Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Feb 28.
Published in final edited form as: J Chem Theory Comput. 2022 Oct 24;18(11):6637–6645. doi: 10.1021/acs.jctc.2c00686

Combining Renormalized Singles GW Methods with the Bethe–Salpeter Equation for Accurate Neutral Excitation Energies

Jiachen Li 1, Dorothea Golze 2, Weitao Yang 3
PMCID: PMC9972216  NIHMSID: NIHMS1872816  PMID: 36279250

Abstract

We apply the renormalized singles (RS) Green’s function in the Bethe–Salpeter equation (BSE)/GW approach to predict accurate neutral excitation energies of molecular systems. The BSE calculations are performed on top of the GRSWRS method, which uses the RS Green’s function also for the computation of the screened Coulomb interaction W. We show that the BSE/GRSWRS approach significantly outperforms BSE/G0W0 for predicting excitation energies of valence, Rydberg, and charge-transfer (CT) excitations by benchmarking the Truhlar–Gagliardi set, Stein CT set, and an atomic Rydberg test set. For the Truhlar–Gagliardi test set, BSE/GRSWRS provides comparable accuracy to time-dependent density functional theory (TDDFT) and is slightly better than BSE starting from eigenvalue self-consistent GW (evGW). For the Stein CT test set, BSE/GRSWRS significantly outperforms BSE/G0W0 and TDDFT with the accuracy comparable to BSE/evGW. We also show that BSE/GRSWRS predicts Rydberg excitation energies of atomic systems well. Besides the excellent accuracy, BSE/GRSWRS largely eliminates the dependence on the choice of the density functional approximation. This work demonstrates that the BSE/GRSWRS approach is accurate and efficient for predicting excitation energies for a broad range of systems, which expands the applicability of the BSE/GW approach.

Graphical Abstract

graphic file with name nihms-1872816-f0001.jpg

1. INTRODUCTION

The accurate prediction of neutral excitation energies from first-principle calculations plays a critical role in guiding the understanding and new developments in chemistry, biochemistry, and material science, such as drug delivery1,2 and design of an organic sensitizer.3,4 In the past decades, many theoretical approaches have been developed to compute accurate excitation energies. Time-dependent density functional theory57 (TDDFT) ranks among the most popular approaches. Because of the good accuracy and the affordable computational cost, TDDFT has been widely implemented in modern quantum chemistry packages to calculate energies, structures, and other properties of excited states for molecular and periodic systems.810 However, TDDFT has several well-known shortcomings. For example, TDDFT with commonly used density functional approximations (DFAs) poorly predicts excitation energies of Rydberg and charge-transfer (CT) excitations.11,12 The failure of TDDFT for these excitations must be attributed to the incorrect description of the long-range behavior using conventional DFAs.11,12 Attempts to address this issue include the usage of range-separated or Coulomb-attenuated functionals1316 and tuning the fraction of the Hartree–Fock (HF)17,18 exchange in DFAs.19 In addition, the accuracy of TDDFT strongly depends on the exchange–correlation (XC) kernel, which is the second derivative of the XC energy from the given DFA with respect to the electron density.10,20 The difference in the calculated excitation energies obtained from TDDFT with different DFAs can exceed 1.0 eV for valence excitations and even exceed 2.0 eV for Rydberg excitations.10,21 Recent developments to improve the accuracy of TDDFT include multireference DFT,22,23 orbital-optimized DFT,24 and mixed-reference spin-flip TDDFT.25

In the last years, the Bethe–Salpeter equation2628 (BSE) formalism in the Green’s function many-body perturbation theory29,30 has become increasingly popular for computing the optical excitations of molecules31 and has been most recently also applied to K-edge transitions.32 BSE is commonly performed on top of a GW calculation30,33,34 using the GW quasiparticle (QP) energies as the input. We denote this approach as BSE/GW. Recently, BSE in combination with the T-matrix approximation for predicting neutral excitation energies has also been reported.35 In the BSE/GW approach, the electron–hole interaction is described by the screened Coulomb interaction W instead of the bare Coulomb interaction. The GW QP energies substantially improve upon the Kohn–Sham (KS) orbital energies for predicting fundamental HOMO–LUMO gaps and ionization potentials in general, which are the critical quantities in BSE to calculate excitation energies. BSE/GW has been successfully applied to calculate excitation energies for systems of different sizes including molecules, solids, and low-dimensional materials.32,3649

However, the BSE/GW approach still suffers from several problems. First, although BSE has a computationally favorable scaling of O(N4) with respect to the system size N, which is the same as in TDDFT, the preceding GW calculation is computationally demanding. In the fully analytical treatment of the GW self-energy, the diagonalization of the random phase approximation (RPA) equation scales as O(N6) and the self-energy evaluation scales as O(N5).34,50 Thus, the computationally expensive GW step dominates the computational cost in a BSE/GW calculation. Many techniques have been developed to reduce the computational cost of GW calculations, for example, the contour deformation technique51 or analytic continuation of the self-energy,52,53 which scales O(N4) for valence states. The computational cost is further reduced in cubic scaling GW implementations, which recently emerged for localized basis set codes5457 enabling GW calculations for large systems of more than thousand atoms.

Second, the accuracy of the commonly used BSE/G0W0 approach has an undesirable dependence on the choice of the DFA,58,59 which is due the perturbative nature of the G0W0 scheme. In BSE/G0W0, the dependence on the DFA starting point is in the range of 0.5 eV for predicting valence excitation energies and can even exceed 1.0 eV for predicting CT excitation energies.59 It has been shown that BSE/G0W0 based on range-separated functionals and tuned hybrid functionals provides more accurate excitation energies32,60 than BSE/G0W0 based on GGA functionals. This dependence can be largely reduced by introducing self-consistency into the GW calculations, such as the eigenvalue-self-consistent GW (evGW) method, where the eigenvalues are iterated in G and W, the quasiparticle-self-consistent GW (qsGW) scheme,c61,62 or the fully self-consistent GW (scGW) approach.63,64 It has been shown that the BSE/evGW approach provides an accuracy comparable to TDDFT for predicting valence excitations and significantly outperforms BSE/G0W0 and TDDFT for predicting CT and Rydberg excitations.38,59,65 The DFA dependence in BSE/evGW is largely washed out.38,59,65 In practice, evGW calculations can converge within a few steps by using the linear mixing method.62 However, evGW might have convergence difficulties for systems with multiple solutions.66,67 In addition, the extra computational cost in evGW is expensive for large systems.

To reduce the computational cost, efforts were recently made to approximate the GW QP energies by improved orbital energies from DFT and to use the latter as the BSE input. For example, we recently employed the localized orbital scaling correction68 (LOSC), which minimizes the delocalization error in DFAs to predict orbital energies. We showed that the BSE/LOSC approach provides good accuracy for predicting excitation energies of different systems59 and yields significantly better results than BSE/G0W0. The BSE/LOSC approach scales as O(N4) complexity. Another computationally cheaper BSE approach is based on Koopmans-compliant (KC) functionals, where the orbital energies are derived from KC functionals and the screened interaction is obtained via direct minimization on top of a maximally localized Wannier function basis.69 The KC-based BSE yields similar accuracy as the BSE/G0W0 method.69

In this work, we applied the recently developed renormalized singles (RS) Green’s function70,71 in BSE/GW to compute accurate excitation energies with affordable computational cost. The idea of the RS Green’s function approach is to compute the HF self-energy with the KS orbitals from DFT7274 instead of the HF orbitals. Because of Brillouin’s theorem,18 there is no singles contribution to the HF self-energy in the perturbation. However, the HF Green’s function is not a good starting point for G0W0 calculations.70,75 Therefore, the HF Hamiltonian is constructed with KS orbitals and diagonalized separately in the occupied and virtual subspaces. This renormalization procedure absorbs all singles contributions into the self-energy to reduce the dependence on the choice of the DFA. The resulting RS Green’s function is constructed with RS orbital energies from the renormalization process, which has the same form as the KS Green’s function. From the viewpoint of the RPA, higher order contributions are also included in the infinite summation by this renormalization process. Compared with the self-consistent GW methods that solve Hedin’s equations iteratively to eliminate the starting point dependence, the one-shot RS process captures all contributions of the single excitations while hardly increasing the computational compared to G0W0. The RS Green’s function has been used in the GW methods and T-matrix methods to predict accurate QP energies for valence and core states.59,70,76 The concept of RS has also been used in the multireference DFT approach23 to describe the static correlation in strongly correlated systems. The RS Green’s function shares similar thinking as the RS excitation in the RPA calculation for correlation energies.7779

The RS Green’s function has been applied in the GW approximation in two flavors. The first one is the GRSW070 method, which uses the RS Green’s function as a new starting point and calculates the screened interaction with the KS Green’s function. We found that the GRSW0 method provides a considerable improvement over G0W0 for predicting valence ionization potentials and electron affinities with a reduced starting point dependence70 but fails to restore the correct physics for deep core excitations.71 The second one is the GRSWRS method,71 which uses the RS Green’s function as a new starting point and calculates the screened interaction with the RS Green’s function. We found that GRSWRS also yields an improvement over G0W0,80 that is, it opens the fundamental gaps compared to G0W0 similar to evGW; see also Section 1 in the Supporting Information. Because of the nonlinear nature of the QP equation, multiple solutions can be found due to the unphysical discontinuities and intruder states,66,67,81 especially for core states. GRSWRS properly separates the main core QP state from the satellites and provides a solution to the QP equation corresponding to the desired core state,71 which is not the case for G0W0 with GGA or hybrid functionals with a low amount of exact exchange. Even though the absolute core-level energies are overestimated by several electronvolts, the relative shifts are predicted with a reasonable accuracy.71 For valence as well as core states, the starting point dependence is significantly reduced compared to G0W0.

In the present work, we benchmark both RS flavors, BSE/GRSW0 and BSE/GRSWRS, for the prediction of neutral excitations of molecular systems, including valence, CT, and Rydberg excitations.

2. THEORY

2.1. RS Green’s Function.

The goal of the RS Green’s function approach is to improve the accuracy of perturbative GW methods by including exactly the exchange self-energy, which is a one-electron contribution, and reduce the starting point dependence of the orbital energies on the DFA. The general idea is to construct the HF self-energy from KS orbitals, followed by a separate diagonalization in the occupied orbital subspace and the virtual orbital subspace.70 Effectively, the RS approach treats the one-body exchange self-energy exactly, or nonperturbatively, by diagonalization, unlike the commonly used G0W0 approach, in which the one-body exchange is included perturbatively along with the many-body correlation self-energy. The RS Green’s function GRS is defined as the solution of the two projected equations in the occupied orbital subspace and the virtual orbital subspace70

P(GRS1)P=P(G01)P+P(ΣHx[G0]vHxc)P (1)

and

Q(GRS1)Q=Q(G01)Q+Q(ΣHx[G0]vHxc)Q (2)

where P=iocc|ψiψi| is the projection into the occupied orbital space and Q = IP is the projection into the virtual orbital space, ψ is the KS orbital, ΣHx is the Hartree-exchange self-energy, and vHxc is the Hartree-exchange–correlation potential. ΣHx[G0] means that the HF self-energy (Hartree and exchange contribution) is constructed from the KS density matrix. Equivalently, the RS Green’s function is obtained by using the DFA density matrix in the HF Hamiltonian, namely, HHF[G0], and solving two projected HF equations in the occupied and virtual subspaces70

P(HHF[G0])P|ψiRS=ϵiRSP|ψiRS (3)

and

Q(HHF[G0])Q|ψaRS=ϵaRSQ|ψaRS (4)

where ϵRS is the RS orbital energy and ψRS is the RS orbital. Here, we use i, j, k, l for occupied orbitals, a, b, c, d for virtual orbitals, p, q, r, s for general orbitals. The subspace diagonalization of the HF Hamiltonian is performed only once. The resulting RS Green’s function is diagonal in the occupied and virtual subspaces70

[GRS(ω)]pq=δpq1ωϵpRS+iηsgn(ϵpRSμ) (5)

Here, μ is the chemical potential and η is the broadening parameter. As shown in eq 5, the RS Green’s function has a similar form as the KS Green’s function except that the KS orbital energies in the denominator are replaced by the RS orbital energies. Therefore, the RS Green’s function can be directly implemented in existing GW codes.

2.2. GRSW0 and GRSWRS.

The RS Green’s function is used in two GW flavors: GRSW070 and GRSWRS.71 In both methods, KS orbitals instead of RS orbitals are used for simplicity70,71 because using the RS orbitals does not change the results significantly. Therefore, the exchange part of the self-energy in GRSW0 and GRSWRS is the same as G0W0. In GRSW0,70 the RS Green’s function is used as the starting point and the screened interaction is calculated with the KS Green’s function, which means the KS Green’s function is used in the RPA calculation. The correlation part of the self-energy in GRSW0 is70

[ΣcGRSW0(ω)]pp=mq|(pqρmKS)|2ωϵqRS(ΩmKSiη˜)sgn(ϵqRSμ) (6)

where ρmKS and ΩmKS are the transition density and the excitation energy from RPA calculated with the KS Green’s function, m is the index for the RPA excitation, and η˜=3η.50

With the self-energy in eq 6, the QP equation for GRSW0 is70

ϵpQP=ϵp0+p|xcGRSW0(ϵpQP)vxc|p (7)

where ϵp0 is the KS orbital energy. In eq 7, the QP energy ϵpQP appears in both sides, which means eq 7 needs to be solved iteratively. To reduce the computational cost, eq 7 can be linearized30,70 as

ϵpQP=ϵp0+ZpGRSW0p|ΣxcGRSW0(ϵpRS)vxc|p (8)

with the factor ZpGRSW0=1[ΣcGRSW0(ω)]ppω|ω=ϵpRS1.

The GRSWRS method71 uses the RS Green’s function as a new starting point and calculates the screened interaction with the RS Green’s function, which means that the RS Green’s function is used in the RPA calculation. The correlation part of the self-energy in GRSWRS71

[ΣcGRSWRS(ω)]pp=mq|(pqρmRS)|2ωϵqRS(ΩmRSiη˜)sgn(ϵqRSμ) (9)

where ρmRS and ΩmRS are the transition density and the excitation energy from RPA calculated with the RS Green’s function. Therefore, the QP equation for GRSWRS is71

ϵpQP=ϵp0+p|ΣxcGRSWRS(ϵpQP)vxc|p (10)

Equation 10 can also be linearized as

ϵpQP=ϵp0+ZpGRSWRSp|ΣxcGRSWRS(ϵpRS)vxc|p (11)

with the factor ZpGRSWRS=(1[ΣcGRSWRS(ω)]ppω|ω=ϵpRS)1.

As shown in ref 34, the linearized QP equation gives small errors for valence QP energy calculations, which are important in BSE. In Section 2 in the Supporting Information, we show that using linearized QP equations defined in eqs 8 and 11 gives small differences around 0.01 eV compared to eqs 7 and 10 for the type of excitations studied here. Therefore, the linearized QP equations are solved to reduce the computational cost in the present work.

2.3. BSE/GRSW0 and BSE/GRSWRS Approaches.

The QP energies obtained from GRSWRS are used in BSE to calculate the excitation energies. With the static approximation for the screened interaction that treats the frequency as zero,31,82,83 the working equation of BSE is a generalized eigenvalue equation,31,82,83 which is similar to the Casida equation in TDDFT6,7

[ABB*A*][XY]=Ω[I00I][XY] (12)

where Ω is the excitation energy. In eq 12 the A, B matrices are defined as

Aia,jb=δijδab(ϵaQPϵiQP)+via,jbWij,ab(ω=0) (13)
Bia,jb=via,bjWib,aj(ω=0) (14)

where v is the Coulomb interaction and W(ω = 0) is the static screened interaction. v is the Coulomb interaction defined as

vpq,rs=dx1dx2ψp*(x1)ψr*(x2)ψq(x1)ψs(x2)|r1r2| (15)

where {ψp} is the set of input orbitals and x is the combined space-spin variable for (r, σ). W is the screened interaction defined as

Wpq,rs=tu(D1)pq,tuvtu,rs (16)

where the dielectric function D is calculated by the static response function χ82,83

Dpq,rs=δprδqsvpq,rsχrs,rs (17)
χia,ia=χai,ai=(ϵiQPϵaQP)1 (18)

The BSE working equation in eq 12 is analogous to the Casida equation6,7 in TDDFT. The only difference is that the BSE kernel replaces the XC kernel. Thus, the scaling of solving eq 12 is O(N4) by using the canonical Davidson algorithm.84,85

3. COMPUTATIONAL DETAILS

We implemented the BSE/GRSW0 and BSE/GRSWRS approaches in the QM4D quantum chemistry package86 and applied them to calculate the excitation energies of different systems. As discussed in ref 59, the application of the Tamm–Dancoff approximation (TDA) improves the accuracy of BSE/G0W0 for predicting both singlet and triplet excitation energies because BSE/G0W0 largely underestimates the excitation energies of molecular systems. However, for BSE/GRSWRS and BSE/evGW, which predict larger excitation energies than BSE/G0W0, using TDA leads to similar or worse triplet results and worsens singlet results as shown in Section 4 in the Supporting Information and ref 38. In addition, as shown in recent studies, using TDA in BSE/GW can lead to blueshifts in nanosized systems8789 and worse accuracy for singlet-triplet energy gaps in organic molecules.36 Therefore, TDA is not used in the present work. We tested three different sets: the comprehensive Truhlar–Gagliardi test set90 that contains singlet, triplet, valence, CT, and Rydberg excitations, the Stein CT test set91 that contains 12 intramolecular CT excitations between an aromatic donor and the tetracyanoethylene acceptor, and a test set for Rydberg excitations that contains three atomic systems. For the Truhlar–Gagliardi test set,90 the aug-cc-pVTZ basis set92,93 was used for all molecules, except for naphthalene, pNA, and DMABN, for which the augcc-pVDZ basis set92,93 was employed. It has been shown that the aug-cc-pVTZ basis sets yield converged neutral optical excitations,46 and even aug-cc-pVDZ results were found to deviate by not more than 0.2–0.3 eV from the basis set limit.46 B-TCNE was excluded due to the high computational cost. Reference values for pNA and DMABN were taken from refs 94 and 95 for the remaining molecules in the Truhlar–Gagliardi test set. The reference values are the theoretical best estimates, for example, using FCI or CCSDTQ.94,95 Geometries were all taken from ref 90. Note that geometries used in ref 95 are slightly different from those in the Truhlar–Gagliardi set. As shown in refs 95 and 96, the difference between theoretical best estimates obtained with two slightly different geometries is around 0.01 eV. Thus, we do not expect the small differences to change the conclusion. For the Stein CT test set,91 the cc-pVDZ92 basis set was used. Because theory best estimates for the Stein CT test set are not available, the experimental values in the gas phase91 were taken as the reference, which can be a source of errors. For the test of Rydberg excitation energies of B+, Be, and Mg, the aug-cc-pVQZ basis set92,93 was employed. Experimental reference values were taken from ref 97. TDDFT calculations were performed with the GAUSSIAN16 A.03 software.98 BSE/G0W0, BSE/GRSW0, BSE/GRSWRS, and evGW calculations were performed with QM4D. QM4D uses Cartesian basis sets and uses the resolution of identity technique52,99,100 to compute two-electron integrals. All basis sets and corresponding fitting basis sets were taken from the basis set exchange.101103

4. RESULTS

4.1. Truhlar–Gagliardi Test Set.

We first examine the performance of the BSE/GRSWRS approach for predicting the excitation energies of molecules from the Truhlar–Gagliardi test set. Excluding B-TCNE, this test set contains 18 valence excitations, two Rydberg excitations, and two CT excitations. The valence excitations in this set refer to n → π* and ππ* excitations. The mean absolute errors (MAEs) and mean signed errors (MSEs) of excitation energies obtained from TDDFT, BSE/G0W0, BSE/GRSW0, BSE/GRSWRS, and BSE/evGW with HF, BLYP, PBE, B3LYP, PBE0, and PBEh(0.75) are shown in Figure 1 and Table 1. The PBE-based hybrid functional PBEh(0.75) that has 75% HF exchange is shown as the optimal starting point for G0W0 to predict IPs in the GW100 set.104 The signed error is defined as the difference between the calculated value and the reference value, that is, EcalcEreference. Because this test set mainly contains valence excitations, TDDFT with conventional DFAs provides good accuracy. However, TDDFT based on the DFA with a large percentage of the HF exchange has large errors and can suffer from triplet instability. As evident from Figure 1, BSE/G0W0 greatly underestimates the excitation energies and provides large MAEs, which were also reported previously.31,59 This error can be attributed to the overscreening problem in G0W0, that is, the screened interaction W is calculated with an underestimated KS gap.31 The underestimated fundamental gap in G0W0 leads to the underestimated optical gap in BSE/G0W0. In addition, BSE/G0W0 has a strong starting point dependence. The difference between MAEs of BSE/G0W0 with GGA and hybrid functionals is larger than 0.7 eV. BSE/G0W0, based on PBEh(0.75), that is the optimal starting point for valence QP energy provides a small MAE of 0.27 eV. However, the percentage of the HF exchange needs to be reoptimized for excitations of different characters and different species.32 Recent work has shown that the optimally tuned range-separated hybrid DFAs can be a good starting point for BSE/G0W0.105 The BSE/GRSW0 approach improves upon BSE/G0W0. Using GRSW0 instead of G0W0, the MAEs are reduced by around 0.6 eV with GGA functionals and by around 0.3 eV with hybrid functionals. However, there is still an undesired starting point dependence in BSE/GRSW0 because the screened interaction in GRSW0 is calculated at the chosen DFA level. The BSE/GRSWRS approach significantly outperforms BSE/G0W0 and BSE/GRSW0. The MAEs from BSE/GRSWRS with conventional DFAs are around 0.4 eV. They are similar to the MAEs from TDDFT with hybrid functionals and are slightly better than the ones from BSE/evGW. BSE/GRSWRS with the optimal starting point PBEh(0.75) provides the smallest MAE of 0.29 eV. As shown in Section 3 in the Supporting Information, fundamental gaps obtained from RS orbital energies are always larger than those obtained from KS orbital energies. By inserting the RS Green’s function into the RPA equation to formulate the screened interaction, BSE/GRSWRS greatly reduces the overscreening error and provides excellent accuracy. The starting point dependence in BSE/GRSWRS is largely reduced, which is similar to BSE/evGW. The different DFAs induce only small changes of less than 0.1 eV in the BSE/GRSWRS MAEs. We find that BSE/GRSWRS, BSE/G0W0, and BSE/evGW yield triplet excitation energies which are significantly too low, which is in agreement with previous work.31

Figure 1.

Figure 1.

MAEs and MSEs of excitation energies in the Truhlar–Gagliardi test set obtained from TDDFT, BSE/G0W0, BSE/GRSW0, BSE/GRSWRS, and BSE/evGW based on HF, BLYP, PBE, B3LYP, PBE0, and PBEh(0.75). Reference values for pNA and DMABN were taken from ref 94 and for the remaining molecules from ref 95. The reference values are the theoretical best estimates. The aug-cc-pVDZ basis set was used for naphthalene, pNA, and DMABN, and the aug-cc-pVTZ basis set was used for the remaining systems. B-TCNE was excluded due to the high computational cost. Total MAEs and total MSEs were calculated by averaging all systems with equal weights. The error for system i is defined as Errori=EicalcEireference.

Table 1.

Total MAEs of Excitation Energies in the Truhlar–Gagliardi Test Set Obtained from TDDFT, BSE/G0W0, BSE/GRSW0, BSE/GRSWRS, and BSE/evGW Based on HF, BLYP, PBE, B3LYP, PBE0, and PBEh(0.75)a

HF BLYP PBE B3LYP PBE0 PBEh
BSE/G0W0 0.76 1.43 1.40 0.99 0.88 0.27
BSE/GRSW0 0.76 0.83 0.82 0.67 0.61 0.28
BSE/GRSWRS 0.76 0.37 0.36 0.35 0.34 0.29
BSE/evGW 0.74 0.48 0.47 0.52 0.53 0.26
TDDFT 1.54 0.52 0.51 0.33 0.31 0.85
a

Reference values for pNA and DMABN were taken from ref 94 and for the remaining molecules from ref 95. The reference values are the theoretical best estimates. The aug-cc-pVDZ basis set was used for naphthalene, pNA, and DMABN, and the aug-cc-pVTZ basis set was used for the remaining systems. B-TCNE was excluded due to the high computational cost. Total MAEs were calculated by averaging all systems with equal weights. The error for system i is defined as Errori=EicalcEireference.

4.2. Stein CT Test Set.

We further study the performance of BSE/GRSWRS in predicting CT excitation energies by testing the Stein CT test set. This test set contains 12 intramolecular CT systems. The MAEs and MSEs of excitation energies obtained from TDDFT, BSE/G0W0, BSE/GRSW0, and BSE/GRSWRS with HF, BLYP, PBE B3LYP, PBE0, and PBEh(0.75) are listed in Table 2. It shows that TDDFT with conventional DFAs fails to predict CT excitation energies due to the incorrect description of the long-range behavior. TDDFT with both GGA and hybrid functionals greatly underestimates the CT excitation energies and gives MAEs larger than 1.0 eV. TDDFT with PBEh(0.75) that has a large percentage of the HF exchange provides a small MAE of 0.19 eV. In addition, TDDFT has a strong starting point dependence. BSE/G0W0 provides improved results over TDDFT because of the correct long-range behavior from the screened interaction. However, BSE/G0W0 still suffers from a strong dependence on the choice of the DFA and yields relatively large errors. As shown in Section 6 in the Supporting Information, BSE/G0W0 always underestimates the excitation energies of CT systems. Slightly larger excitation energies can be obtained when using the TDA.59 The BSE/GRSW0 approach improves again upon BSE/G0W0. Compared with BSE/G0W0, the MAEs of BSE/GRSW0 are 1.0 and 0.4 eV smaller with GGA and hybrid functionals, respectively. The dependence on the DFA is reduced to only 0.09 eV in the BSE/GRSW0 scheme. The BSE/GRSWRS approach provides the most accurate results with the smallest starting point dependence. The MAEs of BSE/GRSWRS with all tested DFAs are only around 0.15 eV, which are comparable to the accuracy of BSE/evGW as reported in ref 41. In addition, the dependence on the choice of the DFA in BSE/GRSWRS is largely eliminated. The difference originating from using different DFAs is only around 0.04 eV.

Table 2.

MAEs and MSEs of CT Excitation Energies in the Stein CT Test Set Obtained from TDDFT, BSE/G0W0, BSE/GRSW0, and BSE/GRSWRS with HF, BLYP, PBE, B3LYP, PBE0, and PBEh(0.75)a

HF
BLYP
PBE
B3LYP
PBE0
PBEh
MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE
TDDFT 0.78 0.78 1.44 −1.44 1.45 −1.45 1.16 −1.16 1.08 −1.08 0.19 0.02
BSE/G0W0 0.10 −0.06 1.28 −1.28 1.31 −1.31 0.74 −0.74 0.65 −0.65 0.14 0.09
BSE/GRSW0 0.10 −0.06 0.35 −0.35 0.35 −0.37 0.29 −0.28 0.29 −0.29 0.16 −0.15
BSE/GRSWRS 0.10 −0.06 0.17 0.14 0.18 0.11 0.14 0.09 0.17 0.14 0.11 −0.07
a

All values in eV. Geometries were taken from ref 91. Experiment values in the gas phase were taken as the reference values.91 Gas-phase experimental references were used. The error for system i is defined as Errori=EicalcEiexperiment The cc-pVDZ basis set was used.

4.3. Rydberg Excitations.

We further investigate the performance of the BSE/GRSW0 and the BSE/GRSWRS approaches for predicting Rydberg excitation energies of atomic systems by testing B+, Be, and Mg. The signed errors of Rydberg excitation energies obtained from TDDFT, BSE/G0W0, BSE/GRSW0, BSE/GRSWRS, and BSE/evGW with HF, BLYP, PBE, B3LYP, and PBE0 are listed in Figure 2. Numerical results are shown in Section 7 in the Supporting Information. Similar to the CT excitation energies, it is well-known that TDDFT with common DFAs has relatively large errors for predicting Rydberg excitation energies.31 The latter largely underestimates with GGA as well as hybrid functionals in TDDFT. Compared to TDDFT, the BSE/G0W0 approach yields only slightly better predictions, reducing the MAE by around 0.1 eV. TDDFT and BSE/G0W0 both show strong starting point dependence. BSE/GRSW0 slightly improves upon BSE/G0W0, reducing the MAE by 0.1 eV with respect to BSE/G0W0. The BSE/GRSWRS approach provides further improvements over BSE/GRSW0. As shown in Section 7 in the Supporting Information and Figure 2, BSE/GRSWRS shows significant improvements over TDDFT for predicting singlet Rydberg excitation energies with reduced errors around 0.4 eV. However, the improvements on triplet Rydberg excitation energies are small. The MAEs of BSE/GRSWRS with all different DFAs are around 0.6 eV, which is close to the BSE/evGW level. The DFA starting point dependence is reduced to round 0.2 and 0.1 eV with BSE/GRSW0 and BSE/GRSWRS, respectively.

Figure 2.

Figure 2.

Signed errors of B+, Be and Mg obtained from TDDFT, BSE/G0W0, BSE/GRSW0, BSE/GRSWRS, and BSE/evGW with HF, BLYP, PBE, B3LYP, and PBE0. All values in eV. Experimental values were taken as the reference values.97 The aug-cc-pVQZ basis set was used. The signed error for system i is defined as Errori=EicalcEiexperiment.

5. CONCLUSIONS

In this work, we applied the BSE formalism on top of the GRSWRS method to calculate valence, Rydberg, and CT excitation energies of molecular systems. The GRSWRS method provides improved fundamental gaps compared to G0W0 and largely reduces the dependence on the choice of the density functional approximation. In the BSE/GRSWRS approach, the QP energies from GRSWRS are used in BSE. For the Truhlar–Gagliardi test set, we found that BSE/GRSWRS provides excellent accuracy for excitations of different characters (valence, CT, and Rydberg excitations) with MAEs around 0.4 eV. The accuracy of BSE/GRSWRS is similar to TDDFT and slightly better than BSE/evGW. Using the Stein CT test set, we further showed that BSE/GRSWRS is significantly more accurate for predicting CT excitation energies than BSE/G0W0 and TDDFT. We also found that the predictions compare well to BSE/evGW results reported in the literature.41 We also showed that BSE/GRSWRS predict accurate Rydberg excitation energies for atomic systems. We found for all three test sets that the dependence on the choice of the DFA is also largely eliminated. The computational cost of BSE/GRSWRS is similar to BSE/G0W0, which has a much lower computational cost than BSE/evGW.

This work demonstrates that the BSE/GRSWRS approach is accurate and efficient for predicting all three types of excitation energies of a broad range of systems. Therefore, the BSE/GRSWRS approach is expected to expand the applicability of the BSE/GW approach.

Supplementary Material

ct2c00686_si_001

ACKNOWLEDGMENTS

J.L. acknowledges the support from the National Institute of General Medical Sciences of the National Institutes of Health under award number R01-GM061870. D.G. acknowledges the Emmy Noether Programme of the German Research Foundation under project number 453275048. W.Y. acknowledges the support from the National Science Foundation (grant no. CHE-1900338).

Footnotes

The authors declare no competing financial interest.

ASSOCIATED CONTENT

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.2c00686.

Fundamental gaps obtained from different GW method, errors of using the linearized QP equation in BSE, fundamental gaps obtained from KS-DFT and KS-DFT with RS, comparison of excitation energies obtained from BSE/GRSWRS with and without the TDA, results of the Truhlar–Gagliardi test set, results of the Stein CT test set, and results of Rydberg excitation energies (PDF)

Contributor Information

Jiachen Li, Department of Chemistry, Duke University, Durham, North Carolina 27708, United States.

Dorothea Golze, Faculty of Chemistry and Food Chemistry, Technische Universität Dresden, 01062 Dresden, Germany.

Weitao Yang, Department of Chemistry, Duke University, Durham, North Carolina 27708, United States.

REFERENCES

  • (1).Velema WA; van der Berg JP; Hansen MJ; Szymanski W; Driessen AJM; Feringa BL Optical Control of Antibacterial Activity. Nat. Chem. 2013, 5, 924–928. [DOI] [PubMed] [Google Scholar]
  • (2).Beharry AA; Sadovski O; Woolley GA Azobenzene Photoswitching without Ultraviolet Light. J. Am. Chem. Soc. 2011, 133, 19684–19687. [DOI] [PubMed] [Google Scholar]
  • (3).Shalabi AS; El Mahdy AM; Taha HO; Soliman KA The Effects of Macrocycle and Anchoring Group Replacements on the Performance of Porphyrin Based Sensitizer: DFT and TD-DFT Study. J. Phys. Chem. Solids 2015, 76, 22–33. [Google Scholar]
  • (4).Zhang C-R; Liu Z-J; Chen Y-H; Chen H-S; Wu Y-Z; Feng W; Wang D-B DFT and TD-DFT Study on Structure and Properties of Organic Dye Sensitizer TA-St-CA. Curr. Appl. Phys. 2010, 10, 77–83. [Google Scholar]
  • (5).Runge E; Gross EKU Density-Functional Theory for Time-Dependent Systems. Phys. Rev. Lett. 1984, 52, 997–1000. [Google Scholar]
  • (6).Casida ME Time-Dependent Density Functional Response Theory for Molecules. Recent Advances in Computational Chemistry; World Scientific, 1995; Vol. 1; pp 155–192. [Google Scholar]
  • (7).Ullrich CA Time-Dependent Density-Functional Theory: Concepts and Applications; OUP Oxford, 2011. [Google Scholar]
  • (8).Casida ME Time-Dependent Density-Functional Theory for Molecules and Molecular Solids. J. Mol. Struct.: THEOCHEM 2009, 914, 3–18. [Google Scholar]
  • (9).Casida M; Ipatov A; Cordova F Linear-Response Time-Dependent Density Functional Theory for Open-Shell Molecules. Time-Dependent Density Functional Theory; Marques MA, Ullrich CA, Nogueira F, Rubio A, Burke K, Gross EKU, Eds.; Lecture Notes in Physics; Springer: Berlin, Heidelberg, 2006; pp 243–257. [Google Scholar]
  • (10).Laurent AD; Jacquemin D TD-DFT Benchmarks: A Review. Int. J. Quantum Chem. 2013, 113, 2019–2039. [Google Scholar]
  • (11).Tozer DJ Relationship between long-range charge-transfer excitation energy error and integer discontinuity in Kohn-Sham theory. J. Chem. Phys. 2003, 119, 12697–12699. [Google Scholar]
  • (12).Dreuw A; Weisman JL; Head-Gordon M Long-Range Charge-Transfer Excited States in Time-Dependent Density Functional Theory Require Non-Local Exchange. J. Chem. Phys. 2003, 119, 2943–2946. [Google Scholar]
  • (13).Leininger T; Stoll H; Werner H-J; Savin A Combining Long-Range Configuration Interaction with Short-Range Density Functionals. Chem. Phys. Lett. 1997, 275, 151–160. [Google Scholar]
  • (14).Besley NA; Peach MJG; Tozer DJ Time-Dependent Density Functional Theory Calculations of near-Edge X-ray Absorption Fine Structure with Short-Range Corrected Functionals. Phys. Chem. Chem. Phys. 2009, 11, 10350–10358. [DOI] [PubMed] [Google Scholar]
  • (15).Peach MJG; Benfield P; Helgaker T; Tozer DJ Excitation Energies in Density Functional Theory: An Evaluation and a Diagnostic Test. J. Chem. Phys. 2008, 128, 044118. [DOI] [PubMed] [Google Scholar]
  • (16).Ai W; Fang W-H; Su NQ The Role of Range-Separated Correlation in Long-Range Corrected Hybrid Functionals. J. Phys. Chem. Lett. 2021, 12, 1207–1213. [DOI] [PubMed] [Google Scholar]
  • (17).Slater JC Note on Hartree’s Method. Phys. Rev. 1930, 35, 210–211. [Google Scholar]
  • (18).Szabo A; Ostlund NS Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory; Courier Corporation, 2012. [Google Scholar]
  • (19).Zhao Y; Truhlar DG Density Functional for Spectroscopy: No Long-Range Self-Interaction Error, Good Performance for Rydberg and Charge-Transfer States, and Better Performance on Average than B3LYP for Ground States. J. Phys. Chem. A 2006, 110, 13126–13130. [DOI] [PubMed] [Google Scholar]
  • (20).Leang SS; Zahariev F; Gordon MS Benchmarking the Performance of Time-Dependent Density Functional Methods. J. Chem. Phys. 2012, 136, 104101. [DOI] [PubMed] [Google Scholar]
  • (21).Kaur J; Ospadov E; Staroverov VN What Is the Accuracy Limit of Adiabatic Linear-Response TDDFT Using Exact Exchange-Correlation Potentials and Approximate Kernels? J. Chem. Theory Comput. 2019, 15, 4956–4964. [DOI] [PubMed] [Google Scholar]
  • (22).Chen Z; Zhang D; Jin Y; Yang Y; Su NQ; Yang W Multireference Density Functional Theory with Generalized Auxiliary Systems for Ground and Excited States. J. Phys. Chem. Lett. 2017, 8, 4479–4485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Li J; Chen Z; Yang W Multireference Density Functional Theory for Describing Ground and Excited States with Renormalized Singles. J. Phys. Chem. Lett. 2022, 13, 894–903. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (24).Hait D; Head-Gordon M Orbital Optimized Density Functional Theory for Electronic Excited States. J. Phys. Chem. Lett. 2021, 12, 4517–4529. [DOI] [PubMed] [Google Scholar]
  • (25).Horbatenko Y; Sadiq S; Lee S; Filatov M; Choi CH Mixed-Reference Spin-Flip Time-Dependent Density Functional Theory (MRSF-TDDFT) as a Simple yet Accurate Method for Diradicals and Diradicaloids. J. Chem. Theory Comput. 2021, 17, 848–859. [DOI] [PubMed] [Google Scholar]
  • (26).Sham LJ; Rice TM Many-Particle Derivation of the Effective-Mass Equation for the Wannier Exciton. Phys. Rev. 1966, 144, 708–714. [Google Scholar]
  • (27).Hanke W; Sham LJ Many-Particle Effects in the Optical Excitations of a Semiconductor. Phys. Rev. Lett. 1979, 43, 387–390. [Google Scholar]
  • (28).Salpeter EE; Bethe HA A Relativistic Equation for Bound-State Problems. Phys. Rev. 1951, 84, 1232–1242. [Google Scholar]
  • (29).Hedin L New Method for Calculating the One-Particle Green’s Function with Application to the Electron-Gas Problem. Phys. Rev. 1965, 139, A796–A823. [Google Scholar]
  • (30).Martin RM; Reining L; Ceperley DM Interacting Electrons; Cambridge University Press, 2016. [Google Scholar]
  • (31).Blase X; Duchemin I; Jacquemin D; Loos P-F The Bethe-Salpeter Equation Formalism: From Physics to Chemistry. J. Phys. Chem. Lett. 2020, 11, 7371–7382. [DOI] [PubMed] [Google Scholar]
  • (32).Yao Y; Golze D; Rinke P; Blum V; Kanai Y All-Electron BSE@GW Method for K-Edge Core Electron Excitation Energies. J. Chem. Theory Comput. 2022, 18, 1569. [DOI] [PubMed] [Google Scholar]
  • (33).Reining L The GW Approximation: Content, Successes and Limitations. WIREs Comput. Mol. Sci 2018, 8, No. e1344. [Google Scholar]
  • (34).Golze D; Dvorak M; Rinke P The GW Compendium: A Practical Guide to Theoretical Photoemission Spectroscopy. Front. Chem. 2019, 7, 377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Loos P-F; Romaniello P Static and dynamic Bethe-Salpeter equations in the T-matrix approximation. J. Chem. Phys. 2022, 156, 164101. [DOI] [PubMed] [Google Scholar]
  • (36).Jacquemin D; Duchemin I; Blondel A; Blase X Benchmark of Bethe-Salpeter for Triplet Excited-States. J. Chem. Theory Comput. 2017, 13, 767–783. [DOI] [PubMed] [Google Scholar]
  • (37).Escudero D; Duchemin I; Blase X; Jacquemin D Modeling the Photochrome-TiO2 Interface with Bethe-Salpeter and Time-Dependent Density Functional Theory Methods. J. Phys. Chem. Lett. 2017, 8, 936–940. [DOI] [PubMed] [Google Scholar]
  • (38).Jacquemin D; Duchemin I; Blase X Is the Bethe-Salpeter Formalism Accurate for Excitation Energies? Comparisons with TD-DFT, CASPT2, and EOM-CCSD. J. Phys. Chem. Lett. 2017, 8, 1524–1529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (39).Azarias C; Duchemin I; Blase X; Jacquemin D Bethe-Salpeter Study of Cationic Dyes: Comparisons with ADC(2) and TD-DFT. J. Chem. Phys. 2017, 146, 034301. [DOI] [PubMed] [Google Scholar]
  • (40).Azarias C; Habert C; Budzák Š; Blase X; Duchemin I; Jacquemin D Calculations of n→π* Transition Energies: Comparisons Between TD-DFT, ADC, CC, CASPT2, and BSE/GW Descriptions. J. Phys. Chem. A 2017, 121, 6122–6134. [DOI] [PubMed] [Google Scholar]
  • (41).Blase X; Attaccalite C Charge-Transfer Excitations in Molecular Donor-Acceptor Complexes within the Many-Body Bethe-Salpeter Approach. Appl. Phys. Lett. 2011, 99, 171909. [Google Scholar]
  • (42).Gatti M; Sottile F; Reining L Electron-Hole Interactions in Correlated Electron Materials: Optical Properties of Vanadium Dioxide from First Principles. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 91, 195137. [Google Scholar]
  • (43).Cudazzo P; Sponza L; Giorgetti C; Reining L; Sottile F; Gatti M Exciton Band Structure in Two-Dimensional Materials. Phys. Rev. Lett. 2016, 116, 066803. [DOI] [PubMed] [Google Scholar]
  • (44).Sottile F; Marsili M; Olevano V; Reining L Efficientab initiocalculations of bound and continuum excitons in the absorption spectra of semiconductors and insulators. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 76, 161103. [Google Scholar]
  • (45).Jiang X; Zheng Q; Lan Z; Saidi WA; Ren X; Zhao J Real-Time GW-BSE Investigations on Spin-Valley Exciton Dynamics in Monolayer Transition Metal Dichalcogenide. Sci. Adv. 2021, 7, No. eabf3759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Liu C; Kloppenburg J; Yao Y; Ren X; Appel H; Kanai Y; Blum V All-Electron Ab Initio Bethe-Salpeter Equation Approach to Neutral Excitations in Molecules with Numeric Atom-Centered Orbitals. J. Chem. Phys. 2020, 152, 044105. [DOI] [PubMed] [Google Scholar]
  • (47).Dvorak M; Golze D; Rinke P Quantum Embedding Theory in the Screened Coulomb Interaction: Combining Configuration Interaction with GW/BSE. Phys. Rev. Mater. 2019, 3, 070801. [Google Scholar]
  • (48).Monino E; Loos P-F Spin-Conserved and Spin-Flip Optical Excitations from the Bethe-Salpeter Equation Formalism. J. Chem. Theory Comput. 2021, 17, 2852–2867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Loos P-F; Blase X Dynamical correction to the Bethe-Salpeter equation beyond the plasmon-pole approximation. J. Chem. Phys. 2020, 153, 114120. [DOI] [PubMed] [Google Scholar]
  • (50).van Setten MJ; Weigend F; Evers F The GW-Method for Quantum Chemistry Applications: Theory and Implementation. J. Chem. Theory Comput. 2013, 9, 232–246. [DOI] [PubMed] [Google Scholar]
  • (51).Golze D; Wilhelm J; van Setten MJ; Rinke P Core-Level Binding Energies from GW: An Efficient Full-Frequency Approach within a Localized Basis. J. Chem. Theory Comput. 2018, 14, 4856–4869. [DOI] [PubMed] [Google Scholar]
  • (52).Ren X; Rinke P; Blum V; Wieferink J; Tkatchenko A; Sanfilippo A; Reuter K; Scheffler M Resolution-of-identity approach to Hartree-Fock, hybrid density functionals, RPA, MP2 andGWwith numeric atom-centered orbital basis functions. New J. Phys. 2012, 14, 053020. [Google Scholar]
  • (53).Duchemin I; Blase X Robust Analytic-Continuation Approach to Many-Body GW Calculations. J. Chem. Theory Comput. 2020, 16, 1742–1756. [DOI] [PubMed] [Google Scholar]
  • (54).Wilhelm J; Golze D; Talirz L; Hutter J; Pignedoli CA Toward GW Calculations on Thousands of Atoms. J. Phys. Chem. Lett. 2018, 9, 306–312. [DOI] [PubMed] [Google Scholar]
  • (55).Förster A; Visscher L Low-Order Scaling G0W0 by Pair Atomic Density Fitting. J. Chem. Theory Comput. 2020, 16, 7381–7399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (56).Wilhelm J; Seewald P; Golze D Low-Scaling GW with Benchmark Accuracy and Application to Phosphorene Nanosheets. J. Chem. Theory Comput. 2021, 17, 1662–1677. [DOI] [PubMed] [Google Scholar]
  • (57).Duchemin I; Blase X Cubic-Scaling All-Electron GW Calculations with a Separable Density-Fitting Space-Time Approach. J. Chem. Theory Comput. 2021, 17, 2383–2393. [DOI] [PubMed] [Google Scholar]
  • (58).Ziaei V; Bredow T GW-BSE Approach on S1 Vertical Transition Energy of Large Charge Transfer Compounds: A Performance Assessment. J. Chem. Phys. 2016, 145, 174305. [DOI] [PubMed] [Google Scholar]
  • (59).Li J; Jin Y; Su NQ; Yang W Combining localized orbital scaling correction and Bethe-Salpeter equation for accurate excitation energies. J. Chem. Phys. 2022, 156, 154101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (60).Bruneval F; Hamed SM; Neaton JB A Systematic Benchmark of the Ab Initio Bethe-Salpeter Equation Approach for Low-Lying Optical Excitations of Small Organic Molecules. J. Chem. Phys. 2015, 142, 244101. [DOI] [PubMed] [Google Scholar]
  • (61).van Schilfgaarde M; Kotani T; Faleev S Quasiparticle Self-Consistent GW Theory. Phys. Rev. Lett. 2006, 96, 226402. [DOI] [PubMed] [Google Scholar]
  • (62).Kaplan F; Harding ME; Seiler C; Weigend F; Evers F; van Setten MJ Quasi-Particle Self-Consistent GW for Molecules. J. Chem. Theory Comput. 2016, 12, 2528–2541. [DOI] [PubMed] [Google Scholar]
  • (63).Caruso F; Rinke P; Ren X; Scheffler M; Rubio A Unified Description of Ground and Excited States of Finite Systems: The Self-Consistent GW Approach. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 081102. [Google Scholar]
  • (64).Caruso F; Rinke P; Ren X; Rubio A; Scheffler M Self-Consistent GW: All-electron Implementation with Localized Basis Functions. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 075105. [Google Scholar]
  • (65).Jacquemin D; Duchemin I; Blase X Assessment of the Convergence of Partially Self-Consistent BSE/GW Calculations. Mol. Phys. 2016, 114, 957–967. [Google Scholar]
  • (66).Véril M; Romaniello P; Berger JA; Loos P-F Unphysical Discontinuities in GW Methods. J. Chem. Theory Comput. 2018, 14, 5220–5228. [DOI] [PubMed] [Google Scholar]
  • (67).Monino E; Loos P-F Unphysical Discontinuities, Intruder States and Regularization in GW Methods. J. Chem. Phys. 2022, 156, 231101. [DOI] [PubMed] [Google Scholar]
  • (68).Li C; Zheng X; Su NQ; Yang W Localized Orbital Scaling Correction for Systematic Elimination of Delocalization Error in Density Functional Approximations. Natl. Sci. Rev. 2018, 5, 203–215. [Google Scholar]
  • (69).Elliott JD; Colonna N; Marsili M; Marzari N; Umari P Koopmans Meets Bethe-Salpeter: Excitonic Optical Spectra without GW. J. Chem. Theory Comput. 2019, 15, 3710–3720. [DOI] [PubMed] [Google Scholar]
  • (70).Jin Y; Su NQ; Yang W Renormalized Singles Green’s Function for Quasi-Particle Calculations beyond the G0W0 Approximation. J. Phys. Chem. Lett. 2019, 10, 447–452. [DOI] [PubMed] [Google Scholar]
  • (71).Li J; Jin Y; Rinke P; Yang W; Golze D Benchmark of GW Methods for Core-Level Binding Energies. 2022, arXiv preprint arXiv:2206.05627 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (72).Hohenberg P; Kohn W Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar]
  • (73).Kohn W; Sham LJ Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar]
  • (74).Parr RG; Weitao Y Density-Functional Theory of Atoms and Molecules; Oxford University Press, 1989. [Google Scholar]
  • (75).Bruneval F; Marques MAL Benchmarking the Starting Points of the GW Approximation for Molecules. J. Chem. Theory Comput. 2013, 9, 324–329. [DOI] [PubMed] [Google Scholar]
  • (76).Li J; Chen Z; Yang W Renormalized Singles Green’s Function in the T-Matrix Approximation for Accurate Quasiparticle Energy Calculation. J. Phys. Chem. Lett. 2021, 12, 6203–6210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (77).Ren X; Tkatchenko A; Rinke P; Scheffler M Beyond the Random-Phase Approximation for the Electron Correlation Energy: The Importance of Single Excitations. Phys. Rev. Lett. 2011, 106, 153003. [DOI] [PubMed] [Google Scholar]
  • (78).Ren X; Rinke P; Scuseria GE; Scheffler M Renormalized Second-Order Perturbation Theory for the Electron Correlation Energy: Concept, Implementation, and Benchmarks. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 88, 035120. [Google Scholar]
  • (79).Paier J; Ren X; Rinke P; Scuseria GE; Grüneis A; Kresse G; Scheffler M Assessment of Correlation Energies Based on the Random-Phase Approximation. New J. Phys. 2012, 14, 043002. [Google Scholar]
  • (80).Li J; Yang W Renormalized Singles with Correlation in GW Green’s Function Theory for Accurate Quasiparticle Energies. 2022, arXiv preprint arXiv:2207.06520 [DOI] [PubMed] [Google Scholar]
  • (81).Loos P-F; Romaniello P; Berger JA Green Functions and Self-Consistency: Insights From the Spherium Model. J. Chem. Theory Comput. 2018, 14, 3071–3082. [DOI] [PubMed] [Google Scholar]
  • (82).Krause K; Klopper W Implementation of the Bethe–Salpeter equation in the TURBOMOLE program. J. Comput. Chem. 2017, 38, 383–388. [DOI] [PubMed] [Google Scholar]
  • (83).Ghosh SK; Chattaraj PK Concepts and Methods in Modern Theoretical Chemistry: Electronic Structure and Reactivity; CRC Press, 2016. [Google Scholar]
  • (84).Stratmann RE; Scuseria GE; Frisch MJ An Efficient Implementation of Time-Dependent Density-Functional Theory for the Calculation of Excitation Energies of Large Molecules. J. Chem. Phys. 1998, 109, 8218–8224. [Google Scholar]
  • (85).Davidson ER The Iterative Calculation of a Few of the Lowest Eigenvalues and Corresponding Eigenvectors of Large Real-Symmetric Matrices. J. Comput. Phys. 1975, 17, 87–94. [Google Scholar]
  • (86).See http://www.qm4d.info, for an in-house program for QM/MM simulations (accessed June 30, 2022).
  • (87).Faber C; Boulanger P; Duchemin I; Attaccalite C; Blase X Many-body Green’s function GW and Bethe-Salpeter study of the optical excitations in a paradigmatic model dipeptide. J. Chem. Phys. 2013, 139, 194308. [DOI] [PubMed] [Google Scholar]
  • (88).Rocca D; Lu D; Galli G Ab initio calculations of optical absorption spectra: Solution of the Bethe-Salpeter equation within density matrix perturbation theory. J. Chem. Phys. 2010, 133, 164109. [DOI] [PubMed] [Google Scholar]
  • (89).Duchemin I; Deutsch T; Blase X Short-Range to Long-Range Charge-Transfer Excitations in the Zincbacteriochlorin-Bacteriochlorin Complex: A Bethe-Salpeter Study. Phys. Rev. Lett. 2012, 109, 167801. [DOI] [PubMed] [Google Scholar]
  • (90).Hoyer CE; Ghosh S; Truhlar DG; Gagliardi L Multiconfiguration Pair-Density Functional Theory Is as Accurate as CASPT2 for Electronic Excitation. J. Phys. Chem. Lett. 2016, 7, 586–591. [DOI] [PubMed] [Google Scholar]
  • (91).Stein T; Kronik L; Baer R Reliable Prediction of Charge Transfer Excitations in Molecular Complexes Using Time-Dependent Density Functional Theory. J. Am. Chem. Soc. 2009, 131, 2818–2820. [DOI] [PubMed] [Google Scholar]
  • (92).Dunning TH Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar]
  • (93).Kendall RA; Dunning TH; Harrison R Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar]
  • (94).Gui X; Holzer C; Klopper W Accuracy Assessment of GW Starting Points for Calculating Molecular Excitation Energies Using the Bethe-Salpeter Formalism. J. Chem. Theory Comput. 2018, 14, 2127–2136. [DOI] [PubMed] [Google Scholar]
  • (95).Véril M; Scemama A; Caffarel M; Lipparini F; Boggio-Pasqua M; Jacquemin D; Loos P-F QUESTDB: A Database of Highly Accurate Excitation Energies for the Electronic Structure Community. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2021, 11, No. e1517. [Google Scholar]
  • (96).Loos P-F; Comin M; Blase X; Jacquemin D Reference Energies for Intramolecular Charge-Transfer Excitations. J. Chem. Theory Comput. 2021, 17, 3666–3686. [DOI] [PubMed] [Google Scholar]
  • (97).Xu X; Yang KR; Truhlar DG Testing Noncollinear Spin-Flip, Collinear Spin-Flip, and Conventional Time-Dependent Density Functional Theory for Predicting Electronic Excitation Energies of Closed-Shell Atoms. J. Chem. Theory Comput. 2014, 10, 2070–2084. [DOI] [PubMed] [Google Scholar]
  • (98).Frisch MJ; et al.Gaussian16, Revision A.03; Gaussian Inc: Wallingford CT. 2016. [Google Scholar]
  • (99).Weigend F Accurate Coulomb-fitting Basis Sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057. [DOI] [PubMed] [Google Scholar]
  • (100).Eichkorn K; Treutler O; Öhm H; Häser M; Ahlrichs R Auxiliary Basis Sets to Approximate Coulomb Potentials. Chem. Phys. Lett. 1995, 240, 283–290. [Google Scholar]
  • (101).Feller D The Role of Databases in Support of Computational Chemistry Calculations. J. Comput. Chem. 1996, 17, 1571–1586. [Google Scholar]
  • (102).Schuchardt KL; Didier BT; Elsethagen T; Sun L; Gurumoorthi V; Chase J; Li J; Windus TL Basis Set Exchange: A Community Database for Computational Sciences. J. Chem. Inf. Model. 2007, 47, 1045–1052. [DOI] [PubMed] [Google Scholar]
  • (103).Pritchard BP; Altarawy D; Didier B; Gibson TD; Windus TL New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [DOI] [PubMed] [Google Scholar]
  • (104).Bruneval F; Dattani N; van Setten MJ The GW Miracle in Many-Body Perturbation Theory for the Ionization Potential of Molecules. Front. Chem. 2021, 9, 749779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (105).McKeon CA; Hamed SM; Bruneval F; Neaton JB An Optimally Tuned Range-Separated Hybrid Starting Point for Ab Initio GW plus Bethe–Salpeter Equation Calculations of Molecules. J. Chem. Phys. 2022, 157, 074103. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ct2c00686_si_001

RESOURCES