A new perturbative approach to canonical equation-of-motion coupled-cluster theory is presented using coupled-cluster perturbation theory. A second-order Møller-Plesset partitioning of the Hamiltonian is used to obtain the well known equation-of-motion many-body perturbation theory equations and two new equation-of-motion methods based on the linear coupled-cluster doubles and linear coupled-cluster singles and doubles wavefunctions. These new methods are benchmarked against very accurate theoretical and experimental spectra from 25 small organic molecules. It is found that the proposed methods have excellent agreement with canonical equation-of-motion coupled-cluster singles and doubles state for state orderings and relative excited state energies as well as acceptable quantitative agreement for absolute excitation energies compared with the best estimate theory and experimental spectra.

Optical spectroscopy is an important and ubiquitous tool in modern experimental studies. Improvements over the years in the use of lasers in optical spectroscopy has brought the field to an impressive level of accuracy and precision that is difficult to match using modern theoretical methods. Even so, theory has an important role to play in optical spectroscopy. By assisting in excited state assignments and predicting the density of excited states expected in a given energy range, computational results can be a powerful tool for spectroscopists.

Widespread use of time dependent density functional theory1 (TD-DFT) for the calculation of excited states will continue for a long time to come due to the attractive low computational scaling inherent to the method. However, reliability and accuracy problems2 in TD-DFT keep the theory from replacing the more computationally expensive but highly accurate multireference configuration interaction3 (MRCI) or equation of motion coupled-cluster4,5 (EOM-CC) theories. The most commonly used form of EOM-CC theory is the equation-of-motion coupled-cluster singles and doubles (EOM-CCSD) variant,6 which limits ground and excited state excitations to singles and doubles only. This approximation is qualitatively consistent and approaches quantitative accuracy in many cases where single excitations are dominant. For doubly excited states, perturbative7 (T) or iterative8 (T-n) triple excitations are required9,10 to obtain quantitative agreement with experiment.

The development of approximate excited state methods based on many-body perturbation and coupled-cluster theory has been an active and fruitful endevour since the emergence of the field.11 Many approaches make use of a mean-field starting reference based on configuration interaction singles12 and add perturbative corrections to that reference.13–18 Other approaches use a many-body treatment of propagator theory such as the algebraic-diagrammatic construction19,20 (ADC) scheme or by using approximate coupled-cluster linear-response21–23 theory with the CCn methods.24 Approximate methods based on equation-of-motion theory can either add a post hoc EOM-CC perturbative correction7,25,26 for triples or an a priori perturbative approximation.27–29 We will use perturbation theory here to develop a new perturbative EOM-CC method that also includes infinite-order effects in the ground state wavefunction. The goal will be to provide a consistent and accurate way to obtain excitation energies from either a linear coupled-cluster doubles (LCCD) or coupled-cluster singles and doubles (LCCSD) ground state wavefunction. Thus providing a route toward obtaining electronically excited states from alternative ansatz ground state wavefunctions such as the related molecular cluster perturbation theory embedding method30 which dissociates to the LCCD limit.

In this work, we will use coupled-cluster perturbation theory31 (CCPT) to derive a general EOM-CCPT framework in Sec. II. The various EOM-CC approximations will then be systematically derived from EOM-CCPT by a choice in the Hamiltonian perturbation partitioning. Numerical results of these approximations compared against very accurate theoretical and experimental spectra are presented in Sec. IV.

Central to coupled-cluster theory4 is the similarity transformation of the electronic Hamiltonian by the exponential wave operator,

(1)

The excitation cluster operators (here limited to single and double excitations) are given by

(2)

and the normal ordered electronic Hamiltonian is

(3)
(4)

Here f is the one-electron Fock matrix, v ̄ the antisymmetrized two-electron integrals, {⋯} denotes normal ordering of the enclosed operators, and the one- ( F ˆ ) and two-particle ( W ˆ ) are defined appropriately. Throughout we reserve the indices i, j, k, … and a, b, c, … for occupied and virtual orbitals, respectively, while the indices p, q, … may refer to either occupied or virtual orbitals. Because the Hamiltonian only contains one- and two-particle operators, the Baker-Campbell-Hausdorff expansion of Eq. (1) naturally truncates after four commutators giving

(5)

Starting with an appropriate single-reference mean-field reference, |ϕ0〉, the coupled-cluster Schrödinger equation is expressed with Eq. (5) as

(6)

The CC correlation energy given by projecting on the left by the reference function

(7)

while the necessary cluster amplitudes are obtained by solving the equations generated by projecting on the left with the auxiliary space

(8)

where 〈ϕg| is the set of g-fold excited determinants

(9)

The conventional approach to computing excited states through the equation-of-motion method,4 which directly computes the kth excitation energy (ωk) relative to the ground state coupled-cluster wavefunction, is also obtained with Eq. (5) as

(10)

Here, ωk is the kth general eigenvalue, and R ˆ ( k ) ( L ˆ ( k ) ) is the kth right (left) general eigenvector solution to the non-Hermitian matrix, H ̄ . The 〈…〉C denotes that only fully connected contractions are included. These eigenvector solutions are defined as the linear excitation and de-excitation operators (limited again to only singles and doubles),

(11)
(12)
(13)
(14)

with the biorthogonalization constraint

(15)

Note that because R ˆ ( k ) is an excitation operator, it will commute with the ground state cluster amplitudes,

(16)

As is always possible in Hamiltonian based perturbation theory we can represent the complete Hamiltonian in terms of a zeroth order Hamiltonian H ˆ 0 and perturbation V ˆ

(17)

where α → 1 is an order parameter used for convenience. Equation (17) can be directly inserted into any coupled-cluster similarity transformation or energy functional and expanded to the desired order in α. Examples other than the usual form given by Eq. (5) would include the Hermitian32,33 ( e T ˆ H ˆ e T ˆ ) or unitary34 ( e τ ˆ H ˆ e τ ˆ , where τ ˆ = T ˆ T ˆ ) expansions, but they do not terminate without truncation.

To facilitate the decomposition into a reference and perturbation operator, we rearrange the Hamiltonian into particle excitation rank form

(18)

where the [n] superscript denotes that the operator changes particle rank from right to left by n. From this point the CC perturbation framework used is the same as we have employed in the past.18,30,31 Here the exponential operator e T ˆ is expanded to the appropriate order in terms of nth order cluster operators T ˆ ( n ) . The computation of these perturbative cluster operators is done order by order using standard perturbation theory. This is easily illustrated by expanding Eq. (5) using the partitioned Hamiltonian of Eq. (17). The similarity transformed Hamiltonian to arbitrary order in α is given by

(19)

Throughout this work we denote the exclusive order in α of an operator with the superscript (n) and we use {n} to denote an expansion inclusive of all orders in α up to n. With this notation, the general nth order similarity transformed Hamiltonian can be expressed as a compact series

(20)

The nth order ground state CCPT cluster amplitudes are completely defined from insertion of the exclusive similarity transformed Hamiltonian, Eq. (19), into (8),

(21)

Using these nth order ground state amplitudes in Eq. (7) will give the n + 1 order correlation energy

(22)

beginning with n = 2. From this the exact coupled-cluster correlation energy, relative to the mean-field energy

(23)

is

(24)

or equivalently

(25)
(26)

To obtain excitation energies relative to the nth order correlation energy without perturbatively expanding the L ˆ and R ˆ operators, we use the nth order inclusive similarity transformed Hamiltonian (Eq. (20)) in Eq. (10) rather than the exclusive Eq. (19), giving

(27)

Contributions from the nth order wavefunction to the spectrum obtained from Eq. (27) are possible. However, any such nth order amplitude will be neglected if it does not contribute to the corresponding nth order energy.

We have used the particle rank conserving partitioning of the Hamiltonian,

(28)
(29)

in previous work18,30,31 to great success. Alternatively we can also choose a partitioning scheme that introduces linear coupling between cluster amplitudes, or the traditional Møller-Plesset (MP) partitioning which we will discuss next.

The standard approach in electronic structure theory is to use generalized many-body perturbation theory (GMBPT),35 based on the MP partitioning of the Hamiltonian, given by

(30)
(31)

Using Eqs. (30) and (31) in Eq. (19), the first two orders of H ̄ ( n ) are, respectively,

(32)

and

(33)

Inserting Eq. (32) into Eq. (21) gives the quintessential GMBPT(1) wavefunction,

(34)
(35)

from which

(36)

The GMBPT(2) wavefunction is obtained from Eq. (33) similarly. The inclusive second-order similarity transformed Hamiltonian,

(37)

is the GMBPT generalization of previous27–29 MBPT approximations to H ̄ . The spin-orbital equations for the H ̄ { 2 } matrix elements27,28 using Einstein notation are

(38)
(39)
(40)
(41)
(42)
(43)
(44)
(45)
(46)
(47)
(48)
(49)
(50)
(51)
(52)
(53)

where P(pq) is the anti-permutation operator

(54)

The difference between GMBPT and MP theory occurs when f i a , f j i , and f b a 0 as they would be for canonical Hartree-Fock (HF) orbitals. The first introduces a first-order singles wavefunction (Eq. (34)), while the second and third terms have to be summed to all order as in coupled-cluster theory or better transformed away by a semi-canonical occupied-occupied and virtual-virtual rotation.

The first-order similarity transformed Hamiltonian using the alternative particle rank conserving partitioning of the Hamiltonian (denoted PR0 for brevity, see Eqs. (28) and (29)) is

(55)

Using Eq. (55) in the amplitude projection Equation (21) yields the decoupled, linear-amplitude wavefunction equations

(56)
(57)

The PR0 wavefunction is infinite-order in W ˆ [ 0 ] , which is an advantage for describing potential energy surfaces31,36 and systematically including intermolecular many-body perturbations.30 The second-order similarity transformed Hamiltonian is

(58)

from which the second-order amplitudes are computed. Starting from a canonical Hartree-Fock reference eliminates T ˆ 1 ( 1 ) in Eqs. (55) and (58), yielding the LCCD similarity transformed Hamiltonian

(59)

In this form, the matrix elements of Eq. (59) are identical to Eq. (37).

In order to add linear coupling between cluster operators of different rank (e.g. coupling singles and doubles) the alternative choice in Hamiltonian partitioning (denoted PR1),

(60)
(61)

can be used. From Eqs. (60) and (61) the first- and second-order similarity transformed Hamiltonians are, in operator form, the same as Eqs. (55) and (58). The motivation for this choice is evident when examining the first-order amplitude equations

(62)
(63)
(64)

As can be seen, the PR1 partitioning choice immediately provides the linearized CCSDT wavefunction ansatz. It is here that we choose to only include linear singles and doubles in the ground state, resulting in

(65)

which is the LCCSD similarity transformed Hamiltonian. As with LCCD, the matrix elements of Eq. (65) are identical to Eq. (37) differing only by the T ˆ amplitudes.

To assess both the quality and general behavior of these EOM approximations when computing vertical singlet excitation energies, we employ two established collections of gas phase molecules (see Figure 1): the set of 24 organic molecules from Schreiber et al.37–39 (referred to hereafter as the Mülheim set) and the set of 11 organic molecules from Caricato et al.2,40 (referred to here as the Yale set). This Mülheim set contains 121 reference singlet excitation energies37 which use molecular geometries obtained with an MP2/6-31G* optimization in which the first row 1s atomic orbitals are dropped. The extent and balanced nature of the Mülheim set readily lends itself to obtaining statistics between different theoretical methods. To compare the approximate EOM-CC methods proposed here, we take our best theoretical reference to be EOM-CCSDT-3,41,56 which is known to give both accurate results and is a systematic EOM-CC type theory which will provide a balanced comparison. The Yale set choice in molecules is a subset of the Mülheim set with the addition of acetaldehyde but considers a different selection of excitation energies (69 in all) whose reference values are taken from gas phase experiments (the experiment literature references can be found in Ref. 2). This comparison to experiment provides a different and valuable perspective with which to augment theory vs. theory examinations. Geometries for the Yale set are obtained from an MP2/6-311+G** optimization.

FIG. 1.

Member molecules of the Mülheim test set. Names denote with an are also members of the Yale set, with acetaldehyde only a member of the Yale set.

FIG. 1.

Member molecules of the Mülheim test set. Names denote with an are also members of the Yale set, with acetaldehyde only a member of the Yale set.

Close modal

Calculations using the Mülheim set typically37 use the TZVP basis set (3s1p on hydrogen 5s3p1d on first row atoms) from Schäfer et al.42 This is a light weight basis set that allows the use of many different kinds of computational methods with only a modest cost in overall accuracy.39 While perfectly acceptable when comparing different methods to each other, this basis is not accurate enough to compare with experimental results. This is especially true when calculating Rydberg states, where the inclusion of many diffuse functions is key.43,44 When computing the excitation energies for the Yale set, we use the doubly diffuse function augmented Dunning basis,45,46 d-aug-cc-pVDZ (4s3p on hydrogen and 5s4p3d on first row atoms). This basis is comparable to the 6-311(3+,3+)G** basis used by Caricato et al. in their calculations2,40,47 with almost no gain in including a third set of diffuse functions.48 

All of the reported electronic structure results were obtained using the serial ACESII49 and parallel Aces450ab initio quantum chemistry packages. Calculations were performed on the University of Florida HiPerGator high performance cluster. Throughout this paper we use RMS to mean “root mean square,” MAD to mean “mean absolute deviation,” SD to mean “standard deviation,” and MD to mean “absolute maximum deviation.”

Developed two decades ago for canonical Hartree-Fock27 EOM-MBPT(2) is a well established perturbation method which has been largely ignored by the community in favor of other approximations such as CC224 or ADC(2).19 The EOM-MBPT(2) method remains relevant because of the non-iterative O(n5) computational cost of the ground-state wavefunction and the general accuracy of MBPT(2). It is also reasonably accurate for equilibrium vertical excitation energies. This accuracy, within the Mülheim test set, is illustrated in the correlation Figures 2(a), 3(a), and Table V. Compared to the standard EOM-CCSD and EOM-CCSDT-3 methods, EOM-MBPT(2) has an MAD of 0.13 eV and 0.31 eV, respectively, consistent with the known ∼0.2 eV triples correction to EOM-CCSD.

FIG. 2.

Comparison of vertical singlet excitations against the Mülheim data set EOM-CCSD values.

FIG. 2.

Comparison of vertical singlet excitations against the Mülheim data set EOM-CCSD values.

Close modal
FIG. 3.

Comparison of vertical singlet excitations against the Mülheim data set EOM-CCSDT-3 values.

FIG. 3.

Comparison of vertical singlet excitations against the Mülheim data set EOM-CCSDT-3 values.

Close modal
TABLE V.

Vertical excitation energy comparison statistics for the singlet Mülheim test set. Units are in electron volts (eV), EOM-CCSD, and EOM-CCSDT-3 reference values were taken from Sous et al.55 

Mülheim singlet set vertical excitation statistics
MBPT(2) LCCD LCCSD CCSD
Relative to EOM-CCSD 
MAD  0.13  0.25  0.35 
RMS  0.15  0.28  0.36 
SD  0.14  0.16  0.13 
MD  0.37  0.55  0.80 
Relative to EOM-CCSDT-3 
MAD  0.31  0.45  0.56  0.21 
RMS  0.38  0.53  0.60  0.29 
SD  0.25  0.28  0.23  0.20 
MD  1.19  1.49  1.37  1.02 
Mülheim singlet set vertical excitation statistics
MBPT(2) LCCD LCCSD CCSD
Relative to EOM-CCSD 
MAD  0.13  0.25  0.35 
RMS  0.15  0.28  0.36 
SD  0.14  0.16  0.13 
MD  0.37  0.55  0.80 
Relative to EOM-CCSDT-3 
MAD  0.31  0.45  0.56  0.21 
RMS  0.38  0.53  0.60  0.29 
SD  0.25  0.28  0.23  0.20 
MD  1.19  1.49  1.37  1.02 

The EOM-MBPT(2) approximation and its Löwdin partitioned variant29,51 was recently benchmarked using the Yale test set (using the 6-311(3+,3+)G** basis set) against EOM-CCSD and experiment by Goings et al.47 Our computed EOM-MBPT(2)/d-aug-cc-pVDZ values, presented in Tables IIV, yield nearly identical results. The EOM-MBPT(2) method is completely satisfactory when compared to experimental Rydberg states (such as with the alkenes and carbonyl’s) but is much less reliable for the nπ and ππ heterocycle valence states. The comparative difficulty in describing valence states is a limitation of EOM theory in general which is only satisfactorily overcome by using methods that include triples or alternative Hamiltonians such as similarity transformed equation-of-motion coupled-cluster theory52–55 (STEOM-CC).

TABLE I.

Vertical singlet excitation energies for a few small alkene molecules computed with the d-aug-cc-pVDZ basis using MP2/6-311+G** geometries. Units are in electron volts (eV), experimental references can be found in Ref. 2.

Yale alkene subset vertical excitations
State MBPT(2) LCCD LCCSD CCSD Expt.
Ethylene 
B3u  7.15  7.43  7.45  7.28  7.11 
B1u  7.76  8.13  8.18  7.97  7.65 
B1g  7.79  8.08  8.10  7.97  7.80 
B2g  7.84  8.12  8.14  7.93  7.90 
Ag  8.19  8.46  8.48  8.31  8.28 
B3u  8.64  8.92  8.94  8.77  8.62 
B3u  8.93  9.20  9.22  9.05  8.90 
B3u  9.05  9.33  9.35  9.17  9.08 
B1g  9.19  9.46  9.49  9.34  9.20 
B1u  9.18  9.47  9.50  9.32  9.33 
B3u  9.84  10.11  10.13  9.96  9.51 
Isobutene 
B1  6.29  6.54  6.56  6.38  6.17 
A1  6.81  7.06  7.10  6.91  6.70 
trans-1,3-butadiene 
Bu  6.12  6.54  6.63  6.29  5.91 
Bg  6.17  6.47  6.49  6.24  6.22 
Au  6.47  6.79  6.80  6.55  6.66 
Bu  7.06  7.39  7.43  7.16  7.07 
Bg  7.28  7.59  7.60  7.36  7.36 
Ag  7.54  7.85  7.85  7.61  7.62 
Bu  8.08  8.39  8.40  8.15  8.00 
MAD  0.09  0.31  0.34  0.14 
RMS  0.12  0.34  0.37  0.19 
MD  0.33  0.63  0.72  0.45 
Yale alkene subset vertical excitations
State MBPT(2) LCCD LCCSD CCSD Expt.
Ethylene 
B3u  7.15  7.43  7.45  7.28  7.11 
B1u  7.76  8.13  8.18  7.97  7.65 
B1g  7.79  8.08  8.10  7.97  7.80 
B2g  7.84  8.12  8.14  7.93  7.90 
Ag  8.19  8.46  8.48  8.31  8.28 
B3u  8.64  8.92  8.94  8.77  8.62 
B3u  8.93  9.20  9.22  9.05  8.90 
B3u  9.05  9.33  9.35  9.17  9.08 
B1g  9.19  9.46  9.49  9.34  9.20 
B1u  9.18  9.47  9.50  9.32  9.33 
B3u  9.84  10.11  10.13  9.96  9.51 
Isobutene 
B1  6.29  6.54  6.56  6.38  6.17 
A1  6.81  7.06  7.10  6.91  6.70 
trans-1,3-butadiene 
Bu  6.12  6.54  6.63  6.29  5.91 
Bg  6.17  6.47  6.49  6.24  6.22 
Au  6.47  6.79  6.80  6.55  6.66 
Bu  7.06  7.39  7.43  7.16  7.07 
Bg  7.28  7.59  7.60  7.36  7.36 
Ag  7.54  7.85  7.85  7.61  7.62 
Bu  8.08  8.39  8.40  8.15  8.00 
MAD  0.09  0.31  0.34  0.14 
RMS  0.12  0.34  0.37  0.19 
MD  0.33  0.63  0.72  0.45 
TABLE IV.

Vertical excitation energy comparison statistics for the Yale test set. Units are in electron volts (eV).

Yale set vertical excitation statistics
MBPT(2) LCCD LCCSD CCSD
Relative to EOM-CCSD 
MAD  0.13  0.17  0.24 
RMS  0.14  0.20  0.25 
SD  0.14  0.16  0.10 
MD  0.31  0.41  0.43 
Relative to expt. 
MAD  0.31  0.41  0.42  0.26 
RMS  0.41  0.50  0.53  0.34 
SD  0.40  0.43  0.37  0.31 
MD  1.04  1.25  1.24  0.97 
State orderings relative to EOM-CCSD 
MAD  0.04  0.04  0.05 
RMS  0.06  0.06  0.08 
SD  0.06  0.06  0.08 
MD  0.14  0.17  0.32 
Yale set vertical excitation statistics
MBPT(2) LCCD LCCSD CCSD
Relative to EOM-CCSD 
MAD  0.13  0.17  0.24 
RMS  0.14  0.20  0.25 
SD  0.14  0.16  0.10 
MD  0.31  0.41  0.43 
Relative to expt. 
MAD  0.31  0.41  0.42  0.26 
RMS  0.41  0.50  0.53  0.34 
SD  0.40  0.43  0.37  0.31 
MD  1.04  1.25  1.24  0.97 
State orderings relative to EOM-CCSD 
MAD  0.04  0.04  0.05 
RMS  0.06  0.06  0.08 
SD  0.06  0.06  0.08 
MD  0.14  0.17  0.32 

By employing the second-order similarity transformed Hamiltonians, Eqs. (59) and (65), we now have an equation-of-motion theory completely consistent with the LCCD (LCCSD) ground state wavefunction. The linear EOM methods have the same iterative O(n6) computational scaling as canonical EOM-CCSD. However, with the removal of the quadratic amplitude intermediates, the computational overhead and I/O demands are significantly reduced. Benchmark data from the Mülheim set show a systematic overestimate of excitation energies for both EOM-LCCD and EOM-LCCSD, with a MAD relative to EOM-CCSD of 0.25 and 0.35 eV, respectively (see the correlation Figures 2, 3, and Table V). The comparison of the linear EOM methods against EOM-CCSDT-3 is less satisfactory with a MAD of 0.45 and 0.56 eV for EOM-LCCD and EOM-LCCSD, respectively.

With an acceptable SD of 0.11 and 0.10 eV compared to EOM-CCSD (see Table IV), the excitation energies and relative state orderings obtained with the linear EOM methods are reliably consistent compared to canonical EOM-CCSD. This is further illustrated in Tables IIII where the systematic and consistent overestimation of the excitation energies is balanced by a nearly exact agreement in state ordering and relative state energy differences with a MAD of 0.04, 0.04, and 0.05 eV for EOM-MBPT(2), LCCD, and LCCSD, respectively (see the bottom of Table IV). For approximate methods, getting these relative energies correct is just as useful as quantitatively accurate excitation energies.

TABLE III.

Vertical singlet excitation energies for a number of single ring azabenzenes computed with the d-aug-cc-pVDZ basis using MP2/6-311+G** geometries. Units are in electron volts (eV), experimental references can be found in Ref. 2.

Yale azabenzenes subset vertical excitations
State MBPT(2) LCCD LCCSD CCSD Expt.
Pyrazine 
B3u  4.52  4.63  4.63  4.33  3.83 
B2u  5.36  5.51  5.50  5.10  4.81 
B2g  6.17  6.26  6.29  6.01  5.46 
B1g  7.14  7.35  7.34  7.07  6.10 
B1u  7.04  7.28  7.33  6.95  6.51 
Pyridazine 
B1  4.20  4.32  4.36  4.03  3.60 
A1  5.59  5.33  5.75  5.33  5.00 
A2  4.74  4.63  4.96  4.63  5.30 
B1  6.73  6.62  6.93  6.62  6.00 
B2  6.36  6.26  6.43  6.26  6.50 
Pyridine 
B1  5.28  5.38  5.46  5.17  4.59 
B2  5.41  5.57  5.61  5.23  4.99 
A2  5.64  5.80  5.92  5.60  5.43 
A1  6.76  6.86  6.92  6.69  6.38 
Pyrimidine 
B1  4.74  4.81  4.97  4.63  3.85 
A2  5.07  5.18  5.38  5.03  4.62 
B2  5.66  5.77  5.89  5.47  5.12 
A2  6.26  6.35  6.51  6.18  5.52 
B1  6.53  6.66  6.83  6.50  5.90 
A1  7.02  7.17  7.37  6.96  6.70 
s-tetrazine 
B3u  2.97  3.08  3.07  2.67  2.25 
Au  4.16  4.30  4.32  3.98  3.40 
Au  6.00  6.13  6.11  5.72  5.00 
B3u  7.11  7.25  7.27  6.95  6.34 
MAD  0.61  0.72  0.78  0.49 
RMS  0.65  0.76  0.82  0.53 
MD  1.04  1.25  1.24  0.97 
Yale azabenzenes subset vertical excitations
State MBPT(2) LCCD LCCSD CCSD Expt.
Pyrazine 
B3u  4.52  4.63  4.63  4.33  3.83 
B2u  5.36  5.51  5.50  5.10  4.81 
B2g  6.17  6.26  6.29  6.01  5.46 
B1g  7.14  7.35  7.34  7.07  6.10 
B1u  7.04  7.28  7.33  6.95  6.51 
Pyridazine 
B1  4.20  4.32  4.36  4.03  3.60 
A1  5.59  5.33  5.75  5.33  5.00 
A2  4.74  4.63  4.96  4.63  5.30 
B1  6.73  6.62  6.93  6.62  6.00 
B2  6.36  6.26  6.43  6.26  6.50 
Pyridine 
B1  5.28  5.38  5.46  5.17  4.59 
B2  5.41  5.57  5.61  5.23  4.99 
A2  5.64  5.80  5.92  5.60  5.43 
A1  6.76  6.86  6.92  6.69  6.38 
Pyrimidine 
B1  4.74  4.81  4.97  4.63  3.85 
A2  5.07  5.18  5.38  5.03  4.62 
B2  5.66  5.77  5.89  5.47  5.12 
A2  6.26  6.35  6.51  6.18  5.52 
B1  6.53  6.66  6.83  6.50  5.90 
A1  7.02  7.17  7.37  6.96  6.70 
s-tetrazine 
B3u  2.97  3.08  3.07  2.67  2.25 
Au  4.16  4.30  4.32  3.98  3.40 
Au  6.00  6.13  6.11  5.72  5.00 
B3u  7.11  7.25  7.27  6.95  6.34 
MAD  0.61  0.72  0.78  0.49 
RMS  0.65  0.76  0.82  0.53 
MD  1.04  1.25  1.24  0.97 

We have expanded the coupled-cluster similarity transformed Hamiltonian using general coupled-cluster perturbation theory to obtain the arbitrary order coupled-cluster perturbation theory effective Hamiltonian, Eq. (19). The inclusive (the sum of all orders 0 to n) form of the Hamiltonian is inserted into the standard electronic-excitation equation-of-motion theory to give the completely general EOM-CCPT (Eq. (27)). The result is a way of using the standard equation-of-motion theory to directly compute excitation energies that are consistent with a given CCPT ground state wavefunction.

By choosing the generalized many-body perturbation theory partitioning of the Hamiltonian, we re-derive the well known EOM-MBPT(2)29 (EOM-CCSD(2)27) equations. Two alternative choices in Hamiltonian partitioning, the particle rank conserving and mixing schemes, are also explored. These two schemes, referred to as PR0 and PR1, generate infinite-order amplitudes that correspond to the linearized coupled-cluster doubles (CCD) and CCSD wavefunction. Using the inclusive similarity transformed Hamiltonian derived from the PR0 and PR1 partitioning schemes provides an equation-of-motion method perturbatively consistent with a LCCD and LCCSD reference wavefunction.

These approximate equation-of-motion methods are benchmarked by employing the Mülheim37–39 and Yale2,40 small organic molecule data sets. We use the diverse Mülheim test set to benchmark our new methods against the canonical EOM-CCSD and EOM-CCSDT-3 methods, while the Yale test set is employed in benchmarking against experimental spectra. Our methods are found to consistently over estimate excitation energies, relative to EOM-CCSD and EOM-CCSDT-3 by ∼0.25 and ∼0.5 eV, respectively. The precise statistics and correlation plots can be found in Table V and Figures 2 and 3. The very good ∼0.1 eV standard deviation compared with EOM-CCSD suggests that the predicted relative spectra will be much more accurate than the absolute excitation energies. This is supported by studying the benchmark values from the Yale set given in Tables I–IV where the presented approximate EOM methods obtain relative state orderings and energies to within ∼0.04 eV of canonical EOM-CCSD. The systematic overestimation compared to the complete EOM-CCSD suggests that the similarity transformed Hamiltonian (and thus the excited state spectrum) based on the linear CC wavefunction is over-approximated compared to the ground-state description leading to an increased separation. Having explored three different CC ground state wavefunctions based on a linear amplitude approximation (and associated linear amplitude similarity transformed Hamiltonian), it is evident that nonlinear amplitude contributions play a significantly larger role in the excited state spectrum than for the ground state energy. Comparison with the CC2 method, which has all product T ˆ 1 and T ˆ 1 T ˆ 2 amplitudes but no quadratic T ˆ 2 terms, is further instructive as the average CC2 excitation energy is between38 EOM-CCSD and CC3. The ideal second order singles and doubles limited EOM approximation would then apparently lay somewhere between CC2 and linear EOM-CC.

TABLE II.

Vertical singlet excitation energies for a few small carbonyl molecules computed with the d-aug-cc-pVDZ basis using MP2/6-311+G** geometries. Units are in electron volts (eV), experimental references can be found in Ref. 2.

Yale carbonyl subset vertical excitations
State MBPT(2) LCCD LCCSD CCSD Expt.
Acetaldehyde 
A″  4.19  4.28  4.57  4.32  4.28 
A′  6.64  6.68  6.94  6.78  6.82 
A′  7.32  7.59  7.61  7.67  7.46 
A′  7.56  7.37  7.82  7.46  7.75 
A′  8.24  8.27  8.51  8.36  8.43 
A′  8.27  8.31  8.54  8.39  8.69 
Acetone 
A2  4.38  4.44  4.75  4.48  4.43 
B2  6.25  6.29  6.54  6.37  6.36 
A2  7.17  7.19  7.44  7.27  7.36 
A1  7.26  7.29  7.55  7.37  7.41 
B2  7.25  7.27  7.51  7.35  7.49 
A1  7.88  7.90  8.15  7.98  7.80 
B2  7.65  7.68  7.93  7.76  8.09 
B1  7.95  7.98  8.22  8.06  8.17 
Formaldehyde 
A2  3.82  3.96  4.23  3.99  4.00 
B2  6.87  6.93  7.17  7.03  7.08 
B2  7.70  7.75  7.96  7.83  7.97 
A1  7.83  7.87  8.10  7.97  8.14 
A2  8.06  8.09  8.32  8.19  8.37 
B2  8.77  8.82  9.05  8.91  8.88 
B1  9.10  9.14  9.05  9.23  9.00 
A2  9.20  9.24  9.45  9.32  9.22 
B2  9.03  9.07  9.29  9.16  9.26 
A1  9.07  9.12  9.35  9.20  9.58 
B2  9.07  9.10  9.32  9.19  9.63 
MAD  0.22  0.18  0.14  0.13 
RMS  0.26  0.23  0.17  0.18 
MD  0.56  0.53  0.35  0.44 
Yale carbonyl subset vertical excitations
State MBPT(2) LCCD LCCSD CCSD Expt.
Acetaldehyde 
A″  4.19  4.28  4.57  4.32  4.28 
A′  6.64  6.68  6.94  6.78  6.82 
A′  7.32  7.59  7.61  7.67  7.46 
A′  7.56  7.37  7.82  7.46  7.75 
A′  8.24  8.27  8.51  8.36  8.43 
A′  8.27  8.31  8.54  8.39  8.69 
Acetone 
A2  4.38  4.44  4.75  4.48  4.43 
B2  6.25  6.29  6.54  6.37  6.36 
A2  7.17  7.19  7.44  7.27  7.36 
A1  7.26  7.29  7.55  7.37  7.41 
B2  7.25  7.27  7.51  7.35  7.49 
A1  7.88  7.90  8.15  7.98  7.80 
B2  7.65  7.68  7.93  7.76  8.09 
B1  7.95  7.98  8.22  8.06  8.17 
Formaldehyde 
A2  3.82  3.96  4.23  3.99  4.00 
B2  6.87  6.93  7.17  7.03  7.08 
B2  7.70  7.75  7.96  7.83  7.97 
A1  7.83  7.87  8.10  7.97  8.14 
A2  8.06  8.09  8.32  8.19  8.37 
B2  8.77  8.82  9.05  8.91  8.88 
B1  9.10  9.14  9.05  9.23  9.00 
A2  9.20  9.24  9.45  9.32  9.22 
B2  9.03  9.07  9.29  9.16  9.26 
A1  9.07  9.12  9.35  9.20  9.58 
B2  9.07  9.10  9.32  9.19  9.63 
MAD  0.22  0.18  0.14  0.13 
RMS  0.26  0.23  0.17  0.18 
MD  0.56  0.53  0.35  0.44 

The authors acknowledge support from the U.S. Air Force Office of Scientific Research Grant No. FA 9550-11-1-0065 and the U.S. Army Research Office DURIP Grant No. W911-12-1-0365 which funded access to the University of Florida Research Computing HiPerGator high performance cluster.

1.
A.
Dreuw
and
M.
Head-Gordon
,
Chem. Rev.
105
,
4009
(
2005
).
2.
M.
Caricato
,
G. W.
Trucks
,
M. J.
Frisch
, and
K. B.
Wiberg
,
J. Chem. Theory Comput.
6
,
370
(
2010
).
3.
P. G.
Szalay
,
T.
Müller
,
G.
Gidofalvi
,
H.
Lischka
, and
R.
Shepard
,
Chem. Rev.
112
,
108
(
2012
).
4.
R. J.
Bartlett
and
M.
Musiał
,
Rev. Mod. Phys.
79
,
291
(
2007
).
5.
K.
Sneskov
and
O.
Christiansen
,
WIREs: Comput. Mol. Sci.
2
,
566
(
2011
).
6.
J. F.
Stanton
and
R. J.
Bartlett
,
J. Chem. Phys.
98
,
7029
(
1993
).
7.
T. J.
Watson
, Jr.
,
V. F.
Lotrich
,
P. G.
Szalay
,
A.
Perera
, and
R. J.
Bartlett
,
J. Phys. Chem. A
117
,
2569
(
2013
).
8.
J. D.
Watts
and
R. J.
Bartlett
,
J. Chem. Phys.
101
,
3073
(
1994
).
9.
J. E.
Del Bene
,
J. D.
Watts
, and
R. J.
Bartlett
,
J. Chem. Phys.
106
,
6051
(
1997
).
10.
J. D.
Watts
and
R. J.
Bartlett
,
Spectrochim. Acta, Part A
55
,
495
(
1999
).
11.
J.
Paldus
,
J.
Čížek
,
M.
Saute
, and
A.
Laforgue
,
Phys. Rev. A
17
,
805
(
1978
).
12.
J.
Foresman
,
M.
Head-Gordon
, and
J.
Pople
,
J. Phys. Chem.
96
,
135
(
1992
).
13.
M.
Head-Gordon
,
R. J.
Rico
,
M.
Oumi
, and
T. J.
Lee
,
Chem. Phys. Lett.
219
,
21
(
1994
).
14.
S.
Hirata
,
J. Chem. Phys.
122
,
094105
(
2005
).
15.
X.
Liu
,
Q.
Ou
,
E.
Alguire
, and
J. E.
Subotnik
,
J Chem. Phys.
138
,
221105
(
2013
).
16.
X.
Liu
and
J. E.
Subotnik
,
J. Chem. Theory Comput.
10
,
1004
(
2014
).
17.
X.
Liu
and
J. E.
Subotnik
,
J. Chem. Theory Comput.
10
,
1835
(
2014
).
18.
J. N.
Byrd
,
V. F.
Lotrich
, and
R. J.
Bartlett
,
J. Chem. Phys.
140
,
234108
(
2014
).
19.
J.
Schirmer
,
Phys. Rev. A
26
,
2395
2416
(
1982
).
20.
A. B.
Trofimov
,
I. L.
Krivdina
,
J.
Weller
, and
J.
Schirmer
,
Chem. Phys.
329
,
1
(
2006
).
21.
H. J.
Monkhorst
,
Int. J. Quantum Chem.
12
,
421
(
1977
).
22.
H.
Sekino
and
R. J.
Bartlett
,
Int. J. Quantum Chem.
18
,
255
(
1984
).
23.
H.
Koch
and
P.
Jörgensen
,
J. Chem. Phys.
93
,
3333
(
1990
).
24.
O.
Christiansen
,
H.
Koch
, and
P.
Jørgensen
,
Chem. Phys. Lett.
243
,
409
(
1995
).
25.
S.
Hirata
,
M.
Nooijen
,
I.
Grabowski
, and
R. J.
Bartlett
,
J. Chem. Phys.
114
,
3919
(
2001
).
26.
S.
Hirata
,
M.
Nooijen
,
I.
Grabowski
, and
R. J.
Bartlett
,
J. Chem. Phys.
115
,
3967
(
2001
).
27.
J. F.
Stanton
and
J.
Gauss
,
J. Chem. Phys.
103
,
1064
(
1995
).
28.
M.
Nooijen
and
J. G.
Snijders
,
J. Chem. Phys.
102
,
1681
(
1995
).
29.
S. R.
Gwaltney
,
M.
Nooijen
, and
R. J.
Bartlett
,
Chem. Phys. Lett.
248
,
189
(
1996
).
30.
J. N.
Byrd
,
N.
Jindal
,
R. W.
Molt
, Jr.
,
R. J.
Bartlett
,
B. A.
Sanders
, and
V. F.
Lotrich
, “
Molecular cluster perturbation theory. I. Formalism
,”
Mol. Phys.
(published online 2015).
31.
R. J.
Bartlett
,
M.
Musial
,
V. F.
Lotrich
, and
T.
Kus
, in
Recent Progress in Coupled-Cluster Methods
, edited by
P.
Carsky
,
J.
Paldus
, and
J.
Pittner
(
Springer
,
Dordrecht
,
2010
), Vol.
11
, Chap. 1, pp.
1
34
.
32.
J.
Paldus
,
J.
Čížek
, and
I.
Shavitt
,
Phys. Rev. A
5
,
50
(
1972
).
33.
R. J.
Bartlett
and
J.
Noga
,
Chem. Phys. Lett.
150
,
29
(
1988
).
34.
R. J.
Bartlett
,
S. A.
Kucharski
, and
J.
Noga
,
Chem. Phys. Lett.
155
,
133
(
1989
).
35.
I.
Shavitt
and
R. J.
Bartlett
,
Many-Body Methods in Chemistry and Physics
(
Cambridge
,
New York
,
2009
).
36.
A. G.
Taube
and
R. J.
Bartlett
,
J. Chem. Phys.
130
,
144112
(
2009
).
37.
M.
Schreiber
,
M. R.
Silva-Junior
,
S. P. A.
Sauer
, and
W.
Thiel
,
J Chem. Phys.
128
,
134110
(
2008
).
38.
S. P. A.
Sauer
,
M.
Schreiber
,
M. R.
Silva-Junior
, and
W.
Thiel
,
J. Chem. Theory Comput.
5
,
555
(
2009
).
39.
M. R. M.
Silva-Junior
,
M. M.
Schreiber
,
S. P. A. S.
Sauer
, and
W. W.
Thiel
,
J Chem. Phys.
133
,
174318
(
2010
).
40.
M.
Caricato
,
G. W.
Trucks
,
M. J.
Frisch
, and
K. B.
Wiberg
,
J. Chem. Theory Comput.
7
,
456
(
2011
).
41.

Partial inclusion of triples56 which has all possible Tˆ1 and Tˆ2 to Tˆ3 with O(n7) scaling, but neglects the Tˆ3Tˆ3 contributions that would introduce an O(n8) step.

42.
A.
Schäfer
,
H.
Horn
, and
R.
Ahlrichs
,
J. Chem. Phys.
97
,
2571
(
1992
).
43.
S. R.
Gwaltney
and
R. J.
Bartlett
,
Chem. Phys. Lett.
241
,
26
(
1995
).
44.
K. B.
Wiberg
,
A. E.
de Oliveira
, and
G.
Trucks
,
J. Phys. Chem. A
106
,
4192
(
2002
).
45.
T.
Dunning
, Jr.
,
J. Chem. Phys.
90
,
1007
(
1989
).
46.
R.
Kendall
,
T.
Dunning
, Jr.
, and
R.
Harrison
,
J. Chem. Phys.
96
,
6796
(
1992
).
47.
J. J.
Goings
,
M.
Caricato
,
M. J.
Frisch
, and
X.
Li
,
J. Chem. Phys.
141
,
164116
(
2014
).
48.

Where the d-aug-cc-pVDZ basis is clearly not as accurate as the 6-311(3+,3+)G** basis is in the case of the highest considered excited state of ethylene and formaldehyde.

49.
J. F.
Stanton
,
J.
Gauss
,
S. A.
Perera
,
A.
Yau
,
J. D.
Watts
,
M.
Nooijen
,
N.
Oliphant
,
P. G.
Szalay
,
W. J.
Lauderdale
,
S. R.
Gwaltney
,
S.
Beck
,
A.
Balková
,
D. E.
Bernholdt
,
K.-K.
Baeck
,
P.
Rozyczko
,
H.
Sekino
,
C.
Huber
,
J.
Pittner
, and
R. J.
Bartlett
, ACESII is a product of the quantum theory project, University of Florida, Integral packages included are VMOL (J. Almölf and P. R. Taylor), VPROPS (P. R. Taylor), and ABACUS (T. Helgaker, H. J. Aa. Jensen, P. Jørgensen, J. Olsen, and P. R. Taylor).
50.
B. A.
Sanders
,
N.
Jindal
,
J. N.
Byrd
,
V. F.
Lotrich
,
D.
Lyakh
,
N.
Flocke
,
A.
Perera
, and
R. J.
Bartlett
, Aces4 pre-alpha release, https://github.com/UFParLab.
51.
P.-O.
Löwdin
,
J. Mol. Spectrosc.
10
,
12
(
1963
).
52.
M.
Nooijen
and
R. J.
Bartlett
,
J. Chem. Phys.
106
,
6441
(
1997
).
53.
M.
Nooijen
and
R. J.
Bartlett
,
J. Chem. Phys.
106
,
6449
(
1997
).
54.
M.
Nooijen
and
R. J.
Bartlett
,
J. Chem. Phys.
107
,
6812
(
1997
).
55.
J.
Sous
,
P.
Goel
, and
M.
Nooijen
,
Mol. Phys.
112
,
616
638
(
2014
).
56.
J. D.
Watts
and
R. J.
Bartlett
,
Chem. Phys. Lett.
258
,
581
588
(
1996
).