Accurate multireference electronic structure calculations are important for constructing potential energy surfaces. Still, even in the case of low-scaling methods, their routine use is limited by the steep growth of the computational and storage costs as the active space grows. This is primarily due to the occurrence of three- and higher-body density matrices or, equivalently, their cumulants. This work examines the effect of various cumulant truncation schemes on the accuracy of the driven similarity renormalization group second-order multireference perturbation theory. We test four different levels of three-body reduced density cumulant truncations that set different classes of cumulant elements to zero. Our test cases include the singlet–triplet gap of CH2, the potential energy curves of the XΣg+1 and AΣu+3 states of N2, and the singlet–triplet splittings of oligoacenes. Our results show that both relative and absolute errors introduced by these cumulant truncations can be as small as 0.5 kcal mol−1 or less. At the same time, the amount of memory required is reduced from O(NA6) to O(NA5), where NA is the number of active orbitals. No additional regularization is needed to prevent the intruder state problem in the cumulant-truncated second-order driven similarity renormalization group multireference perturbation theory methods.

Affordable and accurate multireference (MR) electronic structure theories are indispensable for constructing potential energy surfaces. Because of their low computational cost, multireference methods that treat dynamical correlation effects via low-order perturbation theory are the most frequently applied.1–9 A major limitation is extending these approaches to treat many active orbitals. For example, the ideal active space for large conjugated hydrocarbons (e.g., acenes) would include all π orbitals, which becomes a bottleneck for systems with more than 10–12 carbon atoms.10,11 Additionally, computations on polynuclear transition metal compounds, which require at least one d shell in the active space for each metal atom, can quickly become prohibitively expensive for more than three metal centers.12,13 A common strategy to reduce the cost of multireference perturbation theories (MRPTs) is to express the perturber states using partially or fully internally contracted configurations. This approach reduces the scaling of a method with respect to the number of active orbitals (NA) from exponential to polynomial; however, it still requires calculating and storing the three- and four-body reduced density matrices (RDMs) of the reference wavefunction, whose storage scales as O(NA6) and O(NA8), respectively. This reduction in computational and storage costs has enabled MRPT computations with as many as 20 active orbitals. Extending multireference computations beyond this limit requires avoiding the costs introduced by the 3- and 4-RDMs.

Strategies for reducing the cost associated with high-order RDMs fall into two main categories: (1) modifying the existing algorithms to eliminate the bottlenecks and (2) approximating the RDMs. The first category encompasses methods such as second-order complete-active-space perturbation theory (CASPT2)1,2 in combination with the density matrix renormalization group (DMRG-CASPT2) as proposed by Kurashige and Yanai14 and later optimized by Wouters and co-workers.15 These works take advantage of the matrix product structure of DMRG to reduce the cost of terms involving the 4-RDM down to that of the 3-RDM terms. Anderson, Shiozaki, and Booth16 reported CASPT2 and strongly contracted (SC) n-electron valence state perturbation theory (NEVPT2)5,7,8 calculations that avoid storing the 4-RDM by combining these methods with full configuration interaction quantum Monte Carlo.17 An alternative formulation of NEVPT2 was proposed by Sharma and Chan, in which the Hylleraas functional is minimized in the space of wavefunctions described by matrix product states.18 While this approach does not require higher-body RDMs, its efficiency is contingent on the representation of the first-order correction of the wavefunction as a matrix product state with a low bond dimension. Chatterjee and Sokolov proposed an intermediate approach to partially contracted NEVPT2.19 This approach avoids the evaluation of the 4-RDM without introducing any approximation and reduces the scaling to that of computing the 3-RDM. Some NEVPT2-based approaches entirely avoid computing the 3-RDM, such as stochastic SC-NEVPT220,21 and the imaginary-time uncontracted NEVPT2 method.22,23 Other approaches have been reported to reduce the cost of CASPT2 and NEVPT2 computations with large active spaces.24–26 Nevertheless, it is important to note that not all multireference approaches require high-order RDMs. For instance, the second-order driven similarity renormalization group MRPT (DSRG-MRPT2) requires only up to the three-body RDM.27 More recent developments like the multireference adiabatic connection formalism28–30 and multiconfigurational pair density functional theory31,32 require only the 2-RDM.

The second category of methods that can reduce the scaling with respect to the number of active orbitals uses approximate RDMs. RDMs can be expressed in terms of connected quantities (reduced density cumulants) of equal and lower rank.33–38 If RDMs are replaced by their cumulant expansion, it is possible to truncate the rank of the cumulants while preserving the correct size extensivity properties of the full theory. Cumulant approximation has been used in the contracted Schrödinger equation method39–42 and was later extended to the spin-free case.43–45 Zgid et al. investigated the performance of the spin-free cumulant approximation within SC-NEVPT2.46 They explored the possibility of approximating the 3- and 4-RDMs in two ways: (1) cumulant approximated SC-NEVPT2 (cu-SC-NEVPT2), and (2) cumulant with diagonals approximated SC-NEVPT2 (cud-SC-NEVPT2). Unfortunately, these two approximations introduce “false intruders.”46 Recently, Guo et al.47,48 proposed a fully internally contracted NEVPT2 and revisited cumulant approximations therein. They observed that the cumulant approximation could introduce negative eigenvalues in the Koopmans matrix. Guo et al. further developed a more stable variant named full rank NEVPT2 (FR-NEVPT2),48 in which rank-reduction is avoided. FR-NEVPT2 was found to be more stable than NEVPT2 for incomplete active space reference wavefunctions, but new contractions involving the 5-RDM appear [O(NA10) elements]. Unfortunately, cumulant approximations still result in inaccurate Koopmans matrix eigenvalues and introduce “false intruders” within FR-NEVPT2.48 Note that the “false intruders” encountered in NEVPT2 arise due to the neglect of high-order RDMs in the energy denominators46 and unlike classical intruders49–51 do not reflect real near-degeneracies of the perturber and reference states. Indeed, the use of Dyall’s modified zeroth-order Hamiltonian52 in NEVPT2 is one of the reasons for the practical absence of intruder states in this method. One solution to the “false intruder” problem is to shift the energy denominators; Zgid et al. found that a large imaginary level shift (1.2 a.u.) is often required to obtain smooth curves,46 leading to less accurate and, at times, even qualitatively wrong results, as observed in other contexts.53,54 Various post-DMRG methods with approximated cumulants have been implemented. Saitow et al. proposed a DMRG-based multireference configuration interaction scheme with a cumulant approximated 4-RDM.55,56 DMRG-CASPT2 combined with cumulant approximations has also been developed and extensively studied.57–59 

In this work, we explore cumulant truncation in the DSRG-MRPT2 formalism to reduce the original scaling with respect to the number of active orbitals [O(NA6)] and enable computations on large active spaces. Since truncation of the cumulants affects only the numerators entering the DSRG-MRPT2 amplitude and energy expressions, we expect that cumulant approximations would not introduce “false intruders” in DSRG-MRPT2. We test the two cumulant approximations considered by Zgid et al.46 and propose two new schemes that retain a larger portion of the cumulants and scale as O(NA4) and O(NA5). We study the energy errors these cumulant truncation schemes introduce on molecular properties, dissociation energies, and spin splittings. This paper proceeds as follows: In Sec. II A, we give an overview of the DSRG-MRPT2 method. In Sec. II B, we define four different levels of cumulant truncations and evaluate the potential for the appearance of “false intruders.” In Sec. II C, we examine the dependence of the energy of the cumulant truncation with respect to unitary transformations of the orbital basis. In Secs. III and IV, three different systems are used to examine the accuracy of cumulant-truncated DSRG-MRPT2. Finally, in Sec. V we summarize our results and discuss possible future research directions.

In this section, we summarize the MR-DSRG method and its second-order approximation. Starting with an MCSCF reference state (Ψ), the molecular spin orbitals {ϕp} are partitioned into core (C), active (A), and virtual (V) subsets of sizes NC, NA, and NV, respectively. We use the indices m, n to denote core spin orbitals, u, v, w, x, y, z to denote active spin orbitals, and e, f to denote virtual spin orbitals. For convenience, we also define hole (H) and particle (P) spin orbital subsets as H = CA and P = AV of dimensions NH = NC + NA and NP = NA + NV, respectively. The labels i, j, k, l indicate hole spin orbitals and a, b, c, d indicate particle spin orbitals. For general spin orbitals (hole or particle), we use the labels p, q, r, s.

The reference state Ψ with respect to which all operators are normal ordered36,44,60–63 is defined as a linear combination of Slater determinants (Φμ) with corresponding coefficients (cμ),
(1)
From Ψ, we can obtain the one-, two-, and three-body reduced density matrices (γux, γuvxy, and γuvwxyz), defined as
(2)
(3)
(4)
as well as the corresponding two- and three-body reduced density cumulants (λuvxy, λuvwxyz), defined in terms of the antisymmetric (wedge) product (∧) as
(5)
(6)
In MR-DSRG, the Hamiltonian Ĥ is gradually block-diagonalized by a continuous unitary (canonical) transformation that is controlled by a time-like flow parameter, s, defined in the range [0, ∞) as
(7)
The operator Â(s) is anti-Hermitian and is typically truncated to a given substitution level n as
(8)
Each k-body component of Â(s) [Âk(s)] is defined in terms of an electron-substitution operator [T̂k(s)],
(9)
where,
(10)
Note that the sum in Eq. (10) excludes terms with all indices belonging to the active subset, as these can be incorporated by changing the coefficients of the reference wavefunction. The MR-DSRG energy and amplitude equations are
(11)
(12)
where [H̄(s)]N is the nondiagonal part of the transformed Hamiltonian (containing the components that will be suppressed)64,65 and R̂(s) is the source operator containing contributions from normal-ordered many-body operators of different ranks
(13)
The source operator is parameterized and determines how the amplitudes depend on s. Note that the parameterization of R̂(s) is not unique, and we adopt the same one used in our previous work, which is determined from a perturbative analysis of the single-reference SRG equations.66 Our previous work27 found that for values of s[0.1,1.0]Eh2, the DSRG-MRPT2 approach achieves the best agreement with full configuration interaction results.
The DSRG-MRPT2 approach begins by partitioning the normal-ordered Hamiltonian into a zeroth-order term Ĥ(0) plus a first-order perturbation Ĥ(1). The zeroth-order Hamiltonian, Ĥ(0)=E0+F̂(0), includes the reference energy (E0) and a one-body operator [F̂(0)], where F̂(0) contains the diagonal blocks of the generalized Fock matrix (fpq=hpq+ijvpiqjγji),
(14)
The quantities hpq=ϕp|ĥ|ϕq and vpqrs=ϕpϕqϕrϕs are respectively one-electron and antisymmetrized two-electron integrals. In DSRG-MRPT2, we work in the basis of semicanonical orbitals, defined in such a way that the core, active, and virtual blocks of the generalized Fock matrix are diagonal. The DSRG-MRPT2 energy is not defined for orbital bases that are not semicanonical; however, a computational procedure can be formulated to compute the energy in a transformed basis.67 In our current implementation, integrals are automatically rotated to the semicanonical basis, even if the input integrals are not from a semicanonical basis. The DSRG-MRPT2 energy equations can be obtained by performing an order-by-order expansion of the MR-DSRG equations. The first nonzero energy correction appears at the second order and is given by
(15)
where we introduced a modified first-order effective Hamiltonian, H̃(1)(s)=Ĥ(1)(s)+R̂(1)(s). Expanding the DSRG-MRPT2 amplitude equations in the same way leads to the following expressions for the first-order amplitudes:
(16)
(17)
In Eqs. (16) and (17), the quantities appearing in the denominators are the orbital energies εp=fpp. After solving for the first-order amplitudes, the second-order energy E(2)(s) can be obtained using an efficient non-iterative procedure that requires up to the three-body reduced density cumulants. The various terms contributing to the DSRG-MRPT2 second-order energy are reported in Ref. 27.
We now describe the truncation schemes used in this work. The three-body reduced density cumulants enter the DSRG-MRPT2 method only in the energy expression via two tensor contractions (in the following, we use Einstein notation, assuming repeated upper and lower indices to be summed over),
(18)
where the effective two-body interaction ṽabij,(1)(s) is defined as
(19)
The most drastic truncation scheme (labeled “λ2”) neglects the entire contribution of λ3 to the energy, resulting in an overall memory scaling of O(NA4). At the next level, the diagonal approximation is the first scheme that retains elements of λ3 and is defined as
(20)
The λ3[d] and λ2 schemes have the same scaling since the former introduces only O(NA3) additional elements. The λ3[1] approximation augments λ3[d] with the elements of λ3 that contain up to one mismatched pair of indices. This scheme also has the same scaling as the λ2 scheme and is defined by
(21)
Finally, the λ3[2] scheme augments λ3[1] with the elements of λ3 that contain up to two mismatched pairs of indices. This scheme reduces the scaling of the parent method from O(NA6) to O(NA5) and is defined as
(22)
These truncation schemes are listed in Table I.
TABLE I.

Cumulant truncation schemes. All indices represent spin orbitals.

Truncation schemeAdditional elements preservedMemory scaling
λ2 None O(NA4) 
λ3[d] λuvwuvw, λuvwvwu, λuvwwuv, … O(NA4) 
λ3[1] λuvwxvw, λuvwvwx, λuvwwxv, … O(NA4) 
λ3[2] λuvwxyw, λuvwywx, λuvwwxy, … O(NA5) 
Truncation schemeAdditional elements preservedMemory scaling
λ2 None O(NA4) 
λ3[d] λuvwuvw, λuvwvwu, λuvwwuv, … O(NA4) 
λ3[1] λuvwxvw, λuvwvwx, λuvwwxv, … O(NA4) 
λ3[2] λuvwxyw, λuvwywx, λuvwwxy, … O(NA5) 
Next, we discuss whether or not “false intruders” can arise from these cumulant truncations. The untruncated DSRG-MRPT2 approach can avoid the intruder state problem since the amplitude equations do not diverge when one or more denominators approach zero. For example, as |Δabij|0, the corresponding first-order double amplitude [tabij,(1)(s)] can be expressed as
(23)
and similarly for the singles amplitudes. Equation (23) shows that for finite values of s, the amplitudes are continuous and well-behaved near a potential intruder state. Since the DSRG-MRPT2 energy equation is a sum of tensor contractions of amplitudes, effective integrals, and cumulants, all of which are bound, the DSRG-MRPT2 energy is also a finite and continuous quantity. Since the three-body reduced density cumulant only appear in the energy expression as multiplicative quantity [see Eqs. (17) and (18)], full or partial truncation will not introduce “false intruders.”
This section discusses formal results regarding the invariance of the cumulant-truncated DSRG-MRPT2 energy under unitary transformations of the orbital basis. Following a previous study,67 we only consider unitary transformations U that leave the energy of the reference state unchanged. Transformations that satisfy this constraint can be constructed as the direct sum of unitary matrices for core (UC), active (UA), and virtual (UV) orbital rotations:
(24)

Since the DSRG-MRPT2 energy is always computed in a semicanonical basis, which uniquely defines orbitals up to an energy degeneracy, we focus on the degenerate invariance property of MR-DSRG methods.67 This form of invariance implies that the energy remains unaffected under unitary rotations that exclusively mix degenerate orbitals within one of the three orbital subspaces. This is a departure from the traditional concept of invariance, which encompasses any unitary transformation compliant with Eq. (24). The source operator utilized in all MR-DSRG methods assumes a semicanonical orbital basis and is not invariant under general unitary transformations. However, since the source operator does not change form when two degenerate orbitals are mixed, it is possible to formulate a DSRG-MRPT2 procedure that, given any orbital basis, can reproduce the results in the semicanonical basis.67,68 Degenerate invariance proves critical as degenerate semicanonical orbitals lack a unique definition, ensuring the DSRG-MRPT2 method consistently provides a unique solution.

When considering the degenerate invariance of the cumulant-truncated DSRG-MRPT2 methods, it is necessary to examine the consequence of cumulant truncation. In general, a unitary transformation of the form given in (24) does not preserve the structure of a cumulant in which certain blocks are set to zero. Consequently, the λ3[d]-, λ3[1]-, and λ3[2]-DSRG-MRPT2 methods inherently lack degenerate invariance. The exception is represented by the λ2-DSRG-MRPT2 method, which retains degenerate invariance because it omits all three-body cumulant terms. We have tested the degenerate invariance of the cumulant-truncated DSRG-MRPT2 on two systems. The first one is the lowest singlet state of N2 in the dissociation limit, where we use a full valence active space (ten electrons in eight orbitals) and test invariance by localizing the 2s and 2p orbitals. For N2, we find that λ3 is zero in the dissociation limit, and consequently, all truncated schemes yield the same energy as the DSRG-MRPT2 and are invariant with respect to unitary transformations of degenerate orbitals (tested by localizing the symmetry-adapted orbitals via the Pipek–Mezey procedure69). The behavior observed for N2 is in line with a previous report by Hanauer and Köhn,70 which noted that the odd-rank cumulants of N2 vanish in the dissociation limit. They explained the phenomenon with their statistical interpretation of cumulants: since all active orbitals are degenerate, the electron distribution is symmetric, resulting in all odd cumulants being equal to zero.70 The second system we examine is a point along the Be + H2 insertion reaction71 where the HOMO and LUMO become degenerate. Computations on this system used the cc-pVDZ basis set72,73 and correlated all electrons. For Be + H2, at the point of orbital degeneracy (ϵHOMO = ϵLUMO) obtained using a minimal active space (two electrons in two orbitals) also gives λ3 = 0, and consequently, the energy is degenerate invariant for all cumulant-truncated DSRG-MRPT2 schemes. If we expand the active space to include the LUMO + 1 and LUMO + 2 orbitals, the three-body cumulant is no longer zero at the point of HOMO–LUMO orbital degeneracy. The impact of the various truncation schemes evaluated by computing the energy difference for the symmetry-adapted and Pipek–Mezey localized orbitals is shown in Table II. Both the DSRG-MRPT2 and λ2-DSRG-MRPT2 are degenerate invariant (in agreement with our analysis), while the other truncated schemes show energy differences of various degrees. The largest impact of the lack of degenerate invariance is found for the λ3[d]-approximation (55 μEh), a large portion of the three-body cumulant contribution to the correlation energy (75 μEh). Including more elements of the three-body cumulant further reduces the energy difference between the symmetry-adapted and localized results by 8.2 (λ3[1]-) and −0.3 (λ3[2]-) μEh.

TABLE II.

Absolute energies for BeH2 with delocalized or localized semicanonical orbital basis. A CAS(2,4) model space and the cc-pVDZ basis set were used for all computations. In this test, the beryllium atom is positioned at the origin of the Cartesian coordinates, while the two hydrogen atoms have coordinates in Å (2.259 98, ±1.500 409, 0).

MethodSymmetry-adapted (Eh)Pipek–Mezey (Eh)
DSRG-MRPT2 −15.588 550 6 −15.588 550 6 
λ2-DSRG-MRPT2 −15.588 475 4 −15.588 475 4 
λ3[d]-DSRG-MRPT2 −15.588 530 7 −15.588 475 9 
λ3[1]-DSRG-MRPT2 −15.588 544 1 −15.588 535 9 
λ3[2]-DSRG-MRPT2 −15.588 550 0 −15.588 550 3 
MethodSymmetry-adapted (Eh)Pipek–Mezey (Eh)
DSRG-MRPT2 −15.588 550 6 −15.588 550 6 
λ2-DSRG-MRPT2 −15.588 475 4 −15.588 475 4 
λ3[d]-DSRG-MRPT2 −15.588 530 7 −15.588 475 9 
λ3[1]-DSRG-MRPT2 −15.588 544 1 −15.588 535 9 
λ3[2]-DSRG-MRPT2 −15.588 550 0 −15.588 550 3 

As part of this work, we have developed pilot implementations of the λ2-, λ3[d]-, λ3[1]-, and λ3[2]-DSRG-MRPT2 truncation schemes using Forte,74 an open-source plugin for the Psi4 ab initio quantum chemistry package.75 To thoroughly study the accuracy of the cumulant-truncated DSRG-MRPT2 methods, we applied them to compute properties of several systems: the singlet–triplet gap of CH2 (A11B13), the potential energy curves of N2 (both the XΣg+1 state and AΣu+3 state), and the adiabatic singlet–triplet splitting of n-acenes (n = 2–4). The CH2 and N2 calculations used the flow parameter value s=1Eh2, following past work.27 To be consistent with our previous paper,76 we used the value s=0.5Eh2 for the n-acenes calculations.

For CH2, state-specific CAS self-consistent field (CASSCF) computations were performed to obtain the optimized molecular orbitals and reference state used in DSRG-MRPT2. The geometry and the basis set (double zeta with Cartesian H 2p polarization function and C 3d polarization function) for the CH2 calculations were taken from Ref. 77. The singlet and triplet calculations used different C 3d polarization functions.77 The CASSCF active space was taken from Ref. 46 and consisted of the 2a1, 3a1, 4a1, 1b1, 2b1, and 1b2 orbitals in C2v symmetry. The 1a1 core orbital was treated as a restricted doubly occupied orbital in the CASSCF procedure (optimized but with fixed occupation) and frozen (i.e., not included) in the correlated computations.

For N2, we used the cc-pVQZ basis.72,78 Optimized molecular orbitals were obtained via state-specific CASSCF, including all 2s and 2p orbitals in the active space (ten electrons and eight orbitals). The 1s core orbitals were treated as a restricted doubly occupied orbitals and, for consistency with previous work, were included in the correlated computations.46 

For the n-acenes calculations, we combined the DSRG-MRPT2 with a selected configuration interaction reference state. Specifically, we used the adaptive CI (ACI) method,76,79,80 a variant of the CIPSI method.81 The geometries of singlet and triplet n-acenes and the corresponding zero-point vibrational energy (ZPVE) corrections were taken from Ref. 76. Restricted Hartree–Fock (RHF) and restricted open-shell HF (ROHF) computations were performed first for the singlet and triplet states, respectively, followed by ACI computations.76,79,80 The ACI computations of the reference used a prescreening threshold of τV = 10−12 Eh (see Ref. 80). Our computations used the cc-pVXZ, (X = D, T, Q) basis sets72,73 with all 1s-like orbitals on C atoms kept frozen in the following DSRG-MRPT2 computations. Density-fitted integrals were used to accelerate our computations.82 We used the corresponding cc-pVXZ-JKFIT auxiliary basis83 for RHF/ROHF computations, and the cc-PVXZ-RI auxiliary basis sets84,85 for the DSRG-MRPT2 computations of the correlation energy. We computed the singlet–triplet splittings in two ways: first, with an unrelaxed approach, whereby the CAS configuration interaction (CASCI) reference wavefunction was kept frozen and the second-order energy correction was computed from Eq. (15). Alternatively, we reoptimized the coefficients cμ in Eq. (1) by diagonalizing H̄ within the set of reference determinants {Φμ},
(25)
and obtained the relaxed energy.76 The reference relaxation effect is important for the coupling of static and dynamical correlation. In this work, all relaxed energies came from the partially relaxed approach, in which the relaxation procedure was performed once for each computation. A second-order Epstein–Nesbet perturbation theory correction86–91 was used in ACI to account for unselected determinants in the active space to improve unrelaxed and relaxed energies.
In our first test case, we consider the adiabatic singlet–triplet gap of CH2. Due to its small magnitude, any imbalance in the treatment of correlation between the singlet and triplet states can lead to inaccurate predictions of the gap, and therefore, it is a stringent test for the cumulant truncations. This system has been used in previous benchmarks,92–96 including the study on cumulant approximated SC-NEVPT2 (cu-SC-NEVPT2)46 and cu-SC-NEVPT2 with diagonals (cud-SC-NEVPT2) methods.46 To quantify the error introduced by cumulant truncation, we define the cumulant error ΔEcu as
(26)
In Eq. (26), the quantities Eapprox. and EMRPT2 can be absolute energies, singlet–triplet gaps, or dissociation energies provided by cumulant-truncated and non-truncated MRPT2 methods (DSRG-MRPT2 and SC-NEVPT2) in this work. We further define the full CI (FCI) error, denoted as ΔEFCI, as the measure of intrinsic error. It quantifies the absolute energy difference between the results obtained from a calculation method and FCI.

We first look at the accuracy of various methods. The predicted adiabatic singlet–triplet gaps of CH2 are shown in Table III. As can be seen in Table III, the FCI error in DSRG-MRPT2 is larger (ΔEFCI = 4.05 kcal mol−1) than SC-NEVPT2 (ΔEFCI = 1.72 kcal mol−1) and CASPT2 (ΔEFCI = 3.66 kcal mol−1). CASSCF gives the smallest error (ΔEFCI = 0.87 kcal mol−1) among all the computations, likely due to error cancellation since it gives the largest absolute energy error for both singlet and triplet states. Next, we examine the error caused by cumulant truncation. Interestingly, the performance of SC-NEVPT2 and DSRG-MRPT2 differs significantly when applying the crudest approximation: The singlet–triplet gap cumulant error for cu-SC-NEVPT2 (ΔEcu = 10.81 kcal mol−1) is considerably larger than that for λ2-DSRG-MRPT2 (ΔEcu = 1.63 kcal mol−1), which is smaller than the intrinsic DSRG-MRPT2 error (ΔEcu = 4.05 kcal mol−1). Higher-level cumulant truncation methods do not show this significant difference, e.g., the cumulant error for both cud-SC-NEVPT2 (ΔEcu = 0.63 kcal mol−1) and λ3[d]-DSRG-MRPT2 (ΔEcu = 1.20 kcal mol−1). The smaller sensitivity with respect to cumulant truncation of the DSRG-MRPT2, when compared to SC-NEVPT2, is consistent with the different ways the density cumulants enter into these methods, with the former depending only linearly with respect to the cumulants.

TABLE III.

Adiabatic singlet–triplet gap of CH2 computed with different methods (see text for basis and geometry).

MethodSinglet energy (Eh)Triplet energy (Eh)Gap (kcal mol−1)
FCIa −39.027 183 −39.046 229 11.95 
CASSCF(6, 6)b −38.945 529 −38.965 954 12.82 
CASPT2a −39.012 184 −39.037 061 15.61 
SC-NEVPT2a −39.006 707 −39.028 498 13.67 
cu-SC-NEVPT2(0.1)a −39.003 499 −39.042 508 24.48 
cud-SC-NEVPT2(0.1)a −39.006 470 −39.029 253 14.30 
λ2-DSRG-MRPT2 −39.007 091 −39.035 190 17.63 
λ3[d]-DSRG-MRPT2 −39.007 993 −39.035 405 17.20 
λ3[1]-DSRG-MRPT2 −39.009 055 −39.035 445 16.56 
λ3[2]-DSRG-MRPT2 −39.009 916 −39.035 714 16.19 
DSRG-MRPT2 −39.009 963 −39.035 460 16.00 
MethodSinglet energy (Eh)Triplet energy (Eh)Gap (kcal mol−1)
FCIa −39.027 183 −39.046 229 11.95 
CASSCF(6, 6)b −38.945 529 −38.965 954 12.82 
CASPT2a −39.012 184 −39.037 061 15.61 
SC-NEVPT2a −39.006 707 −39.028 498 13.67 
cu-SC-NEVPT2(0.1)a −39.003 499 −39.042 508 24.48 
cud-SC-NEVPT2(0.1)a −39.006 470 −39.029 253 14.30 
λ2-DSRG-MRPT2 −39.007 091 −39.035 190 17.63 
λ3[d]-DSRG-MRPT2 −39.007 993 −39.035 405 17.20 
λ3[1]-DSRG-MRPT2 −39.009 055 −39.035 445 16.56 
λ3[2]-DSRG-MRPT2 −39.009 916 −39.035 714 16.19 
DSRG-MRPT2 −39.009 963 −39.035 460 16.00 
a

Results taken from Ref. 46.

b

Use conventional notation of CASSCF(e, o), for e active electrons and o active orbitals.

To examine in more detail the connection between cumulant structure and the accuracy of the approximate methods, in Fig. 1, we plot the nonredundant tensor elements of the three-body reduced density cumulant (λuvwxyz) retained in the parent and truncated DSRG-MRPT2 schemes. This plot represents λuvwxyz as a matrix indexed by the upper and lower tensor indices (compounded) for the ααα and ααβ spin cases for the singlet and triplet states. The βββ spin case is the same as the ααα case, and the αββ case is equivalent to the ααβ case, with elements ordered differently. See Figs. S1 and S2 in the supplementary material for the βββ and the αββ cases. λ2-DSRG-MRPT2 is not included in the figure since all tensor elements are zero. These plots show that the largest elements are included in λ3[2]-DSRG-MRPT2. Still, many are not included in the simpler truncation schemes, explaining the small cumulant error introduced by λ3[2]-DSRG-MRPT2 for the CH2 singlet–triplet gap. A quantitative argument can be made by analyzing the total number of significant elements of the three-body reduced density cumulant (|λuvwxyz|>104). In Table IV, we report this number for the full and truncated DSRG-MRPT2. In the full DSRG-MRPT2, only 3068 elements of λuvwxyz are significant (less than 2% of the total number). The λ3[2]-DSRG-MRPT2 can retain a large fraction of these important elements (2452) even though it requires storing fewer elements [69% of the total, with O(NA5) storage cost]. This truncation can therefore lead to significant storage and computational cost savings for large active spaces, bringing a potential reduction in cost compared to conventional computations of O(NA).

FIG. 1.

The nonredundant tensor elements of three-body reduced density cumulant for singlet (first and third rows) and triplet (second and fourth rows) states CH2 retained in the parent and truncated DSRG-MRPT2 schemes for the ααα (the first two rows) and ααβ (the last two rows) spin cases. Row vectors and column vectors are sorted by symmetry in the order: A1, A2, B1, B2. Tensor elements with absolute value less than 10−7 are not shown in these plots.

FIG. 1.

The nonredundant tensor elements of three-body reduced density cumulant for singlet (first and third rows) and triplet (second and fourth rows) states CH2 retained in the parent and truncated DSRG-MRPT2 schemes for the ααα (the first two rows) and ααβ (the last two rows) spin cases. Row vectors and column vectors are sorted by symmetry in the order: A1, A2, B1, B2. Tensor elements with absolute value less than 10−7 are not shown in these plots.

Close modal
TABLE IV.

The number of significant (|λuvwxyz|>104) and total tensor elements of λ3 entering in the CH2 DSRG-MRPT2 computations using different cumulant truncation schemes. The numbers reported account for the four spin cases (ααα, βββ, ααβ, αββ) and do not take into account symmetries of the three-body reduced density cumulant.

MethodSingletTriplet
SignificantTotalaSignificantTotala
λ3[d]-DSRG-MRPT2 244 2 784 652 2 784 
λ3[1]-DSRG-MRPT2 820 38 604 1724 38 604 
λ3[2]-DSRG-MRPT2 2452 129 324 4596 129 324 
DSRG-MRPT2 3068 186 624 5716 186 624 
MethodSingletTriplet
SignificantTotalaSignificantTotala
λ3[d]-DSRG-MRPT2 244 2 784 652 2 784 
λ3[1]-DSRG-MRPT2 820 38 604 1724 38 604 
λ3[2]-DSRG-MRPT2 2452 129 324 4596 129 324 
DSRG-MRPT2 3068 186 624 5716 186 624 
a

Number of total tensor elements for a given truncation scheme.

To evaluate the accuracy of the truncated DSRG-MRPT2 methods in describing potential energy curves, we computed the lowest singlet (XΣg+1) and the lowest triplet (AΣu+3) states of N2 as a function of bond length. In cumulant-truncated SC-NEVPT2,46 “false intruders” arise at many geometries along the potential energy curve for both bare (i.e., without level shift) cu-SC-NEVPT2 and bare cud-SC-NEVPT2. A large imaginary level shift (1.2 a.u.) is required to correct those singularities arising from the cumulant approximation. Therefore, this stricter test case is useful to assess if cumulant truncations introduce the intruder state problem in DSRG-MRPT2 and how they affect the potential energy curve.

Potential energy curves and the corresponding cumulant errors for the XΣg+1 and AΣu+3 states computed with the parent and cumulant-truncated DSRG-MRPT2 schemes are shown in Fig. 2. For all cumulant-truncated DSRG methods, the potential energy curve is everywhere continuous and smooth. Cumulant truncations are expected to impact the absolute energy the most when the neglected parts of the cumulants are the largest, which always happens in the recoupling region. The largest cumulant errors for the singlet and triplet states observed for λ2-DSRG-MRPT2 (ΔEcu = 11.20 and 3.05 kcal mol−1) gradually decrease going from the λ3[d]Ecu = 6.13 and 3.01 kcal mol−1) to the λ3[2]Ecu = 0.19 and 0.16 kcal mol−1) approximations. It is encouraging to see that the accuracy of the absolute energy can be systematically improved by using more accurate cumulant truncation schemes and that the cumulant error for λ3[2]-DSRG-MRPT2 is always less than 0.2 kcal mol−1. Cumulant errors in the absolute energy are smallest at the stretched geometry. We attribute this to the small magnitude of the three-body reduced density cumulant elements, as measured by the sum of the tensor Frobenius norm (F-norm) of the four different spin cases (ααα, βββ, ααβ, αββ). The F-norm is negligible at the stretched geometry for both the XΣg+1 (0.023) and the AΣu+3 (0.013) states but not for the other geometries (e.g., 0.437 for the XΣg+1 state in the recoupling region and 1.277 for the AΣu+3 state at the equilibrium geometry). As a consequence, at the stretched geometry, even the crudest cumulant truncation scheme, λ2-DSRG-MRPT2, introduces relatively small absolute errors (0.25 and 0.08 kcal mol−1 for the XΣg+1 and AΣu+3 states, respectively).

FIG. 2.

The potential energy curves and corresponding cumulant errors for both the XΣg+1 and the AΣu+3 states of N2 using the parent and truncated DSRG-MRPT2 schemes.

FIG. 2.

The potential energy curves and corresponding cumulant errors for both the XΣg+1 and the AΣu+3 states of N2 using the parent and truncated DSRG-MRPT2 schemes.

Close modal

To evaluate how many significant tensor elements are used by each truncation method, we plot the nonredundant elements in the three-body reduced density cumulant as a matrix for both the XΣg+1 and the AΣu+3 states of N2 using the method described in the CH2 section. Figures S3–S10 (see the supplementary material) show the nonredundant tensor elements for the XΣg+1 and the AΣu+3 states of N2 at the equilibrium geometry (1.10 and 1.29 Å for the XΣg+1 and the AΣu+3 states, respectively), in the recoupling region (1.75 Å), and at the stretched geometry (3.00 Å). The results are similar to those found for CH2: The largest elements are included in λ3[2]-DSRG-MRPT2, but many are not included in the simpler truncation schemes. Table V shows the total number of significant elements (|λuvwxyz|>104) and total number of elements of λ3 entering in the XΣg+1 and the AΣu+3 states for DSRG-MRPT2 computations at different geometries. Unsurprisingly, most significant elements are missing in λ3[d]-DSRG-MRPT2 and λ3[1]-DSRG-MRPT2, while λ3[2]-DSRG-MRPT2 (storing 59% of total elements) can include more than 70% of them in most cases, except for the AΣu+3 state stretched geometry calculation (66%).

TABLE V.

The number of significant (|λuvwxyz|>104) and total tensor elements of λ3 entering in the N2 DSRG-MRPT2 computations on the XΣg+1 and the AΣu+3 states using different cumulant truncation schemes. The numbers reported account for the four spin cases (ααα, βββ, ααβ, αββ) and do not take into account symmetries of the three-body reduced density cumulant.

N2 XΣg+1
1.10 (Å)1.75 (Å)3.00 (Å)
SignificantTotalSignificantTotalSignificantTotal
λ3[d]-DSRG-MRPT2 1 432 6 976 1 788 6 976 1080 6 976 
λ3[1]-DSRG-MRPT2 2 224 137 904 3 804 137 904 1216 137 904 
λ3[2]-DSRG-MRPT2 8 888 616 816 14 940 616 816 5808 616 816 
DSRG-MRPT2 12 640 1 048 576 21 460 1 048 576 8352 1 048 576 
N2 XΣg+1
1.10 (Å)1.75 (Å)3.00 (Å)
SignificantTotalSignificantTotalSignificantTotal
λ3[d]-DSRG-MRPT2 1 432 6 976 1 788 6 976 1080 6 976 
λ3[1]-DSRG-MRPT2 2 224 137 904 3 804 137 904 1216 137 904 
λ3[2]-DSRG-MRPT2 8 888 616 816 14 940 616 816 5808 616 816 
DSRG-MRPT2 12 640 1 048 576 21 460 1 048 576 8352 1 048 576 
N2 AΣu+3
 1.29 (Å) 1.75 (Å) 3.00 (Å) 
 Significant Total Significant Total Significant Total 
λ3[d]-DSRG-MRPT2 1 788 6 976 1 644 6 976 1064 6 976 
λ3[1]-DSRG-MRPT2 2 716 137 904 3 140 137 904 1088 137 904 
λ3[2]-DSRG-MRPT2 11 716 616 816 14 204 616 816 4984 616 816 
DSRG-MRPT2 16 012 1 048 576 20 084 1 048 576 7592 1 048 576 
N2 AΣu+3
 1.29 (Å) 1.75 (Å) 3.00 (Å) 
 Significant Total Significant Total Significant Total 
λ3[d]-DSRG-MRPT2 1 788 6 976 1 644 6 976 1064 6 976 
λ3[1]-DSRG-MRPT2 2 716 137 904 3 140 137 904 1088 137 904 
λ3[2]-DSRG-MRPT2 11 716 616 816 14 204 616 816 4984 616 816 
DSRG-MRPT2 16 012 1 048 576 20 084 1 048 576 7592 1 048 576 

We now turn to the impact of our cumulant truncation schemes on the dissociation energy and spectroscopic constants. The results are shown in Table VI. Let us first consider the effect of cumulant truncation on the ground state De as predicted by SC-NEVPT2. The error with respect to experiment without truncation is 0.14 eV, and the cu approximation scheme with a 1.2 a.u. imaginary level shift reduces the error to only 0.03 eV. As shown in Ref. 46, this improvement mainly comes from error cancellation resulting from a significant vertical shift of the potential energy curve in the dissociation region. The untruncated DSRG-MRPT2 displays a larger error than SC-NEVPT2 and CASPT2 in the dissociation energy (0.62 eV). However, as shown in Fig. 2, results are consistent across the various cumulant truncation schemes, smoothly converging to the untruncated values as the cumulant truncation is improved.

TABLE VI.

Spectroscopic constants for the nitrogen molecule computed from different cumulant-truncated MR theories for the lowest singlet and triplet states.

N2 XΣg+1
Methodre (Å)De (eV)ωe (cm−1)
Expt.a 1.0977 9.91 2359 
CASSCFa 1.1069 9.23 2496 
CASPT2a 1.1012 9.51 2454 
SC-NEVPT2a 1.1021 9.77 2460 
cu-SC-NEVPT2(0.0)a 1.1537 11.24 3930 
cu-SC-NEVPT2(1.2)a 1.0980 9.94 2470 
cud-SC-NEVPT2(0.0)a 1.1002 9.87 2466 
cud-SC-NEVPT2(1.2)a 1.0997 9.98 2467 
λ2-DSRG-MRPT2 1.0995 9.12 2349 
λ3[d]-DSRG-MRPT2 1.0993 9.18 2350 
λ3[1]-DSRG-MRPT2 1.0993 9.19 2350 
λ3[2]-DSRG-MRPT2 1.0998 9.30 2345 
DSRG-MRPT2 1.0997 9.29 2345 
N2 AΣu+3
Expt.a 1.2866 3.68 1461 
CASSCFa 1.3027 2.79 1548 
CASPT2a 1.2879 3.56 1513 
SC-NEVPT2a 1.2905 3.54 1522 
cud-SC-NEVPT2(0.0)a 1.2922 3.78 1521 
cud-SC-NEVPT2(0.7)a 1.2917 3.79 1522 
cud-SC-NEVPT2(1.2)a 1.2892 3.82 1524 
λ2-DSRG-MRPT2 1.2852 3.35 1465 
λ3[d]-DSRG-MRPT2 1.2839 3.27 1466 
λ3[1]-DSRG-MRPT2 1.2839 3.27 1465 
λ3[2]-DSRG-MRPT2 1.2865 3.37 1457 
DSRG-MRPT2 1.2866 3.37 1456 
N2 XΣg+1
Methodre (Å)De (eV)ωe (cm−1)
Expt.a 1.0977 9.91 2359 
CASSCFa 1.1069 9.23 2496 
CASPT2a 1.1012 9.51 2454 
SC-NEVPT2a 1.1021 9.77 2460 
cu-SC-NEVPT2(0.0)a 1.1537 11.24 3930 
cu-SC-NEVPT2(1.2)a 1.0980 9.94 2470 
cud-SC-NEVPT2(0.0)a 1.1002 9.87 2466 
cud-SC-NEVPT2(1.2)a 1.0997 9.98 2467 
λ2-DSRG-MRPT2 1.0995 9.12 2349 
λ3[d]-DSRG-MRPT2 1.0993 9.18 2350 
λ3[1]-DSRG-MRPT2 1.0993 9.19 2350 
λ3[2]-DSRG-MRPT2 1.0998 9.30 2345 
DSRG-MRPT2 1.0997 9.29 2345 
N2 AΣu+3
Expt.a 1.2866 3.68 1461 
CASSCFa 1.3027 2.79 1548 
CASPT2a 1.2879 3.56 1513 
SC-NEVPT2a 1.2905 3.54 1522 
cud-SC-NEVPT2(0.0)a 1.2922 3.78 1521 
cud-SC-NEVPT2(0.7)a 1.2917 3.79 1522 
cud-SC-NEVPT2(1.2)a 1.2892 3.82 1524 
λ2-DSRG-MRPT2 1.2852 3.35 1465 
λ3[d]-DSRG-MRPT2 1.2839 3.27 1466 
λ3[1]-DSRG-MRPT2 1.2839 3.27 1465 
λ3[2]-DSRG-MRPT2 1.2865 3.37 1457 
DSRG-MRPT2 1.2866 3.37 1456 
a

Value taken from Ref. 46.

Spectroscopic constants for the parent and cumulant-truncated DSRG-MRPT2 were obtained by using DSRG-MRPT2 analytic energy gradients97 for the XΣg+1 and AΣu+3 states of N2. The smoothness of potential energy curves of the parent and cumulant-truncated DSRG-MRPT2 theories is critical for obtaining meaningful properties. The maximum absolute differences for the bond length (re) and the harmonic vibrational frequency (ωe) for all DSRG-MRPT2 methods examined are 0.0020 Å and 14 cm−1 for the XΣg+1 state and 0.0027 Å and 5 cm−1 for the AΣu+3 state. Our results suggest that the crudest cumulant-truncated DSRG-MPRT2 scheme, λ2-DSRG-MRPT2, could already be an efficient method for geometry optimization and frequency calculations. In Table VII, we further compared the predicted adiabatic singlet–triplet gap of N2 using the parent and cumulant-truncated DSRG-MRPT2 methods with the experimental value. All energies reported do not include zero-point energy corrections. While the gap predicted by the DSRG-MRPT2 deviates from experiment by ∼0.3 eV, all the cumulant truncation schemes that include part of the three-body density cumulant are within 0.01 eV from the full DSRG-MRPT2.

TABLE VII.

Adiabatic singlet–triplet gap of N2 computed with DSRG-MRPT2 (see text for basis and geometry). All energies reported do not include zero-point vibrational energy correction.

MethodSinglet energy (Eh)Triplet energy (Eh)Gap (eV)
Expt. N/A N/A 6.224a 
λ2-DSRG-MRPT2 −109.431 124 −109.219 225 5.766 
λ3[d]-DSRG-MRPT2 −109.433 466 −109.216 279 5.910 
λ3[1]-DSRG-MRPT2 −109.433 664 −109.216 462 5.910 
λ3[2]-DSRG-MRPT2 −109.437 921 −109.219 956 5.931 
DSRG-MRPT2 −109.437 361 −109.219 795 5.920 
MethodSinglet energy (Eh)Triplet energy (Eh)Gap (eV)
Expt. N/A N/A 6.224a 
λ2-DSRG-MRPT2 −109.431 124 −109.219 225 5.766 
λ3[d]-DSRG-MRPT2 −109.433 466 −109.216 279 5.910 
λ3[1]-DSRG-MRPT2 −109.433 664 −109.216 462 5.910 
λ3[2]-DSRG-MRPT2 −109.437 921 −109.219 956 5.931 
DSRG-MRPT2 −109.437 361 −109.219 795 5.920 
a

Value taken from Ref. 98.

In the last test case, we study the performance of the cumulant truncations for the case of large active spaces (up to 18 electrons and 18 orbitals). Oligoacenes, or n-acenes (where n is the number of linearly fused benzene rings), have been studied extensively by chemists because of their semiconducting and optical properties.99–109 In particular, their singlet–triplet splittings have been of interest to theorists.79,99,110–122 Table VIII provides adiabatic singlet–triplet splittings for the first three members of the acene series using both unrelaxed and relaxed references. Our results show that all cumulant-truncated DSRG-MRPT2 schemes, including the crudest λ2-DSRG-MRPT2 method, can preserve the accuracy of the parent method across different levels of basis sets, with a maximum cumulant error of 0.2 kcal mol−1 for the energy gap. Contrary to expectations, we observed no specific trend for the performance of different cumulant truncations in the singlet–triplet splittings. To better understand the origin of this phenomenon, we compared the absolute energies of all computations. Additional data, including absolute energies of the singlet and triplet states of the acenes, are provided in Table S1 in the supplementary material. Figure 3 shows cumulant errors in the absolute energy of the acenes for computations with the cc-pVQZ basis set. Consistent with our previous test cases, these data show that the absolute energy can be systematically improved using higher-level cumulant truncations and that near-perfect cancellation of the cumulant errors happens across the truncation schemes. As the most promising cumulant-truncated DSRG-MRPT2 method, λ3[2]-DSRG-MRPT2 shows significant computational savings by storing 50.8%, 39.7%, and 32.5% of total elements for 2-acene, 3-acene, and 4-acene, respectively. Table S2 compares the computed adiabatic singlet–triplet splittings of the acene series to experimental values using the cc-pVQZ basis set and relaxed reference. These data include ZPVE corrections taken from our previous paper76 (−3.4, −2.3, and −1.8 kcal mol−1 for 2-, 3-, and 4-acene, respectively). In this case, the DSRG-MRPT2 underestimates the experimental gaps, a phenomenon already observed in our previous study.76 

TABLE VIII.

Adiabatic singlet–triplet splittings of the acene series (kcal mol−1) with unrelaxed and relaxed references and computed with different cumulant truncation schemes. None of the values include ZPVE correction.

cc-pVDZcc-pVTZcc-pVQZ
na234234234
 CAS(n, n)b (10, 10) (14, 14) (18, 18) (10, 10) (14, 14) (18, 18) (10, 10) (14, 14) (18, 18) 
Nbfc 790 1070 1350 790 1070 1350 790 1070 1350 
σd 0.0 0.0 1.0 0.0 0.0 1.0 0.0 0.0 1.0 
Unrelaxed λ2 63.6 41.5 27.6 62.0 43.3 28.8 63.1 44.3 29.2 
λ3[d] 63.6 41.5 27.6 62.0 43.3 28.8 63.1 44.3 29.2 
λ3[1] 63.6 41.5 27.6 62.0 43.2 28.8 63.1 44.3 29.2 
λ3[2] 63.5 41.5 27.6 62.1 43.3 28.8 63.1 44.3 29.2 
Full 63.7 41.7 27.8 62.1 43.3 28.9 63.1 44.4 29.3 
Relaxed λ2 62.3 40.2 26.9 61.2 41.9 27.7 62.2 43.1 28.3 
λ3[d] 62.3 40.1 26.9 61.2 41.9 27.7 62.2 43.1 28.3 
λ3[1] 62.3 40.1 26.9 61.2 41.9 27.7 62.2 43.1 28.3 
λ3[2] 62.2 40.1 26.9 61.3 41.9 27.7 62.2 43.1 28.3 
Full 62.3 40.3 27.0 61.4 42.0 27.8 62.2 43.2 28.3 
cc-pVDZcc-pVTZcc-pVQZ
na234234234
 CAS(n, n)b (10, 10) (14, 14) (18, 18) (10, 10) (14, 14) (18, 18) (10, 10) (14, 14) (18, 18) 
Nbfc 790 1070 1350 790 1070 1350 790 1070 1350 
σd 0.0 0.0 1.0 0.0 0.0 1.0 0.0 0.0 1.0 
Unrelaxed λ2 63.6 41.5 27.6 62.0 43.3 28.8 63.1 44.3 29.2 
λ3[d] 63.6 41.5 27.6 62.0 43.3 28.8 63.1 44.3 29.2 
λ3[1] 63.6 41.5 27.6 62.0 43.2 28.8 63.1 44.3 29.2 
λ3[2] 63.5 41.5 27.6 62.1 43.3 28.8 63.1 44.3 29.2 
Full 63.7 41.7 27.8 62.1 43.3 28.9 63.1 44.4 29.3 
Relaxed λ2 62.3 40.2 26.9 61.2 41.9 27.7 62.2 43.1 28.3 
λ3[d] 62.3 40.1 26.9 61.2 41.9 27.7 62.2 43.1 28.3 
λ3[1] 62.3 40.1 26.9 61.2 41.9 27.7 62.2 43.1 28.3 
λ3[2] 62.2 40.1 26.9 61.3 41.9 27.7 62.2 43.1 28.3 
Full 62.3 40.3 27.0 61.4 42.0 27.8 62.2 43.2 28.3 
a

Number of linearly fused benzene rings (n).

b

Use conventional notation of CAS(e, o), for e active electrons and o active orbitals.

c

Number of nonfrozen orbitals.

d

The ACI energy importance criteria (in mEh). See Ref. 76.

FIG. 3.

Cumulant errors (kcal mol−1) for the absolute energy of the singlet and triplet computations of the acene series with both unrelaxed and relaxed references using the cc-pVQZ basis set.

FIG. 3.

Cumulant errors (kcal mol−1) for the absolute energy of the singlet and triplet computations of the acene series with both unrelaxed and relaxed references using the cc-pVQZ basis set.

Close modal

In this work, we have explored the possibility of truncating the three-body reduced density cumulant in the second-order multireference perturbation theory based on the driven similarity renormalization group (DSRG-MRPT2). We tested the two cumulant truncations examined in the work of Zgid et al.46 (λ2-DSRG-MRPT2 and λ3[d]-DSRG-MRPT2) and also proposed two new cumulant truncations (λ3[1]-DSRG-MRPT2 and λ3[2]-DSRG-MRPT2). In contrast to SC-NEVPT2, which requires a large imaginary level shift to deal with “false intruders” along potential energy curves, we demonstrated that cumulant-truncated versions of the DSRG-MRPT2 preserve the smoothness of the potential energy curves. Our results also show that the accuracy of cumulant-truncated DSRG-MRPT2 can be systematically improved with higher-level truncations, and λ3[2]-DSRG-MRPT2 is particularly promising since the error introduced from this truncation scheme in energy differences is found to generally be less than 0.5 kcal mol−1. At the same time, the memory cost decreases from O(NA6) to O(NA5). Simpler approximation methods could be applied to reduce the memory and computational cost of energy gap calculations due to near-perfect error cancellations observed when the two states have similar molecular geometries. A careful analysis is necessary when applying cumulant truncation methods to evaluate potential energy surfaces as less expensive schemes may introduce non-negligible errors. The cumulant truncation schemes discussed in this work are particularly useful to reduce the storage and computational cost in applications with large active spaces (NA > 10–20). Further extension of the DSRG-MRPT2 that will enable applications to large systems will also need to address the scaling of the DSRG-MRPT2 with respect to the number of core and virtual orbitals via methods that can take advantage of the locality of dynamical electron correlation. Combining these two strategies is expected to open the door to future applications of multireference methods and Hamiltonian downfolding for quantum computing110 to large systems with large active spaces.

See the supplementary material for (1) plots of the nonredundant tensor elements of the three-body cumulants of CH2 and N2 and (2) adiabatic singlet–triplet splittings of the oligoacenes corrected for zero-point vibrational energy effects.

This work was supported by the U.S. National Science Foundation under Award No. CHE-2312105 and by a Camille Dreyfus Teacher-Scholar Award (No. TC-18-045).

The authors have no conflicts to disclose.

Shuhang Li: Data curation (lead); Investigation (lead); Software (lead); Writing – original draft (lead); Writing – review & editing (equal). Jonathon P. Misiewicz: Supervision (equal); Writing – review & editing (equal). Francesco A. Evangelista: Conceptualization (lead); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article and its supplementary material.

1.
K.
Andersson
,
P.-Å.
Malmqvist
,
B. O.
Roos
,
A. J.
Sadlej
, and
K.
Wolinski
,
J. Phys. Chem.
94
,
5483
(
1990
).
2.
K.
Andersson
,
P.-Å.
Malmqvist
, and
B. O.
Roos
,
J. Chem. Phys.
96
,
1218
(
1992
).
4.
P. M.
Kozlowski
and
E. R.
Davidson
,
J. Chem. Phys.
100
,
3672
(
1994
).
5.
C.
Angeli
,
R.
Cimiraglia
,
S.
Evangelisti
,
T.
Leininger
, and
J.-P.
Malrieu
,
J. Chem. Phys.
114
,
10252
(
2001
).
6.
C.
Angeli
,
R.
Cimiraglia
, and
J.-P.
Malrieu
,
Chem. Phys. Lett.
350
,
297
(
2001
).
7.
C.
Angeli
,
R.
Cimiraglia
, and
J.-P.
Malrieu
,
J. Chem. Phys.
117
,
9138
(
2002
).
8.
C.
Angeli
,
M.
Pastore
, and
R.
Cimiraglia
,
Theor. Chem. Acc.
117
,
743
(
2007
).
9.
R. K.
Chaudhuri
,
K. F.
Freed
,
G.
Hose
,
P.
Piecuch
,
K.
Kowalski
,
M.
Włoch
,
S.
Chattopadhyay
,
D.
Mukherjee
,
Z.
Rolik
,
Á.
Szabados
,
G.
Tóth
, and
P. R.
Surján
,
J. Chem. Phys.
122
,
134105
(
2005
).
10.
M.
Bendikov
,
H. M.
Duong
,
K.
Starkey
,
K.
Houk
,
E. A.
Carter
, and
F.
Wudl
,
J. Am. Chem. Soc.
126
,
7416
(
2004
).
11.
Y.
Kawashima
,
T.
Hashimoto
,
H.
Nakano
, and
K.
Hirao
,
Theor. Chem. Acc.
102
,
49
(
1999
).
12.
K. H.
Marti
,
I. M.
Ondík
,
G.
Moritz
, and
M.
Reiher
,
J. Chem. Phys.
128
,
014104
(
2008
).
13.
Y.
Kurashige
and
T.
Yanai
,
J. Chem. Phys.
130
,
234114
(
2009
).
14.
Y.
Kurashige
and
T.
Yanai
,
J. Chem. Phys.
135
,
094104
(
2011
).
15.
S.
Wouters
,
V.
Van Speybroeck
, and
D.
Van Neck
,
J. Chem. Phys.
145
,
054120
(
2016
).
16.
R. J.
Anderson
,
T.
Shiozaki
, and
G. H.
Booth
,
J. Chem. Phys.
152
,
054101
(
2020
).
17.
G. H.
Booth
,
A. J.
Thom
, and
A.
Alavi
,
J. Chem. Phys.
131
,
054106
(
2009
).
18.
S.
Sharma
and
G. K.
Chan
,
J. Chem. Phys.
141
,
111101
(
2014
).
19.
K.
Chatterjee
and
A. Y.
Sokolov
,
J. Chem. Theory Comput.
16
,
6343
(
2020
).
20.
A.
Mahajan
,
N. S.
Blunt
,
I.
Sabzevari
, and
S.
Sharma
,
J. Chem. Phys.
151
,
211102
(
2019
).
21.
N. S.
Blunt
,
A.
Mahajan
, and
S.
Sharma
,
J. Chem. Phys.
153
,
164120
(
2020
).
22.
A. Y.
Sokolov
and
G. K.
Chan
,
J. Chem. Phys.
144
,
064102
(
2016
).
23.
A. Y.
Sokolov
,
S.
Guo
,
E.
Ronca
, and
G. K.-L.
Chan
,
J. Chem. Phys.
146
,
244102
(
2017
).
24.
P.
Celani
and
H.-J.
Werner
,
J. Chem. Phys.
112
,
5546
(
2000
).
25.
S.
Sharma
,
G.
Knizia
,
S.
Guo
, and
A.
Alavi
,
J. Chem. Theory Comput.
13
,
488
(
2017
).
26.
E.
Giner
,
C.
Angeli
,
Y.
Garniron
,
A.
Scemama
, and
J.-P.
Malrieu
,
J. Chem. Phys.
146
,
224108
(
2017
).
27.
C.
Li
and
F. A.
Evangelista
,
J. Chem. Theory Comput.
11
,
2097
(
2015
).
29.
E.
Pastorczak
and
K.
Pernal
,
J. Chem. Theory Comput.
14
,
3493
(
2018
).
30.
E.
Maradzike
,
M.
Hapka
,
K.
Pernal
, and
A. E.
DePrince
III
,
J. Chem. Theory Comput.
16
,
4351
(
2020
).
31.
G.
Li Manni
,
R. K.
Carlson
,
S.
Luo
,
D.
Ma
,
J.
Olsen
,
D. G.
Truhlar
, and
L.
Gagliardi
,
J. Chem. Theory Comput.
10
,
3669
(
2014
).
32.
P.
Sharma
,
J. J.
Bao
,
D. G.
Truhlar
, and
L.
Gagliardi
,
Annu. Rev. Phys. Chem.
72
,
541
(
2021
).
33.
F. E.
Harris
,
Int. J. Quantum Chem.
90
,
105
(
2002
).
34.
M.
Nooijen
,
M.
Wladyslawski
, and
A.
Hazra
,
J. Chem. Phys.
118
,
4832
(
2003
).
35.
L.
Kong
and
E. F.
Valeev
,
J. Chem. Phys.
134
,
214109
(
2011
).
36.
37.
D.
Mukherjee
and
W.
Kutzelnigg
,
J. Chem. Phys.
114
,
2047
(
2001
).
38.
J. M.
Herbert
and
J. E.
Harriman
,
Cumulants, Extensivity, and the Connected Formulation of the Contracted Schrödinger Equation
(
Wiley Online Library
,
2007
), pp.
261
292
.
39.
F.
Colmenero
and
C.
Valdemoro
,
Phys. Rev. A
47
,
979
(
1993
).
41.
D. A.
Mazziotti
,
Phys. Rev. A
57
,
4219
(
1998
).
42.
D. A.
Mazziotti
,
Reduced-Density-Matrix Mechanics: With Applications to Many-Electron Atoms and Molecules
(
Wiley Online Library
,
2007
), Vol.
134
.
43.
W.
Kutzelnigg
and
D.
Mukherjee
,
J. Chem. Phys.
110
,
2800
(
1999
).
44.
K.
Shamasundar
,
J. Chem. Phys.
131
,
174109
(
2009
).
45.
W.
Kutzelnigg
,
K.
Shamasundar
, and
D.
Mukherjee
,
Mol. Phys.
108
,
433
(
2010
).
46.
D.
Zgid
,
D.
Ghosh
,
E.
Neuscamman
, and
G. K.-L.
Chan
,
J. Chem. Phys.
130
,
194107
(
2009
).
47.
Y.
Guo
,
K.
Sivalingam
, and
F.
Neese
,
J. Chem. Phys.
154
,
214111
(
2021
).
48.
Y.
Guo
,
K.
Sivalingam
,
C.
Kollmar
, and
F.
Neese
,
J. Chem. Phys.
154
,
214113
(
2021
).
49.
S.
Evangelisti
,
J. P.
Daudey
, and
J. P.
Malrieu
,
Phys. Rev. A
35
,
4930
(
1987
).
50.
K.
Kowalski
and
P.
Piecuch
,
Phys. Rev. A
61
,
052506
(
2000
).
51.
52.
K. G.
Dyall
,
J. Chem. Phys.
102
,
4909
(
1995
).
53.
C.
Camacho
,
H. A.
Witek
, and
S.
Yamamoto
,
J. Comput. Chem.
30
,
468
(
2009
).
54.
C.
Camacho
,
R.
Cimiraglia
, and
H. A.
Witek
,
Phys. Chem. Chem. Phys.
12
,
5058
(
2010
).
55.
M.
Saitow
,
Y.
Kurashige
, and
T.
Yanai
,
J. Chem. Phys.
139
,
044118
(
2013
).
56.
M.
Saitow
,
Y.
Kurashige
, and
T.
Yanai
,
J. Chem. Theory Comput.
11
,
5120
(
2015
).
57.
Y.
Kurashige
,
J.
Chalupskỳ
,
T. N.
Lan
, and
T.
Yanai
,
J. Chem. Phys.
141
,
174111
(
2014
).
58.
Q. M.
Phung
,
S.
Wouters
, and
K.
Pierloot
,
J. Chem. Theory Comput.
12
,
4352
(
2016
).
59.
N.
Nakatani
and
S.
Guo
,
J. Chem. Phys.
146
,
094102
(
2017
).
60.
W.
Kutzelnigg
and
D.
Mukherjee
,
J. Chem. Phys.
107
,
432
(
1997
).
61.
U. S.
Mahapatra
,
B.
Datta
,
B.
Bandyopadhyay
, and
D.
Mukherjee
,
Adv. Quantum Chem.
30
,
163
(
1998
).
62.
L.
Kong
,
M.
Nooijen
, and
D.
Mukherjee
,
J. Chem. Phys.
132
,
234107
(
2010
).
63.
D.
Sinha
,
R.
Maitra
, and
D.
Mukherjee
,
Comput. Theor. Chem.
1003
,
62
(
2013
).
64.
W.
Kutzelnigg
, in
Recent Progress in Coupled Cluster Methods
,
Challenges and Advances in Computational Chemistry and Physics
, edited by
P.
Čársky
,
J.
Paldus
, and
J.
Pittner
(
Springer Netherlands
,
2010
), Vol.
11
, pp.
299
356
.
65.
W.
Kutzelnigg
,
Int. J. Quantum Chem.
109
,
3858
(
2009
).
66.
F. A.
Evangelista
,
J. Chem. Phys.
141
,
054109
(
2014
).
67.
C.
Li
and
F. A.
Evangelista
,
Annu. Rev. Phys. Chem.
70
,
245
(
2019
).
68.
C.
Li
and
F. A.
Evangelista
,
J. Chem. Phys.
144
,
164114
(
2016
).
69.
J.
Pipek
and
P. G.
Mezey
,
J. Chem. Phys.
90
,
4916
(
1989
).
70.
M.
Hanauer
and
A.
Köhn
,
Chem. Phys.
401
,
50
(
2012
).
71.
G. D.
Purvis
III
,
R.
Shepard
,
F. B.
Brown
, and
R. J.
Bartlett
,
Int. J. Quantum Chem.
23
,
835
(
1983
).
72.
T. H.
Dunning
, Jr.
,
J. Chem. Phys.
90
,
1007
(
1989
).
73.
D. E.
Woon
and
T. H.
Dunning
, Jr.
,
J. Chem. Phys.
98
,
1358
(
1993
).
74.
Forte, a suite of quantum chemistry methods for strongly correlated electrons. For current version see https://github.com/evangelistalab/forte,
2020
.
75.
D. G.
Smith
,
L. A.
Burns
,
A. C.
Simmonett
,
R. M.
Parrish
,
M. C.
Schieber
,
R.
Galvelis
,
P.
Kraus
,
H.
Kruse
,
R.
Di Remigio
,
A.
Alenaizan
et al,
J. Chem. Phys.
152
,
184108
(
2020
).
76.
J. B.
Schriber
,
K. P.
Hannon
,
C.
Li
, and
F. A.
Evangelista
,
J. Chem. Theory Comput.
14
,
6295
(
2018
).
77.
C. W.
Bauschlicher
, Jr.
and
P. R.
Taylor
,
J. Chem. Phys.
85
,
6510
(
1986
).
78.
K. L.
Schuchardt
,
B. T.
Didier
,
T.
Elsethagen
,
L.
Sun
,
V.
Gurumoorthi
,
J.
Chase
,
J.
Li
, and
T. L.
Windus
,
J. Chem. Inf. Model.
47
,
1045
(
2007
).
79.
J. B.
Schriber
and
F. A.
Evangelista
,
J. Chem. Phys.
144
,
161106
(
2016
).
80.
J. B.
Schriber
and
F. A.
Evangelista
,
J. Chem. Theory Comput.
13
,
5354
(
2017
).
81.
B.
Huron
,
J.
Malrieu
, and
P.
Rancurel
,
J. Chem. Phys.
58
,
5745
(
1973
).
82.
K. P.
Hannon
,
C.
Li
, and
F. A.
Evangelista
,
J. Chem. Phys.
144
,
204111
(
2016
).
83.
F.
Weigend
,
Phys. Chem. Chem. Phys.
4
,
4285
(
2002
).
84.
F.
Weigend
,
A.
Köhn
, and
C.
Hättig
,
J. Chem. Phys.
116
,
3175
(
2002
).
85.
C.
Hättig
,
Phys. Chem. Chem. Phys.
7
,
59
(
2005
).
86.
A. A.
Holmes
,
N. M.
Tubman
, and
C. J.
Umrigar
,
J. Chem. Theory Comput.
12
,
3674
(
2016
).
87.
S.
Sharma
,
A. A.
Holmes
,
G.
Jeanmairet
,
A.
Alavi
, and
C. J.
Umrigar
,
J. Chem. Theory Comput.
13
,
1595
(
2017
).
88.
Y.
Garniron
,
A.
Scemama
,
P.-F.
Loos
, and
M.
Caffarel
,
J. Chem. Phys.
147
,
034101
(
2017
).
89.
P.-F.
Loos
,
A.
Scemama
,
A.
Blondel
,
Y.
Garniron
,
M.
Caffarel
, and
D.
Jacquemin
,
J. Chem. Theory Comput.
14
,
4360
(
2018
).
90.
J.
Li
,
M.
Otten
,
A. A.
Holmes
,
S.
Sharma
, and
C. J.
Umrigar
,
J. Chem. Phys.
149
,
214110
(
2018
).
91.
N. M.
Tubman
,
C. D.
Freeman
,
D. S.
Levine
,
D.
Hait
,
M.
Head-Gordon
, and
K. B.
Whaley
,
J. Chem. Theory Comput.
16
,
2139
(
2020
).
92.
R. W.
Havenith
,
P. R.
Taylor
,
C.
Angeli
,
R.
Cimiraglia
, and
K.
Ruud
,
J. Chem. Phys.
120
,
4619
(
2004
).
93.
J.-Q.
Fan
,
W.-Y.
Zhang
,
Q.
Ren
, and
F.
Chen
,
Mol. Phys.
120
,
e2048109
(
2022
).
94.
M.
Hodecker
and
A.
Dreuw
,
J. Chem. Phys.
153
,
084112
(
2020
).
95.
E.
Giner
,
A.
Scemama
,
J.
Toulouse
, and
P.-F.
Loos
,
J. Chem. Phys.
151
,
144118
(
2019
).
96.
A. D.
Chien
,
A. A.
Holmes
,
M.
Otten
,
C. J.
Umrigar
,
S.
Sharma
, and
P. M.
Zimmerman
,
J. Phys. Chem. A
122
,
2714
(
2018
).
97.
S.
Wang
,
C.
Li
, and
F. A.
Evangelista
,
J. Chem. Theory Comput.
17
,
7666
(
2021
).
98.
NIST Chemistry WebBook, NIST Standard Reference Database Number 69
, edited by
P. J.
Linstrom
and
W. G.
Mallard
(
National Institute of Standards and Technology
,
Gaithersburg, MD
,
2022
).
99.
H. F.
Bettinger
,
Pure Appl. Chem.
82
,
905
(
2010
).
100.
Y.
Yang
,
E. R.
Davidson
, and
W.
Yang
,
Proc. Natl. Acad. Sci. U.S.A.
113
,
E5098
(
2016
).
101.
I.
Paci
,
J. C.
Johnson
,
X.
Chen
,
G.
Rana
,
D.
Popović
,
D. E.
David
,
A. J.
Nozik
,
M. A.
Ratner
, and
J.
Michl
,
J. Am. Chem. Soc.
128
,
16546
(
2006
).
102.
H.
Marciniak
,
M.
Fiebig
,
M.
Huth
,
S.
Schiefer
,
B.
Nickel
,
F.
Selmaier
, and
S.
Lochbrunner
,
Phys. Rev. Lett.
99
,
176402
(
2007
).
103.
T. S.
Kuhlman
,
J.
Kongsted
,
K. V.
Mikkelsen
,
K. B.
Møller
, and
T. I.
Sølling
,
J. Am. Chem. Soc.
132
,
3431
(
2010
).
104.
P. M.
Zimmerman
,
F.
Bell
,
D.
Casanova
, and
M.
Head-Gordon
,
J. Am. Chem. Soc.
133
,
19944
(
2011
).
105.
M. W. B.
Wilson
,
A.
Rao
,
B.
Ehrler
, and
R. H.
Friend
,
Acc. Chem. Res.
46
,
1330
(
2013
).
106.
S.
Izadnia
,
D. W.
Schönleber
,
A.
Eisfeld
,
A.
Ruf
,
A. C.
LaForge
, and
F.
Stienkemeier
,
J. Phys. Chem. Lett.
8
,
2068
(
2017
).
107.
M.
Wibowo
,
R.
Broer
, and
R. W. A.
Havenith
,
Comput. Theor. Chem.
1116
,
190
(
2017
).
108.
S.
Refaely-Abramson
,
F. H.
da Jornada
,
S. G.
Louie
, and
J. B.
Neaton
,
Phys. Rev. Lett.
119
,
267401
(
2017
).
109.
L. A.
Martínez-Martínez
,
M.
Du
,
R.
Florentino Ribeiro
,
S.
Kéna-Cohen
, and
J.
Yuen-Zhou
,
J. Phys. Chem. Lett.
9
,
1951
(
2018
).
110.
R.
Huang
,
C.
Li
, and
F. A.
Evangelista
,
PRX Quantum
4
,
020313
(
2023
).
111.
J.
Hachmann
,
J. J.
Dorando
,
M.
Avilés
, and
G. K.-L.
Chan
,
J. Chem. Phys.
127
,
134309
(
2007
).
112.
B.
Hajgató
,
D.
Szieberth
,
P.
Geerlings
,
F.
De Proft
, and
M. S.
Deleuze
,
J. Chem. Phys.
131
,
224321
(
2009
).
113.
B.
Hajgató
,
M.
Huzak
, and
M. S.
Deleuze
,
J. Phys. Chem. A
115
,
9282
(
2011
).
114.
W.
Mizukami
,
Y.
Kurashige
, and
T.
Yanai
,
J. Chem. Theory Comput.
9
,
401
(
2013
).
115.
J. P.
Malrieu
,
R.
Caballol
,
C. J.
Calzado
,
C.
De Graaf
, and
N.
Guihery
,
Chem. Rev.
114
,
429
(
2014
).
116.
S.
Lehtola
,
N. M.
Tubman
,
K. B.
Whaley
, and
M.
Head-Gordon
,
J. Chem. Phys.
147
,
154105
(
2017
).
117.
S.
Lehtola
,
J.
Parkhill
, and
M.
Head-Gordon
,
J. Chem. Phys.
145
,
134110
(
2016
).
118.
C.-N.
Yeh
and
J.-D.
Chai
,
Sci. Rep.
6
,
30562
(
2016
).
119.
J.
Fosso-Tande
,
D. R.
Nascimento
, and
I. A.
Eugene DePrince
,
Mol. Phys.
114
,
423
(
2016
).
120.
J.
Lee
,
D. W.
Small
,
E.
Epifanovsky
, and
M.
Head-Gordon
,
J. Chem. Theory Comput.
13
,
602
(
2017
).
121.
S.
Battaglia
,
N.
Faginas-Lago
,
D.
Andrae
,
S.
Evangelisti
, and
T.
Leininger
,
J. Phys. Chem. A
121
,
3746
(
2017
).
122.
N.
Dupuy
and
M.
Casula
,
J. Chem. Phys.
148
,
134112
(
2018
).

Supplementary Material