Diagrammatically size-consistent and basis-set-free vibrational coupled-cluster (XVCC) theory for both zero-point energies and transition frequencies of a molecule, the latter through the equation-of-motion (EOM) formalism, is defined for an nth-order Taylor-series potential energy surface (PES). Quantum-field-theoretical tools (the rules of normal-ordered second quantization and Feynman–Goldstone diagrams) for deriving their working equations are established. The equations of XVCC and EOM-XVCC including up to the mth-order excitation operators are derived and implemented with the aid of computer algebra in the range of 1 ≤ m ≤ 8. Algorithm optimizations known as strength reduction, intermediate reuse, and factorization are carried out before code generation, reducing the cost scaling of the mth-order XVCC and EOM-XVCC in an nth-order Taylor-series PES (mn) to the optimal value of O(Nm+⌊n/2⌋), where N is the number of modes. The calculated zero-point energies and frequencies of fundamentals, overtones, and combinations as well as Fermi-resonant modes display rapid and nearly monotonic convergence with m towards the exact values for the PES. The theory with the same excitation rank as the truncation order of the Taylor-series PES (m = n) seems to strike the best cost-accuracy balance, achieving the accuracy of a few tenths of cm−1 for transitions involving (m − 3) modes and of a few cm−1 for those involving (m − 2) modes. The relationships between XVCC and the vibrational coupled-cluster theories of Prasad and coworkers and of Christiansen and coworkers as well as the size-extensive vibrational self-consistent-field and many-body perturbation theories are also elucidated.

Coupled-cluster (CC) theory1,2 was introduced by Coester and Kümmel3,4 in nuclear physics and soon adopted by Čížek and coworkers5,6 in the field of quantum chemistry, where it was fully developed into the most accurate, versatile, and widely applicable electronic structure theory.7–9 Being rigorously (i.e., diagrammatically) size consistent,10,11 it was applied to problems in solid-state physics.12–18 It also witnessed a revival in nuclear physics,19,20 as the Hamiltonian for nucleons became increasingly more accurate. Another area where the applicability of this theory is being explored is quantum nuclear dynamics in a bound potential energy surface (PES), i.e., anharmonic vibrational structure theory.21,22

Following its applications to model PES’s,23–25 vibrational CC (VCC) theory was pioneered and extensively developed by Prasad and coworkers for general molecules with truncated Taylor-series PES’s.26–34 They considered both zero-point energies and transition frequencies using the equation-of-motion (EOM)27,32 or multireference29,34 formalism. They also calculated spectra and band intensities in either the time-dependent26,28,30,31 or independent33 framework. These methods are based on a single Hartree product of harmonic-oscillator wave functions as a reference, of which the mode frequencies and coordinate center are varied to minimize its energy. Their working equations are derived using the second-quantized ladder operators, which are valid for harmonic oscillators. The reference wave function adopted can be identified as that of the size-extensive vibrational self-consistent-field (XVSCF) method with an anharmonic geometry correction, i.e., XVSCF[n] (n is the rank of the Taylor-series PES).35 

Christiansen and coworkers36–42 proposed alternative formulations of VCC theory and its excited-state analog, VCC linear response (VCCLR). They may be viewed as a direct extension of vibrational correlation theories pioneered by Bowman and coworkers and by Ratner and coworkers, such as vibrational configuration-interaction (VCI)43 and vibrational Møller–Plesset perturbation (VMP) theories,44 and share similar design philosophy. They all rely on a single-reference Hartree product of one-mode functions (“modals”) along normal coordinates furnished by vibrational self-consistent-field (VSCF) theory.45–47 These modals are determined so as to minimize the VSCF energy and are numerically defined on a grid or a basis set. These vibrational correlation theories tend to be implemented in symbolic (as opposed to matrix-algebraic) algorithms, in which excited Hartree products are manipulated during a calculation (not during formula derivation) with the Hamiltonian matrix elements in modals evaluated by quadrature.

In the meantime, it has been shown that VSCF theory is not diagrammatically size consistent.48 Unlike a truncated CI for electrons, which gives zero electron-correlation energy per electron in the thermodynamic limit, VSCF yields a finite and thus physically meaningful anharmonic correction in this limit. Nevertheless, this correction is due to a minute portion of a large number of calculated contributions, most of which are shown to vanish in the same limit as they correspond to disconnected diagrams. This renders VSCF an exceedingly inefficient method for large molecules and solids. When such vanishing contributions are erased from the VSCF equation, we arrive at a streamlined, rigorously size-consistent, but less accurate variant of VSCF, which we call XVSCF(n) or XVSCF[n],35,49,50 depending on whether an anharmonic geometry correction is suppressed (in the former) or not (in the latter). XVSCF is no longer exact for any one-dimensional potential, which VSCF is, in the infinite basis-set limit. Instead, each modal of XVSCF (or that of VSCF in the thermodynamic limit) is shown to be a harmonic-oscillator wave function with its frequency and center dressed with mean-field anharmonic effects,48,49 which is consistent with Makri’s theorem.51 Since the Hamiltonian matrix elements in harmonic-oscillator modals are expressed simply with force constants, XVSCF and vibrational correlation theories based on it can be defined compactly with force constants and excitation amplitudes. Grids, basis sets, and quadrature are no longer necessary.

Insofar as the defining feature of CC theory is its size consistency, which also ensures its uniformly high accuracy across system sizes, its formalism should be stipulated with many-body-theoretical techniques such as normal-ordered second quantization and Feynman–Goldstone diagrams and be based on a rigorously size-consistent reference wave function. A normal-ordered second-quantized form of the vibrational Hamiltonian has been introduced recently.52 Remarkably, its constant term and excitation amplitudes are identified as the energy, gradient, and frequency expressions of XVSCF and of the closely related self-consistent-phonon (SCP) method53–57 (but not of VSCF). This is reminiscent of the normal-ordered electronic Hamiltonian, in which the expressions of the Hartree–Fock (HF) energy and Fock operator naturally emerge.2 This observation highlights the central place XVSCF and SCP occupy in anharmonic vibrational structure theory for large molecules and solids and suggests that they provide an ideal reference wave function for vibrational correlation theories for such systems. The establishment of XVSCF,35,49,50 along with the elucidation of their relationship to the normal-ordered vibrational Hamiltonian,52 therefore, sets the stage for a more complete development of a VCC theory.

In this article, we present a rigorously (i.e., diagrammatically) size-consistent VCC theory for the ground and excited states, the latter through the EOM formalism. We call them XVCCm and EOM-XVCCm theories, respectively, where m stands for the highest connected excitation rank considered. We propose a rational and expedient method of deriving their algebraic definitions using the rules of normal-ordered second quantization and Feynman–Goldstone diagrams. We demonstrate its utility through the derivation of low-order XVCC and EOM-XVCC equations. The formalisms thus derived are valid for any single-reference Hartree product of harmonic-oscillator modals along delocalized, rectilinear coordinates, such as XVSCF(n),49 XVSCF[n],35 and SCP (Refs. 53–57) as well as the harmonic approximation. Some diagrammatic contributions are shown to vanish in the XVSCF and SCP reference, making them more favorable, although higher-order XVCC and EOM-XVCC are increasingly insensitive to the reference choice. XVCC and EOM-XVCC with the XVSCF[n] reference are found to be equivalent to the VCC formalism first developed by Prasad and coworkers,27,32 shedding light on the relationship between the two VCC theories and XVCC theory: the VCC theory of Prasad and coworkers and XVCC theory may be diagrammatically size-consistent and infinite-basis counterparts of the VCC theory of Christiansen and coworkers in a truncated Taylor-series PES, if no other approximations are made.

We furthermore derive and implement high-order XVCCm and EOM-XVCCm equations up to m = 8 or octuple excitations, with the aid of computerized derivation, optimization, and code synthesis. We apply these methods to the zero-point energies and frequencies of the fundamentals, overtones, and combinations as well as Fermi-resonant modes of the water and formaldehyde molecules, using quartic force fields (QFF’s).58 A comparative analysis is performed for the results of these methods and of XVSCF(n), XVSCF[n], size-extensive second-order vibrational many-body perturbation theory (XVMP2),59 and full VCI. It shows that XVCCm and EOM-XVCCm with m equal to the truncation rank of the Taylor-series PES are uniformly accurate and stable for both resonant and nonresonant modes and are likely the methods with the greatest practical utility of all considered.

We express the mth-order XVCC (XVCCm) wave function for the zero-point (ground) state (labeled by subscript “0”) as

(1)

where |Φ0〉 is a reference wave function. The latter, in turn, is a Hartree product of harmonic-oscillator modals (which include but are not limited to those from the harmonic approximation to the PES),

(2)

where |ni〉 denotes the harmonic-oscillator wave function with quantum number ni and frequency ωi along the ith normal, first-order Dyson,50 or other delocalized, rectilinear coordinate. Operator Tˆ is the sum of the single, double, through m-fold cluster excitation operators,

(3)

with

(4)
(5)

where τi1i2in is an excitation amplitude, the value of which is to be determined by solving the amplitude equations (see below). Operator aˆi and its adjoint aˆi are the ladder operators for the harmonic-oscillator modals, i.e.,

(6)
(7)

The braces {⋯} in Eq. (5) bring the operators enclosed by them into a normal order.52 Two or more of the mode indices (i1, i2, …, in) in Eq. (4) are permitted to be coincident.

We require that the wave function of Eq. (1) satisfy the vibrational Schrödinger equation in the space of Hartree products reachable by Tˆ from |Φ0〉,

(8)

and

(9)

for all ν1, …, νn and each n (1 ≤ nm). Here, E0(m) is the XVCCm total energy and

(10)

The number of the amplitude equations expressed collectively by Eq. (9) is equal to the number of unknown excitation amplitudes {τ} in Eq. (4). Furthermore, it can be shown below that E0(m), which is also unknown, is eliminated from Eq. (9). Hence, by solving Eq. (9), we obtain {τ}. Substituting the latter in the energy equation [Eq. (8)], we determine E0(m).

The reference wave function |Φ0〉 can be furnished by any of the following: XVSCF(n),49 XVSCF[n],35 SCP,53–57 or the harmonic approximation. They differ from one another in the coordinates, their center, and the modal frequencies,50 but they all share the same form: the Hartree product of harmonic-oscillator modals in delocalized, rectilinear coordinates. The XVCCm formalisms are essentially unchanged by the choice of the references, except that some terms may vanish in some references. The foregoing definition of XVCCm and the working equations derived in the following are valid for any of them.

The rather restricted form of reference wave functions is dictated by the fact that every modal of a Hartree product in delocalized, rectilinear coordinates (such as normal coordinates) reduces to a harmonic oscillator in the thermodynamic limit, as shown both analytically48 and numerically49 in the case of VSCF and also in a more general sense by Makri.51 This makes it convenient to express the formalisms of vibrational correlation methods in terms of the ladder operators and force constants (as opposed to molecular integrals) and also renders the methods free of numerically defined basis functions, as discussed below. Its drawback is a lack of choices of coordinates60–62 and the poor performance of the reference methods for strongly anharmonic potentials such as a Morse potential and a double-well potential, for overtones and combinations, and for anharmonic resonances. As shown below, some of these issues are ameliorated by XVCC, but others (such as a double-well problem) are unlikely to be resolved.

One may notice at this stage a striking difference between electronic CC and XVCC. Since the one-mode basis functions of XVCC are harmonic-oscillator wave functions, which need not be stored numerically, XVCC can take account of an infinite number of them implicitly. This means that, even when the Hamiltonian is truncated after the n-body term (n ≥ 3) for a molecule with N vibrational degrees of freedom, neither XVCCn nor XVCCN is exact mathematically, though XVCCm is shown below to be rapidly convergent with m in practice. In other words, XVCC is an infinite-basis or basis-set-free method despite the finiteness of its equations. This is where XVCC and Christiansen’s VCC (Ref. 36) also differ from each other fundamentally. The excitation space reached by Tˆ is limited by the number of basis functions included in VCC, whereas it is not in XVCC.

The VCC ansatz of Prasad and coworkers27,32 is the same as above, but it is predicated on a reference Hartree product of Gaussian modals along normal coordinates, whose centers and widths are varied to minimize its energy. This reference method is identified as the XVSCF[n] method.35 Therefore, XVCCm with the XVSCF[n] reference corresponds to the VCC theory of Prasad and coworkers,27,32 although XVCCm is defined for more general reference Hartree products, including XVSCF(n), SCP, and the harmonic approximation.

Here, we illustrate an algebraic derivation of the working equations of XVCCm using the normal-ordered second quantization and Wick’s theorem.52 We use, as an example, XVCC1 or XVCC with single excitations (XVCCS) for a vibrational Hamiltonian with a QFF.

This Hamiltonian can be written in the normal-ordered second-quantized form52 as

(11)

with

(12)
(13)
(14)
(15)

where it should be understood that there is term-by-term correspondence between the left- and right-hand sides of each equation.

The constant term in Eq. (11), E0(0), is the total ground-state energy expression of XVSCF,35,49,52 regardless of the actual reference method adopted in the calculation, and is given by

(16)

where V0 is the value of the PES at the minimum, ωi is again the frequency of the ith modal, and F’s are the force constants. The Hamiltonian amplitudes, {W}, are given by

(17)
(18)
(19)
(20)

in a QFF. These amplitudes are invariant to any permutation of indices. With the XVSCF(4) reference, Wii+=ωi holds. In the XVSCF[4] reference, additionally, Wi = 0 is true. Furthermore, Wij+=0 for ij with the SCP reference. The fact that these key quantities of the XVSCF methods naturally emerge in the expression of the normal-ordered Hamiltonian underscores their fundamental significance in many-body vibrational theory.52 

From Eq. (8), we find the XVCCS energy equation to be

(21)

where E0(1) is the XVCCS total energy. The Taylor series of the exponential operator is truncated after the Tˆ14 term because the higher-order terms cannot contribute in a QFF (see below). The XVCCS amplitude equation, Eq. (9), similarly reduces to

(22)

in a QFF. Here, the Taylor series can be terminated after the Tˆ15 term (see below).

Wick’s theorem states that the vacuum expectation value of a product of two or more normal-ordered series of operators is the sum of all full contractions with no internal contractions.2,52,63 It also states that the vacuum expectation value of just one normal-ordered series is zero.2,52,63

Using this theorem, one can evaluate the first term of Eq. (21) as

(23)

The second term has only one nonzero full contraction,

(24)

The third term has two nonzero full contractions,

(25)

The fourth and fifth terms are evaluated similarly, yielding

(26)

and

(27)

Since the Hamiltonian with a QFF has only up to four-body operators, the Tˆ15 term and higher in the exponential operator cannot produce a full contraction.

Together, the XVCCS energy is given by

(28)

It may be noticed that each of the second and subsequent terms is connected in the sense that every factor has at least one common summation index with another factor. This ensures the extensivity of E0(1).10,11

The connectedness of the energy equation is a universal feature of XVCCm for all m. For future convenience, therefore, we introduce an XVCCm energy equation alternative to Eq. (8), which reads

(29)

where HˆN=HˆE0(0), ΔE0(m)=E0(m)E0(0), and (⋯)C deletes terms that are disconnected. We call this the energy equation of XVCCm.

The XVCCS amplitude equation can likewise be derived by Wick’s theorem. The first term of Eq. (22) has one nonzero full contraction,

(30)

The second term has two nonzero full contractions,

(31)

The third term has four such contractions,

(32)

The remainders are evaluated similarly, yielding

(33)
(34)
(35)

The higher-order terms in the exponential operator cannot produce a full contraction and are zero.

Combining these, we find the XVCCS amplitude equation to read

(36)

Note that the first five terms in the right-hand side are disconnected in the sense that there is a factor (τν) that has no common summation index with any other factor; each of them is, therefore, a simple product of two factors.

Factoring these disconnected terms, we can rewrite the above equation as

(37)
(38)

where we use the fact that the sum in the parenthesis is equal to E0(1) as per Eq. (28). We can, therefore, further simplify the amplitude equation into

(39)

Notice that this equation consists of connected terms only.

In fact, we can recast the amplitude equations of XVCCm in general into a form that is equivalent to, but more convenient than Eq. (9),

(40)

for all ν1, …, νn and all n (1 ≤ nm). We call this the Tn amplitude equation of XVCCm. The diagrammatic derivations of XVCC and the EOM-XVCC formulations to be discussed below are based on Eqs. (29) and (40).

In the foregoing derivation, the connected structure of the XVCC energy and amplitude equations is only illustrated with the XVCCS example. In  Appendix A, we document an alternative derivation that proves more rigorously the connectedness of these equations at any order.1,2

The equations of XVCC can be derived even more expediently with Feynman–Goldstone diagrams, which is equivalent to the normal-ordered second-quantized derivation; a diagram corresponds to all topologically related Wick’s contractions. Here, we introduce the rules of drawing closed diagrams for XVCC energy equations and open diagrams for XVCC amplitude equations as well as the rules of translating them into algebraic expressions. These rules differ from those for XVMP energies and self-energies.64 We illustrate their application to XVCCm with m = 1 (XVCCS) and m = 2 (XVCCSD) in this subsection.

We recast each term in the normal-ordered Hamiltonian, HˆN, into diagrammatic forms, as shown in Fig. 1. Each diagram represents one of the normal-ordered operators defined by Eqs. (12)–(15). The cluster excitation operators, {Tˆn}, are expressed diagrammatically in Fig. 2.

FIG. 1.

Diagrammatic representation of the normal-ordered Hamiltonian HˆN with a QFF.

FIG. 1.

Diagrammatic representation of the normal-ordered Hamiltonian HˆN with a QFF.

Close modal
FIG. 2.

Diagrammatic representation of the cluster excitation operators.

FIG. 2.

Diagrammatic representation of the cluster excitation operators.

Close modal

Table I lists the rules to generate diagrams of the XVCCm energy and amplitude equations in an nth-order Taylor-series PES. Applying these rules, we find the energy diagrams of XVCCS in a QFF (n = 4) to be those in Fig. 3. The energy diagrams of XVCCSD in a QFF are the union of Figs. 3 and 4. Hence,

(41)
(42)

where the summands are the algebraic interpretation of the diagrams. Since each Tˆk vertex in an energy diagram has only upgoing lines (signifying the fact that they describe an excitation process), all of its lines have no choice but to connect with the HˆN vertex, making the whole diagram always connected when closed, guaranteeing its thermodynamic extensivity as per the extensive diagram theorem.10,11,65

TABLE I.

Rules to draw all closed diagrams in the energy equation and all open diagrams in the Tl amplitude equation of XVCCm in an nth-order Taylor-series PES.

(1) Draw an HˆN vertex having up to n lines. 
(2) Draw 0 to nTˆk vertices (1 ≤ km) beneath the HˆN vertex. 
(3) Connect the lines of the Tˆk vertices with the lines of the HˆN 
 vertex without altering (upgoing or downgoing) line directions, 
 leaving l external (i.e., dangling) lines to form a connected 
 diagram, which is either closed (l = 0) or open (l > 0). 
(4) Repeat steps (1)–(3) to enumerate all topologically distinct 
 diagrams. If permutation of Tˆ vertices allows two diagrams 
 to overlap, they are deemed topologically equivalent. 
(1) Draw an HˆN vertex having up to n lines. 
(2) Draw 0 to nTˆk vertices (1 ≤ km) beneath the HˆN vertex. 
(3) Connect the lines of the Tˆk vertices with the lines of the HˆN 
 vertex without altering (upgoing or downgoing) line directions, 
 leaving l external (i.e., dangling) lines to form a connected 
 diagram, which is either closed (l = 0) or open (l > 0). 
(4) Repeat steps (1)–(3) to enumerate all topologically distinct 
 diagrams. If permutation of Tˆ vertices allows two diagrams 
 to overlap, they are deemed topologically equivalent. 
FIG. 3.

Diagrams in the energy equation of XVCCS in a QFF.

FIG. 3.

Diagrams in the energy equation of XVCCS in a QFF.

Close modal
FIG. 4.

Additional diagrams in the energy equation of XVCCSD in a QFF.

FIG. 4.

Additional diagrams in the energy equation of XVCCSD in a QFF.

Close modal

Table II provides a set of rules to translate these diagrams into algebraic expressions. For example, diagram E4 of Fig. 3 can be labeled as shown in Fig. 5 using rules 1 through 5 of the table. Following the remaining rules, we find the algebraic expression of the whole diagram to be

(43)

where the 1/4! factor comes from rule 8 because there are four equivalent Tˆ1 vertices. It is straightforward to confirm that the algebraic interpretations of E1, E2, and E3 obtained with Table II agree exactly with the second, third, and fourth terms of the XVCCS energy expression [Eq. (28)], respectively.

TABLE II.

Rules to translate a closed or open diagram of XVCC and EOM-XVCC into an algebraic expression.

(1) Label each internal line with a mode index (i, j, k, …). 
 An internal line is a line terminated by vertices at both ends. 
(2) Label each external line with a mode index (ν, μ, λ, …). 
 An external line is a line terminated by a vertex at only one end. 
(3) Associate each HˆN vertex with a Hamiltonian amplitude W
 The indices of W are given by the labels of the lines connected 
 to that vertex. A two-line HˆN vertex is associated with W 
 if both lines are upgoing or downgoing; with W+ otherwise. 
(4) Associate each Tˆk vertex with an excitation amplitude τ
 The indices of τ are given by the labels of the lines connected 
 to that vertex. 
(5) Associate each Rˆk vertex with an excitation amplitude r
 The indices of r are given by the labels of the lines connected 
 to that vertex. 
(6) Sum over all internal line indices (i, j, k, …). 
(7) Sum over all permutations of external line indices (ν, μ, λ, …). 
(8) For each set of n equivalent vertices, multiply by a factor of 
 (1/n!). If permutation of two vertices leaves the diagram 
 unchanged topologically, the two vertices are equivalent. 
(9) For each set of m equivalent lines, multiply by a factor of 
 (1/m!). If permutation of two lines leaves the diagram 
 unchanged topologically, the two lines are equivalent. 
(1) Label each internal line with a mode index (i, j, k, …). 
 An internal line is a line terminated by vertices at both ends. 
(2) Label each external line with a mode index (ν, μ, λ, …). 
 An external line is a line terminated by a vertex at only one end. 
(3) Associate each HˆN vertex with a Hamiltonian amplitude W
 The indices of W are given by the labels of the lines connected 
 to that vertex. A two-line HˆN vertex is associated with W 
 if both lines are upgoing or downgoing; with W+ otherwise. 
(4) Associate each Tˆk vertex with an excitation amplitude τ
 The indices of τ are given by the labels of the lines connected 
 to that vertex. 
(5) Associate each Rˆk vertex with an excitation amplitude r
 The indices of r are given by the labels of the lines connected 
 to that vertex. 
(6) Sum over all internal line indices (i, j, k, …). 
(7) Sum over all permutations of external line indices (ν, μ, λ, …). 
(8) For each set of n equivalent vertices, multiply by a factor of 
 (1/n!). If permutation of two vertices leaves the diagram 
 unchanged topologically, the two vertices are equivalent. 
(9) For each set of m equivalent lines, multiply by a factor of 
 (1/m!). If permutation of two lines leaves the diagram 
 unchanged topologically, the two lines are equivalent. 
FIG. 5.

Diagram E4 with labels.

FIG. 5.

Diagram E4 with labels.

Close modal

Likewise, the application of the rules in Table II to the diagrams in Fig. 4 leads to the XVCCSD energy equation,

(44)
(45)

The factor of (1/2!)3 in the penultimate term (diagram E7) arises from the fact that there are one pair of equivalent Tˆ2 vertices and two pairs of equivalent lines.

Table I also outlines rules to enumerate all diagrams in the XVCCm amplitude equations. Unlike energy diagrams, a diagram in the Tl amplitude equation is open with l external lines. Such diagrams can be disconnected, in which case they must be deleted as per rule 3.

Figure 6 lists all four connected diagrams of the XVCCS T1 amplitude equation in a QFF, which can thus be written as

(46)

where Si is the algebraic interpretation of the corresponding diagram. The T1 amplitude equation of XVCCSD in a QFF is diagrammatically expressed by Figs. 6 and 7. The T2 amplitude equation of the same method is defined with the diagrams in Fig. 8. The T1 and T2 amplitude equations of XVCCSD are thus formally written as

(47)
(48)
FIG. 6.

Diagrams in the T1 amplitude equation of XVCCS in a QFF.

FIG. 6.

Diagrams in the T1 amplitude equation of XVCCS in a QFF.

Close modal
FIG. 7.

Additional diagrams in the T1 amplitude equation of XVCCSD in a QFF.

FIG. 7.

Additional diagrams in the T1 amplitude equation of XVCCSD in a QFF.

Close modal
FIG. 8.

Diagrams in the T2 amplitude equation of XVCCSD in a QFF.

FIG. 8.

Diagrams in the T2 amplitude equation of XVCCSD in a QFF.

Close modal

The interpretation rules in Table II apply to open diagrams in these amplitude equations also. For example, lines and vertices in diagram S4 can be labeled as shown in Fig. 9. Its algebraic expression then reads

(49)

Applying the same rules to diagrams S1, S2, and S3 and substituting the results in Eq. (46) reproduces the T1 amplitude equation [Eq. (39)] derived algebraically.

FIG. 9.

Diagram S4 with labels.

FIG. 9.

Diagram S4 with labels.

Close modal

The T1 amplitude equation of XVCCSD in a QFF is likewise obtained directly from Figs. 6 and 7,

(50)
(51)

where T1(1) is given algebraically in Eq. (39).

Similarly, the T2 amplitude equation of the same method can directly result from Fig. 8 and reads

(52)
(53)

where Pˆ(ν1νn) yields the sum of n! permutations of indices {ν1, …, νn} in the term it acts on (cf. rule 7 of Table II). For instance,

(54)

In the same way, the diagrammatic and algebraic energy and amplitude equations of XVCCm with m = 3 (XVCCSDT) in a QFF are derived. They are documented in  Appendix B. For 4 ≤ m ≤ 8, diagrams and equations were generated with the aid of computer algebra based on the rules provided in Tables I and II and are documented in the supplementary material.66 

In a vibrational analysis, it is usually the frequencies (excitation energies) that are of greater interest than the zero-point energies (total energies in the ground state). Here, we stipulate and implement an excited-state XVCC theory using the EOM formalism,67–72 which is equivalent to CC linear response73–78 for excitation energies.

The EOM-XVCCm theory can be viewed as a CI for excited states using the XVCC effective Hamiltonian H̄N defined as

(55)

where Tˆ is truncated after the m-fold cluster excitation operator [Eq. (3)] and all of its amplitudes should have been determined by a preceding XVCCm calculation.

Let Rˆ be a linear excitation operator truncated after the same rank,

(56)

with

(57)
(58)

where ri1i2in is an excitation amplitude. We require that the excited-state wave function Rˆ|Φ0 satisfy the Schrödinger equation with the modified Hamiltonian H̄N in the space of Hartree products reachable by Rˆ from |Φ0〉. That is,

(59)
(60)

for all ν1, …, νn and all n (1 ≤ nm), where ΔEk(m)=Ek(m)E0(0) and Ek(m) is the total EOM-XVCCm energy of the kth excited state. The constant term r0 is formally necessary in Eq. (56) for excited states of the same symmetry as the ground state. Although Rˆ and its amplitudes vary with k and should thus carry a k identifier, it is suppressed for brevity.

Equations (59) and (60) define EOM-XVCCm as an eigenvalue problem with a non-Hermitian matrix of H̄N, which thus has both left and right eigenvectors with Rˆ corresponding to the latter. They can, however, be further simplified by using the fact that T amplitudes in H̄N satisfy the XVCCm equations. Equations (29) and (40) indicate that |Φ0〉 satisfies the Schrödinger equation with H̄N in the same Hartree-product space. This implies

(61)
(62)

because Rˆ|Φ0{Rˆ|Φν1ν2νn} spans the space no greater than |Φ0〉∪{|Φν1ν2νn〉}. Subtracting these from Eqs. (59) and (60), we obtain

(63)
(64)

where ωk(m)=ΔEk(m)ΔE0(m) and is the EOM-XVCCm transition energy (frequency) to the kth excited state.

Equation (64) is the ansatz for EOM-XVCCm for frequencies, which can also be written as

(65)

for all ν1, …, νn and all n (1 ≤ nm). We call this the Rn amplitude equation of EOM-XVCCm. The nested connectedness requirement is always met when the term is simply connected because both Tˆ and Rˆ are excitation operators and can only directly connect with HˆN, but not with each other. The connectedness requirement, however, deletes all terms involving r0 because the latter has no index (or a line in its diagrammatic representation) to be shared with other factors.

Equation (65), therefore, refers to a set of equations whose number is equal to the number of unknowns (excluding r0) minus one (because ωk(m) is also unknown) and, with an appropriate normalization condition, can determine ωk(m) and all the amplitudes in Rˆ except r0. The value of r0, in turn, can be determined by Eq. (63), though it is not necessary if only frequencies are sought. Equation (65) satisfies the intensive diagram theorem,10,11 ensuring the size consistency of EOM-XVCCm for total excited-state energies and frequencies, despite its CI-like structure.

Just like the electronic counterpart, EOM-XVCCm can also be derived by applying the first-order time-dependent perturbation (linear-response) theory to XVCCm. This derivation (not repeated here because it is identical to the electronic case78) elegantly explains why the truncation ranks of Tˆ and Rˆ should be the same. EOM-XVCCm with the XVSCF[n] reference is the same as excited-state VCC of Prasad and coworkers.27,32

Here, we illustrate the normal-ordered second-quantized derivation of the amplitude equations of EOM-XVCCm using m = 1 (EOM-XVCCS) in a QFF as an example. The working equations of EOM-XVCCm with m = 2 (EOM-XVCCSD) are documented in  Appendix C.

The left-hand side of Eq. (65) in the case of m = 1 reads

(66)

where the nested connectedness means simple connectedness in this case and terms containing higher powers of Tˆ1 cannot yield any connected contributions.

The first term of Eq. (66) has only one full contraction,

(67)

which is connected. The next term has two connected full contractions,

(68)

The last term yields the connected full contraction,

(69)

Together, we obtain the R1 amplitude equation of EOM-XVCCS in a QFF as

(70)

The diagrammatic forms of the Rˆn operators are given in Fig. 10. They are isomorphic to the diagrams of the Tˆn operators with the only difference being the colors of the vertices. The rules to generate all diagrams entering the amplitude equations of EOM-XVCCm are given in Table III. The rules for interpreting these diagrams algebraically are provided in Table II.

FIG. 10.

Diagrammatic representation of the linear excitation operators.

FIG. 10.

Diagrammatic representation of the linear excitation operators.

Close modal
TABLE III.

Rules to draw all open diagrams in the Rl amplitude equation of EOM-XVCCm in an nth-order Taylor-series PES.

(1) Draw an HˆN vertex having up to n lines. 
(2) Draw a Rˆh vertex (1 ≤ hm) and zero to (n1)Tˆk vertices 
 (1 ≤ km) beneath the HˆN vertex. 
(3) Connect the lines of the Rˆh and Tˆk vertices with the lines of 
 the HˆN vertex without altering (upgoing or downgoing) line 
 directions, leaving l external (i.e., dangling) lines to form 
 a connected diagram. 
(4) Repeat steps (1)–(3) to enumerate all topologically distinct 
 diagrams. If permutation of Tˆ vertices allows two 
 diagrams to overlap, they are deemed topologically equivalent. 
(1) Draw an HˆN vertex having up to n lines. 
(2) Draw a Rˆh vertex (1 ≤ hm) and zero to (n1)Tˆk vertices 
 (1 ≤ km) beneath the HˆN vertex. 
(3) Connect the lines of the Rˆh and Tˆk vertices with the lines of 
 the HˆN vertex without altering (upgoing or downgoing) line 
 directions, leaving l external (i.e., dangling) lines to form 
 a connected diagram. 
(4) Repeat steps (1)–(3) to enumerate all topologically distinct 
 diagrams. If permutation of Tˆ vertices allows two 
 diagrams to overlap, they are deemed topologically equivalent. 

As an illustration, we derive the R1 amplitude equation of EOM-XVCCS in a QFF diagrammatically. Diagrams in this equation are drawn in Fig. 11. The R1 amplitude equation can thus be formally written as

(71)

Diagram S3 may be labeled as shown in Fig. 12 according to the rules in Table II. The corresponding algebraic translation is

(72)

where the 1/2! factor is due to a pair of equivalent Tˆ1 vertices. The other two diagrams can be interpreted similarly, yielding the same R1 amplitude equation as Eq. (70).

FIG. 11.

Diagrams in the R1 amplitude equation of EOM-XVCCS in a QFF.

FIG. 11.

Diagrams in the R1 amplitude equation of EOM-XVCCS in a QFF.

Close modal
FIG. 12.

Diagram S3 with labels.

FIG. 12.

Diagram S3 with labels.

Close modal

The diagrammatic and algebraic equations of EOM- XVCCm for m = 2 through 8 were generated with the aid of computer algebra based on the rules provided in Tables II and III. The EOM-XVCCSD equations are documented in  Appendix C, while the EOM-XVCCm equations for 3 ≤ m ≤ 8 are documented in the supplementary material.66 

The XVCCm and EOM-XVCCm methods with m = 1 through 8 in a QFF were implemented. The Hamiltonian employed was the pure vibrational one, neglecting any coupling between vibrations and rotations. The XVSCF(4), XVSCF[4], or the harmonic reference wave function was used (“4” standing for the QFF). The choice of the reference wave function is indicated by a suffix as in XVCCm (4), XVCCm[4], or XVCCm{4}, respectively; “{n}” suffix indicates the use of a harmonic reference with the nth-order PES.

The Tn amplitude equations of XVCCm were solved with a standard Newton–Krylov algorithm,79 implemented in igmres, available from scipy. The initial values of τ were zero. The Rn amplitude equations of EOM-XVCCm with m ≥ 2 were solved with the iterative Arnoldi method, a non-Hermitian generalization of the Lanczos method,80–82 available in arpack.83 The eigenvalue convergence criterion was set to machine precision. The Arnoldi method does not require the matrix elements of H̄N stored in memory; instead, it evaluates the left-hand sides of the Rn amplitude equations [Eq. (65)] directly for several trial r vectors. This allows a small subset of low-lying excited-state roots to be determined in a memory-efficient manner. On the other hand, the EOM-XVCCS problem was solved by diagonalization of the H̄N matrix, solving Eqs. (59) and (60) for all excited-state energies instead.

The XVCCm and EOM-XVCCm diagrammatic equations were derived by a symbolic algebra program using the rules presented in Tables I and III. These diagrammatic equations were then translated by the same symbolic algebra program into the algebraic form according to the interpretation rules given in Table II. Terms that originate from a single diagram and are thus related by a permutation operator were identified and consolidated. When the resulting expressions were evaluated literally, the operation costs of XVCCm and EOM-XVCCm in an nth-order Taylor-series PES would scale as steeply as O(Nm+n), where N is the number of modes. To reduce the cost, three types of algorithm optimizations were performed computationally before the automatic synthesis of c++ codes was carried out.84,85 The optimizations considered in this work were the strength reduction, intermediate reuse, and factorization.84 The details of each optimization procedure are given below.

Every tensor has a high degree of index permutation symmetry such as τi1i2i3 = τi1i3i2 = τi2i1i3 = τi2i3i1 = τi3i1i2 = τi3i2i1. Exploiting it can compress the memory spaces of the m-fold cluster and linear excitation amplitudes and of mth-order force constants by a factor of m! and the operation cost of their contractions to an even greater degree. Nevertheless, the compression factor of m! in XVCCm is much smaller than (m!)2 in electronic CC theory with connected m-fold excitations, and our preliminary implementation of the algorithm that takes into account this symmetry did not demonstrate any gain in computational speed owing to an enormous cost overhead associated with complicated data layout. Therefore, in this study, we neglected this symmetry.

The algebraic equations of XVCC and EOM-XVCC are sums of contractions of tensors (force constants W, T amplitudes, and R amplitudes). The operation cost of evaluating a contraction (product) of more than two tensors is greatly reduced by carrying out a series of binary tensor contractions. Suppose a tensor contraction E = ABCD, where A through E are tensors of some dimension (including possibly scalars). This product can be evaluated stepwise as

(73)

where I1 and I2 are the so-called intermediate tensors. Alternatively, the same product may be evaluated as

(74)

Both orders of contraction give the same final result (E) but at generally different operation and memory costs. Strength reduction determines the sequential contraction order with the minimal operation cost for each product by blanket search; it neglects parallel contraction such as (AB)(CD).

The result of this procedure is depicted in Fig. 13 for diagram T17, which occurs in the T3 amplitude equation of XVCCSDT (see  Appendix B). It is algebraically defined as

(75)

A literal evaluation of this equation involves an O(N7) arithmetic operations, as it must perform N4 multiplications and additions to evaluate each of N3 elements of T17, where N is the number of modes. However, the same result can be obtained at an O(N4) cost by breaking the quaternary tensor contraction into three separate binary tensor contractions carried out in the following order. First, the force-constant tensor is multiplied with τl, which has no external index,

(76)

This step involves N4 multiplications and additions that define an intermediate, T17a, of dimension N3. This intermediate is then multiplied by τjkν at an O(N4) cost to define another intermediate, T17b, of dimension N2,

(77)

Finally, T17 is obtained by contraction of intermediate T17b with τiμλ at an O(N4) cost,

(78)

Therefore, the cost of evaluating diagram T17 is reduced from O(N7) to O(N4).

FIG. 13.

Strength reduction of diagram T17 in the T3 amplitude equation of XVCCSDT (see  Appendix B).

FIG. 13.

Strength reduction of diagram T17 in the T3 amplitude equation of XVCCSDT (see  Appendix B).

Close modal

This optimization simply stores an intermediate appearing two or more times in the whole equation and reuses it without recalculating. Consider the following sum-of-product expression of tensors: E = A(BC) + D(BC). If the strength reduction procedure suggests the optimal contraction orders specified by the parentheses, intermediate I1 = BC is calculated only once and stored for reuse. Hence, the whole sum is evaluated as

(79)

The numbers of multiplications and additions as a function of N required to evaluate all the amplitude equations of XVCCm and EOM-XVCCm with no optimization and with strength reduction and intermediate reuse are compared in Table IV. For all methods with m ≥ 2, strength reduction leads to a great reduction in the operation cost from O(Nm+4) to O(Nm+2). Generally, for an nth-order PES, the optimal scaling of XVCCm and EOM-XVCCm is O(Nm+⌊n/2⌋) when mn and is no greater than O(Nn+⌊(m−1)/2⌋) when m < n. A proof of this statement is given in  Appendix D.

TABLE IV.

The number of multiplications and additions as a function of the number of modes (N) required in evaluating all the amplitude equations of XVCC or EOM-XVCC implemented with no optimization, with strength reduction and intermediate reuse, or with strength reduction, intermediate reuse, and factorization. The three leading terms only are shown.

MethodNo optimizationStrength reductionFactorizationa
XVCCS 5N4 + 4N3 + 2N2 2N4 + 4N3 + 8N2 2N4 + 6N3 + 4N2 
XVCCSD 11N6 + 31N5 + 34N4 13N4 + 54N3 + 36N2 13N4 + 44N3 + 25N2 
XVCCSDT 60N7 + 119N6 + 110N5 11N5 + 165N4 + 96N3 11N5 + 86N4 + 70N3 
XVCCSDTQ 241N8 + 360N7 + 282N6 24N6 + 425N5 + 225N4 24N6 + 172N5 + 109N4 
XVCC5 799N9 + 1053N8 + 720N7 104N7 + 1061N6 + 488N5 82N7 + 342N6 + 194N5 
XVCC6 2577N10 + 3086N9 + 1888N8 175N8 + 3079N7 + 1134N6 143N8 + 1135N7 + 366N6 
XVCC7 8398N11 + 9246N10 + 5070N9 285N9 + 8902N8 + 3156N7 241N9 + 4392N8 + 1161N7 
XVCC8 28 099N12 + 28 380N11 + 13 989N10 468N10 + 26 641N9 + 8991N8 410N10 + 18 021N9 + 4420N8 
EOM-XVCCS 5N4 + 3N3 + 2N2 2N4 + 4N3 + 7N2 2N4 + 4N3 + 7N2 
EOM-XVCCSD 30N6 + 70N5 + 54N4 21N4 + 97N3 + 63N2 21N4 + 57N3 + 52N2 
EOM-XVCCSDT 183N7 + 286N6 + 220N5 15N5 + 343N4 + 189N3 15N5 + 131N4 + 99N3 
EOM-XVCCSDTQ 729N8 + 916N7 + 598N6 34N6 + 965N5 + 474N4 34N6 + 311N5 + 170N4 
EOM-XVCC5 2490N9 + 2794N8 + 1614N7 207N7 + 2649N6 + 1096N5 143N7 + 696N6 + 356N5 
EOM-XVCC6 8295N10 + 8533N9 + 4432N8 348N8 + 8164N7 + 2782N6 254N8 + 2527N7 + 731N6 
EOM-XVCC7 27 975N11 + 26 506N10 + 12 457N9 569N9 + 24 962N8 + 8309N7 439N9 + 11 231N8 + 2570N7 
EOM-XVCC8 96 603N12 + 83 992N11 + 35 934N10 934N10 + 78 522N9 + 25 129N8 762N10 + 51 289N9 + 11 274N8 
MethodNo optimizationStrength reductionFactorizationa
XVCCS 5N4 + 4N3 + 2N2 2N4 + 4N3 + 8N2 2N4 + 6N3 + 4N2 
XVCCSD 11N6 + 31N5 + 34N4 13N4 + 54N3 + 36N2 13N4 + 44N3 + 25N2 
XVCCSDT 60N7 + 119N6 + 110N5 11N5 + 165N4 + 96N3 11N5 + 86N4 + 70N3 
XVCCSDTQ 241N8 + 360N7 + 282N6 24N6 + 425N5 + 225N4 24N6 + 172N5 + 109N4 
XVCC5 799N9 + 1053N8 + 720N7 104N7 + 1061N6 + 488N5 82N7 + 342N6 + 194N5 
XVCC6 2577N10 + 3086N9 + 1888N8 175N8 + 3079N7 + 1134N6 143N8 + 1135N7 + 366N6 
XVCC7 8398N11 + 9246N10 + 5070N9 285N9 + 8902N8 + 3156N7 241N9 + 4392N8 + 1161N7 
XVCC8 28 099N12 + 28 380N11 + 13 989N10 468N10 + 26 641N9 + 8991N8 410N10 + 18 021N9 + 4420N8 
EOM-XVCCS 5N4 + 3N3 + 2N2 2N4 + 4N3 + 7N2 2N4 + 4N3 + 7N2 
EOM-XVCCSD 30N6 + 70N5 + 54N4 21N4 + 97N3 + 63N2 21N4 + 57N3 + 52N2 
EOM-XVCCSDT 183N7 + 286N6 + 220N5 15N5 + 343N4 + 189N3 15N5 + 131N4 + 99N3 
EOM-XVCCSDTQ 729N8 + 916N7 + 598N6 34N6 + 965N5 + 474N4 34N6 + 311N5 + 170N4 
EOM-XVCC5 2490N9 + 2794N8 + 1614N7 207N7 + 2649N6 + 1096N5 143N7 + 696N6 + 356N5 
EOM-XVCC6 8295N10 + 8533N9 + 4432N8 348N8 + 8164N7 + 2782N6 254N8 + 2527N7 + 731N6 
EOM-XVCC7 27 975N11 + 26 506N10 + 12 457N9 569N9 + 24 962N8 + 8309N7 439N9 + 11 231N8 + 2570N7 
EOM-XVCC8 96 603N12 + 83 992N11 + 35 934N10 934N10 + 78 522N9 + 25 129N8 762N10 + 51 289N9 + 11 274N8 
a

Factorization leads to a cost increase in XVCCS and thus was turned off in its implementation.

The operation cost of XVCC and EOM-XVCC can be further reduced by factorization. Let us consider the sum-of-product tensor equation, E = A(BC) + A(BD), where the parentheses indicate the optimal order of contraction for each product. This sum can be evaluated as

(80)

Two contractions by the common tensor B in I2 = BC + BD are factored as I2 = B(C + D), reducing the multiplication cost by half. Also, two contractions by A are also merged into one, further halving the multiplication cost. This procedure should not be confused with intermediate reuse.

Factorability of equations depends on the contraction order of each product, making strength reduction and factorization a coupled optimization problem. Since a complete solution of this coupled problem is exponentially complex, we approach it as two separate, sequential optimizations: strength reduction is performed first for each product, whereupon factorization is carried out by blanket search without altering the order of contraction. Hence, a sum of intermediates multiplied by the same tensor and the same permutation operator is factored. The resulting new intermediates are sums of tensors and they are recursively factored. In this work, the permutation operator itself is not decomposed into a product of smaller permutation operators, unlike in previous work for electronic CC.84 

An example of the procedure is given below and drawn in Fig. 14,

(81)
(82)
(83)
(84)
(85)

where the parentheses specify the optimal order of contraction for each product determined by the preceding strength reduction step. Factorization does not alter these orders. When evaluating the sum of these five terms, the common factor τ is pulled out,

(86)

where the intermediate D6+8+9+17+23 is the sum of three intermediates,

(87)

with the subscripts indicating the diagrammatic origin of the intermediates. The first intermediate does not lend itself to any further factorization and is given by

(88)

The second intermediate is further factored by the common tensor τj and is evaluated as

(89)
(90)

The third intermediate is also further factored as

(91)
(92)
FIG. 14.

Factorable diagrams in the T2 amplitude equation of XVCCSDT (see  Appendix B).

FIG. 14.

Factorable diagrams in the T2 amplitude equation of XVCCSDT (see  Appendix B).

Close modal

The operation costs of XVCCm and EOM-XVCCm after factorization (in addition to strength reduction and intermediate reuse) are listed in Table IV. It should be understood that intermediate reuse (combined with factorization) applies to common intermediates appearing two or more times that are either a product or a sum-of-product of tensors. Since factorization is more likely for the final contraction step in our implementation, it is especially effective when the final step is the most costly. Factorization does not alter the overall scaling of the cost function, but generally reduces the prefactor multiplying the leading-order term, while it may increase or decrease prefactors of sub-leading-order terms.

Factorization is more important for EOM-XVCC and for higher values of m. In particular, for m > 4, factorization lowers the prefactors on the leading-order term of the cost functions. For m = 3 or 4, only the prefactors of the second and third leading-order terms of the cost functions are reduced. For m = 1 and 2, the cost is dominated by the contraction of the force-constant tensor with a tensor of T amplitudes, which occurs first in each product (see  Appendix D); therefore, factorization is found least effective there.

In Fig. 15, the CPU time spent in one cycle of the iterative solution of the XVCCm or EOM-XVCCm equation (3 ≤ m ≤ 5) in a model QFF is plotted as a function of the number of modes N. Each plot falls accurately on the asymptotic cost-scaling line of O(Nm+2) for N ≥ 20, suggesting that the leading-order terms in the cost function dominate the cost. For instance, the XVCCSDTQ and EOM-XVCCSDTQ methods with the aforementioned optimizations are remarkably efficient, taking only on the order of ten seconds per iteration for N = 35, despite the fact that their algorithms do not exploit the index permutation symmetry. These costs are expected to be negligible as compared with the cost associated with generating all higher-order force constants at the electronic structure level appropriate for these accurate vibrational methods.

FIG. 15.

CPU time (in seconds) spent in one cycle of the iterative solution of the XVCCm or EOM-XVCCm equation for a system with N modes in a QFF.

FIG. 15.

CPU time (in seconds) spent in one cycle of the iterative solution of the XVCCm or EOM-XVCCm equation for a system with N modes in a QFF.

Close modal

The XVCCm and EOM-XVCCm methods were applied to the vibrational zero-point energies and frequencies of the water and formaldehyde molecules. Their equilibrium structure, normal modes, and QFF’s were calculated at the MP2/aug-cc-pVTZ level using nwchem86 and sindo.87 The results were compared with those obtained from full VCI calculations with mavi88 using the 20 lowest-lying harmonic-oscillator wave functions of each mode as the basis set. The VCI results were converged to within 0.1 cm−1 of the exact solutions for the QFF. Our results were also compared with the XVMP2(4) and XVMP2[4] results,59 obtained either in the frequency-independent approximation to the self-energy or by solving the Dyson equation with the frequency-dependent self-energy.

Table V compiles the errors in the XVCCm zero-point energy (E0) and EOM-XVCCm frequencies (ν) from the VCI results for the water molecule. These errors decrease nearly monotonically with the excitation rank m, reaching the accuracy of a few cm−1 at m = 3 (CCSDT) for both E0 and ν. At m = 4 (CCSDTQ), the results are generally within tenths of cm−1 of the VCI results with a few exceptions. The results for m ≥ 5 are given in Table VI and are converged with accuracy higher than necessary, e.g., within 0.001 cm−1 at m = 8.

TABLE V.

The calculated zero-point energy (E0) and frequencies (νi) of the fundamental transitions (in cm−1) of the water molecule. Except for VCI, the errors from VCI are shown.

MethodaE0ν1ν2ν3
VCI 4642.4 3682.4 1556.6 3791.5 
Harmonic 108.4 139.7 71.7 156.4 
XVSCF(4) 104.5 280.2 −17.9 289.4 
XVSCF[4] 21.5 105.3 11.7 116.1 
XVCCS{4} 21.9 196.5 −1.2 206.7 
XVCCS(4) 26.6 197.2 −2.4 207.0 
XVCCS[4] 21.5 105.3 11.7 116.1 
XVCCSD{4} 22.1 33.6 7.9 34.3 
XVCCSD(4) 23.8 38.3 8.5 38.3 
XVCCSD[4] 21.5 26.0 6.8 26.6 
XVCCSDT{4} 0.2 −1.7 2.2 −2.0 
XVCCSDT(4) 0.1 −2.4 2.2 −2.7 
XVCCSDT[4] 0.4 −1.7 2.4 −1.9 
XVCCSDTQ{4} 0.0 −0.7 0.2 −0.7 
XVCCSDTQ(4) −0.0 −1.0 0.1 −1.0 
XVCCSDTQ[4] 0.0 −0.1 0.1 −0.1 
XVMP2(4)b 4.5 −24.6 3.8 −28.3 
XVMP2[4]b −1.3 −18.9 2.4 −20.3 
XVMP2(4)c 4.5 −14.3 3.8 −20.8 
XVMP2[4]c −1.3 −14.8 2.5 −17.0 
MethodaE0ν1ν2ν3
VCI 4642.4 3682.4 1556.6 3791.5 
Harmonic 108.4 139.7 71.7 156.4 
XVSCF(4) 104.5 280.2 −17.9 289.4 
XVSCF[4] 21.5 105.3 11.7 116.1 
XVCCS{4} 21.9 196.5 −1.2 206.7 
XVCCS(4) 26.6 197.2 −2.4 207.0 
XVCCS[4] 21.5 105.3 11.7 116.1 
XVCCSD{4} 22.1 33.6 7.9 34.3 
XVCCSD(4) 23.8 38.3 8.5 38.3 
XVCCSD[4] 21.5 26.0 6.8 26.6 
XVCCSDT{4} 0.2 −1.7 2.2 −2.0 
XVCCSDT(4) 0.1 −2.4 2.2 −2.7 
XVCCSDT[4] 0.4 −1.7 2.4 −1.9 
XVCCSDTQ{4} 0.0 −0.7 0.2 −0.7 
XVCCSDTQ(4) −0.0 −1.0 0.1 −1.0 
XVCCSDTQ[4] 0.0 −0.1 0.1 −0.1 
XVMP2(4)b 4.5 −24.6 3.8 −28.3 
XVMP2[4]b −1.3 −18.9 2.4 −20.3 
XVMP2(4)c 4.5 −14.3 3.8 −20.8 
XVMP2[4]c −1.3 −14.8 2.5 −17.0 
a

XVCC for zero-point energy and EOM-XVCC for frequencies. Suffix “{4}” means the use of a harmonic reference, “(4)” means the use of a XVSCF(4) reference, and “[4]” means the use of a XVSCF[4] reference.

b

In the frequency-independent approximation.59 

c

Solutions of the Dyson equation with the frequency-dependent self-energy.59 

TABLE VI.

The calculated zero-point energy (E0) and frequencies (νi) of the fundamental transitions (in cm−1) of the water molecule. The errors from XVCC8 with the corresponding reference wave function are shown.

MethodaE0ν1ν2ν3
XVCC5{4} −0.02 −0.05 0.02 −0.06 
XVCC5(4) −0.03 −0.12 0.01 −0.12 
XVCC5[4] −0.03 0.00 0.02 0.00 
XVCC6{4} −0.001 0.004 0.004 0.005 
XVCC6(4) −0.001 0.005 0.002 0.006 
XVCC6[4] −0.001 0.006 0.003 0.007 
XVCC7{4} −0.0004 0.003 0.001 0.003 
XVCC7(4) −0.0002 0.004 0.001 0.005 
XVCC7[4] −0.0005 −0.001 0.002 −0.001 
XVCC8{4} 0  0  
XVCC8(4) 0  0  
XVCC8[4] 0  0  
MethodaE0ν1ν2ν3
XVCC5{4} −0.02 −0.05 0.02 −0.06 
XVCC5(4) −0.03 −0.12 0.01 −0.12 
XVCC5[4] −0.03 0.00 0.02 0.00 
XVCC6{4} −0.001 0.004 0.004 0.005 
XVCC6(4) −0.001 0.005 0.002 0.006 
XVCC6[4] −0.001 0.006 0.003 0.007 
XVCC7{4} −0.0004 0.003 0.001 0.003 
XVCC7(4) −0.0002 0.004 0.001 0.005 
XVCC7[4] −0.0005 −0.001 0.002 −0.001 
XVCC8{4} 0  0  
XVCC8(4) 0  0  
XVCC8[4] 0  0  
a

XVCC for zero-point energy and EOM-XVCC for frequencies. Suffix “{4}” means the use of a harmonic reference, “(4)” means the use of a XVSCF(4) reference, and “[4]” means the use of a XVSCF[4] reference.

Note that the VCI results in this table were obtained with 20 basis functions per mode. The number of Hartree products involved in such a calculation grows exponentially with size or as O(bN), where b is the number of basis functions and N is the number of modes. The numbers of the T and R amplitudes in XVCCm and EOM-XVCCm, on the other hand, increase only polynomially or as O(Nm). The values of bN for VCI and Nm for XVCCSDTQ (N = 3, b = 20, and m = 4) are 8000 and 81, respectively, implying a much severer limitation in VCI’s applicability to larger molecules. The number of unknowns in Christiansen’s VCC (Ref. 36) grows as O(bm). Therefore, it too suffers from the consequence of having a finite number of basis functions, unlike XVCCm, which is inherently an infinite-basis method.

The accuracy of E0 from XVCCm is relatively insensitive to the choice of the reference wave function, despite much larger differences in the results of the reference methods themselves: harmonic approximation, XVSCF(4), and XVSCF[4]. This is in line with a similar observation72 made in the electronic CC results for correlation energies, which is understood to be due to the single-excitation operator which corrects any deficiency in the reference wave function. Likewise, the XVCCS(4) and XVCCS[4] results of E0 are close to each other, while the XVSCF(4) results differ considerably from those of XVSCF[4]. This difference is due to the anharmonic shift in the geometry, neglected in XVSCF(4) but included in XVSCF[4]. Diagrammatically, this contribution is largely represented by E1 of Fig. 3. The fact that this diagram is found among the XVCCS energy equation means that the effect of anharmonic geometry shift is included in XVCCS regardless of the reference wave function.

XVSCF[4] determines and adopts the first-order Dyson geometry50 as the center of coordinates, where Wi = 0 for all i. The T1 amplitude equation of XVCCS[4], Eq. (39), then has a trivial solution of τi = 0 for all i. This renders the XVCCS correction, E0(1)E0(0) in Eq. (28), zero, making the XVCCS[4] and XVSCF[4] values of E0 identical to each other. This expected behavior is observed in Table V. Furthermore, substituting τi = 0 into the R1 amplitude equation of EOM-XVCCS [Eq. (70)], we find the EOM-XVCCS[4] frequencies to be the eigenvalues of the Wij+ matrix, whose eigenvectors are known to define the first-order Dyson coordinates.50,52 These eigenvalues are usually, but not always, close to the XVSCF[4] frequencies,50 the latter being the diagonal elements (Wii+) of the same matrix. This is also observed in Table V. Therefore, XVCCS can take into account the leading-order effects in E0 and ν from anharmonic corrections both in geometry and coordinates.

It may be noted that XVMP2 performs considerably better than XVCCSD for E0. This is in striking contrast with the electronic counterparts; CCSD has been observed to consistently outperform MP2.1,2 This difference can be traced to the difference in the rank of the Hamiltonian in the two cases: the electronic Hamiltonian is a two-body operator, whereas the vibrational one is an n-body operator, where n is the rank of the Taylor-series PES (n = 4 in this calculation). The most important anharmonic correction to E0 is diagram E9 in Fig. 16 (in  Appendix B), which involves cubic force constants and first appears at XVCCSDT. This contribution is to a great extent accounted for by the isomorphic diagram in XVMP2 labeled 2C′ in Ref. 59. This explains why only at m ≥ 3 does XVCCm outperform XVMP2. It is generally recommended to set the value of m equal to n, if not higher, because the overall cost of the calculation is dominated by the O(Nn) step of the force-constant evaluation.

FIG. 16.

Additional diagrams in the energy equation of XVCCSDT in a QFF.

FIG. 16.

Additional diagrams in the energy equation of XVCCSDT in a QFF.

Close modal

The rather small change in E0 from XVCCS to XVCCSD can also be understood as follows. The chief contribution of XVCCSD to E0, diagram E5 of Fig. 4, is nearly zero in the XVSCF reference, where Wii=0 holds.52 When the normal coordinates are close to the first-order Dyson coordinates, which is usually the case, Wij=Wij+0 is also true for ij, making the remaining (i.e., off-diagonal) contributions to E5 insignificant. Furthermore, the smallness of Wij implies the smallness of the T2 amplitudes [cf. Eq. (53)]. Although XVCCSD uses cubic force constants in E6, they are multiplied by these small T2 amplitudes and can hardly alter E0 of XVCCS. The frequencies from EOM-XVCCSD, on the other hand, are considerably better than EOM-XVCCS and are similar to those of XVMP2. This may be ascribed to diagram S5 in Fig. 20 (in  Appendix C) included in EOM-XVCCSD, which can describe the same essential cubic corrections accounted for by self-energy diagrams 2e′ and 2f′ (Ref. 59) of XVMP2.

FIG. 20.

Additional diagrams in the R1 amplitude equation of EOM-XVCCSD in a QFF.

FIG. 20.

Additional diagrams in the R1 amplitude equation of EOM-XVCCSD in a QFF.

Close modal

For the zero-point energies and frequencies that are not affected by any resonance, we observe the following general order of accuracy: harmonic ≈ XVSCF(n) < XVCCS{n} ≈ XVCCS(n) < XVSCF[n] ≈ XVCCS[n] < XVCCSD < XVMP2 < XVCCSDT < XVCCSDTQ ≤ XVCC5 ≤ XVCC6 ≤ XVCC7 ≤ XVCC8, which differs from the order established in electronic-structure theory.1,2 In a QFF (n = 4), XVCCSDTQ or m = 4 already achieves accuracy high enough for practical purposes and XVCCm with m > n seem to have diminishing returns. Here, XVCC stands for both XVCC and EOM-XVCC and the effects of the reference wave functions and the frequency-independent approximation in XVMP2 are insignificant because there are no resonances.

The XVCCm and EOM-XVCCm results for the formaldehyde molecule are summarized in Tables VII and VIII. The calculated zero-point energy and frequencies of the fundamentals (except for ν1 and ν5) confirm the general observations made above for the water molecule.

TABLE VII.

The calculated zero-point energy (E0) and frequencies (νi) of the fundamental transitions (in cm−1) of the formaldehyde molecule. Except for VCI, the errors from VCI are shown.

MethodaE0ν1ν2ν3ν4ν5ν3ν6ν2ν6ν6
VCI 5813.6 2837.2 1719.8 1501.5 1160.7 2873.3 2702.8 2989.2 1237.5 
Harmonic 74.3 136.2 33.2 38.6 36.2 174.2 … … 29.4 
XVSCF(4) 73.0 185.0 41.0 10.8 −13.7 216.4 … … −3.3 
XVSCF[4] 22.9 79.0 20.4 9.8 10.6 106.7 … … 11.2 
XVCCS{4} 24.0 134.8 31.9 9.4 −1.3 164.7 … … 4.0 
XVCCS(4) 25.5 134.1 31.9 9.7 −0.8 163.6 … … 4.4 
XVCCS[4] 22.9 79.1 20.9 9.3 10.6 106.7 … … 11.1 
XVCCSD{4} 23.5 30.1 6.7 5.1 6.7 38.1 31.6 34.6 6.4 
XVCCSD(4) 24.1 31.6 6.6 5.4 7.5 38.9 32.2 35.4 6.8 
XVCCSD[4] 22.9 24.2 6.0 4.2 6.0 31.0 35.5 33.0 5.6 
XVCCSDT{4} 0.6 0.3 0.5 1.2 1.7 1.9 7.3 2.2 1.4 
XVCCSDT(4) 0.6 −0.0 0.5 1.2 1.7 1.8 7.9 2.4 1.4 
XVCCSDT[4] 0.8 0.1 0.5 1.2 1.9 1.3 6.1 1.4 1.4 
XVCCSDTQ{4} 0.0 −0.2 0.0 0.1 0.1 0.5 2.4 0.9 0.1 
XVCCSDTQ(4) 0.0 −0.3 0.0 0.1 0.1 0.4 2.4 0.9 0.1 
XVCCSDTQ[4] 0.1 0.2 0.0 0.1 0.0 0.7 2.3 1.0 0.0 
XVMP2(4)b 1.3 −190.9 0.7 0.9 2.7 33.0 … … 3.0 
XVMP2[4]b −1.0 −15.9 0.1 1.3 1.6 −414.3 … … 1.8 
XVMP2(4)c 1.3 −6.6 1.4 1.0 2.6 −0.2 6.6 29.3 3.0 
XVMP2[4]c −1.0 −8.9 0.3 1.4 1.5 2.2 18.0 24.8 1.8 
MethodaE0ν1ν2ν3ν4ν5ν3ν6ν2ν6ν6
VCI 5813.6 2837.2 1719.8 1501.5 1160.7 2873.3 2702.8 2989.2 1237.5 
Harmonic 74.3 136.2 33.2 38.6 36.2 174.2 … … 29.4 
XVSCF(4) 73.0 185.0 41.0 10.8 −13.7 216.4 … … −3.3 
XVSCF[4] 22.9 79.0 20.4 9.8 10.6 106.7 … … 11.2 
XVCCS{4} 24.0 134.8 31.9 9.4 −1.3 164.7 … … 4.0 
XVCCS(4) 25.5 134.1 31.9 9.7 −0.8 163.6 … … 4.4 
XVCCS[4] 22.9 79.1 20.9 9.3 10.6 106.7 … … 11.1 
XVCCSD{4} 23.5 30.1 6.7 5.1 6.7 38.1 31.6 34.6 6.4 
XVCCSD(4) 24.1 31.6 6.6 5.4 7.5 38.9 32.2 35.4 6.8 
XVCCSD[4] 22.9 24.2 6.0 4.2 6.0 31.0 35.5 33.0 5.6 
XVCCSDT{4} 0.6 0.3 0.5 1.2 1.7 1.9 7.3 2.2 1.4 
XVCCSDT(4) 0.6 −0.0 0.5 1.2 1.7 1.8 7.9 2.4 1.4 
XVCCSDT[4] 0.8 0.1 0.5 1.2 1.9 1.3 6.1 1.4 1.4 
XVCCSDTQ{4} 0.0 −0.2 0.0 0.1 0.1 0.5 2.4 0.9 0.1 
XVCCSDTQ(4) 0.0 −0.3 0.0 0.1 0.1 0.4 2.4 0.9 0.1 
XVCCSDTQ[4] 0.1 0.2 0.0 0.1 0.0 0.7 2.3 1.0 0.0 
XVMP2(4)b 1.3 −190.9 0.7 0.9 2.7 33.0 … … 3.0 
XVMP2[4]b −1.0 −15.9 0.1 1.3 1.6 −414.3 … … 1.8 
XVMP2(4)c 1.3 −6.6 1.4 1.0 2.6 −0.2 6.6 29.3 3.0 
XVMP2[4]c −1.0 −8.9 0.3 1.4 1.5 2.2 18.0 24.8 1.8 
a

XVCC for zero-point energy and EOM-XVCC for frequencies. Suffix “{4}” means the use of a harmonic reference, “(4)” means the use of a XVSCF(4) reference, and “[4]” means the use of a XVSCF[4] reference.

b

In the frequency-independent approximation.59 

c

Solutions of the Dyson equation with the frequency-dependent self-energy.59 

TABLE VIII.

The calculated zero-point energy (E0) and frequencies (νi) of the fundamental transitions (in cm−1) of the formaldehyde molecule. The errors from XVCC8 with the corresponding reference wave function are shown.

MethodaE0ν1ν2ν3ν4ν5ν3ν6ν2ν6ν6
XVCC5{4} −0.02 −0.03 0.00 0.01 0.00 0.03 0.33  0.06 0.01 
XVCC5(4) −0.02 −0.06 0.00 0.01 0.00 0.01 0.32 0.05 0.01 
XVCC5[4] −0.02 0.02 0.00 0.01 −0.00 0.04 0.26 0.06 0.01 
XVCC6{4} −0.001 0.000 0.001 0.001 −0.001 0.005 0.067 0.011 0.001 
XVCC6(4) −0.001 −0.002 0.000 0.001 −0.001 0.003 0.062 0.010 0.001 
XVCC6[4] −0.000 0.002 0.000 0.000 −0.002 0.006 0.052 0.008 0.000 
XVCC7{4} −0.0003 0.0012 0.0002 0.0004 0.0003 0.0010 0.0131 0.0021 0.0005 
XVCC7(4) −0.0002 0.0014 0.0002 0.0004 0.0003 0.0010 0.0124 0.0022 0.0005 
XVCC7[4] −0.0004 −0.0011 0.0002 0.0006 0.0007 −0.0005 0.0100 0.0004 0.0007 
XVCC8{4} 
XVCC8(4) 
XVCC8[4] 
MethodaE0ν1ν2ν3ν4ν5ν3ν6ν2ν6ν6
XVCC5{4} −0.02 −0.03 0.00 0.01 0.00 0.03 0.33  0.06 0.01 
XVCC5(4) −0.02 −0.06 0.00 0.01 0.00 0.01 0.32 0.05 0.01 
XVCC5[4] −0.02 0.02 0.00 0.01 −0.00 0.04 0.26 0.06 0.01 
XVCC6{4} −0.001 0.000 0.001 0.001 −0.001 0.005 0.067 0.011 0.001 
XVCC6(4) −0.001 −0.002 0.000 0.001 −0.001 0.003 0.062 0.010 0.001 
XVCC6[4] −0.000 0.002 0.000 0.000 −0.002 0.006 0.052 0.008 0.000 
XVCC7{4} −0.0003 0.0012 0.0002 0.0004 0.0003 0.0010 0.0131 0.0021 0.0005 
XVCC7(4) −0.0002 0.0014 0.0002 0.0004 0.0003 0.0010 0.0124 0.0022 0.0005 
XVCC7[4] −0.0004 −0.0011 0.0002 0.0006 0.0007 −0.0005 0.0100 0.0004 0.0007 
XVCC8{4} 
XVCC8(4) 
XVCC8[4] 
a

XVCC for zero-point energy and EOM-XVCC for frequencies. Suffix “{4}” means the use of a harmonic reference, “(4)” means the use of a XVSCF(4) reference, and “[4]” means the use of a XVSCF[4] reference.

The ν5 fundamental is known to undergo Fermi resonances with combinations ν3ν6 and ν2ν6. For this mode, XVMP2[4] in the frequency-independent approximation shows a sign of divergent perturbation series, with a large error of −414 cm−1 (the VMP2 result is even worse).59 XVMP2(4) in the same approximation, on the other hand, is reasonable (with an error of 33 cm−1) but breaks down unexpectedly for nonresonant ν1 with an error of −191 cm−1.59 Therefore, the behavior of XVMP2 in the frequency-independent approximation is rather unpredictable. This method also lacks roots for Fermi-resonant counterparts.

Solving the Dyson equation with the frequency-dependent XVMP2 self-energy rectifies the aforementioned problems of the frequency-independent approximation.59 With this approach, the frequency of ν5 is reproduced within 3 cm−1 of VCI and the roots corresponding to Fermi-resonant counterparts, ν3ν6 and ν2ν6, can be located within 30 cm−1. The calculated results of ν1 also regain stability with an error of less than 10 cm−1. The price one pays is that a numerically cumbersome pole search needs to be performed along each mode.

Without any pole search, EOM-XVCC yields results that display stable convergence and have generally high accuracy for these strongly correlated modes. EOM-XVCCSDT seems to systematically outperform XVMP2 with the frequency-dependent self-energy for all listed modes including the Fermi resonances. EOM-XVCCSDTQ yields near-exact results with the largest error being merely 2.4 cm−1. Remarkably, EOM-XVCCSD, though lacking an Rˆ3 operator to fully describe cubic anharmonicity, locates all three Fermi-resonant transitions (ν5, ν3ν6, and ν2ν6) with uniform, medium accuracy. XVSCF and EOM-XVCCS, as expected, do not have roots corresponding to ν3ν6 and ν2ν6 and give rather poor results for ν1 and ν5.

The combinations and overtones accessible by XVMP2 with the frequency-dependent self-energy are limited to those that are in resonance with a fundamental, which may be an advantage in some cases, but a disadvantage in others. EOM-XVCCm does not have such a limitation and has roots for all low-lying states involving up to m modes and is accurate for those involving up to (m − 1) modes. This is analogous to the electronic case: EOM-CCSD has roots for up to two-electron excited states and is accurate for one-electron excitation energies; for accurate results for two-electron excitation energies, one needs to use EOM-CCSDT.89 Table IX attests to this general trend for two- and three-mode combinations or overtones. The errors in EOM-XVCCSD for two-mode excitations and those in EOM-XVCCSDT for three-mode excitations are on the same order of magnitude as those in EOM-XVCCS for nonresonant fundamentals (ν2, ν3, ν4, and ν6). EOM-XVCCSDTQ for three-mode excitations, EOM- XVCCSDT for two-mode excitations, and EOM-XVCCSD for nonresonant fundamentals have comparable, high accuracy.

TABLE IX.

The calculated frequencies of selected overtones and combinations (in cm−1) of the formaldehyde molecule. Except for VCI, the errors from VCI are shown.

Method(000200)(000101)(000002)(001100)(010100)(002000)(000300)(000201)
VCI 2318.1 2399.6 2472.5 2660.6 2871.2 3000.8 3464.2 3555.2 
EOM-XVCCSD[4] 35.7 28.3 31.0 77.7 38.3 23.2 … … 
EOM-XVCCSDT[4] 7.6 6.0 5.3 5.7 4.5 4.7 62.5 49.6 
EOM-XVCCSDTQ[4] 2.1 2.0 2.1 2.0 1.8 2.0 14.2 10.9 
EOM-XVCC5[4] 0.1 0.1 0.2 0.1 0.1 0.2 3.3 3.0 
EOM-XVCC6[4] 0.0 0.0 0.0 0.0 0.0 0.0 0.5 0.3 
EOM-XVCC7[4] 0.0 0.0 0.0 0.0 0.0 0.0 0.1 0.1 
EOM-XVCC8[4] 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 
Method(000200)(000101)(000002)(001100)(010100)(002000)(000300)(000201)
VCI 2318.1 2399.6 2472.5 2660.6 2871.2 3000.8 3464.2 3555.2 
EOM-XVCCSD[4] 35.7 28.3 31.0 77.7 38.3 23.2 … … 
EOM-XVCCSDT[4] 7.6 6.0 5.3 5.7 4.5 4.7 62.5 49.6 
EOM-XVCCSDTQ[4] 2.1 2.0 2.1 2.0 1.8 2.0 14.2 10.9 
EOM-XVCC5[4] 0.1 0.1 0.2 0.1 0.1 0.2 3.3 3.0 
EOM-XVCC6[4] 0.0 0.0 0.0 0.0 0.0 0.0 0.5 0.3 
EOM-XVCC7[4] 0.0 0.0 0.0 0.0 0.0 0.0 0.1 0.1 
EOM-XVCC8[4] 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 

For transition frequencies to strongly correlated states involving at least two modes (Fermi resonances, combinations, and overtones), the following is the order of accuracy: XVCCSD < XVMP2 with pole search < XVCCSDT < XVCCSDTQ < XVCC5 < XVCC6 < XVCC7 ≤ XVCC8. The methods not mentioned in the list either do not have the corresponding roots or tend to display signs of nonconvergence.

We have presented an accurate, stable, efficient, and widely applicable vibrational CC (XVCC) theory for both zero-point energies and transition frequencies, the latter through the EOM formalism, of a molecule with an anharmonic PES. It relies on a Hartree product of harmonic-oscillator modals as a reference wave function and a truncated Taylor-series PES. They lead to streamlined sets of equations defining XVCC and EOM-XVCC, which mirror exactly those of electronic CC and EOM-CC theories. XVCC and EOM-XVCC are diagrammatically size consistent, free of modal basis sets, and thus an inherently infinite-basis theory.

We have established the rules for enumerating Feynman– Goldstone diagrams for XVCC and EOM-XVCC theories and the ones for translating them into algebraic equations. This method of derivation is not only expedient but also insightful because the diagrammatic topology can more immediately determine size consistency of a theory or the lack thereof as well as the relationships among various diagrammatic theories. The rules are justified by the underlying normal-ordered second-quantized method of derivation, which gives the same equations as the diagrammatic derivation in each case.

XVCC and EOM-XVCC theories are general in the sense that their formalisms are unchanged for any single-reference Hartree product of harmonic-oscillator modals, including the wave functions of XVSCF(n), XVSCF[n], SCP, or the harmonic approximation. Reflecting the centrality of XVSCF (which includes SCP) as the size-consistent single-Hartree-product method in many-body vibrational theory, some diagrammatic contributions in XVCC and EOM-XVCC vanish in this reference, simplifying their equations further. Low-order XVCC and EOM-XVCC results are also the same or very close to the XVSCF results for the same reason.

The XVCCm and EOM-XVCCm methods up to m = 8 have been formulated and implemented by computer algebra of the aforementioned diagram logic as well as of algorithmic optimizations such as strength reduction, intermediate reuse, and factorization. The numerical tests of these methods show that the results are smoothly and rapidly convergent towards the exact values. This is in contrast to XVMP2 in the frequency-independent approximation to the self-energy, which displays signs of divergent perturbation series not only for resonant modes but also for nonresonant modes. While this problem is rectified by solving the Dyson equation with the frequency-dependent self-energy in XVMP2, EOM-XVCC may be said to achieve the same with a more well-studied diagonalization algorithm instead of a pole search. XVCCm and EOM-XVCCm with m = n, where n is the truncation rank of the Taylor-series PES, seem to be ones with the best cost-accuracy balance and hence the greatest practical utility.

The VCC theory of Prasad and coworkers27,32 can be identified as XVCC and EOM-XVCC with the XVSCF[n] reference. Our work is, therefore, built upon the previous studies of Prasad and coworkers,27,32 but it generalizes the theories to a variety of reference wave functions, establishes the normal-ordered, diagrammatic, and automated derivations, and implements higher-order methods up to connected octuple excitations using computerized code optimization and synthesis. Christiansen’s VCC theory,36 on the other hand, relies on VSCF and modals that are numerically defined and differs from XVCC; unlike XVCC, VCC likely retains many terms that are not size consistent and is expected to be fundamentally less efficient for large molecules and solids despite the exponential parameterization of its wave function. In VCC, the rank of disconnected excitations is also limited by the number of basis functions, which is not the case with basis-set-free XVCC.

VSCF is exact for any one-dimensional potential in the infinite basis-set limit, whereas XVSCF captures only a small portion of its anharmonic effect. For solids, VSCF reduces to XVSCF, the former giving the same results as the latter, but at a much greater computational cost. This suggests that Christiansen’s VCC may also perform well for small molecules with particularly strong anharmonicity, while reducing to more streamlined XVCC or the VCC theory of Prasad and coworkers in the thermodynamic limit. Our numerical results show that, for Morse-like anharmonicity, overtones and combinations, and anharmonic resonances, XVCC can nevertheless rectify the deficiency of XVSCF for small molecules. However, XVSCF and XVCC are expected to not work well for a double-well potential, for instance.

Furthermore, rectilinear coordinates such as normal coordinates, also underlying XVCC, may become awkward for large-amplitude vibrations. The methods that are suitable for such strong correlation and handle different coordinates flexibly include the multiconfiguration time-dependent Hartree (MCTDH) method,90 which is variational, but may lack size consistency.91 XVMP2 and XVCC are designed primarily for weak correlation in the ground state and are nonvariational and diagrammatically size consistent.91 Excited states are usually strongly correlated but can be handled by either class of methods.59 Therefore, it may be said that XVCC is not well suited for strong correlation in the ground state, but it is an excellent—arguably the best—size-consistent method for weak correlation in the ground state and for strong and weak correlation in excited states.

This work is supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Grant No. DE-FG02-11ER16211. J.A.F. is supported by the National Science Foundation Graduate Research Fellowship under Grant No. DGE-1144245. The authors thank Matthew R. Hermes for useful discussions and supplying the QFF’s and VCI results used in this work.

As before, we demand the XVCC wave function, eTˆ|Φ0, satisfy the vibrational Schrödinger equation,

(A1)

but only in the space of Hartree products reachable by Tˆ from |Φ0〉. Prior to projecting this equation onto this space, we first act eTˆ on both sides of the equation,

(A2)

The projection of this equation then yields the energy and amplitude equations, which read

(A3)
(A4)

for all ν1, …, νn and each n (1 ≤ nm). It may be understood that the projection space 〈Φ0|∪{〈Φν1ν2νn|} used in the derivation in the main text spans the same space Φ0|eTˆ{Φν1ν2νn|eTˆ} as above and, therefore, the two derivations lead to one and the same theory.

The similarity-transformed Hamiltonian, eTˆHˆNeTˆ, can be rewritten using the Baker–Campbell–Hausdorff expansion1,2 as

(A5)

which terminates with the n-fold commutator, where n is the truncation rank of the Taylor-series PES. The commutators immediately guarantee the diagrammatically connected structure of each surviving term. Therefore, we may write

(A6)

underscoring the equivalence of Eqs. (A3) and (A4) with Eqs. (29) and (40), respectively.

Here, we document the algebraic and diagrammatic forms of the XVCCSDT energy and amplitude equations in a QFF. The diagrams in the energy equation that are not already listed in the XVCCSD energy equation are shown in Fig. 16, and the corresponding algebraic expressions are given in Table X. As the numerical results illustrate (see above), the inclusion of connected triple excitations is critical for highly accurate zero-point energies. This is attributed to the importance of cubic force constants and diagram E9. The T1 and T2 amplitude equations are given by

(B1)
(B2)

where T1(2) and T2(2) are defined by Eqs. (51) and (53), respectively. The additional diagrams in the T1 and T2 amplitude equations are shown in Figs. 17 and 18. The algebraic interpretation of the diagrams are given in Tables XI and XII. The T3 amplitude equation is written as

(B3)

The diagrammatic form of this equation is shown in Fig. 19 with the corresponding algebraic interpretation in Table XIII.

TABLE X.

The energy equation of XVCCSDT in a QFF. Einstein summation convention is adopted.

ΔE0(3)=ΔE0(2)+13!Wijkτijk+13!Wijklτiτjkl 
ΔE0(3)=ΔE0(2)+13!Wijkτijk+13!Wijklτiτjkl 
FIG. 17.

Additional diagrams in the T1 amplitude equation of XVCCSDT in a QFF.

FIG. 17.

Additional diagrams in the T1 amplitude equation of XVCCSDT in a QFF.

Close modal
FIG. 18.

Additional diagrams in the T2 amplitude equation of XVCCSDT in a QFF.

FIG. 18.

Additional diagrams in the T2 amplitude equation of XVCCSDT in a QFF.

Close modal
TABLE XI.

The T1 amplitude equation of XVCCSDT in a QFF. Einstein summation convention is adopted.

T1(3)T1(2)+12!Wijτijν+13!Wνijkτijk+12!Wijkτiτjkν 
+12!2Wijklτijτklν+13!Wijklτiντjkl+12!2Wijklτiτjτklν 
= 0 
T1(3)T1(2)+12!Wijτijν+13!Wνijkτijk+12!Wijkτiτjkν 
+12!2Wijklτijτklν+13!Wijklτiντjkl+12!2Wijklτiτjτklν 
= 0 
TABLE XII.

The T2 amplitude equation of XVCCSDT in a QFF. Einstein summation convention is adopted.

T2(3) ≡ T2(2)+Pˆ(νμ)(12!Wiτiνμ+12!Wνijτijμ+12!Wijτiτjνμ 
  +12!Wνijkτiτjkμ+12!Wijkτiντjkμ+12!2Wijkτijτkνμ 
  +12!2Wijkτiτjτkνμ+12!3!Wijklτijkτlνμ+12!3Wijklτijντklμ 
  +12!2Wijklτiτjkτlνμ+12!Wijklτiτjντklμ+12!3!Wijklτiτjτkτlνμ) 
 
T2(3) ≡ T2(2)+Pˆ(νμ)(12!Wiτiνμ+12!Wνijτijμ+12!Wijτiτjνμ 
  +12!Wνijkτiτjkμ+12!Wijkτiντjkμ+12!2Wijkτijτkνμ 
  +12!2Wijkτiτjτkνμ+12!3!Wijklτijkτlνμ+12!3Wijklτijντklμ 
  +12!2Wijklτiτjkτlνμ+12!Wijklτiτjντklμ+12!3!Wijklτiτjτkτlνμ) 
 
FIG. 19.

Diagrams in the T3 amplitude equation of XVCCSDT in a QFF.

FIG. 19.

Diagrams in the T3 amplitude equation of XVCCSDT in a QFF.

Close modal
TABLE XIII.

The T3 amplitude equation of XVCCSDT in a QFF. Einstein summation convention is adopted.

T3(3) ≡ Pˆ(νμλ)(13!Wνμλ+13!Wνμλiτi+12!Wνμiτiλ+12!Wνi+τiμλ 
  +12!2Wνμijτijλ+12!Wνμijτiτjλ+12!Wνijτiμτjλ+12!Wνijτiτjμλ 
  +12!Wijτiντjμλ+12!Wνijkτiμτjkλ+12!2Wνijkτijτkμλ 
  +12!2Wijkτijντkμλ+12!Wνijkτiτjμτkλ+13!Wijkτiντjμτkλ 
  +12!2Wνijkτiτjτkμλ+12!Wijkτiτjντkμλ+12!2Wijklτiτjkντlμλ 
  +12!2Wijklτiντjμτklλ+12!2Wijklτijτkντlμλ+13!Wijklτiτjντkμτlλ 
  +12!2Wijklτiτjτkντlμλ) 
 
T3(3) ≡ Pˆ(νμλ)(13!Wνμλ+13!Wνμλiτi+12!Wνμiτiλ+12!Wνi+τiμλ 
  +12!2Wνμijτijλ+12!Wνμijτiτjλ+12!Wνijτiμτjλ+12!Wνijτiτjμλ 
  +12!Wijτiντjμλ+12!Wνijkτiμτjkλ+12!2Wνijkτijτkμλ 
  +12!2Wijkτijντkμλ+12!Wνijkτiτjμτkλ+13!Wijkτiντjμτkλ 
  +12!2Wνijkτiτjτkμλ+12!Wijkτiτjντkμλ+12!2Wijklτiτjkντlμλ 
  +12!2Wijklτiντjμτklλ+12!2Wijklτijτkντlμλ+13!Wijklτiτjντkμτlλ 
  +12!2Wijklτiτjτkντlμλ) 
 

The R1 and R2 amplitude equations of EOM-XVCCSD in a QFF are written as

(C1)
(C2)

where R1(1) is defined by Eq. (70). The additional diagrams in the R1 amplitude equation are depicted in Fig. 20. Their corresponding algebraic interpretation is given in Table XIV. The diagrammatic form of the R2 amplitude equation is drawn in Fig. 21, while the algebraic form is given in Table XV. EOM-XVCCSD provides the first sizable correction to the vibrational frequencies. This is attributed to the inclusion of diagrams such as S5, which account for the important cubic force constants that are not multiplied by near-zero T2 amplitudes. In particular, using a XVSCF[4] reference, EOM-XVCCS yields frequencies that are equivalent to the eigenvalues of the Wij+ matrix, completely neglecting cubic force constants. This is rectified by EOM-XVCCSD giving results with accuracy comparable to XVMP2.

TABLE XIV.

The R1 amplitude equation of EOM-XVCCSD in a QFF. Einstein summation convention is adopted.

R1(2) ≡ R1(1)+Wiriν+12!Wνijrij+Wijτjνri+Wijτjriν 
  +12!Wνijkτjkri+12!Wνijkτkrij+12!Wijkτjkriν+12!Wijkτkνrij 
  +Wijkτjτkνri+12!Wijkτjτkriν+12!Wijklτjkτlνri 
  +12!Wijklτkτlνrij+12!Wijklτjτklriν+12!Wijklτjτkτlνri 
  +13!Wijklτjτkτlriν 
 ωk(2)rν 
R1(2) ≡ R1(1)+Wiriν+12!Wνijrij+Wijτjνri+Wijτjriν 
  +12!Wνijkτjkri+12!Wνijkτkrij+12!Wijkτjkriν+12!Wijkτkνrij 
  +Wijkτjτkνri+12!Wijkτjτkriν+12!Wijklτjkτlνri 
  +12!Wijklτkτlνrij+12!Wijklτjτklriν+12!Wijklτjτkτlνri 
  +13!Wijklτjτkτlriν 
 ωk(2)rν 
FIG. 21.

Diagrams in the R2 amplitude equation of EOM-XVCCSD in a QFF.

FIG. 21.

Diagrams in the R2 amplitude equation of EOM-XVCCSD in a QFF.

Close modal
TABLE XV.

The R2 amplitude equation of EOM-XVCCSD in a QFF. Einstein summation convention is adopted.

R2(2) ≡ Pˆ(νμ)(12!Wνμiri+Wνi+riμ+12!2Wνμijrij+12!Wνμijτjri 
  +Wνijτjμri+Wνijτjriμ+Wijτjμriν+12!Wνijkτjkriμ 
  +12!Wνijkτkμrij+Wνijkτjτkμri+12!Wijkτjντkμri 
  +12!Wνijkτjτkriμ+Wijkτjτkμriν+12!Wijklτjkτlμriν 
  +12!2Wijklτkντlμrij+12!Wijklτjτkντlμri+12!Wijklτjτkτlμriν) 
 ωk(2)rνμ 
R2(2) ≡ Pˆ(νμ)(12!Wνμiri+Wνi+riμ+12!2Wνμijrij+12!Wνμijτjri 
  +Wνijτjμri+Wνijτjriμ+Wijτjμriν+12!Wνijkτjkriμ 
  +12!Wνijkτkμrij+Wνijkτjτkμri+12!Wijkτjντkμri 
  +12!Wνijkτjτkriμ+Wijkτjτkμriν+12!Wijklτjkτlμriν 
  +12!2Wijklτkντlμrij+12!Wijklτjτkντlμri+12!Wijklτjτkτlμriν) 
 ωk(2)rνμ 

Theorem.

The peak operation cost of XVCCm or EOM-XVCCm in an nth-order PES can be expressed as O(NC), where N is the number of vibrational degrees of freedom and C is an integer. The lowest value of C (Cmin) is m + ⌊n/2⌋ for mn and is not greater than n + ⌊(m − 1)/2⌋ for m < n.

Remark.

The proof is given below separately for mn and for m < n. In both cases, we first show that the aforementioned values of C are an upper bound; for any tensor contraction appearing in XVCCm or EOM-XVCCm, there is at least one contraction order whose cost function rank does not exceed the respective value. For mn, we additionally show that the aforementioned value of C is also a lower bound; there is always one tensor contraction whose minimal cost function rank is exactly this value. Together they prove that the expression of C given above is minimal for mn.

In what follows, we call an R amplitude of EOM-XVCC a T amplitude because they need not be distinguished in this context. We also only need to consider sequential contraction orders in which the force-constant tensor is always contracted first without losing rigor of the proof. Let T(k) be the kth T amplitude to be contracted, having Text(k) external lines and Tint(k) internal lines. Let I(k) be the kth intermediate. The symbol I(k) also denotes its size, i.e., the total number of lines. I(0) is the force-constant tensor, i.e., I(0) = n.

In the contraction of I(k−1) with T(k), the number of summation indexes is Tint(k), because two T amplitudes never share common internal lines and, therefore, all internal lines of T(k) must connect with I(k−1). The size of the product, i.e., the kth intermediate, is I(k1)+Text(k)Tint(k) and the cost function rank of the contraction is I(k1)+Text(k).

The theorem applies not only to XVCCm and EOM-XVCCm but also to other CC and EOM-CC methods for electrons and nucleons.

Lemma.

Let h be the total number of T amplitudes in a tensor contraction. Given intermediate I(i) (0 ≤ ih), there is at least one contraction order of remaining (hi) T amplitudes in which none of the subsequent intermediates has size exceeding max(I(i), I(h)).

Proof of the lemma.
The size of the kth intermediate is given by
(D1)
Let T(k) amplitudes be ordered as follows: T(k)’s with negative values of the difference in the external and internal lines (Text(k)Tint(k)) are contracted first, next by T(k)’s with zero difference, and finally by T(k)’s with positive values of the difference. Then, I(k) as a function of k for this contraction order is a convex function and thus cannot exceed the greater of their values at the end points, I(k) ≤ max(I(i), I(h)).

Proof of the theorem for mn.

First, we show that C = m + ⌊n/2⌋ is an upper bound. We consider the following two cases: (1) there is at least one T(k) with Tint(k)n/2; (2) otherwise.

In case 1, the T amplitude with Tint(1)n/2 is contracted first with the force-constant tensor. The cost function rank of this contraction is given by

(D2)

where we use Text(1)+Tint(1)m in the first inequality. The size of the resulting intermediate is

(D3)

where Eq. (D2) is used in the first inequality.

Next, we consider case 2 and the second and subsequent contractions in case 1 together. In the former, Tint(k)n/2 for all k. In the latter, Tint(k)n/2 for all k ≥ 2. Here, we contract T amplitudes in one of the orders of the lemma. As per the lemma, this guarantees I(k)m in both cases [see Eq. (D3) in case 1]. The sizes of the kth and (k − 1)th intermediates are related to each other by

(D4)

Therefore, the cost function rank of the kth contraction is

(D5)

Equations (D2) and (D5) prove Cminm + ⌊n/2⌋.

Second, we show that the same value of C is a lower bound. There is always a binary product of the force-constant tensor, I(0), and a Tm amplitude with ⌈n/2⌉ internal lines. Note that ⌈n/2⌉ cannot exceed n or m because mn ≥ 1. Furthermore, the number of external lines is m + n − 2⌈n/2⌉ ≤ m. Hence, this is a valid tensor contraction. The cost function rank of this contraction is

(D6)

which proves Cminm + ⌊n/2⌋.

Together, we obtain Cmin = m + ⌊n/2⌋ for mn.

Remark.

This is an example of the optimization problem in computer science known as the matrix chain multiplication.84,92,93 In this case (mn), we have shown that there is an analytical solution that minimizes the exponent of the leading-order cost function and its value.

Proof of the theorem for m  <  n.

We contract T amplitudes in one of the orders of the lemma. As per the lemma, this ensures I(k)n. We consider the following three cases in each contraction: (1) Text(k)<Tint(k), (2) Text(k)>Tint(k), and (3) Text(k)=Tint(k).

In case 1, since Text(k)+Tint(k)m, we have Text(k)(m1)/2. The cost function rank of the kth contraction is

(D7)

In case 2, similarly, Tint(k)(m1)/2 holds. The size of the kth intermediate is given by

(D8)

which implies

(D9)

In case 3, m must be even and Text(k)=Tint(k)=m/2. The intermediate resulting from the contraction of this T amplitude has the size,

(D10)

The first inequality follows from the fact that I(l) is a convex function with its minimum occurring at l = k, where Text(k)Tint(k)=0. The cost function rank of this step is

(D11)

Equations (D7), (D9), and (D11) prove Cminn + ⌊(m − 1)/2⌋ for m < n.

Remark.

This upper-bound expression for m < n happens to agree with the optimal cost function ranks for all cases of m < n in Table IV. However, we have identified many other combinations of m < n whose optimal ranks are lower.

1.
R. J.
Bartlett
and
M.
Musiał
,
Rev. Mod. Phys.
79
,
291
(
2007
).
2.
I.
Shavitt
and
R. J.
Bartlett
,
Many-Body Methods in Chemistry and Physics: MBPT and Coupled-Cluster Theory
(
Cambridge University Press
,
Cambridge
,
2009
).
4.
F.
Coester
and
H.
Kümmel
,
Nucl. Phys.
9
,
225
(
1958
).
5.
J.
Čížek
,
J. Chem. Phys.
45
,
4256
(
1966
).
6.
J.
Paldus
,
J.
Čížek
, and
I.
Shavitt
,
Phys. Rev. A
5
,
50
(
1972
).
7.
J. A.
Pople
,
R.
Krishnan
,
H. B.
Schlegel
, and
J. S.
Binkley
,
Int. J. Quantum Chem.
14
,
545
(
1978
).
8.
R. J.
Bartlett
and
G. D.
Purvis
,
Int. J. Quantum Chem.
14
,
561
(
1978
).
9.
G. D.
Purvis
III
and
R. J.
Bartlett
,
J. Chem. Phys.
76
,
1910
(
1982
).
10.
S.
Hirata
,
Theor. Chem. Acc.
129
,
727
(
2011
).
11.
S.
Hirata
,
M.
Keçeli
,
Y.
Ohnishi
,
O.
Sode
, and
K.
Yagi
,
Annu. Rev. Phys. Chem.
63
,
131
(
2012
).
12.
R. F.
Bishop
,
Theor. Chim. Acta
80
,
95
(
1991
).
13.
W.
Förner
,
Int. J. Quantum Chem.
43
,
221
(
1992
).
14.
M.
Yu
,
S.
Kalvoda
, and
M.
Dolg
,
Chem. Phys.
224
,
121
(
1997
).
15.
P.
Reinhardt
,
Theor. Chem. Acc.
104
,
426
(
2000
).
16.
S.
Hirata
,
I.
Grabowski
,
M.
Tobita
, and
R. J.
Bartlett
,
Chem. Phys. Lett.
345
,
475
(
2001
).
17.
S.
Hirata
,
R.
Podeszwa
,
M.
Tobita
, and
R. J.
Bartlett
,
J. Chem. Phys.
120
,
2581
(
2004
).
18.
J. J.
Shepherd
,
T. M.
Henderson
, and
G. E.
Scuseria
,
J. Chem. Phys.
140
,
124102
(
2014
).
19.
D. J.
Dean
,
J. R.
Gour
,
G.
Hagen
,
M.
Hjorth-Jensen
,
K.
Kowalski
,
T.
Papenbrock
,
P.
Piecuch
, and
M.
Wloch
,
Nucl. Phys. A
752
,
299C
(
2005
).
20.
P.
Piecuch
,
M.
Wloch
,
J. R.
Gour
,
D. J.
Dean
,
M.
Hjorth-Jensen
, and
T.
Papenbrock
,
AIP Conf. Proc.
777
,
28
(
2005
).
21.
O.
Christiansen
,
Phys. Chem. Chem. Phys.
9
,
2942
(
2007
).
22.
M. R.
Hermes
and
S.
Hirata
,
Int. Rev. Phys. Chem.
34
,
71
(
2015
).
23.
U. B.
Kaulfuss
and
M.
Altenbokum
,
Phys. Rev. D
33
,
3658
(
1986
).
24.
R. F.
Bishop
and
M. F.
Flynn
,
Phys. Rev. A
38
,
2211
(
1988
).
25.
R. F.
Bishop
,
M. C.
Bosca
, and
M. F.
Flynn
,
Phys. Rev. A
40
,
3484
(
1989
).
26.
M. D.
Prasad
,
J. Chem. Phys.
88
,
7005
(
1988
).
27.
V.
Nagalakshmi
,
V.
Lakshminarayana
,
G.
Sumithra
, and
M. D.
Prasad
,
Chem. Phys. Lett.
217
,
279
(
1994
).
28.
G. M.
Sastry
and
M. D.
Prasad
,
Chem. Phys. Lett.
228
,
213
(
1994
).
29.
G.
Latha
and
M. D.
Prasad
,
Chem. Phys. Lett.
241
,
215
(
1995
).
30.
G. S.
Latha
and
M. D.
Prasad
,
J. Chem. Phys.
105
,
2972
(
1996
).
31.
M. D.
Prasad
,
Int. J. Mol. Sci.
3
,
447
(
2002
).
32.
S.
Banik
,
S.
Pal
, and
M. D.
Prasad
,
J. Chem. Phys.
129
,
134111
(
2008
).
33.
S.
Banik
,
S.
Pal
, and
M. D.
Prasad
,
J. Chem. Theory Comput.
6
,
3198
(
2010
).
34.
S.
Banik
,
S.
Pal
, and
M. D.
Prasad
,
J. Chem. Phys.
137
,
114108
(
2012
).
35.
M. R.
Hermes
,
M.
Keçeli
, and
S.
Hirata
,
J. Chem. Phys.
136
,
234109
(
2012
).
36.
O.
Christiansen
,
J. Chem. Phys.
120
,
2149
(
2004
).
37.
O.
Christiansen
,
J. Chem. Phys.
122
,
194105
(
2005
).
38.
P.
Seidler
and
O.
Christiansen
,
J. Chem. Phys.
126
,
204101
(
2007
).
39.
P.
Seidler
,
M. B.
Hansen
, and
O.
Christiansen
,
J. Chem. Phys.
128
,
154113
(
2008
).
40.
P.
Seidler
,
E.
Matito
, and
O.
Christiansen
,
J. Chem. Phys.
131
,
034115
(
2009
).
41.
P.
Seidler
and
O.
Christiansen
,
J. Chem. Phys.
131
,
234109
(
2009
).
42.
P.
Seidler
,
M.
Sparta
, and
O.
Christiansen
,
J. Chem. Phys.
134
,
054119
(
2011
).
43.
K. M.
Christoffel
and
J. M.
Bowman
,
Chem. Phys. Lett.
85
,
220
(
1982
).
44.
L. S.
Norris
,
M. A.
Ratner
,
A. E.
Roitberg
, and
R. B.
Gerber
,
J. Chem. Phys.
105
,
11261
(
1996
).
45.
J. M.
Bowman
,
J. Chem. Phys.
68
,
608
(
1978
).
46.
M. A.
Ratner
and
R. B.
Gerber
,
J. Phys. Chem.
90
,
20
(
1986
).
47.
J. M.
Bowman
,
Acc. Chem. Res.
19
,
202
(
1986
).
48.
S.
Hirata
,
M.
Keçeli
, and
K.
Yagi
,
J. Chem. Phys.
133
,
034109
(
2010
).
49.
M.
Keçeli
and
S.
Hirata
,
J. Chem. Phys.
135
,
134108
(
2011
).
50.
M. R.
Hermes
and
S.
Hirata
,
J. Phys. Chem. A
117
,
7179
(
2013
).
51.
N.
Makri
,
J. Phys. Chem. B
103
,
2823
(
1999
).
52.
S.
Hirata
and
M. R.
Hermes
,
J. Chem. Phys.
141
,
184111
(
2014
).
53.
54.
T. R.
Koehler
,
Phys. Rev. Lett.
17
,
89
(
1966
).
55.
P.
Choquard
,
The Anharmonic Crystal
(
Benjamin
,
New York
,
1967
).
56.
N. S.
Gillis
,
N. R.
Werthamer
, and
T. R.
Koehler
,
Phys. Rev.
165
,
951
(
1968
).
57.
T.
Luty
,
A.
van der Avoird
,
R. M.
Berns
, and
T.
Wasiutynski
,
J. Chem. Phys.
75
,
1451
(
1981
).
58.
K.
Yagi
,
K.
Hirao
,
T.
Taketsugu
,
M. W.
Schmidt
, and
M. S.
Gordon
,
J. Chem. Phys.
121
,
1383
(
2004
).
59.
M. R.
Hermes
and
S.
Hirata
,
J. Chem. Phys.
139
,
034111
(
2013
).
60.
K.
Yagi
,
M.
Keçeli
, and
S.
Hirata
,
J. Chem. Phys.
137
,
204118
(
2012
).
61.
B.
Thomsen
,
K.
Yagi
, and
O.
Christiansen
,
J. Chem. Phys.
140
,
034111
(
2014
).
62.
X.
Cheng
and
R. P.
Steele
,
J. Chem. Phys.
141
,
104105
(
2014
).
63.
64.
M. R.
Hermes
and
S.
Hirata
,
J. Chem. Phys.
141
,
084105
(
2014
).
65.
J.
Goldstone
,
Proc. R. Soc. A
239
,
267
(
1957
).
66.
See supplementary material at http://dx.doi.org/10.1063/1.4931472 for the higher-order XVCC and EOM-XVCC equations.
69.
H.
Sekino
and
R. J.
Bartlett
,
Int. J. Quantum Chem.
S18
,
255
(
1984
).
70.
J.
Geertsen
,
M.
Rittby
, and
R. J.
Bartlett
,
Chem. Phys. Lett.
164
,
57
(
1989
).
71.
D. C.
Comeau
and
R. J.
Bartlett
,
Chem. Phys. Lett.
207
,
414
(
1993
).
72.
J. F.
Stanton
and
R. J.
Bartlett
,
J. Chem. Phys.
98
,
7029
(
1993
).
73.
H. J.
Monkhorst
,
Int. J. Quantum Chem.
12
(
S11
),
421
(
1977
).
74.
S.
Ghosh
,
D.
Mukherjee
, and
S.
Bhattacharyya
,
Mol. Phys.
43
,
173
(
1981
).
75.
M.
Takahashi
and
J.
Paldus
,
J. Chem. Phys.
85
,
1486
(
1986
).
76.
H.
Koch
and
P.
Jørgensen
,
J. Chem. Phys.
93
,
3333
(
1990
).
77.
H.
Koch
,
H. J. A.
Jensen
,
P.
Jørgensen
, and
T.
Helgaker
,
J. Chem. Phys.
93
,
3345
(
1990
).
78.
R. J.
Rico
and
M.
Head-Gordon
,
Chem. Phys. Lett.
213
,
224
(
1993
).
79.
D.
Knoll
and
D.
Keyes
,
J. Comput. Phys.
193
,
357
(
2004
).
80.
C.
Lanczos
,
J. Res. Natl. Bur. Stand.
45
,
255
(
1950
).
81.
E. R.
Davidson
,
J. Comput. Phys.
17
,
87
(
1975
).
82.
K.
Hirao
and
H.
Nakatsuji
,
J. Comput. Phys.
45
,
246
(
1982
).
83.
D.
Sorensen
,
R.
Lehoucq
,
C.
Yang
, and
K.
Maschhoff
,
arpack
(
Rice University
,
1996-2008
).
84.
S.
Hirata
,
J. Phys. Chem. A
107
,
9887
(
2003
).
85.
T.
Shiozaki
,
M.
Kamiya
,
S.
Hirata
, and
E. F.
Valeev
,
Phys. Chem. Chem. Phys.
10
,
3358
(
2008
).
86.
M.
Valiev
,
E. J.
Bylaska
,
N.
Govind
,
K.
Kowalski
,
T. P.
Straatsma
,
H. J. J.
Van Dam
,
D.
Wang
,
J.
Nieplocha
,
E.
Apra
,
T. L.
Windus
, and
W. A.
de Jong
,
Comput. Phys. Commun.
181
,
1477
(
2010
).
87.
K.
Yagi
,
sindo
(
The University of Tokyo
,
2006
).
88.
M.
Keçeli
,
mavi
(
University of Illinois at Urbana-Champaign
,
2011
).
89.
S.
Hirata
,
M.
Nooijen
, and
R. J.
Bartlett
,
Chem. Phys. Lett.
326
,
255
(
2000
).
90.
M. H.
Beck
,
A.
Jäckle
,
G. A.
Worth
, and
H. D.
Meyer
,
Phys. Rep.
324
,
1
(
2000
).
91.
S.
Hirata
and
I.
Grabowski
,
Theor. Chem. Acc.
133
,
1440
(
2014
).
92.
A. A.
Auer
,
G.
Baumgartner
,
D. E.
Bernholdt
,
A.
Bibireata
,
V.
Choppella
,
D.
Cociorva
,
X. Y.
Gao
,
R.
Harrison
,
S.
Krishnamoorthy
,
S.
Krishnan
,
C. C.
Lam
,
Q. D.
Lu
,
M.
Nooijen
,
R.
Pitzer
,
J.
Ramanujam
,
P.
Sadayappan
, and
A.
Sibiryakov
,
Mol. Phys.
104
,
211
(
2006
).
93.
A.
Hartono
,
Q. D.
Lu
,
T.
Henretty
,
S.
Krishnamoorthy
,
H. J.
Zhang
,
G.
Baumgartner
,
D. E.
Bernholdt
,
M.
Nooijen
,
R.
Pitzer
,
J.
Ramanujam
, and
P.
Sadayappan
,
J. Phys. Chem. A
113
,
12715
(
2009
).

Supplementary Material