https://github.com/jamesETsmith/hci_nuc_gradients.

All calculations used for extrapolation use ϵ2 = 5·10−8 hartree and ten different values of ϵ1 ranging from 5·10−5 to 10−5 hartree. The procedure for the extrapolation is as follows:

  1. Run all SHCI calculations.

  2. Collect the PT correction (E2) (our independent variable) and total energy (ESHCI).

  3. Fit linear curves to ESHCI as a function of E2.

  4. Estimate the error in this extrapolated energy as the standard deviation of the curve fitting parameter representing the y-intercept.

Here, we provide the numerical data for Fig. 5 along with ⟨S2⟩ and the RMSD compared to the singlet and triplet geometries from Ortuño and Cramer in Table VI.

TABLE VI.

Comparing the DFT results as a function of starting geometry and functional. The RMSD values are calculated relative to broken symmetry geometries Ortuño and Cramer reported. All singlets are compared to their “BS(1,1)” geometry, and triplets were compared to “BS(1,3).”

FunctionalMultiplicityStarting geometryEnergy (hartree)Fe spin densityŜ2RMSD (Å)
M06-2X A −2506.345 333 1.949 442 1.9345 1.1·10−1 
M06-2X C −2506.364 616 3.721 854 3.9329 3.3·10−1 
M06-L A −2506.768 403 1.454 747 1.2123 6.6·10−4 
M06-L C −2506.753 225 3.120 475 2.9871 4.4·10−3 
MN15 A −2505.529 508 1.767 367 1.6642 3.3·10−2 
MN15 C −2505.526 531 2.040 751 2.5562 7.8·10−1 
FunctionalMultiplicityStarting geometryEnergy (hartree)Fe spin densityŜ2RMSD (Å)
M06-2X A −2506.345 333 1.949 442 1.9345 1.1·10−1 
M06-2X C −2506.364 616 3.721 854 3.9329 3.3·10−1 
M06-L A −2506.768 403 1.454 747 1.2123 6.6·10−4 
M06-L C −2506.753 225 3.120 475 2.9871 4.4·10−3 
MN15 A −2505.529 508 1.767 367 1.6642 3.3·10−2 
MN15 C −2505.526 531 2.040 751 2.5562 7.8·10−1 

Since the UDFT and UHF calculations used to prepare orbitals for the multireference calculations were so sensitive to their optimization settings, we surveyed a large number of these settings to find the lowest energy state possible. We emphasize that such care is needed for this species because MCSCF optimizations are sensitive to the initial orbitals for this species. We examined six different optimization options for the DFT stage and three for the UHF stage leading to a total of 108 ways to prepare the orbitals. In Table VII, we show a subset of these calculations where the final UHF produced a stable wave function with respect to the orbital rotation parameters.

TABLE VII.

A summary of the various stable UHF solutions as a function of geometry, multiplicity, as well as SCF optimization options. The “Fast Newton” strategies are heuristic options implemented in PySCF where a second order optimization is performed on a minimal basis and then projected to the desired basis to accelerate the calculation. No stable solutions were found using DIIS in the UHF optimization.

MultiplicityGeometryDFT opt. strategyUHF opt. strategyFe spin densityS2Energy (hartree)
Fast Newton+ADIIS NEWTON −2.50 3.16 −2497.225 037 
ADIIS NEWTON −2.50 3.16 −2497.225 036 
NEWTON NEWTON 2.50 3.19 −2497.224 119 
Fast Newton+ADIIS ADIIS −2.01 2.69 −2497.219 975 
ADIIS ADIIS −2.01 2.70 −2497.219 970 
Fast Newton+DIIS NEWTON 7.50 × 10−2 2.75 −2497.188 196 
DIIS NEWTON −9.79 × 10−2 2.75 −2497.188 062 
Fast Newton+DIIS ADIIS −3.90 × 10−4 2.63 −2497.187 794 
DIIS ADIIS 9.10 × 10−4 2.61 −2497.175 954 
ADIIS NEWTON −1.98 3.60 −2497.207 645 
NEWTON NEWTON −1.78 4.16 −2497.200 086 
Fast Newton+ADIIS NEWTON 1.77 4.13 −2497.199 389 
Fast Newton+DIIS NEWTON −1.77 4.13 −2497.199 389 
Fast Newton+DIIS ADIIS −1.76 4.12 −2497.198 453 
NEWTON ADIIS −1.80 3.89 −2497.194 584 
Fast Newton+ADIIS ADIIS 1.80 3.88 −2497.194 527 
DIIS NEWTON −1.33 × 10−2 3.54 −2497.176 719 
DIIS ADIIS 7.18 × 10−3 3.34 −2497.175 636 
ADIIS NEWTON 2.06 3.81 −2497.206 466 
ADIIS ADIIS 2.05 3.58 −2497.205 565 
DIIS NEWTON 1.25 × 10−1 4.10 −2497.204 954 
Fast Newton+DIIS NEWTON 1.88 4.61 −2497.196 780 
NEWTON NEWTON 1.88 4.61 −2497.196 780 
Fast Newton+DIIS NEWTON 3.76 5.42 −2497.264 950 
NEWTON NEWTON 3.76 5.42 −2497.264 950 
Fast Newton+ADIIS NEWTON 3.76 5.42 −2497.264 950 
ADIIS NEWTON 3.76 5.42 −2497.264 950 
DIIS NEWTON 1.84 5.11 −2497.200 823 
DIIS ADIIS 1.99 4.33 −2497.196 357 
MultiplicityGeometryDFT opt. strategyUHF opt. strategyFe spin densityS2Energy (hartree)
Fast Newton+ADIIS NEWTON −2.50 3.16 −2497.225 037 
ADIIS NEWTON −2.50 3.16 −2497.225 036 
NEWTON NEWTON 2.50 3.19 −2497.224 119 
Fast Newton+ADIIS ADIIS −2.01 2.69 −2497.219 975 
ADIIS ADIIS −2.01 2.70 −2497.219 970 
Fast Newton+DIIS NEWTON 7.50 × 10−2 2.75 −2497.188 196 
DIIS NEWTON −9.79 × 10−2 2.75 −2497.188 062 
Fast Newton+DIIS ADIIS −3.90 × 10−4 2.63 −2497.187 794 
DIIS ADIIS 9.10 × 10−4 2.61 −2497.175 954 
ADIIS NEWTON −1.98 3.60 −2497.207 645 
NEWTON NEWTON −1.78 4.16 −2497.200 086 
Fast Newton+ADIIS NEWTON 1.77 4.13 −2497.199 389 
Fast Newton+DIIS NEWTON −1.77 4.13 −2497.199 389 
Fast Newton+DIIS ADIIS −1.76 4.12 −2497.198 453 
NEWTON ADIIS −1.80 3.89 −2497.194 584 
Fast Newton+ADIIS ADIIS 1.80 3.88 −2497.194 527 
DIIS NEWTON −1.33 × 10−2 3.54 −2497.176 719 
DIIS ADIIS 7.18 × 10−3 3.34 −2497.175 636 
ADIIS NEWTON 2.06 3.81 −2497.206 466 
ADIIS ADIIS 2.05 3.58 −2497.205 565 
DIIS NEWTON 1.25 × 10−1 4.10 −2497.204 954 
Fast Newton+DIIS NEWTON 1.88 4.61 −2497.196 780 
NEWTON NEWTON 1.88 4.61 −2497.196 780 
Fast Newton+DIIS NEWTON 3.76 5.42 −2497.264 950 
NEWTON NEWTON 3.76 5.42 −2497.264 950 
Fast Newton+ADIIS NEWTON 3.76 5.42 −2497.264 950 
ADIIS NEWTON 3.76 5.42 −2497.264 950 
DIIS NEWTON 1.84 5.11 −2497.200 823 
DIIS ADIIS 1.99 4.33 −2497.196 357 

We investigated the sensitivity of the vHCISCF orbitals to the HCI parameter, ϵ1, analogous to the analysis from Smith et al.56 This process involves two steps: first, converging the vHCISCF orbitals with respect to several different values of ϵ1 and second, running an SHCI calculation using these orbitals with identical settings. Despite the fact that Fe(PDI) is more multireference than the Fe(porphyrin) system studies in Ref. 56, we find that the orbitals are relatively insensitive to ϵ1 and show the results in Fig. 9.

FIG. 9.

We report the sensitivity of the vHCISCF orbitals with respect to ϵ1. In the two step procedure, we first optimize the vHCISCF orbitals and then run final ”tight” SHCI calculations using those orbitals to obtain more accurate energy. ϵ1 on the x-axis is for the vHCISCF calculations and indicates the accuracy/quality of the wave function with more accurate wave functions on the left. Our results show that the SHCI energy is relatively insensitive to ϵ1 used during the vHCISCF optimization.

FIG. 9.

We report the sensitivity of the vHCISCF orbitals with respect to ϵ1. In the two step procedure, we first optimize the vHCISCF orbitals and then run final ”tight” SHCI calculations using those orbitals to obtain more accurate energy. ϵ1 on the x-axis is for the vHCISCF calculations and indicates the accuracy/quality of the wave function with more accurate wave functions on the left. Our results show that the SHCI energy is relatively insensitive to ϵ1 used during the vHCISCF optimization.

Close modal

We show the error between several theoretical results and the experimentally determined bond lengths for a subset of bond lengths near the central Fe atom (Figs. 1012). For all states, we compare vHCISCF to M06-L and for (1,A), (3,A), and (3,C) and we compare to MN15 and M06-2X as well, since those calculations were already performed during our earlier analysis. In general, vHCISCF performs similarly to the ensemble of DFT results but does not typically provide the most accurate bond lengths.

FIG. 10.

Comparing errors in initial and final geometry compared to experimentally obtained bond lengths for state (1,C).74 

FIG. 10.

Comparing errors in initial and final geometry compared to experimentally obtained bond lengths for state (1,C).74 

Close modal
FIG. 11.

Comparing errors in initial and final geometry compared to experimentally obtained bond lengths for state (3,A).74 

FIG. 11.

Comparing errors in initial and final geometry compared to experimentally obtained bond lengths for state (3,A).74 

Close modal
FIG. 12.

Comparing errors in initial and final geometry compared to experimentally obtained bond lengths for state (3,C).74 

FIG. 12.

Comparing errors in initial and final geometry compared to experimentally obtained bond lengths for state (3,C).74 

Close modal
1.
B. O.
Roos
,
P. R.
Taylor
, and
P. E. M.
Siegbahn
, “
A complete active space SCF method (CASSCF) using a density matrix formulated super-CI approach
,”
Chem. Phys.
48
,
157
(
1980
).
2.
J.
Olsen
,
B. O.
Roos
,
P.
Jørgensen
, and
H. J. A.
Jensen
, “
Determinant based configuration interaction algorithms for complete and restricted configuration interaction spaces
,”
J. Chem. Phys.
89
,
2185
2192
(
1988
).
3.
P. Å.
Malmqvist
,
A.
Rendell
, and
B. O.
Roos
, “
The restricted active space self-consistent-field method, implemented with a split graph unitary group approach
,”
J. Phys. Chem.
94
,
5477
5482
(
1990
).
4.
K. D.
Vogiatzis
,
D.
Ma
,
J.
Olsen
,
L.
Gagliardi
, and
W. A.
De Jong
, “
Pushing configuration-interaction to the limit: Towards massively parallel MCSCF calculations
,”
J. Chem. Phys.
147
,
184111
(
2017
); arXiv:1707.04346.
5.
P.
Celani
and
H.-J.
Werner
, “
Multireference perturbation theory for large restricted and selected active space reference wave functions
,”
J. Chem. Phys.
112
,
5546
5557
(
2000
).
6.
T.
Fleig
,
J.
Olsen
, and
C. M.
Marian
, “
The generalized active space concept for the relativistic treatment of electron correlation. I. Kramers-restricted two-component configuration interaction
,”
J. Chem. Phys.
114
,
4775
4790
(
2001
).
7.
D.
Ma
,
G.
Li Manni
, and
L.
Gagliardi
, “
The generalized active space concept in multiconfigurational self-consistent field methods
,”
J. Chem. Phys.
135
,
044128
(
2011
).
8.
S. R.
White
, “
Density matrix formulation for quantum renormalization groups
,”
Phys. Rev. Lett.
69
,
2863
2866
(
1992
).
9.
S. R.
White
, “
Density-matrix algorithms for quantum renormalization groups
,”
Phys. Rev. B
48
,
10345
(
1993
).
10.
G.
Fano
,
F.
Ortolani
, and
L.
Ziosi
, “
The density matrix renormalization group method. Application to the PPP model of a cyclic polyene chain
,”
J. Chem. Phys.
108
,
9246
9252
(
1998
); arXiv:9803071 [cond-mat].
11.
S. R.
White
and
R. L.
Martin
, “
Ab initio quantum chemistry using the density matrix renormalization group
,”
J. Chem. Phys.
110
,
4127
(
1999
).
12.
U.
Schollwöck
, “
The density-matrix renormalization group
,”
Rev. Mod. Phys.
77
,
259
315
(
2005
); arXiv:0409292 [cond-mat].
13.
S.
Szalay
,
M.
Pfeffer
,
V.
Murg
,
G.
Barcza
,
F.
Verstraete
,
R.
Schneider
, and
Ö.
Legeza
, “
Tensor product methods and entanglement optimization for ab initio quantum chemistry
,”
Int. J. Quantum Chem.
115
,
1342
1391
(
2015
); arXiv:1412.5829.
14.
U.
Schollwöck
, “
The density-matrix renormalization group in the age of matrix product states
,”
Ann. Phys.
326
,
96
192
(
2011
); arXiv:1008.3477.
15.
S.
Daul
,
I.
Ciofini
,
C.
Daul
, and
S. R.
White
, “
Quantum chemistry using the density matrix renormalization group
,”
J. Chem. Phys.
115
,
6815
6821
(
2001
).
16.
G. K.-L.
Chan
and
M.
Head-Gordon
, “
Highly correlated calculations with a polynomial cost algorithm: A study of the density matrix renormalization group
,”
J. Chem. Phys.
116
,
4462
4476
(
2002
).
17.
G.
Moritz
and
M.
Reiher
, “
Construction of environment states in quantum-chemical density-matrix renormalization group calculations
,”
J. Chem. Phys.
124
,
034103
(
2006
).
18.
D.
Zgid
and
M.
Nooijen
, “
Obtaining the two-body density matrix in the density matrix renormalization group method
,”
J. Chem. Phys.
128
,
144115
(
2008
).
19.
H. G.
Luo
,
M. P.
Qin
, and
T.
Xiang
, “
Optimizing Hartree-Fock orbitals by the density-matrix renormalization group
,”
Phys. Rev. B
81
,
235129
(
2010
); arXiv:1002.1287.
20.
K. H.
Marti
and
M.
Reiher
, “
The density matrix renormalization group algorithm in quantum chemistry
,”
Z. Phys. Chem.
224
,
583
599
(
2010
).
21.
G. K.-L.
Chan
and
S.
Sharma
, “
The density matrix renormalization group in quantum chemistry
,”
Annu. Rev. Phys. Chem.
62
,
465
481
(
2011
).
22.
Y.
Kurashige
and
T.
Yanai
, “
Second-order perturbation theory with a density matrix renormalization group self-consistent field reference function: Theory and application to the study of chromium dimer
,”
J. Chem. Phys.
135
,
094104
(
2011
).
23.
S.
Sharma
and
G. K.-L.
Chan
, “
Spin-adapted density matrix renormalization group algorithms for quantum chemistry
,”
J. Chem. Phys.
136
,
124121
(
2012
).
24.
S.
Sharma
,
K.
Sivalingam
,
F.
Neese
, and
G. K.-L.
Chan
, “
Low-energy spectrum of iron–sulfur clusters directly from many-particle quantum mechanics
,”
Nat. Chem.
6
,
927
(
2014
); arXiv:1408.5080.
25.
Y.
Kurashige
,
G. K.-L.
Chan
, and
T.
Yanai
, “
Entangled quantum electronic wavefunctions of the Mn4CaO5 cluster in photosystem II
,”
Nat. Chem.
5
,
660
(
2013
).
26.
S.
Wouters
,
T.
Bogaerts
,
P.
Van Der Voort
,
V.
Van Speybroeck
, and
D.
Van Neck
, “
Communication: DMRG-SCF study of the singlet, triplet, and quintet states of oxo-Mn(Salen)
,”
J. Chem. Phys.
140
,
241103
(
2014
).
27.
S.
Keller
and
M.
Reiher
, “
Spin-adapted matrix product states and operators
,”
J. Chem. Phys.
144
,
134101
(
2016
).
28.
Y.
Kurashige
, “
Multireference electron correlation methods with density matrix renormalization group reference functions
,”
Mol. Phys.
112
,
1485
1494
(
2014
).
29.
T.
Yanai
,
Y.
Kurashige
,
W.
Mizukami
,
J.
Chalupský
,
T. N.
Lan
, and
M.
Saitow
, “
Density matrix renormalization group for ab initio calculations and associated dynamic correlation methods: A review of theory and applications
,”
Int. J. Quantum Chem.
115
,
283
299
(
2015
).
30.
J.
Ivanic
and
K.
Ruedenberg
, “
Identification of deadwood in configuration spaces through general direct configuration interaction
,”
Theor. Chem. Acc.
106
,
339
351
(
2001
).
31.
B.
Huron
,
J. P.
Malrieu
, and
P.
Rancurel
, “
Iterative perturbation calculations of ground and excited state energies from multiconfigurational zeroth-order wavefunctions
,”
J. Chem. Phys.
58
,
5745
5759
(
1973
).
32.
R. J.
Buenker
and
S. D.
Peyerimhoff
, “
Individualized configuration selection in CI calculations with subsequent energy extrapolation
,”
Theor. Chim. Acta
35
,
33
58
(
1974
).
33.
F. A.
Evangelista
,
J. P.
Daudey
, and
J. P.
Malrieu
, “
Convergence of an improved CIPSI algorithm
,”
Chem. Phys.
75
,
91
102
(
1983
).
34.
R. J.
Harrison
, “
Approximating full configuration interaction with selected configuration interaction and perturbation theory
,”
J. Chem. Phys.
94
,
5021
5031
(
1991
).
35.
M. M.
Steiner
,
W.
Wenzel
,
K. G.
Wilson
, and
J. W.
Wilkins
, “
The efficient treatment of higher excitations in CI calculations. A comparison of exact and approximate results
,”
Chem. Phys. Lett.
231
,
263
268
(
1994
).
36.
W.
Wenzel
,
M. M.
Steiner
, and
K. G.
Wilson
, “
Multireference basis-set reduction
,”
Int. J. Quantum Chem.
60
,
1325
1330
(
1996
).
37.
F.
Neese
, “
A spectroscopy oriented configuration interaction procedure
,”
J. Chem. Phys.
119
,
9428
9443
(
2003
).
38.
M. L.
Abrams
and
C. D.
Sherrill
, “
Important configurations in configuration interaction and coupled-cluster wave functions
,”
Chem. Phys. Lett.
412
,
121
124
(
2005
).
39.
L.
Bytautas
and
K.
Ruedenberg
, “
A priori identification of configurational deadwood
,”
Chem. Phys.
356
,
64
75
(
2009
).
40.
F. A.
Evangelista
, “
A driven similarity renormalization group approach to quantum many-body problems
,”
J. Chem. Phys.
141
,
054109
(
2014
).
41.
P. J.
Knowles
, “
Compressive sampling in configuration interaction wavefunctions
,”
Mol. Phys.
113
,
1655
1660
(
2015
).
42.
J. B.
Schriber
and
F. A.
Evangelista
, “
Communication: An adaptive configuration interaction approach for strongly correlated electrons with tunable accuracy
,”
J. Chem. Phys.
144
,
161106
(
2016
); arXiv:1603.08063.
43.
W.
Liu
and
M. R.
Hoffmann
, “
ICI: Iterative CI toward full CI
,”
J. Chem. Theory Comput.
12
,
1169
1178
(
2016
).
44.
M.
Caffarel
,
T.
Applencourt
,
E.
Giner
, and
A.
Scemama
, “
Using CIPSI nodes in diffusion Monte Carlo
,”
ACS Symp. Ser.
1234
,
15
46
(
2016
); arXiv:1607.06742.
45.
N. M.
Tubman
,
J.
Lee
,
T. Y.
Takeshita
,
M.
Head-Gordon
, and
K. B.
Whaley
, “
A deterministic alternative to the full configuration interaction quantum Monte Carlo method
,”
J. Chem. Phys.
145
,
044112
(
2016
); arXiv:1603.02686.
46.
Y.
Garniron
,
A.
Scemama
,
P.-F.
Loos
, and
M.
Caffarel
, “
Hybrid stochastic-deterministic calculation of the second-order perturbative contribution of multireference perturbation theory
,”
J. Chem. Phys.
147
,
034101
(
2017
).
47.
N. M.
Tubman
,
D. S.
Levine
,
D.
Hait
,
M.
Head-Gordon
, and
K. B.
Whaley
, “
An efficient deterministic perturbation theory for selected configuration interaction methods
,” arXiv:1808.02049 (
2018
).
48.
N. M.
Tubman
,
C. D.
Freeman
,
D. S.
Levine
,
D.
Hait
,
M.
Head-Gordon
, and
K. B.
Whaley
, “
Modern approaches to exact diagonalization and selected configuration interaction with the adaptive sampling CI method
,”
J. Chem. Theory Comput.
16
,
2139
2159
(
2020
); arXiv:1807.00821.
49.
D. S.
Levine
,
D.
Hait
,
N. M.
Tubman
,
S.
Lehtola
,
K. B.
Whaley
, and
M.
Head-Gordon
, “
CASSCF with extremely large active spaces using the adaptive sampling configuration interaction method
,”
J. Chem. Theory Comput.
16
,
2340
2354
(
2020
); arXiv:1912.08379.
50.
G. H.
Booth
,
A. J. W.
Thom
, and
A.
Alavi
, “
Fermion Monte Carlo without fixed nodes: A game of life, death, and annihilation in slater determinant space
,”
J. Chem. Phys.
131
,
054106
(
2009
).
51.
D.
Cleland
,
G. H.
Booth
, and
A.
Alavi
, “
Communications: Survival of the fittest: Accelerating convergence in full configuration-interaction quantum Monte Carlo
,”
J. Chem. Phys.
132
,
041103
(
2010
).
52.
F. R.
Petruzielo
,
A. A.
Holmes
,
H. J.
Changlani
,
M. P.
Nightingale
, and
C. J.
Umrigar
, “
Semistochastic projector Monte Carlo method
,”
Phys. Rev. Lett.
109
,
230201
(
2012
); arXiv:1207.6138v2.
53.
R. E.
Thomas
,
Q.
Sun
,
A.
Alavi
, and
G. H.
Booth
, “
Stochastic multiconfigurational self-consistent field theory
,”
J. Chem. Theory Comput.
11
,
5316
5325
(
2015
); arXiv:1510.03635.
54.
A. A.
Holmes
,
N. M.
Tubman
, and
C. J.
Umrigar
, “
Heat-bath configuration interaction: An efficient selected configuration interaction algorithm inspired by heat-bath sampling
,”
J. Chem. Theory Comput.
12
,
3674
3680
(
2016
).
55.
S.
Sharma
,
A. A.
Holmes
,
G.
Jeanmairet
,
A.
Alavi
, and
C. J.
Umrigar
, “
Semistochastic heat-bath configuration interaction method: Selected configuration interaction with semistochastic perturbation theory
,”
J. Chem. Theory Comput.
13
,
1595
1604
(
2017
); arXiv:1610.06660v2.
56.
J. E. T.
Smith
,
B.
Mussard
,
A. A.
Holmes
, and
S.
Sharma
, “
Cheap and near exact CASSCF with large active spaces
,”
J. Chem. Theory Comput.
13
,
5468
5478
(
2017
); arXiv:1708.07544.
57.
B. F. E.
Curchod
and
T. J.
Martínez
, “
Ab initio nonadiabatic quantum molecular dynamics
,”
Chem. Rev.
118
,
3305
3336
(
2018
).
58.
R.
Guareschi
and
C.
Filippi
, “
Ground- and excited-state geometry optimization of small organic molecules with quantum Monte Carlo
,”
J. Chem. Theory Comput.
9
,
5513
5525
(
2013
).
59.
F.
Liu
,
Y.
Kurashige
,
T.
Yanai
, and
K.
Morokuma
, “
Multireference ab initio density matrix renormalization group (DMRG)-CASSCF and DMRG-CASPT2 study on the photochromic ring opening of spiropyran
,”
J. Chem. Theory Comput.
9
,
4462
4469
(
2013
).
60.
W.
Hu
and
G. K.-L.
Chan
, “
Excited-state geometry optimization with the density matrix renormalization group, as applied to polyenes
,”
J. Chem. Theory Comput.
11
,
3000
3009
(
2015
); arXiv:1502.07731.
61.
E.
Maradzike
,
G.
Gidofalvi
,
J. M.
Turney
,
H. F.
Schaefer
, and
A. E.
DePrince
, “
Analytic energy gradients for variational two-electron reduced-density-matrix-driven complete active space self-consistent field theory
,”
J. Chem. Theory Comput.
13
,
4113
4122
(
2017
).
62.
Y.
Ma
,
S.
Knecht
, and
M.
Reiher
, “
Multiconfigurational effects in theoretical resonance Raman spectra
,”
ChemPhysChem
18
,
384
393
(
2017
).
63.
A. W.
Schlimgen
and
D. A.
Mazziotti
, “
Analytical gradients of variational reduced-density-matrix and wavefunction-based methods from an overlap-reweighted semidefinite program
,”
J. Chem. Phys.
149
,
164111
(
2018
).
64.
M.
Dash
,
S.
Moroni
,
A.
Scemama
, and
C.
Filippi
, “
Perturbatively selected configuration-interaction wave functions for efficient geometry optimization in quantum Monte Carlo
,”
J. Chem. Theory Comput.
14
,
4176
4182
(
2018
); arXiv:1804.09610.
65.
M.
Dash
,
J.
Feldt
,
S.
Moroni
,
A.
Scemama
, and
C.
Filippi
, “
Excited states with selected configuration interaction-quantum Monte Carlo: Chemically accurate excitation energies and geometries
,”
J. Chem. Theory Comput.
15
,
4896
4906
(
2019
).
66.
L.
Freitag
,
Y.
Ma
,
A.
Baiardi
,
S.
Knecht
, and
M.
Reiher
, “
Approximate analytical gradients and nonadiabatic couplings for the state-average density matrix renormalization group self-consistent-field method
,”
J. Chem. Theory Comput.
15
,
6724
6737
(
2019
); arXiv:1905.01558.
67.
J. W.
Park
, “
Second-order orbital optimization with large active space using adaptive sampling configuration interaction (ASCI) and its application to molecular geometry optimization
,”
J. Chem. Theory Comput.
17
,
1522
1534
(
2021
).
68.
J. W.
Park
, “
Near-exact CASSCF-level geometry optimization with a large active space using adaptive sampling configuration interaction self-consistent field corrected with second-order perturbation theory (ASCI-SCF-PT2)
,”
J. Chem. Theory Comput.
17
,
4092
4104
(
2021
).
69.
Q.
Sun
, “
Libcint: An efficient general integral library for Gaussian basis functions
,”
J. Comput. Chem.
36
,
1664
1671
(
2015
); arXiv:1412.0649.
70.
Q.
Sun
,
J.
Yang
, and
G. K.-L.
Chan
, “
A general second order complete active space self-consistent-field solver for large-scale systems
,”
Chem. Phys. Lett.
683
,
291
299
(
2017
); arXiv:1610.08394.
71.
Q.
Sun
,
T. C.
Berkelbach
,
N. S.
Blunt
,
G. H.
Booth
,
S.
Guo
,
Z.
Li
,
J.
Liu
,
J. D.
McClain
,
E. R.
Sayfutyarova
,
S.
Sharma
,
S.
Wouters
, and
G. K. L.
Chan
, “
PySCF: The python-based simulations of chemistry framework
,”
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
8
,
e1340
(
2018
); arXiv:1701.08223.
72.
Q.
Sun
,
X.
Zhang
,
S.
Banerjee
,
P.
Bao
,
M.
Barbry
,
N. S.
Blunt
,
N. A.
Bogdanov
,
G. H.
Booth
,
J.
Chen
,
Z.-H.
Cui
,
J. J.
Eriksen
,
Y.
Gao
,
S.
Guo
,
J.
Hermann
,
M. R.
Hermes
,
K.
Koh
,
P.
Koval
,
S.
Lehtola
,
Z.
Li
,
J.
Liu
,
N.
Mardirossian
,
J. D.
McClain
,
M.
Motta
,
B.
Mussard
,
H. Q.
Pham
,
A.
Pulkin
,
W.
Purwanto
,
P. J.
Robinson
,
E.
Ronca
,
E. R.
Sayfutyarova
,
M.
Scheurer
,
H. F.
Schurkus
,
J. E. T.
Smith
,
C.
Sun
,
S.-N.
Sun
,
S.
Upadhyay
,
L. K.
Wagner
,
X.
Wang
,
A.
White
,
J. D.
Whitfield
,
M. J.
Williamson
,
S.
Wouters
,
J.
Yang
,
J. M.
Yu
,
T.
Zhu
,
T. C.
Berkelbach
,
S.
Sharma
,
A. Y.
Sokolov
, and
G. K.-L.
Chan
, “
Recent developments in the PySCF program package
,”
J. Chem. Phys.
153
,
024109
(
2020
); arXiv:2002.12531.
73.
M. A.
Ortuño
and
C. J.
Cramer
, “
Multireference electronic structures of Fe-pyridine(diimine) complexes over multiple oxidation states
,”
J. Phys. Chem. A
121
,
5932
5939
(
2017
).
74.
S. C. E.
Stieber
,
C.
Milsmann
,
J. M.
Hoyt
,
Z. R.
Turner
,
K. D.
Finkelstein
,
K.
Wieghardt
,
S.
Debeer
, and
P. J.
Chirik
, “
Bis(imino)pyridine iron dinitrogen compounds revisited: Differences in electronic structure between four- and five-coordinate derivatives
,”
Inorg. Chem.
51
,
3770
3785
(
2012
).
75.
P. S.
Epstein
, “
The Stark effect from the point of view of Schroedinger’s quantum theory
,”
Phys. Rev.
28
,
695
710
(
1926
).
76.
R. K.
Nesbet
, “
Configuration interaction in orbital theories
,”
Proc. R. Soc. London, Ser. A
230
,
312
321
(
1955
).
77.
Y.
Garniron
,
T.
Applencourt
,
K.
Gasperich
,
A.
Benali
,
A.
Ferté
,
J.
Paquier
,
B.
Pradines
,
R.
Assaraf
,
P.
Reinhardt
,
J.
Toulouse
,
P.
Barbaresco
,
N.
Renon
,
G.
David
,
J.-P.
Malrieu
,
M.
Véril
,
M.
Caffarel
,
P.-F.
Loos
,
E.
Giner
, and
A.
Scemama
, “
Quantum package 2.0: An open-source determinant-driven suite of programs
,”
J. Chem. Theory Comput.
15
,
3591
3609
(
2019
); arXiv:1902.08154.
78.
J.
Li
,
M.
Otten
,
A. A.
Holmes
,
S.
Sharma
, and
C. J.
Umrigar
, “
Fast semistochastic heat-bath configuration interaction
,”
J. Chem. Phys.
149
,
214110
(
2018
); arXiv:1809.04600.
79.
Y.
Yamaguchi
and
H. F.
Schaefer
, “
Analytic derivative methods in molecular electronic structure theory: A new dimension to quantum chemistry and its applications to spectroscopy
,” in
Handbook of High-Resolution Spectroscopy
(
John Wiley & Sons
,
2011
), pp.
325
362
.
80.
N. C.
Handy
and
H. F.
Schaefer
, “
On the evaluation of analytic energy derivatives for correlated wave functions
,”
J. Chem. Phys.
81
,
5031
5033
(
1984
).
81.
M. J.
Frisch
,
G. W.
Trucks
,
H. B.
Schlegel
,
G. E.
Scuseria
,
M. A.
Robb
,
J. R.
Cheeseman
,
G.
Scalmani
,
V.
Barone
,
G. A.
Petersson
,
H.
Nakatsuji
,
X.
Li
,
M.
Caricato
,
A. V.
Marenich
,
J.
Bloino
,
B. G.
Janesko
,
R.
Gomperts
,
B.
Mennucci
,
H. P.
Hratchian
,
J. V.
Ortiz
,
A. F.
Izmaylov
,
J. L.
Sonnenberg
,
D.
Williams-Young
,
F.
Ding
,
F.
Lipparini
,
F.
Egidi
,
J.
Goings
,
B.
Peng
,
A.
Petrone
,
T.
Henderson
,
D.
Ranasinghe
,
V. G.
Zakrzewski
,
J.
Gao
,
N.
Rega
,
G.
Zheng
,
W.
Liang
,
M.
Hada
,
M.
Ehara
,
K.
Toyota
,
R.
Fukuda
,
J.
Hasegawa
,
M.
Ishida
,
T.
Nakajima
,
Y.
Honda
,
O.
Kitao
,
H.
Nakai
,
T.
Vreven
,
K.
Throssell
,
J. A. J.
Montgomery
,
J. E.
Peralta
,
F.
Ogliaro
,
M. J.
Bearpark
,
J. J.
Heyd
,
E. N.
Brothers
,
K. N.
Kudin
,
V. N.
Staroverov
,
T. A.
Keith
,
R.
Kobayashi
,
J.
Normand
,
K.
Raghavachari
,
A. P.
Rendell
,
J. C.
Burant
,
S. S.
Iyengar
,
J.
Tomasi
,
M.
Cossi
,
J. M.
Millam
,
M.
Klene
,
C.
Adamo
,
R.
Cammi
,
J. W.
Ochterski
,
R. L.
Martin
,
K.
Morokuma
,
O.
Farkas
,
J. B.
Foresman
, and
D. J.
Fox
, “Gaussian 16 Revision C.01 Release Notes,”
2016
.
82.
E.
Epifanovsky
,
A. T. B.
Gilbert
,
X.
Feng
,
J.
Lee
,
Y.
Mao
,
N.
Mardirossian
,
P.
Pokhilko
,
A. F.
White
,
M. P.
Coons
,
A. L.
Dempwolff
,
Z.
Gan
,
D.
Hait
,
P. R.
Horn
,
L. D.
Jacobson
,
I.
Kaliman
,
J.
Kussmann
,
A. W.
Lange
,
K. U.
Lao
,
D. S.
Levine
,
J.
Liu
,
S. C.
McKenzie
,
A. F.
Morrison
,
K. D.
Nanda
,
F.
Plasser
,
D. R.
Rehn
,
M. L.
Vidal
,
Z.-Q.
You
,
Y.
Zhu
,
B.
Alam
,
B. J.
Albrecht
,
A.
Aldossary
,
E.
Alguire
,
J. H.
Andersen
,
V.
Athavale
,
D.
Barton
,
K.
Begam
,
A.
Behn
,
N.
Bellonzi
,
Y. A.
Bernard
,
E. J.
Berquist
,
H. G. A.
Burton
,
A.
Carreras
,
K.
Carter-Fenk
,
R.
Chakraborty
,
A. D.
Chien
,
K. D.
Closser
,
V.
Cofer-Shabica
,
S.
Dasgupta
,
M.
De Wergifosse
,
J.
Deng
,
M.
Diedenhofen
,
H.
Do
,
S.
Ehlert
,
P.-T.
Fang
,
S.
Fatehi
,
Q.
Feng
,
T.
Friedhoff
,
J.
Gayvert
,
Q.
Ge
,
G.
Gidofalvi
,
M.
Goldey
,
J.
Gomes
,
C. E.
González-Espinoza
,
S.
Gulania
,
A. O.
Gunina
,
M. W. D.
Hanson-Heine
,
P. H. P.
Harbach
,
A.
Hauser
,
M. F.
Herbst
,
M.
Hernández Vera
,
M.
Hodecker
,
Z. C.
Holden
,
S.
Houck
,
X.
Huang
,
K.
Hui
,
B. C.
Huynh
,
M.
Ivanov
,
Á.
Jász
,
H.
Ji
,
H.
Jiang
,
B.
Kaduk
,
S.
Kähler
,
K.
Khistyaev
,
J.
Kim
,
G.
Kis
,
P.
Klunzinger
,
Z.
Koczor-Benda
,
J. H.
Koh
,
D.
Kosenkov
,
L.
Koulias
,
T.
Kowalczyk
,
C. M.
Krauter
,
K.
Kue
,
A.
Kunitsa
,
T.
Kus
,
I.
Ladjánszki
,
A.
Landau
,
K. V.
Lawler
,
D.
Lefrancois
,
S.
Lehtola
,
R. R.
Li
,
Y.-P.
Li
,
J.
Liang
,
M.
Liebenthal
,
H.-H.
Lin
,
Y.-S.
Lin
,
F.
Liu
,
K.-Y.
Liu
,
M.
Loipersberger
,
A.
Luenser
,
A.
Manjanath
,
P.
Manohar
,
E.
Mansoor
,
S. F.
Manzer
,
S.-P.
Mao
,
A. V.
Marenich
,
T.
Markovich
,
S.
Mason
,
S. A.
Maurer
,
P. F.
McLaughlin
,
M. F. S. J.
Menger
,
J.-M.
Mewes
,
S. A.
Mewes
,
P.
Morgante
,
J. W.
Mullinax
,
K. J.
Oosterbaan
,
G.
Paran
,
A. C.
Paul
,
S. K.
Paul
,
F.
Pavošević
,
Z.
Pei
,
S.
Prager
,
E. I.
Proynov
,
Á.
Rák
,
E.
Ramos-Cordoba
,
B.
Rana
,
A. E.
Rask
,
A.
Rettig
,
R. M.
Richard
,
F.
Rob
,
E.
Rossomme
,
T.
Scheele
,
M.
Scheurer
,
M.
Schneider
,
N.
Sergueev
,
S. M.
Sharada
,
W.
Skomorowski
,
D. W.
Small
,
C. J.
Stein
,
Y.-C.
Su
,
E. J.
Sundstrom
,
Z.
Tao
,
J.
Thirman
,
G. J.
Tornai
,
T.
Tsuchimochi
,
N. M.
Tubman
,
S. P.
Veccham
,
O.
Vydrov
,
J.
Wenzel
,
J.
Witte
,
A.
Yamada
,
K.
Yao
,
S.
Yeganeh
,
S. R.
Yost
,
A.
Zech
,
I. Y.
Zhang
,
X.
Zhang
,
Y.
Zhang
,
D.
Zuev
,
A.
Aspuru-Guzik
,
A. T.
Bell
,
N. A.
Besley
,
K. B.
Bravaya
,
B. R.
Brooks
,
D.
Casanova
,
J.-D.
Chai
,
S.
Coriani
,
C. J.
Cramer
,
G.
Cserey
,
A. E.
Deprince
,
R. A.
Distasio
,
A.
Dreuw
,
B. D.
Dunietz
,
T. R.
Furlani
,
W. A.
Goddard
,
S.
Hammes-Schiffer
,
T.
Head-Gordon
,
W. J.
Hehre
,
C.-P.
Hsu
,
T.-C.
Jagau
,
Y.
Jung
,
A.
Klamt
,
J.
Kong
,
D. S.
Lambrecht
,
W.
Liang
,
N. J.
Mayhall
,
C. W.
McCurdy
,
J. B.
Neaton
,
C.
Ochsenfeld
,
J. A.
Parkhill
,
R.
Peverati
,
V. A.
Rassolov
,
Y.
Shao
,
L. V.
Slipchenko
,
T.
Stauch
,
R. P.
Steele
,
J. E.
Subotnik
,
A. J. W.
Thom
,
A.
Tkatchenko
,
D. G.
Truhlar
,
T.
Van Voorhis
,
T. A.
Wesolowski
,
K. B.
Whaley
,
H. L.
Woodcock
,
P. M.
Zimmerman
,
S.
Faraji
,
P. M. W.
Gill
,
M.
Head-Gordon
,
J. M.
Herbert
, and
A. I.
Krylov
, “
Software for the frontiers of quantum chemistry: An overview of developments in the Q-Chem 5 package
,”
J. Chem. Phys.
155
,
084801
(
2021
).
83.
L.-P.
Wang
and
C.
Song
, “
Geometry optimization made simple with translation and rotation coordinates
,”
J. Chem. Phys.
144
,
214108
(
2016
).
84.
T. H.
Dunning
, “
Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen
,”
J. Chem. Phys.
90
,
1007
1023
(
1989
).
85.
J. P.
Perdew
,
M.
Ernzerhof
, and
K.
Burke
, “
Rationale for mixing exact exchange with density functional approximations
,”
J. Chem. Phys.
105
,
9982
9985
(
1996
).
86.
C.
Adamo
and
V.
Barone
, “
Toward reliable density functional methods without adjustable parameters: The PBE0 model
,”
J. Chem. Phys.
110
,
6158
6170
(
1999
).
87.
S.
Keller
,
K.
Boguslawski
,
T.
Janowski
,
M.
Reiher
, and
P.
Pulay
, “
Selection of active spaces for multiconfigurational wavefunctions
,”
J. Chem. Phys.
142
,
244104
(
2015
).
88.
J. M.
Bofill
and
P.
Pulay
, “
The unrestricted natural orbital-complete active space (UNO-CAS) method: An inexpensive alternative to the complete active space-self-consistent-field (CAS-SCF) method
,”
J. Chem. Inf. Model.
90
,
3637
(
1989
).
89.
P.
Pulay
, “
Convergence acceleration of iterative sequences. The case of SCF iteration
,”
Chem. Phys. Lett.
73
,
393
398
(
1980
).
90.
X.
Hu
and
W.
Yang
, “
Accelerating self-consistent field convergence with the augmented Roothaan-Hall energy function
,”
J. Chem. Phys.
132
,
054109
(
2010
).
91.
W.
Kabsch
, “
A discussion of the solution for the best rotation to relate two sets of vectors
,”
Acta Crystallogr., Sect. A: Found. Crystallogr.
32
,
922
923
(
1976
).
92.
J. C.
Kromann
, “RMSD,”
2020
.
93.
A. W.
Schlimgen
and
D. A.
Mazziotti
, “
Static and dynamic electron correlation in the ligand noninnocent oxidation of nickel dithiolates
,”
J. Phys. Chem. A
121
,
9377
9384
(
2017
).
94.
D. A.
Kreplin
,
P. J.
Knowles
, and
H.-J.
Werner
, “
Second-order MCSCF optimization revisited. I. Improved algorithms for fast and robust second-order CASSCF convergence
,”
J. Chem. Phys.
150
,
194106
(
2019
).
95.
Y.
Yao
and
C. J.
Umrigar
, “
Orbital optimization in selected configuration interaction methods
,”
J. Chem. Theory Comput.
17
,
4183
4194
(
2021
); arXiv:2104.02587.
96.
Y.
Zhao
and
D. G.
Truhlar
, “
A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions
,”
J. Chem. Phys.
125
,
194101
(
2006
).
97.
Y.
Zhao
and
D. G.
Truhlar
, “
The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other function
,”
Theor. Chem. Acc.
120
,
215
241
(
2008
).
98.
H. S.
Yu
,
X.
He
,
S. L.
Li
, and
D. G.
Truhlar
, “
MN15: A Kohn-Sham global-hybrid exchange-correlation density functional with broad accuracy for multi-reference and single-reference systems and noncovalent interactions
,”
Chem. Sci.
7
,
5032
5051
(
2016
).
99.
J. W.
Mullinax
,
E.
Epifanovsky
,
G.
Gidofalvi
, and
A. E.
DePrince
, “
Analytic energy gradients for variational two-electron reduced-density matrix methods within the density fitting approximation
,”
J. Chem. Theory Comput.
15
,
276
289
(
2019
).
100.
P. M.
Zimmerman
and
A. E.
Rask
, “
Evaluation of full valence correlation energies and gradients
,”
J. Chem. Phys.
150
,
244117
(
2019
).