Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jun 19.
Published in final edited form as: J Chem Theory Comput. 2020 Dec 28;17(1):240–254. doi: 10.1021/acs.jctc.0c01015

Block-Localized Excitation for Excimer Complex and Diabatic Coupling

Peng Bao 1, Christian P Hettich 2, Qiang Shi 3, Jiali Gao 4
PMCID: PMC8214332  NIHMSID: NIHMS1713894  PMID: 33370101

Abstract

We describe a block-localized excitation (BLE) method to carry out constrained optimization of block-localized orbitals for constructing valence bond-like, diabatic excited configurations using multistate density functional theory (MSDFT). The method is an extension of the previous block-localized wave function method through a fragment-based ΔSCF approach to optimize excited determinants within a molecular complex. In BLE, both the number of electrons and the electronic spin of different fragments in a whole system can be constrained, whereas electrostatic, exchange, and polarization interactions among different blocks can be fully taken into account of. To avoid optimization collapse to unwanted states, a ΔSCF projection scheme and a maximum overlap of wave function approach have been presented. The method is illustrated by the excimer complex of two naphthalene molecules. With a minimum of eight spin-adapted configurational state functions, it was found that the inversion of La and Lb states near the optimal structure of the excimer complex is correctly produced, which is in quantitative agreement with DMRG-CASPT2 calculations and experiments. Trends in the computed transfer integrals associated with excited-state energy transfer both in the singlet and triplet states are discussed. The results suggest that MSDFT may be used as an efficient approach to treat intermolecular interactions in excited states with a minimal active space (MAS) for interpretation of the results and for dynamic simulations, although the selection of a small active space is often system dependent.

Graphical Abstract

graphic file with name nihms-1713894-f0001.jpg

1. INTRODUCTION

It is often more convenient to use a diabatic representation of localized initial and final states than the delocalized adiabatic ground and excited states to determine reaction rates and to model solvent reorganization in electron transfer (ET) and excited-state energy transfer (EET) processes.1,2 Furthermore, intermolecular interactions of aromatic compounds can have drastic changes in the excited state, sometimes forming stronger excimer complexes than that in the ground state.3 The subsequent nonadiabatic dynamics of the excimer complexes can undergo radiative emission or conical intersection to return to the ground state.2 In this case, calculations of nonadiabatic couplings require derivatives of the wave function, and the derivative coupling terms may become singular at the crossing points.4 In contrast, the potential energy surfaces of diabatic states are smooth.5 However, diabatic states are not unique,6 and many computational techniques have been described to approximate diabatic states.4,714 To this end, we have proposed a generalized diabatic at construction (GDAC) approach on the basis of an effective valence bond (VB) classification,15,16 which maximally constrains these diabatic states in bonding characters to the asymptotic dissociation limit for the processes of interest. GDAC diabatic states are physically interpretable configurations that form an active space to describe the adiabatic potential energy surfaces. Therefore, the validity and adequacy of GDAC diabatic states can be verified by comparison of experimental observables with properties determined from their resulting adiabatic states. In this article, we present a block-localized excitation (BLE) method to define locally excited configurations for treating excited energy transfer and for describing intermolecular interactions of excimer complexes.

VB theory naturally has orbital constraints, which can be generalized to represent diabatic states. The large number of configurations in ab initio VB theory prevents its practical use on large and condensed phase systems,17 though modern approaches are improving its applicability.1820 On the other hand, block-localized wave function (BLW) theory allows for the construction of an effective VB-like configuration using a single determinant.21 In BLW, a diabatic state is represented in terms of block-localized orbitals (BLO), an effective representation of the ab initio VB configurations within a molecular fragment.2225 The BLOs can be optimized in wave function theory (WFT)21,22,24 and using Kohn–Sham density functional theory (KS-DFT).26,27 In this regard, the block-localized Kohn–Sham orbitals are strictly localized on a given fragment by construction that is constrained in orbital space.22,25 Related localization methods include that developed by Stoll and coworkers,28 the self-consistent field for molecular interactions (SCF-MI),2931 locally projected molecular orbitals,32,33 absolutely localized molecular orbitals,34,35 and gradient methods with respect to orbital coefficients.28,3639 The BLW configurations can be used in nonorthogonal configuration interaction (NOCI) either in WFT,40 called mixed molecular orbital and VB theory,23,41 or in multistate density functional theory (MSDFT),27,4244 to yield the adiabatic ground and excited states. MSDFT has been used in a number of applications,15,16,27,43,4547 including proton-coupled electron transfer (PCET)48 and singlet fission49 with approximate local-excitations.

A closely related approach is constrained density functional theory-configuration interaction (CDFT-CI) based on density confinement in a certain spatial domain.5052 CDFT-CI has been used to estimate electronic couplings for ET and EET reactions, but the results heavily depend on basis functions used (refs 53 and 54 and references there in). Recently, non-Aufbau occupation of the orbitals was employed to construct localized singlet excited state in singlet fission.55 One question here is physically constraining electrons within a spatial region on a property that is in fact distributed in the full space. Consequently, the computation also depends on the particular population method used because it is not a physical observable,54 and the constrained integration number of electrons actually is a composite from the wave functions from different target states.56 Another related approach is the Frenkel–Davydov exciton approach using states derived from time-dependent density functional theory calculations for isolated monomers,5759 with which both dipole-coupling and Dexter exchange interactions are directly determined as in the study of singlet fission using MSDFT49 and PCET.48 In the latter case, local states are variationally optimized in the presence of the environment through a QM/MM interaction Hamiltonian.49

In this work, we extended the block-localization method to treat excited states constrained within individual fragments. Similar to previous works for the ground state, the main focus is to obtain BLOs in the excited states. In principle, such calculations can be performed using CASSCF or time-dependent density functional theory (TDDFT) within a fixed molecular fragment.60 However, a single-determinant representation rather than a multiconfigurational approach is desired to keep spirit of computational efficiency of KS-DFT for the ground state in MSDFT for the excited state. Here, we find a ΔSCF type of models suitable for our aim to construct localized excited states.61 Of course, the performance of ΔSCF methods is rather system-dependent, encountering optimization collapse to the ground state or to undesired configurations. To resolve these problems, a number of techniques have been developed such as the maximum overlap (MOM),62 improved MOM,63 the MOM square methods,64 and square gradient minimization (SGM).65 These approaches can be directly adopted to construct diabatic states in the active space for MSDFT. In particular, the present BLE approach is a variation of ΔSCF methods, which is illustrated by the intermolecular interactions of a naphthalene dimer complex both in its first triplet and in two singlet excited states.6669 The latter features an interesting curve crossing of two low-lying singlet states as the two aromatic compounds are in close proximity. This was clearly resolved recently using a DMRG-CASPT2 approach including the full active space of 20 electrons in 20 orbitals, but such a curve crossing is only revealed by adding post-SCF perturbation corrections for dynamic correlation.68

The remainder of this article is organized as follows. In Section 2, we present the theory and computational details of a BLE method to obtain single-determinant representation of excited configurations in the MSDFT active space. An SCF projection method and an MOM of wave function procedure through singular value decomposition are described to ensure ΔSCF convergence. Numerical results on the naphthalene excimer complex are then presented in Section 3, where we apply the present method to determine the transfer integrals associated with excited energy transfer both in the lowest singlet and in the triplet states, and binding interactions of excimer formation in these two states. Finally, the article is concluded with summary remarks in Section 4.

2. THEORY

2.1. BLW Method for the Ground State.

For completeness, we begin with a brief review of the BLW method for the ground state, which will be used later to construct diabatic states of locally excited configurations. In BLW, the determinant wave function is written as follows21,25

Ψu=NuA^{Φ1Φ2ΦK} (1)

where Nu is the normalization constant, A^ is the antisymmetrization operator, K is the number of blocks, and ΦA denotes a product of spin-orbitals of block A, ΦA=φ1Aαφ1AβφnAαAαφnAβAβ, which are localized by construction

ϕA=χTA (2)

where φA is a matrix of BLOs in block A and χ a row vector of basis orbitals, and TA is a matrix of orbital coefficients. In general, BLOs from different blocks are nonorthogonal.

The total coefficient matrix T is block-diagonal in the original BLW method when different blocks have different basis orbitals. In this work, that restriction has been lifted such that different blocks can share a group of common basis functions. The total density matrix D is given by27

D=T(TST)1T (3)

where S = χχ is the overlap matrix of the atomic orbitals. Then, the energy of such a block-constrained system is given by

E=Tr(Dh)+12Tr(DJD)+Exc(ρ) (4)

where h and J are the usual one-electron and Coulomb integrals, and Exc(ρ) is the exchange–correlation energy in standard KS-DFT. The constrained optimization of the BLOs can be achieved using the Broyden–Fletcher–Goldfarb–Shanno updating scheme70 with the energy gradient given as follows28,3639

δE=Tr(δDF)=2Tr[(ISDα)FαTα(TαSTα)1δTα+(ISDβ)FβTβ(TβSTβ)1δTβ] (5)

where F is the Fock matrix.

The gradient optimization method can be applied to closed shell, restricted open shell, and unrestricted open shell cases. Both analytic orbital gradient and numerical orbital Hessian have been implemented in our program, though the latter is time-consuming. The gradient optimization method is used when common basis functions are used in different blocks.37,38 Three different forms of SCF equations have been described in the literature.27,2935,71 In Appendix A, we provide another direct derivation and highlight the properties of the non-orthogonal block SCF equations. These methods are derived on the basis of zero energy gradient, and their computational costs are essentially the same when the number of the blocks is not large. In this work, the approach described in ref 27 is used for the ground state calculations, and is extended to the localized orbitals for excited states (below).

To accelerate the convergence of the SCF, we employ the DIIS (direct inversion in the iterative subspace) method72 by using energy gradients (Appendix A) as the error vectors and updating the projected Fock matrix and the effective overlap matrix for each block. We have also implemented a DIIS by updating the Fock matrix and coefficients of the whole system. Efficiencies of the two methods are found to be similar for the naphthalene dimer system.

2.2. Block-Localized Excitation.

In the present study, we are interested in obtaining a computationally efficient representation of locally excited configurations within a given block. Traditional methods for excited states are usually of multiconfigurational nature, even for the simplest excited state approaches such as configuration interaction with single excitations (CIS) and TDDFT. Clearly, a multiconfigurational wave function technique within a fixed block would have not been desirable for a DFT-based method for the intended applications. To this end, we employ a ΔSCF-like approach to represent locally excited configurations.61 ΔSCF is a single determinant technique, with which the exchange–correlation functional developed for KS-DFT can be directly used, keeping in mind that the present Kohn–Sham determinant using BLO is constrained in the framework presented in Section 2.1. In a ΔSCF scheme, the localized orbitals can be optimized in a way that is very similar to the ground state Hartree–Fock–Roothaan equation. However, it is known that the original ΔSCF method often suffers from SCF convergence,61 and when it converges, an unwanted excited state or the ground state could be obtained.

A number of local excitation ΔSCF methods have been reported,7386 including the MOM method62 and MOM squared,64 the SGM,65 the local SCF method for core-excited states,74 and a constricted variational DFT (SCF-CV-DFT) by one electron excitation to the virtual orbital space.75 The single determinant approach can also be achieved by orthogonality constraints on the excited state. The “big shift” method was formulated by setting a large value (more than 1010 au) to the diagonal element of the Fock matrix corresponding to ground state orbitals.76 Similarly, an asymptotic projection method was suggested by adding an infinite projector to the frozen orbital in the Hamiltonian operator.77 Yet, another alternative is the guided SCF approach by transformation of the Fock matrix from the atom orbitals (AO) basis to the excited state orbitals basis.78 Similar to the BLW method, but enforcing orbital orthogonality between the core and the valence is the projection configuration interaction method.79

In this work, we adopt a ΔSCF approach to represent locally excited non-Aufbau configurations and employ two techniques to ensure SCF convergence. In the first method, we project the ΔSCF equations to the BLO space of the ground state in each iteration to retain the desired order of orbitals for the subsequent SCF iteration. Either the HF or KS-DFT method can be used. The approach in this study can be summarized below, making use of a general case of a single electron excitation from an occupied BLO j to a virtual BLO b. For clarity, only the α spin is shown.

  1. We obtain BLOs of the ground state by solving the generalized secular equation F0T0 = ST0E0, where the superscript emphasizes ground state orbitals.

  2. The initial guess determinant corresponding to the excitation φj0(α)φb0(α) is constructed by switching the occupied and virtual BLOs: Ψb0BLE=|(φ10φ20φb0φn0)(φn+10φj0φm0)α. Here, the two pairs of parentheses denote occupied and virtual spaces, and there is no change in the occupation of β electrons.

  3. To solve the eigen equation f|φ〉 = |φε, we set |φ〉 = χT0T, and left multiply φ0|=(χT0). This has the effect of diagonalizing the Fock-like matrix in the BLO basis of the ground state, (T0†FT0)T = (T0†ST0)TEBLE, to yield a new set of orbitals T for the block-localized excited determinant. After diagonalization, it is not needed to resort the eigenvectors according to the energy of each orbital as the order of ground state molecular orbitals is kept.

  4. The effective Fock matrix for the BLE configuration can be obtained in the AO basis using the transformation matrix Tnew = T0TBLE for the next SCF iteration in step 3 until convergence, where TBLE is T with the orbitals φj(α) → φb(α) swapped. Usually, good convergence can be obtained by projecting the ground state orbitals. We can also use the orbitals produced in a previous iteration to do the projection. Alternatively, we can use a permutated T0 to realize the ΔSCF excitation, then follow the process above without eigenvector permutation.

The above procedure is general in that any combinations of α and β excitations, including multiple excitations can be included. We note that the guided SCF method also employs a similar projection method.78 The determinant that is optimized using the method presented above is not orthogonal to the ground state, but the overlap with the ground-state determinant is usually small. To obtain a localized excited state that is orthogonal to the localized ground state, the particle orbital can be obtained from the projection of unoccupied orbital space after obtaining the hole orbital from a full space projection. The derivation within localized constraint is presented in Appendix A. Furthermore, the adiabatic ground and excited states can be obtained using MSDFT through configuration interaction among these determinant configurations (see the next section) such that conical intersection between the ground and first excited states can be obtained.43

The second method that we have implemented to improve the ΔSCF convergence is based on the maximum overlap of wave function (MOW) with the initial determinant for the excited configuration, denoted as Tb0BLE=(φ10,,φb0,,φn0)α. All steps in the MOW procedure are identical to those outlined above, except step 4 to choose the correct n occupied orbitals of the BLE configuration. Here, at the Ith iteration, the goal is to construct a determinant for a specific BLE configuration that maximizes the overlap with the initial guess Ψb0BLE :maxφ{Ψ0bBLEΨib=|T0bBLESTIb|}. It would be unfeasible to try out all Cmn possibilities by brute force. Instead, we choose the n orbitals that maximize its Frobenius norm, T0bBLESTIbF=injnOi0Ij2, which can be easily accomplished by singular value decomposition. This is equivalent to an approach called MOM squared.62,64

With the above ΔSCF projection and the MOW method, BLOs for an excited configuration within a given block A can be obtained using the approach described in Subsection 2.1. Because the excited determinants are fully localized, there is no charge transfer contribution between different blocks. We have adopted an unrestricted open-shell treatment both in the ΔSCF projection and in the MOW method, which yield identical results at convergence. All results in the examples illustrated below are obtained by using the block-localized ΔSCF projection scheme.

2.3. MSDFT and Electronic Coupling of BLE Diabatic States.

The adiabatic ground and excited states of the delocalized system can be obtained using MSDFT through NOCI by solving the generalized secular equation,23,27,41,42,44,87 for which the BLE configurations form a minimal active space (MAS) for the adiabatic excited states of interest

HC=SCE (6)

where E is a diagonal matrix of adiabatic energies, C is a matrix of the CI-coefficient, H is the CI-Hamiltonian, and S is the overlap matrix, which is determined using the separately optimized Kohn–Sham BLE-determinants (SAB=ΨABLEΨBBLE).23 The adiabatic energies of the ground and excited states are given using27,42

EIMS[ρI]=AcAI2HABLE[ρA]+2A>BcAIcBIHABTDF[ρAB,ΨABLE,ΨBBLE] (7)

In MSDFT, the diagonal matrix elements of the CI Hamiltonian HABLE[ρA] are the energies determined using the methods described in Section 2.1 for the ground state determinant and in Section 2.2 for block-localized excited configurations. Each off-diagonal element consists of two terms: a standard interaction term between two nonorthogonal determinants, ΨABLE and ΨBBLE, and a transition density functional (TDF) energy15,16,27,42,44

HABTDF=ΨABLE|H|ΨBBLE+EABTDF[ρAB,ΨABLE,ΨBBLE] (8)

where EABTDF is the dynamic correlation contribution to the TDF, or simply, the TDF that depends on the transition density ρAB as well as the determinant functions ΨABLE and ΨBBLE. TDF is a novel class of correlation functional, which does not exist in KS-DFT, accounting for dynamic correlation in electronic coupling and playing a role in removing any double counting of electron correlation in a multiconfigurational representation of DFT.43 As in KS-DFT, the exact functional form of EABTDF is not known, but under special conditions, such as the spin coupling of two unpaired electrons forming a pair of spin-adapted singlet and triplet states, the value of EABTDF can be obtained consistently with a particular KS-DFT functional approximation to the two spin-mixed determinant configurations.43,88 This is achieved by enforcing the energy degeneracy of the high (Ms = 1) and low (Ms = 0) spin projections of the triplet states proposed by Ziegler et al.61 A related approach is the spin-restricted density functional theory for biradical calcualtions.8991 Here, the correlation energy of the high-spin projection state EcKS[ρ(|1,1)], which can be adequately represented by a single determinant in KS-DFT, along with that of the mixed low-spin determinant EcKS[ρ(|1,0)], is used to define the magnitude of the correlation contribution43

EABTDF[ρAB,ΨABLE,ΨBBLE]=EcKS[ρ(1,1>)]EcKS[ρ(1,0>)] (9)

In general applications, one possible approximation for EcTDF is a determinant-weighted average of correlation energies of the two interacting states43

EABTDFΨABLE|H|ΨBBLEEAHF[ΨABLE]+EBHF[ΨBBLE][EAc(ρA)+EBc(ρB)] (10)

where EAHF[ΨABLE] and EBHF[ΨBBLE] are HF energies determined by using the KS-determinants for the BLE configurations, and EAc(ρA) and EBc(ρB) are the corresponding correlation energies. Alternatively, the TDF energy can be approximated by an overlap scaled average of the correlation energies of the two diabatic states.27,42

EABTDF12MABKS[EAc(ρA)+EBc(ρB)] (11)

where MABKS is the overlap between two Kohn–Sham determinants. The numerical results from these approximate approaches are similar.

These block-localized configurations may be considered as charge- and energy-localized diabatic states, and their electronic coupling elements are related to the transfer integrals for electron and excited energy transfer applications. Within the MSDFT framework and CDFT-CI, the ET integral between any two diabatic states can be calculated.27,54,56 Because the BLE configurational states are generally non-orthogonal, the effective coupling term VAB between states A and B can be calculated through Löwdin orthogonalization

VAB=HABMAB(HA+HB)/21MAB2 (12)

where MAB is the overlap integral of states A and B.

3. COMPUTATIONAL DETAILS

The naphthalene monomer structure was optimized with the D2h symmetry using Gaussian-16 at the B3LYP/6–31+G(d) level of theory, following the work of Shirai et al.68 Then, the monomer structure is kept fixed and is used to construct a face-to-face dimer configuration at various separation distances, ranging from 2.9 to 20 Å. A structure at 50 Å was also built to verify asymptotic energies. It is known that the face-to-face structure of the excimer complex has the lowest energy in both the singlet and triplet states.6669,9294 Further, upon relaxation from the monomer geometry, the Franck–Condon approximation suggests that the diabatic couplings should not change significantly.67,69 In BLE calculations, the dimer is separated into two blocks, each containing one naphthalene molecule. Single-point energy calculations are performed using MSDFT, for which a total of 16 block-localized configurations plus the adiabatic ground state have been obtained using block-localized KS-DFT. The PBE0 functional and the cc-pVDZ basis set are used in all calculations.95 For the excitations on both monomers and with all three combinations of source and target orbitals, the separately optimized open-shell singlet determinants with αβ and βα spin combinations are converged to the same set of BLO, based on which the spin-coupled singlet and triplet states are obtained. This is equivalent to switch α and β electron spins of an open-shell configuration to yield a spin pair. The all-spin up triplet configuration (Ms = 1) at each geometry was obtained by separate open-shell calculations. The energy for the Ms = 1 state is used to determine the correlation energy term EABTDF for the TDF of spin-pair coupling (eq 6). This makes the orbitals not entirely compatible, but the energy difference is less than 0.1 eV, smaller than uncertainties from the method itself, compared with the results obtained using a common set of orbitals. Although this is not a conventional procedure for spin-adapted configurations,96 it is not without precedents and we have used it in a number of applications.15,16,42,43 This is also consistent with the use of a single determinant to treat open-shell systems using KS-DFT,97,98 yielding lower energy than that adopting a restricted open-shell approach. All other computational details are the same as in other studies.

The potential energies curves for the naphthalene dimer have also been computed using MSDFT with the M06–2X density functional and the cc-pVDZ basis, yielding the same trends with the benefit that additional and empirical dispersion corrections are not needed. The results are given in the Supporting Information.

The method described in the previous section has been implemented in a locally modified version of the GAMESS-US program,99 with which all calculations have been performed.

4. RESULTS AND DISCUSSION

An important application of the present BLE method is to understand dimer interactions of aromatic species in the excited states and to determine the electronic coupling in excitation energy transfer (EET) processes from one part of the molecular system to another.100 We first apply the BLE method to determine the diabatic and adiabatic potential energy curves for the naphthalene dimer in the face-to-face configuration. This is followed by calculations of the EET couplings both in the singlet and triplet states. This system is chosen because it has been extensively studied and there is a challenging curve-crossing that is only obtained with a method balancing static and dynamic correlation.6669,9294,101 Here, we chose the face-to-face stacked parallel configuration in our study, corresponding to the minimum-energy configuration in the first singlet and triplet excited states of the dimer complex (excimer).6769,92,102104

4.1. Excimer Interactions in the Singlet State.

Shirai et al.68 reported a most comprehensive investigation of the naphthalene excimer complex in the singlet state using DMRG-CASPT2 with cc-pVDZ and cc-pVTZ basis functions, which revealed an inversion of the 1La and 1Lb states along the dimer separation path. In that study, the full π electrons in an active space of [20e, 20o] were used. Dynamic correlation is critical for the excimer calculations, without which the qualitative trends of the two low-lying states are wrong at the DMRG-CASSCF level.

MSDFT follows a dynamic-then-static (DTS) ansatz, in which dynamic correlation is incorporated in the optimization of BLOs through Kohn–Sham density functionals for each individual determinant configuration.27,42 Moreover, each configuration in the active space for MSDFT is separately optimized, further introducing orbital relaxation. Consequently, it is expected that a minimal number of essential configurations would be sufficient to describe the properties of the relevant low-energy states. Following the insights provided by Shirai et al. on the naphthalene excimer complex, we tested the idea of a MAS for the two low-lying excited states of naphthalene excimer by MSDFT (Figure 1). This includes three local singly excited configurations for each monomer (eqs 1320), plus a pair of charge transfer states between two monomers (eqs 19 and 20), which are depicted in Figure 1.

ΩA*Bs23s0=12{ΨA*B(S23¯S0)ΨA*B(S2¯3S0)} (13)
ΩA*Bs24s0=12{ΨA*B(S24¯S0)ΨA*B(S2¯4S0)} (14)
ΩA*Bs13s0=12{ΨA*B(S13¯S0)ΨA*B(S1¯3S0)} (15)
ΩAB*s0s23=12{ΨAB*(S0S23¯)ΨAB*(S0S2¯3)} (16)
ΩAB*s0s24=12{ΨAB*(S0S24¯)ΨAB*(S0S2¯4)} (17)
ΩAB*s0s13=12{ΨAB*(S0S13¯)ΨAB*(S0S1¯3)} (18)
ΩAB+=12{ΨAB+(DαDβ)ΨAB+(DβDα)} (19)
ΩA+B=12{ΨA+B(DβDα)ΨA+B(DαDβ)} (20)

Figure 1.

Figure 1.

Schematic depiction of (a) block-locally excited configurations of monomer A* in the presence of monomer B in its ground state, (b) block-locally excited configurations of monomer B* in the presence of monomer A in its ground state, and (c) two charge transfer states between monomers A and B. Each yellow block shows one block-localized determinant configuration of the whole dimer, and the left and right halves within it show the four relevant fragment orbitals of the two monomers, respectively. The four BLOs used in the figure include the two highest occupied fragment orbitals, φ1 and φ2, and the two lowest unoccupied fragment orbitals, φ3 and φ4, of the excited monomer. The locally excited monomer is indicated by a star. Only singlet spin-adapted states are shown, whereas the symmetric (in-phase) combination of two determinants corresponds to a triplet state.

In eqs 1320, the notation Sia¯ or Si¯a denotes a single excitation within a given singlet block in which the β-spin of an electron in the corresponding orbital is represented by an over-hat, and D specifies a doublet block with α and β spin of the unpaired electrons in the occupied (subscript) and virtual (superscript) orbitals (with the reference ground-state configuration, S0). In short, eqs 1320 represent single excitations in a blocked 4-electron and 4-orbital space minus the largest energy transition, in which each individual determinant is separately optimized using BLE to yield spin-adapted configurational state functions for singlet (as written) and triplet states (symmetric combinations). Note that MAS is not necessarily a block-box selection in that a great deal of insights of the configuration space ought to be known in the first place, such as the naphthalene dimer in the present study. Furthermore, the configurational state functions in MAS are not equivalent to the orthogonal configurations in CASSCF. However, MAS, when verified by comparison of its resulting adiabatic energies with experiment, can be quite useful for interpretation from energy decomposition analysis25 and for carrying out dynamic simulations of nonadiabatic processes.

We begin with the low-lying excited states of a naphthalene monomer, whose active space in MSDFT can be represented using eqs 1315 (omitting the orbitals on fragment B). There are two singlet excited states of similar energies, corresponding to 1B3u(1Lb) and 1B2u(1Lb) symmetry.105 The latter is of HOMO → LUMO character given by eq 13 (Figure 1a), and the complementary symmetric combination of the two determinants, {ΨA*B(S23¯S0)+ΨA*B(S2¯3S0)}/2, corresponds to the MS = 0 multiplet of the first triplet excited state |1,0〉. The MS = 1 state of the triplet (|1,1〉) can be adequately treated by KS-DFT, and a vertical excitation energy of 3.13 eV above the ground state is obtained with PBE0/cc-pVDZ. The value is in accord with experiment (3.0 eV), and is used to define61 the TDF for spin-coupling interaction (eq 9).43 We obtained a vertical excitation energy of 4.52 eV 1La singlet state (eq 10) from MSDFT@PBE0, falling in the range of experimental observations (4.45106,107 and 4.7 eV).108

For the diabatic coupling between the symmetry-allowed spin-adapted states E1415TDF (eqs 14 and 15), we used the average of the two spin-coupling values determined with eq 9106,107 because the transition energies are nearly degenerate for HOMO – 1 → LUMO (4.48 eV) and HOMO → LUMO + 1 (4.51 eV) single excitations.68 This yields an energy for 1Lb of 4.14 eV in comparison with the experimental value of 3.97 EV.106,107 The E1415TDF value for naphthalene is used for the local (monomer) 1Lb excitation in the dimer complex at all geometries because it is intrinsically a local (monomer) property. For comparison, CASPT2/cc-pVTZ calculations yielded vertical excitation energies of 3.83 and 4.34 eV for the 1Lb and 1La states,68 whereas 4.03 and 4.56 eV were obtained by Roos and coworkers using natural atomic orbital basis functions.109 Overall, we found that with an MAS of three spin-adapted diabatic states in MSDFT, the computed energies of the three low-lying excited states (two singlet and one triplet) are within 0.2 eV of experimental values.106,107

Second, Figure 2 shows the potential energy curves for the ground state and three low-lying excited states determined from MSDFT@PBE0 calculations (solid curves), along with the corresponding curves for the singlet states by DMRG-CASPT2 (dashed curves) and DMRG-CASSCF (dotted curves) from ref 68. Because the energies for the excited states are overestimated by 0.3–0.5 eV from DMRG-CASPT2 calculations of the naphthalene monomer, this trend is transferred to the potential energy curves of the dimer complex. At large interfragment separations (>3.5 Å), the two low-lying singlet states of naphthalene in the excimer complex follow the same order in energy as that of the monomer: 1La(1B3g)>1Lb(1B2g). However, the charge-transfer (CT) state (Figure 1c and eqs 19 and 20) is strongly mixed in 1La+ at short distances (Figure 3), stabilizing intermolecular interactions between two naphthalene molecules in the excimer complex. This leads to an inversion of the 1La and 1Lb energy levels to bring 1La as the first singlet excited state of the excimer complex. MSDFT calculations show that the singlet energies are inverted at a distance of about 3.9 Å (Figure 2), somewhat longer than the DMRG-CASPT2 prediction (distances of inversion at 3.26–3.62 Å using, respectively, cc-pVDZ and cc-pVTZ basis sets).68 The qualitative trend of these two low-lying states are critically dependent on the active space used in CASPT2 calculations; the correct order is only obtained when the active space is greater than [12e, 12o]. The order of energy levels is not correctly obtained using the single-reference EOM-CCSD/cc-pVTZ method.68 Furthermore, the work of Shirai et al. demonstrated that dynamic correlation is absolutely necessary for the excimer complex, without which 1Lb remains the lowest state incorrectly at all separations, below 1La by as much as 1.6 eV using DMRG-CASSCF [20e, 20o].

Figure 2.

Figure 2.

Computed potential energy curves of the adiabatic ground and the first triplet (T1) and first two singlet (S1) excited states from MSDFT along with those determined by using DMRG-CASPT2 (DMRG) and DMRG-CASSCF (CAS) with 20 electrons in 20 orbitals for the singlet states (ref 68). The PBE0 functional is used in MSDFT calculations.

Figure 3.

Figure 3.

Computed adiabatic singlet excited states, along with BLE diabatic states, excitonic resonance (ER) diabatic states, and CT resonance diabatic states. Energies are given relative to the separated naphthalene molecules in the ground state. All computations have been performed using MSDFT with the PBE0 density functional (MSDFT@PBE0) and the cc-pVDZ basis set.

An anonymous referee noted that the energy of the 1Lb state is more repulsive at short distances than that determined by DMRG-PT2, which may reflect CT effects not included in the small active space used in this study. While CT effects are captured in the 1La state, high-lying CT states at longer separations may indeed become competitive to make contributions at a short interfragment distance, and this has been observed in the study of the photochemical dissociation of LiH.15

Third, using the active states defined by eqs 1320, we obtained vertical emission (fluorescent) energies of 3.0 and 4.0 eV from 1La and 1Lb, respectively, which are similar to DMRG-CASPT2 values of 3.25 and 3.65 eV (Supporting Information, see ref 68), and in reasonable accordance with an experimental value of 3.13 eV for the first state.110 These values are dependent on the structures of the excimer complex, which were not optimized in the present study. We obtained a binding energy of 1.44 eV for the 1La state of the excimer (including Grimme’s D3 dispersion correction applied to all states), stronger than that predicted using DMRG-CASPT2 (1.23 and 1.33 eV using cc-pVDZ and cc-pVTZ basis sets) and a value of 0.99 eV using SOS-CIS(D0)/aug-cc-pVDZ.69 Basis set superposition errors (BSSE), which decrease the binding energy by 0.33 and 0.17 eV for the two basis sets, have been noted.68 In principle, the present BLE configurations in MSDFT do not have BSSE by construction, which was pioneered by Gianinetti et al.2931 The binding energy is much weaker for the 1La state at about 0.36 eV from MSDFT, comparable to the DMRG-CASPT2 value (ca. 0.5 eV).68

Finally, it is of interest to note that only eight spin-adapted configurational state functions (16 determinants) plus the ground state are used in the present MSDFT calculations, whereas the DMRG study included all π-electron in a [20e, 20o] active space, consisting of more than 3 × 1010 configurations.68 Both qualitative trends as well as quantitative results from MSDFT with an MAS are at least of the quality of DMRG-CASPT2. It is sobering to note that without perturbation correction for dynamic correlation, 1La from DMRG-CASSCF [20e, 20o] is more than 2 eV too high than the experimental value at large dimer separation, while the result for 1Lb is nearly perfect because there is little dynamic correlation (Figure 2). The agreement highlights the usefulness of a DTS approach for balancing dynamic and static correlation in excited state calculations at a cost comparable to KS-DFT for the ground state.

The nature of resonance stabilization of 1La relative to that of 1Lb is revealed by comparing diabatic states in the two symmetry groups in Figure 3. For 1Lb(11B2g), the BLEs of each naphthalene molecule in the presence of the second monomer are constructed as follows

|A*B(Lb)=a24AΩA*BS24S0a13AΩA*BS13S0 (21)
|AB*(Lb)=a24BΩAB*S0S24a13BΩAB*S0S13 (22)

where the coefficients aiaA/B are determined by using MSDFT (eqs 14 and 15 for |A*B〉 and eqs 17 and 18 for |AB*〉). These two diabatic states are completely degenerate shown as a single curve in Figure 3 (red). Interestingly, the presence of the second monomer lowers the BLE state |A*B(Lb)〉 by 0.30 eV relative to an isolated naphthalene, largely because of the empirical D3 dispersion term. The resonance delocalization of these two BLEs is weak, resulting in a pair of adiabatic states 1Lb and 1Lb+ (blue dashed curves), whose energy splitting is only 0.17 eV at the minimum of the complex (3.4 Å), gaining only 0.06 eV in binding energy for the lower state 1Lb. This is consistent with the results from DMRG calculations with and without the PT2 correction (Figure 2).

In contrast, it is no longer physically meaningful to define a BLE (HOMO → LUMO) diabatic state for 1La(11B3g) because CT configurations dominate coupling interactions and it is not a local (monomer) property. Of course, one can nevertheless construct a BLE for the “covalent” diabatic configuration (not shown) similar to eqs 18 and 19, and its energy monotonically decreases as the dimer distance shortens. In this case, we construct two pairs of resonance states from the local HOMO → LUMO excited configurations, ER, and from the CT configurations across monomer boundaries, charge-transfer resonance (CR)

(La±)ER=12{ΩA*BS23S0±ΩAB*S0S23} (23)
(CT)CR±=12{ΩAB+±ΩA+B} (24)

Figure 3 shows that the resonance splitting (0.88 eV) in (La±)ER is significantly greater than that in the 1Lb state, gaining a resonance stabilization energy of 0.65 eV, and the resonance propagates to long dimer separations without vanishing at 10 Å (black and gray curves). The interaction between the CT configurations is less pronounced (green and orange curves). Importantly, electronic coupling of the BLE resonance diabatic state (La)ER and that of the CT state (CT)CR lowers the complexation energy of the BLE resonance state by 0.79 eV and shifts the minimum of the ER state from 3.3 to 3.1 Å (Figure 3). Therefore, both the resonance delocalization of BLE and CT stabilization contributes significantly, with the latter playing a slightly greater role (55 vs 45%) in the total excimer binding energy.

4.2. Excimer in the Triplet State.

The computed first triplet excited state is shown in Figure 2 and its diabatic state composition is separated in Figure 4. Although a number of higher energy states, both singlet and triplet, can be obtained, other diabatic configurations may or may not be needed to fully characterize these adiabatic states. Because the main goal is to illustrate the BLE optimization method to define configurational states for MSDFT calculations, these states are not further analyzed and shown. Although there has been controversy on the existence of a triplet excimer of naphthalene,67,69,111114 the work of Pabst et al. convincingly demonstrated that the triplet state excimer is as important and as stable as that for the singlet configurations.67 The reason is simple from the present diabatic perspective because the excimer complex in the triplet state 3La has an equivalent manifold of interactions as that of the excimer complex in the singlet state 1La, except that electronic couplings in the triplet states are somewhat weaker. Consequently, a binding energy of 0.85 eV is obtained for the excimer complex at a separation of 3.2 Å from MSDFT calculations. Thus, the triplet excimer also forms a strong complex. This may be compared with a value of 0.56 eV predicted by Pabst et al. using CC2/cc-pVTZ after the BSSE correction,67 and values of 0.79 and 0.60 eV with frozen and relaxed monomer structures by Kim.69 The interfragment distance for the triplet excimer was found to be 3.08 Å in the work of Pablst et al. Another key spectroscopic quantity of interest is the adiabatic phosphorescence energy, which is estimated to be 2.27 eV from MSDFT, in reasonable agreement with the result of Pabst et al. (2.54 eV).67 Phosphorescence in the present face-to-face configuration is symmetry forbidden, but deviations from the minimum structure could produce emissions and experimental observations of naphthalene triplet excimers show an energy of 2.3–2.4 eV in solution.113,114

Figure 4.

Figure 4.

Computed adiabatic triplet excited state (T1) along with BLE diabatic states, ER diabatic states, and CT resonance diabatic states. Energies are given relative to the separated monomers in the ground state. All computations have been performed using MSDFT with the PBE0 density functional (MSDFT@PBE0) and the cc-pVDZ basis set.

The present study also yield a singlet–triplet (S–T) energy gap, which is 1.0 eV between 3La and 1Lb for naphthalene, and 0.79 eV between 3La and 1La in the excimer complex. Thus, there is a small red-shift in the energy gap of 0.21 eV between the monomer and excimer complexes. S–T energy gaps of 1.16 and 1.38 were obtained by Kim for naphthalene in the ground monomer geometry and relaxed structure, respectively, which are correspondingly red-shifted by 0.2 in the monomer and 0.3 eV in the excimer complex.69

4.3. Excitonic Coupling between Singlet Excited States of Naphthalene Molecules.

EET plays an important role in many artificial and natural molecular devices such as organic light-emitting diodes, photovoltaics, and the photosynthetic system.115,116 In the weak coupling limit, the EET rate constant can be calculated using the Fermi’s golden rule4

kEET=2πħ|V|2δ(EiEf) (25)

where V is the electronic coupling between the initial state Ei and final state Ef. Therefore, the electronic coupling parameter is a key quantity to determine the EET rate constants.

The excitonic coupling between two locally excited states (covalent configuration) in the 1La state has been determined in the range of 2.5–20 Å between dimer separation (Figure 5), which may be considered to associate with singlet excitation energy transfer (SEET) from the initial (eq 13) to the final (eq 16) states.101 At distances greater than 6 Å, the excitonic coupling is reciprocally proportional to the cube of the separation distance, indicating that the energy resonance in the excited state is dominated by long-range electrostatic interactions, consistent with a Förster dipole–dipole interaction model.4 At close proximity, Dexter exchange coupling becomes important and deviation from the linearity of long-range behavior occurs. Figure 5 shows that the effect of the basis set size on the computed BLE coupling is small at all distances, suggesting that the present BLE approach is a robust procedure for modeling exciton coupling in the singlet state.

Figure 5.

Figure 5.

Computed transfer integral (V) for the excited energy transfer in the first singlet excited state between two naphthalene molecules as a function of the intermolecular separation. The PBE0 density functional is used in all calculations with different basis sets. Spin-flip using orbitals optimized in the triplet-state (T-SF) with the 6–31G(d) basis set was used; ionic configurations determined by perturbation theory (IC-PT); and IC by solving an eigenvalue equation (IC-EIG). ICs were calculated using the aug-cc-pVDZ basis set. The slopes were fitted using data in the range from 8 to 20 Å.

The SEET coupling parameters have also been computed on the basis of locally excited diabatic states constructed using BLOs optimized in the triplet state (Ms = 1) followed by an αβ spin flip (T-SF). In these calculations, the 6–31G(d) basis set was used. The computed electronic coupling values are not sensitive with respect to orbital optimization in the singlet or triplet states (Figure 5). We further examined the effect of ionic configurations (ICs) given in eqs 19 and 20, which could be important for excited energy transfer through a super-exchange mechanism.49 Two approaches were used to determine IC contributions. Using a perturbation theory (IC-PT) with the aug-cc-pVDZ basis set,117 we noticed significant deviations from results obtained using other models at short interchromophore distances. Not surprisingly, perturbation theory is expected to be valid when contributions from ICs are small at long distances. The IC contribution can be determined in the initial (A excitation) and final (B excitation) states using MSDFT (IC-EIG). Figure 5 shows that IC contributions cause a modest enhancement to the computed transfer integrals obtained directly using locally excited configurations at short distances, whereas at long ranges this effect diminishes beyond 5 Å.

4.4. Excitonic Coupling in the Triplet Excited States.

In contrast to the singlet state, excitonic coupling in the triplet excitation energy transfer (TEET) process is dominated by short-range coupling. Using MSDFT, both spin multiplets of the locally excited triplet state can be defined. In particular, the Ms = 0 state is defined by two BLE determinants through symmetric combination complementary of the singlet state in eqs 13 and 16.

ΩA*BT23S0=12{ΨA*B(S23¯S0)+ΨA*B(S2¯3S0)} (26)
ΩAB*S0T23=12{ΨAB*(S0S23¯)+ΨAB*(S0S2¯3)} (27)

Of course, the MS = 1 state can be treated by a single determinant. We note that the condition for energy degeneracy of these spin multiplets defines the spin-coupling correlation energy of the TDF along with the PBE0 correlation energy.43

The calculated TEET excitonic couplings are shown in Figure 6 for both the Ms = 0 and Ms = 1 triplet states. We first compare the “1 + 1”, “BLW”, and “relaxed” methods for Ms = 1 states and the BLE method for Ms = 0 states with the 6–31G(d) basis set. “1 + 1” means that the localized wave function is composed directly of the wave functions of two blocks. “Relaxed” means that the wave functions are relaxed without any restriction starting from BLW. The couplings of the “relaxed” method are the same starting either from BLW or “1 + 1”. The couplings of “1 + 1”, “BLW”, and BLE methods have small differences because the polarization interaction between two blocks is small for the eclipsed naphthalene molecules at certain distances. The “relaxed” method did not give a smooth curve because the obtained wave function may deviate from the optimized state when the distance is less than 4.5 Å. Therefore, the “relax” method may not be stable when the interaction between two blocks is strong.

Figure 6.

Figure 6.

Computed electronic coupling vs monomer separation distance between two stacked locally excited naphthalene molecules in the first triplet state, both in the MS = 0 and MS = 1 spin-projection states. In all calculations, the hybrid PBE0 functional is used with basis functions indicated in the figure or in the text. T-SF with the 6–31G(d) basis set was used; IC-PT; IC-EIG; 1 + 1: localized state directly composed of fragment orbitals for the monomers in the triplet state; and relax: variationally optimized locally excited states using the 6–31G(d) basis set. ICs were calculated using the aug-cc-pVDZ basis set. The exponential parameters in the figures are 2.60 Å obtained using IC-EIG at distances of 3.5–7 Å and 2.48 Å−1 without ICs at distances of 2.5–7 Å.

In contrast to SEET, where the computed transfer integrals do not show dependence on basis functions, Figure 6 reveals that inclusion of diffuse functions in computing the EET coupling is critical. Charge transfer contributions are also important, enhancing the coupling term by a factor of 2 over that without including ICs. Both the perturbation approach and eigenvalue solution yielded similar results except at distances shorter than 3 Å. The coupling values obtained with diffuse functions decay exponentially, VTEET eβr, with β values of 2.48 Å−1 without ICs, and 2.60 Å−1 with the inclusion of ICs. These observations are consistent with the results reported in refs 101, 117120.

The exponential decay of the TEET coupling is similar to ET coupling because the TEET can be viewed as two electron exchanges with different spins.4 We further calculated the couplings of ET and hole transfer (HT) at the aug-cc-pVDZ level from 3.5 to 7 Å using MSDDFT for two naphthalene molecules. The β values of ET and HT are 1.00 and 1.45 Å−1, respectively. Thus, the β value of TEET is almost the sum of the β values of ET and HT.

Equation 8 shows that the computed electronic coupling is proportional to the overlap of the two diabatic states when the overlap is small and orbital variations can be ignored.

VAB=MAB1MAB2(FABHAA/2HBB/2)MABCAB (28)

The approximation VAB = CMAB, where MAB is the overlap integral between determinants A and B, has been applied to ET.121 This trend is indeed reproduced using MSDFT along with BLE states in Figure 7, in which all methods exhibit nearly perfect linear regression (R2 > 0.99 from exponential root-mean-square logarithmic error analysis).121,122

Figure 7.

Figure 7.

Correlations between the coupling constants and the wavefunction overlaps for TEET. The labels for all the curves are the same as those in Figure 6.

5. CONCLUSIONS

In this work, we describe a BLE method on the basis of fragment-block localization and ΔSCF optimization such that a single-determinant representation for excited diabatic states can be approximated using density functional theory. We examined two optimization approaches to ensure the convergence of the block-localized molecular orbitals for a BLE, including a ΔSCF projection and an MOW technique. Within the framework of MSDFT, the use of the present BLE method in the construction of an effective, MAS is illustrated by the excimer complex formation of two naphthalene molecules both in the singlet and triplet states. Furthermore, excited energy transfer integrals in the singlet (SEET) and triplet (TEET) excited states of the naphthalene dimer are determined using block-localized excited diabatic states.

Numerical results show that the BLE method can provide a reasonable description of the transfer integrals associated with excited energy transfer in the singlet and triplet states. Using an MAS of eight spin-adapted configurations, the present MSDFT method correctly produced qualitatively and quantitatively the trends of an energy-level inversion involving the two lowest singlet excited states 1La and 1Lb, in good accordance with the DMRG-CASPT2 results. The latter approach used a large, complete active space of all 20 π electrons in 20 orbitals, consisting of more than 3 × 1010 configurations. Yet, without perturbation correction, it was still not possible to produce the correct order, with the 1Lb state below that of 1La by as much as 1.6 eV for the excimer complex. Dynamic correlation proved to be absolutely critical to bring 1La below 1Lb in DMRG-CASPT2. The good performance of MAS-MSDFT in comparison with DMRG-CASPT2 illustrates the effectiveness of a DTS ansatz to reduce the size of the active space in excited state calculations. It should be reminded that the choice of an MAS is not necessarily a black-box approach and an understanding of the nature of the relevant excited states is required. The benefits of this representation include physical insights gained from a diabatic perspective and its connection to classical theories such as the Marcus theory for ET. Although the calculations were illustrated on a model system of two identical monomers, the method can certainly be applied to heterodimers and complex molecular blocks in photoreceptor proteins.46,47

Supplementary Material

Supporting Information of BLE

ACKNOWLEDGMENTS

This work was partially supported by NSFC (grant nos 21673246 and 21933011) and the Beijing Municipal Science and Technology Commission (grant no. Z191100007219009) at the CAS, and Shenzhen Municipal Science and Technology Innovation Commission (KQTD2017-0330155106581) and the National Natural Science Foundation of China (grant number 21533003) at SZBL. The study was completed for the excimer complex at Minnesota, which was partially supported by the National Institutes of Health (grant number GM046736).

APPENDIX A

Calculation of the Block-Localized Ground and Excited State Energy Gradients

The operator form of the ground state energy gradient for RHF is given using44

EφA|=4(1ρ)f|φ˜A (A1)

where the reciprocal orbitals |φ˜ are

|φ˜=|φAφAφAφAφAφA1=|φAφA(φAφAφAφAφAφAφAφA)1 (A2)

here eq 〈φ|φ〉 is the overlap matrix of the molecular orbitals, not an integral value. By using the method of blockwise matrix inversion123

(ABCD)1=((ABD1C)1(ABD1C)1BD1D1C(ABD1C)1D1C(ABD1C)1BD1+D1) (A3)

we set

(φAφAφAφAφAφA1φAφA)1=φA|(1ρA)|φA1=α (A4)

After transformations, the following expressions are obtained

|φ˜A=|(1ρA)|φAα,(1(1ρA)|φAαφA|)φA|φA1 (A5)
ρ=|φ˜φAφA|=(1ρA)|φAαφA|(1ρA)+ρA (A6)
EφA|=4(1ρ)f|φ˜A=(1(1ρA)|φAαφA|)(1ρA)f(1ρA)|φAα (A7)

We can further set α as a unit matrix to make (1ρA)|ϕA orthonormalized

α=φA|(1ρA)(1ρA)|φA1=ΦAΦA1=I (A8)

and obtain the energy gradient for block A

EφA|=4(1(1ρA)|φAφA|)(1ρA)f(1ρA)|φA (A9)

At the point of the lowest energy, the energy gradient is zero. We obtain the SCF equation with

φiA|(1ρA)f(1ρA)|φiA=εiA (A10)

The SCF equation (in refs 30 and 34) can then be obtained by putting eq A11 into the above SCF equation. Because α is a unit matrix, eq A6 now becomes

ρ=(1ρA)ρA(1ρA)+ρA=ρAx+ρA (A11)

Equation S5 can then be obtained by inserting eqs A11 into S1.

We have the property

(1ρA)|φA=|φAρA|φA=0 (A12)

If we set |ΦiA=(1ρA)|φiA, then  φAΦiA=0. By multiplying φiA| on the left of eq 8, using the idempotency relation (1ρA)(1ρA)=(1ρA), and keeping ΦA orthogonal, we obtain

ΦiA|f|ΦiA=ΦiAΦiAεiA=εiA (A13)
f|ΦiA=εiA|ΦA (A14)

The above derivation from eqs A12A14 is a reversed process of the generalized Phillips–Kleiman pseudopotential derivation in ref 75. This means that the projected wavefuction of block A orthogonalized to all other blocks is the eigenfuction of the whole system.

To constrain the excited state particle orbitals |φA〉 orthogonal to the ground state occupied orbitals |φA0, we need φA(1ρA0)φA0=0. The following equation is solved with a Lagrange multiplier λ

LφA|=(1(1ρA)|φAφA|)(1ρA)f(1ρA)|φAλ(1ρA0)|φA0=0 (A15)

By multiplying φA0| on the left, the Lagrange multiplier λ can be obtained

λ=φA0|(1(1ρA)|φAφA|)(1ρA)f(1ρA)|φA (A16)

The equation then becomes

(1(1ρA0)|φA0φA0|)(1(1ρA)|φAφA|)(1ρA)f(1ρA)|φA=0 (A17)

We can further rewrite the above equation into a symmetric form

ρ(1ρA)f(1ρA)ρ|φiA=ρ(1ρA)|φiAεiA (A18)

where ρ=1(1ρA0)|φA0φA0|=(1ρA0)|φAv0φAv0| by using normalization property of complete nonorthogonal basis, |φAv0 is ground state unoccupied orbitals. To solve this eigenequation, we multiply φAv0|=(χTAv0) on the left and set |φA=χTA=χTAv0TA to project the equation to ground state unoccupied orbital space. The following equation is then obtained

(TAv0FATAv0)TA=(TAv0SATAv0)TAEA (A19)

After the diagonalization, the particle orbitals will be selected according to the order of ground state unoccupied orbitals.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.0c01015.

Total and relative energies for the ground and excited states and the Hamiltonian and overlap matrices used in MSDFT calculations for the naphthalene monomer as well as empirical dispersion energies (PDF)

Complete contact information is available at: https://pubs.acs.org/10.1021/acs.jctc.0c01015

The authors declare no competing financial interest.

Contributor Information

Peng Bao, Beijing National Laboratory for Molecular Sciences, State Key Laboratory for Structural Chemistry of Unstable and Stable Species, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China;.

Christian P. Hettich, Department of Chemistry and Supercomputing Institute, University of Minnesota, Minneapolis, Minnesota 55455, United States

Qiang Shi, Beijing National Laboratory for Molecular Sciences, State Key Laboratory for Structural Chemistry of Unstable and Stable Species, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China; University of the Chinese Academy of Sciences, Beijing 100049, China;.

Jiali Gao, Department of Chemistry and Supercomputing Institute, University of Minnesota, Minneapolis, Minnesota 55455, United States; Institute of Systems and Physical Biology, Shenzhen Bay Laboratory, Shenzhen 518055, China; Beijing University Shenzhen Graduate School, Shenzhen 518055, China;.

REFERENCES

  • (1).Baer M Beyond Born–Oppenheimer: Electronic Nonadiabatic Coupling Terms and Conical Intersections; Wiley: New York, 2006. [Google Scholar]
  • (2).May V; Kühn O Charge and Energy Transfer Dynamics in Molecular Systems, 3rd ed.; rev. and enl. ed.; Wiley-VCH: Weinheim, 2011; p xix, 562 p. [Google Scholar]
  • (3).Saigusa H; Lim EC Excimer formation in van der Waals dimers and clusters of aromatic molecules. Acc. Chem. Res 1996, 29, 171–178. [Google Scholar]
  • (4).Hsu C-P The Electronic Couplings in Electron Transfer and Excitation Energy Transfer. Acc. Chem. Res 2009, 42, 509–518. [DOI] [PubMed] [Google Scholar]
  • (5).Koppel H Diabatic representation: methods for the constrution of diabatic electronic states. In Conical Intersections: Electronic Structure, Dynamics and Spectroscopy; Domcke W, Yarkony DR, Koppel H, Eds.; World Scientific Publishing Co.: Hackensack, NJ, 2004; pp 175–204. [Google Scholar]
  • (6).Mead CA; Truhlar DG Conditions for the definition of a strictly diabatic electronic basis for molecular systems. J. Chem. Phys 1982, 77, 6090–6098. [Google Scholar]
  • (7).Subotnik JE; Yeganeh S; Cave RJ; Ratner MA Constructing diabatic states from adiabatic states: Extending generalized Mulliken-Hush to multiple charge centers with Boys localization. J. Chem. Phys 2008, 129, 244101. [DOI] [PubMed] [Google Scholar]
  • (8).Zhang A-J; Zhang P-Y; Chu T-S; Han K-L; He G-Z Quantum dynamical study of the electronic nonadiabaticity in the D +DBr Br(Br*)+D2 reaction on new diabatic potential energy surfaces. J. Chem. Phys 2012, 137, 194305. [DOI] [PubMed] [Google Scholar]
  • (9).Ou Q; Bellchambers GD; Furche F; Subotnik JE First-order derivative couplings between excited states from adiabatic TDDFT response theory. J. Chem. Phys 2015, 142, 064114. [DOI] [PubMed] [Google Scholar]
  • (10).Zhu X; Yarkony DR Constructing diabatic representations using adiabatic and approximate diabatic data - Coping with diabolical singularities. J. Chem. Phys 2016, 144, 044104. [DOI] [PubMed] [Google Scholar]
  • (11).Nakamura H; Truhlar DG Extension of the fourfold way for calculation of global diabatic potential energy surfaces of complex, multiarrangement, non-Born-Oppenheimer systems: Application to HNCO(S-0,S-1). J. Chem. Phys 2003, 118, 6816–6829. [Google Scholar]
  • (12).Hoyer CE; Parker K; Gagliardi L; Truhlar DG The DQ and DQ Phi electronic structure diabatization methods: Validation for general applications. J. Chem. Phys 2016, 144, 194101. [DOI] [PubMed] [Google Scholar]
  • (13).Domcke W; Yarkony DR Role of Conical Intersections in Molecular Spectroscopy and Photoinduced Chemical Dynamics. Annu. Rev. Phys. Chem 2012, 63, 325–352. [DOI] [PubMed] [Google Scholar]
  • (14).Li SL; Truhlar DG; Schmidt MW; Gordon MS Model space diabatization for quantum photochemistry. J. Chem. Phys 2015, 142, 064106. [DOI] [PubMed] [Google Scholar]
  • (15).Grofe A; Qu Z; Truhlar DG; Li H; Gao J Diabatic-At-Construction Method for Diabatic and Adiabatic Ground and Excited States Based on Multistate Density Functional Theory. J. Chem. Theory Comput 2017, 13, 1176–1187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Liu M; Chen X; Grofe A; Gao J Diabatic States at Construction (DAC) through Generalized Singular Value Decomposition. J. Phys. Chem. Lett 2018, 9, 6038–6046. [DOI] [PubMed] [Google Scholar]
  • (17).Gerratt J; Cooper DL; Karadakov PB; Raimondi M Modern valence bond theory. Chem. Soc. Rev 1997, 26, 87–100. [Google Scholar]
  • (18).Wu W; Su P; Shaik S; Hiberty PC Classical Valence Bond Approach by Modern Methods. Chem. Rev 2011, 111, 7557–7593. [DOI] [PubMed] [Google Scholar]
  • (19).Small DW; Lawler KV; Head-Gordon M Coupled Cluster Valence Bond Method: Efficient Computer Implementation and Application to Multiple Bond Dissociations and Strong Correlations in the Acenes. J. Chem. Theory Comput 2014, 10, 2027–2040. [DOI] [PubMed] [Google Scholar]
  • (20).Lee J; Small DW; Head-Gordon M Open-shell coupled-cluster valence-bond theory augmented with an independent amplitude approximation for three-pair correlations: Application to a model oxygen-evolving complex and single molecular magnet. J. Chem. Phys 2018, 149, 244121. [DOI] [PubMed] [Google Scholar]
  • (21).Mo Y; Peyerimhoff SD Theoretical analysis of electronic delocalization. J. Chem. Phys 1998, 109, 1687–1697. [Google Scholar]
  • (22).Mo Y; Gao J; Peyerimhoff SD Energy decomposition analysis of intermolecular interactions using a block-localized wave function approach. J. Chem. Phys 2000, 112, 5530–5538. [Google Scholar]
  • (23).Mo Y; Gao J An Ab Initio Molecular Orbital-Valence Bond (MOVB) Method for Simulating Chemical Reactions in Solution. J. Phys. Chem. A 2000, 104, 3012–3020. [Google Scholar]
  • (24).Song L; Gao J On the Construction of Diabatic and Adiabatic Potential Energy Surfaces Based on Ab Initio Valence Bond Theory. J. Phys. Chem. A 2008, 112, 12925–12935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Mo Y; Bao P; Gao J Energy decomposition analysis based on a block-localized wavefunction and multistate density functional theory. Phys. Chem. Chem. Phys 2011, 13, 6760–6775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (26).Mo Y; Song L; Lin Y Block-Localized Wavefunction (BLW) Method at the Density Functional Theory (DFT) Level. J. Phys. Chem. A 2007, 111, 8291–8301. [DOI] [PubMed] [Google Scholar]
  • (27).Cembran A; Song L; Mo Y; Gao J Block-localized density functional theory (BLDFT), diabatic coupling, and its use in valence bond theory for representing reactive potential energy surfaces. J. Chem. Theory Comput 2009, 5, 2702–2716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (28).Stoll H; Wagenblast G; Preuβ H On the use of local basis sets for localized molecular orbitals. Theor. Chim. Acta 1980, 57, 169–178. [Google Scholar]
  • (29).Gianinetti E; Raimondi M; Tornaghi E Modification of the Roothaan equations to exclude BSSE from molecular interactions calculations. Int. J. Quantum Chem 1996, 60, 157–166. [Google Scholar]
  • (30).Gianinetti E; Vandoni I; Famulari A; Raimondi M Extension of the SCF-MI method to the case of K fragments one of which is an open-shell system. Adv. Quantum Chem 1998, 31, 251–266. [Google Scholar]
  • (31).Raimondi M; Famulari A; Specchio R; Sironi M; Moroni F; Gianinetti E Ab initio non-orthogonal approaches to the computation of weak interactions and of localized molecular orbitals for QM/MM procedures. Theochem 2001, 573, 25–42. [Google Scholar]
  • (32).Nagata T; Takahashi O; Saito K; Iwata S Basis set superposition error free self-consistent field method for molecular interaction in multi-component systems: Projection operator formalism. J. Chem. Phys 2001, 115, 3553–3560. [Google Scholar]
  • (33).Ferenczy GG; Adams WH Optimization of selected molecular orbitals in group basis sets. J. Chem. Phys 2009, 130, 134108. [DOI] [PubMed] [Google Scholar]
  • (34).Khaliullin RZ; Head-Gordon M; Bell AT An efficient self-consistent field method for large systems of weakly interacting components. J. Chem. Phys 2006, 124, 204105. [DOI] [PubMed] [Google Scholar]
  • (35).Khaliullin RZ; Head-Gordon M; Bell AT Theoretical Study of Solvent Effects on the Thermodynamics of Iron(III) [Tetrakis(pentafluorophenyl)]porphyrin Chloride Dissociation. J. Phys. Chem. B 2007, 111, 10992–10998. [DOI] [PubMed] [Google Scholar]
  • (36).Smits GF; Altona C Calculation and Properties of Non-Orthogonal, Strictly Local Molecular-Orbitals. Theor. Chim. Acta 1985, 67, 461–475. [Google Scholar]
  • (37).Couty M; Bayse CA; Hall MB Extremely localized molecular orbitals (ELMO): a non-orthogonal Hartree-Fock method. Theor. Chem. Acc 1997, 97, 96–109. [Google Scholar]
  • (38).Sorakubo K; Yanai T; Nakayama K; Kamiya M; Nakano H; Hirao K A non-orthogonal Kohn-Sham method using partially fixed molecular orbitals. Theor. Chem. Acc 2003, 110, 328–337. [Google Scholar]
  • (39).Song L; Song J; Mo Y; Wu W An efficient algorithm for energy gradients and orbital optimization in valence bond theory. J. Comput. Chem 2009, 30, 399–406. [DOI] [PubMed] [Google Scholar]
  • (40).Sundstrom EJ; Head-Gordon M Non-orthogonal configuration interaction for the calculation of multielectron excited states. J. Chem. Phys 2014, 140, 114103. [DOI] [PubMed] [Google Scholar]
  • (41).Mo Y; Gao J Ab initio QM/MM simulations with a molecular orbital-valence bond (MOVB) method: application to an SN2 reaction in water. J. Comput. Chem 2000, 21, 1458–1469. [Google Scholar]
  • (42).Gao J; Grofe A; Ren H; Bao P Beyond Kohn–Sham Approximation: Hybrid Multistate Wave Function and Density Functional Theory. J. Phys. Chem. Lett 2016, 7, 5143–5149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (43).Grofe A; Chen X; Liu W; Gao J Spin-Multiplet Components and Energy Splittings by Multistate Density Functional Theory. J. Phys. Chem. Lett 2017, 8, 4838–4845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (44).Chen X; Gao J Fragment Exchange Potential for Realizing Pauli Deformation of Interfragment Interactions. J. Phys. Chem. Lett 2020, 11, 4008–4016. [DOI] [PubMed] [Google Scholar]
  • (45).Asgari P; Hua Y; Bokka A; Thiamsiri C; Prasitwatcharakorn W; Karedath A; Chen X; Sardar S; Yum K; Leem G; Pierce BS; Nam K; Gao J; Jeon J Catalytic hydrogen atom transfer from hydrosilanes to vinylarenes for hydrosilylation and polymerization. Nat. Catal 2019, 2, 164–173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (46).Li H; Wang Y; Ye M; Li S; Li D; Ren H; Wang M; Du L; Li H; Veglia G; Gao J; Weng Y Dynamical and allosteric regulation of photoprotection in light harvesting complex II. Sci. China: Chem 2020, 63, 1121–1133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).Li X; Ren H; Kundu M; Liu Z; Zhong FW; Wang L; Gao J; Zhong D A leap in quantum efficiency through light harvesting in photoreceptor UVR8. Nat. Commun 2020, 11, 4316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (48).Cembran A; Provorse MR; Wang C; Wu W; Gao J The Third Dimension of a More O’Ferrall-Jencks Diagram for Hydrogen Atom Transfer in the Isoelectronic Hydrogen Exchange Reactions of (PhX)(2)H-center dot with X = O, NH, and CH2. J. Chem. Theory Comput 2012, 8, 4347–4358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (49).Chan W-L; Berkelbach TC; Provorse MR; Monahan NR; Tritsch JR; Hybertsen MS; Reichman DR; Gao J; Zhu X-Y The quantum coherent mechanism for singlet fission: Experiment and theory. Acc. Chem. Res 2013, 46, 1321–1329. [DOI] [PubMed] [Google Scholar]
  • (50).Wu Q; Van Voorhis T Constrained Density Functional Theory and Its Application in Long-Range Electron Transfer. J. Chem. Theory Comput 2006, 2, 765–774. [DOI] [PubMed] [Google Scholar]
  • (51).Wu Q; Cheng C-L; Van Voorhis T Configuration interaction based on constrained density functional theory: A multireference method. J. Chem. Phys 2007, 127, 164119. [DOI] [PubMed] [Google Scholar]
  • (52).Van Voorhis T; Kowalczyk T; Kaduk B; Wang L-P; Cheng C-L; Wu Q The Diabatic Picture of Electron Transfer, Reaction Barriers, and Molecular Dynamics. Annu. Rev. Phys. Chem 2010, 61, 149–170. [DOI] [PubMed] [Google Scholar]
  • (53).Kaduk B; Kowalczyk T; Van Voorhis T Constrained Density Functional Theory. Chem. Rev 2012, 112, 321–370. [DOI] [PubMed] [Google Scholar]
  • (54).Mavros MG; Van Voorhis T Communication: CDFT-CI couplings can be unreliable when there is fractional charge transfer. J. Chem. Phys 2015, 143, 231102. [DOI] [PubMed] [Google Scholar]
  • (55).Yost SR; Lee J; Wilson MWB; Wu T; McMahon DP; Parkhurst RR; Thompson NJ; Congreve DN; Rao A; Johnson K; Sfeir MY; Bawendi MG; Swager TM; Friend RH; Baldo MA; Van Voorhis T A transferable model for singlet-fission kinetics. Nat. Chem 2014, 6, 492–497. [DOI] [PubMed] [Google Scholar]
  • (56).Ren H; Provorse MR; Bao P; Qu Z; Gao J Multistate Density Functional Theory for Effective Diabatic Electronic Coupling. J. Phys. Chem. Lett 2016, 7, 2286–2293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (57).Morrison AF; You Z-Q; Herbert JM Ab Initio Implementation of the Frenkel-Davydov Exciton Model: A Naturally Parallelizable Approach to Computing Collective Excitations in Crystals and Aggregates. J. Chem. Theory Comput 2014, 10, 5366–5376. [DOI] [PubMed] [Google Scholar]
  • (58).Morrison AF; Herbert JM Low-Scaling Quantum Chemistry Approach to Excited-State Properties via an ab Initio Exciton Model: Application to Excitation Energy Transfer in a Self-Assembled Nanotube. J. Phys. Chem. Lett 2015, 6, 4390–4396. [DOI] [PubMed] [Google Scholar]
  • (59).Herbert JM; Zhang X; Morrison AF; Liu J Beyond Time-Dependent Density Functional Theory Using Only Single Excitations: Methods for Computational Studies of Excited States in Complex Systems. Acc. Chem. Res 2016, 49, 931–941. [DOI] [PubMed] [Google Scholar]
  • (60).Runge E; Gross EKU Density-functional theory for time-dependent systems. Phys. Rev. Lett 1984, 52, 997–1000. [Google Scholar]
  • (61).Ziegler T; Rauk A; Baerends EJ Calculation of Multiplet Energies by Hartree-Fock-Slater Method. Theor. Chim. Acta 1977, 43, 261–271. [Google Scholar]
  • (62).Gilbert ATB; Besley NA; Gill PMW Self-Consistent Field Calculations of Excited States Using the Maximum Overlap Method (MOM). J. Phys. Chem. A 2008, 112, 13164–13171. [DOI] [PubMed] [Google Scholar]
  • (63).Barca GMJ; Gilbert ATB; Gill PMW Simple Models for Difficult Electronic Excitations. J. Chem. Theory Comput 2018, 14, 1501–1509. [DOI] [PubMed] [Google Scholar]
  • (64).Liu J; Zhang Y; Liu W Photoexcitation of Light-Harvesting C-P-C-60 Triads: A FLMO-TD-DFT Study. J. Chem. Theory Comput 2014, 10, 2436–2448. [DOI] [PubMed] [Google Scholar]
  • (65).Hait D; Head-Gordon M Excited State Orbital Optimization via Minimizing the Square of the Gradient: General Approach and Application to Singly and Doubly Excited States via Density Functional Theory. J. Chem. Theory Comput 2020, 16, 1699–1710. [DOI] [PubMed] [Google Scholar]
  • (66).East ALL; Lim EC Naphthalene dimer: Electronic states, excimers, and triplet decay. J. Chem. Phys 2000, 113, 8981–8994. [Google Scholar]
  • (67).Pabst M; Lunkenheimer B; Köhn A The Triplet Excimer of Naphthalene: A Model System for Triplet-Triplet Interactions and Its Spectral Properties. J. Phys. Chem. C 2011, 115, 8335–8344. [Google Scholar]
  • (68).Shirai S; Kurashige Y; Yanai T Computational Evidence of Inversion of L-1(a) and L-1(b)-Derived Excited States in Naphthalene Excimer Formation from ab Initio Multireference Theory with Large Active Space: DMRG-CASPT2 Study. J. Chem. Theory Comput 2016, 12, 2366–2372. [DOI] [PubMed] [Google Scholar]
  • (69).Kim D Effects of Intermolecular Interactions on the Singlet-Triplet Energy Difference: A Theoretical Study of the Formation of Excimers in Acene Molecules. J. Phys. Chem. C 2015, 119, 12690–12697. [Google Scholar]
  • (70).Liu DC; Nocedal J On the Limited Memory Bfgs Method for Large-Scale Optimization. Math Program 1989, 45, 503–528. [Google Scholar]
  • (71).Nagata T; Iwata S Perturbation expansion theory corrected from basis set superposition error. I. Locally projected excited orbitals and single excitations. J. Chem. Phys 2004, 120, 3555–3562. [DOI] [PubMed] [Google Scholar]
  • (72).Famulari A; Calderoni G; Moroni F; Raimondi M; Karadakov PB Application of the DIIS technique to improve the convergence properties of the SCF-MI algorithm. Theochem 2001, 549, 95–99. [Google Scholar]
  • (73).Ferré N; Assfeld X Application of the local self-consistent-field method to core-ionized and core-excited molecules, polymers, and proteins: True orthogonality between ground and excited states. J. Chem. Phys 2002, 117, 4119–4125. [Google Scholar]
  • (74).Gavnholt J; Olsen T; Engelund M; Schiotz J Delta self-consistent field method to obtain potential energy surfaces of excited molecules on surfaces. Phys. Rev. B: Condens. Matter Mater. Phys 2008, 78, 075441. [Google Scholar]
  • (75).Ziegler T; Seth M; Krykunov M; Autschbach J; Wang F On the relation between time-dependent and variational density functional theory approaches for the determination of excitation energies and transition moments. J. Chem. Phys 2009, 130, 154102. [DOI] [PubMed] [Google Scholar]
  • (76).De Mello P. C. a.; Hehenberger M; Zernert MC Converging SCF Calculations on Excited-States. Int. J. Quantum Chem 1982, 21, 251–258. [Google Scholar]
  • (77).Glushkov VN; Gidopoulos NI; Wilson S Alternative Technique for the Constrained Variational Problem Based on an Asymptotic Projection Method: I. Basics. Front. Quantum Syst. Chem. Phys 2008, 18, 429. [Google Scholar]
  • (78).Pen B; Van Kuiken BE; Ding FZ; Li XS A Guided Self-Consistent-Field Method for Excited-State Wave Function Optimization: Applications to Ligand-Field Transitions in Transition-Metal Complexes. J. Chem. Theory Comput 2013, 9, 3933–3938. [DOI] [PubMed] [Google Scholar]
  • (79).Surján PR Orthogonality constrained excited states. Chem. Phys. Lett 2000, 325, 120–126. [Google Scholar]
  • (80).Glushkov VN; Assfeld X Orthogonality-constrained Hartree-Fock and perturbation theory for high-spin open-shell excited states. Theor. Chem. Acc 2016, 135, 3. [Google Scholar]
  • (81).Evangelista FA; Shushkov P; Tully JC Orthogonality Constrained Density Functional Theory for Electronic Excited States. J. Phys. Chem. A 2013, 117, 7378–7392. [DOI] [PubMed] [Google Scholar]
  • (82).Derricotte WD; Evangelista FA Simulation of X-ray absorption spectra with orthogonality constrained density functional theory. Phys. Chem. Chem. Phys 2015, 17, 14360–14374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (83).Tassi M; Theophilou I; Thanos S Hartree-Fock calculation for excited states. Int. J. Quantum Chem 2013, 113, 690–693. [Google Scholar]
  • (84).Tassi M; Theophilou I; Thanos S Double excitations from modified Hartree Fock subsequent minimization scheme. J. Chem. Phys 2013, 138, 124107. [DOI] [PubMed] [Google Scholar]
  • (85).Theophilou I; Tassi M; Thanos S Charge transfer excitations from excited state Hartree-Fock subsequent minimization scheme. J. Chem. Phys 2014, 140, 164102. [DOI] [PubMed] [Google Scholar]
  • (86).Carter-Fenk K; Herbert JM State-Targeted Energy Projection: A Simple and Robust Approach to Orbital Relaxation of Non-Aufbau Self-Consistent Field Solutions. J. Chem. Theory Comput 2020, 16, 5067–5082. [DOI] [PubMed] [Google Scholar]
  • (87).Cembran A; Payaka A; Lin Y.-l.; Xie W; Mo Y; Song L; Gao J A Non-Orthogonal Block-Localized Effective Hamiltonian Approach for Chemical and Enzymatic Reactions. J. Chem. Theory Comput 2010, 6, 2242–2251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (88).Yang L; Grofe A; Reimers J; Gao JL Multistate density functional theory applied with 3 unpaired electrons in 3 orbitals: The singdoublet and tripdoublet states of the ethylene cation. Chem. Phys. Lett 2019, 736, 136803. [Google Scholar]
  • (89).Filatov M; Shaik S Spin-restricted density functional approach to the open-shell problem. Chem. Phys. Lett 1998, 288, 689–697. [Google Scholar]
  • (90).Filatov M; Shaik S A spin-restricted ensemble-referenced Kohn-Sham method and its application to diradicaloid situations. Chem. Phys. Lett 1999, 304, 429–437. [Google Scholar]
  • (91).Filatov M; Shaik S Diradicaloids: Description by the spin-restricted, ensemble-referenced Kohn-Sham density functional method. J. Phys. Chem. A 2000, 104, 6628–6636. [Google Scholar]
  • (92).Dubinets NO; Safonov AA; Bagaturyants AA Structures and Binding Energies of the Naphthalene Dimer in Its Ground and Excited States. J. Phys. Chem. A 2016, 120, 2779–2782. [DOI] [PubMed] [Google Scholar]
  • (93).Climent C; Barbatti M; Wolf MO; Bardeen CJ; Casanova D The photophysics of naphthalene dimers controlled by sulfur bridge oxidation. Chem. Sci 2017, 8, 4941–4950. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (94).Tretiak S; Zhang WM; Chernyak V; Mukamel S Excitonic couplings and electronic coherence in bridged naphthalene dimers. Proc. Natl. Acad. Sci. U.S.A 1999, 96, 13003–13008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (95).Perdew JP; Burke K; Ernzerhof M Generalized gradient approximation made simple. Phys. Rev. Lett 1996, 77, 3865–3868. [DOI] [PubMed] [Google Scholar]
  • (96).Pokhilko P; Izmodenov D; Krylov AI Extension of frozen natural orbital approximation to open-shell references: Theory, implementation, and application to single-molecule magnets. J. Chem. Phys 2020, 152, 034105. [DOI] [PubMed] [Google Scholar]
  • (97).Pople JA; Gill PMW; Handy NC Spin-Unrestricted Character of Kohn-Sham Orbitals for Open-Shell Systems. Int. J. Quantum Chem 1995, 56, 303–305. [Google Scholar]
  • (98).Ess DH; Cook TC Unrestricted Prescriptions for Open-Shell Singlet Diradicals: Using Economical Ab Initio and Density Functional Theory to Calculate Singlet-Triplet Gaps and Bond Dissociation Curves. J. Phys. Chem. A 2012, 116, 4922–4929. [DOI] [PubMed] [Google Scholar]
  • (99).Schmidt MW; Baldridge KK; Boatz JA; Elbert ST; Gordon MS; Jensen JH; Koseki S; Matsunaga N; Nguyen KA; Su S; Windus TL; Dupuis M; Montgomery JA General atomic and molecular electronic structure system. J. Comput. Chem 1993, 14, 1347–1363. [Google Scholar]
  • (100).Speiser S Photophysics and mechanisms of intramolecular electronic energy transfer in bichromophoric molecular systems: Solution and supersonic jet studies. Chem. Rev 1996, 96, 1953–1976. [DOI] [PubMed] [Google Scholar]
  • (101).Hsu C-P; You Z-Q; Chen H-C Characterization of the short-range couplings in excitation energy transfer. J. Phys. Chem. C 2008, 112, 1204–1212. [Google Scholar]
  • (102).Chandra AK; Lim EC Semiempirical Theory of Excimer Luminescence. J. Chem. Phys 1968, 48, 2589. [Google Scholar]
  • (103).Kołaski M; Arunkumar CR; Kim KS Aromatic Excimers: Ab Initio and TD-DFT Study. J. Chem. Theory Comput 2013, 9, 847–856. [DOI] [PubMed] [Google Scholar]
  • (104).Shirai S; Iwata S; Tani T; Inagaki S Ab Initio Studies of Aromatic Excimers Using Multiconfiguration Quasi-Degenerate Perturbation Theory. J. Phys. Chem. A 2011, 115, 7687–7699. [DOI] [PubMed] [Google Scholar]
  • (105).Schreiber M; Silva MR; Sauer SPA; Thiel W Benchmarks for electronically excited states: CASPT2, CC2, CCSD, and CC3. J. Chem. Phys 2008, 128, 134110. [DOI] [PubMed] [Google Scholar]
  • (106).George GA; Morris GC Intensity of Absorption of Naphthalene from 30 000 Cm-1 to 53 000 Cm-1. J. Mol. Spectrosc 1968, 26, 67–71. [Google Scholar]
  • (107).Huebner RH; Meilczarek SR; Kuyatt CE Electron Energy-Loss Spectroscopy of Naphthalene Vapor. Chem. Phys. Lett 1972, 16, 464–469. [Google Scholar]
  • (108).Mcconkey JW; Trajmar S; Man KF; Ratliff JM Excitation of Naphthalene by Electron-Impact. J. Phys. B: At., Mol. Opt. Phys 1992, 25, 2197–2204. [Google Scholar]
  • (109).Matos J; Matos BO; Roos BO A CASSCF Study of the Singlet Singlet and Triplet Triplet Spectroscopy of Naphthalene. Theor. Chim. Acta 1988, 74, 363–379. [Google Scholar]
  • (110).Azumi T; Mcglynn SP Energy of Excimer Luminescence .I. Reconsideration of Excimer Processes. J. Chem. Phys 1964, 41, 3131–3138. [Google Scholar]
  • (111).Lim EC Molecular Triplet Excimers. Acc. Chem. Res 1987, 20, 8–17. [Google Scholar]
  • (112).Chandra AK; Lim EC Semiempirical Theory of Excimer Luminescence .2. Comparison with Previous Theories and Consideration of Transition Probability and Stability of Excimer Triplet State. J. Chem. Phys 1968, 49, 5066–5072. [Google Scholar]
  • (113).Langelaar J; Rettschnick RPH; Lambooy AMF; Hoytink GJ Monomer and excimer phosphorescence of phenanthrene and naphthalene in solution. Chem. Phys. Lett 1968, 1, 609–612. [Google Scholar]
  • (114).Takemura T; Aikawa M; Baba H; Shindo Y Kinetic Study of Triplet Excimer Formation in Fluid Solution by Means of Phosphorimetry. J. Am. Chem. Soc 1976, 98, 2205–2210. [Google Scholar]
  • (115).Holten D; Bocian DF; Lindsey JS Probing electronic communication in covalently linked multiporphyrin arrays. A guide to the rational design of molecular photonic devices. Acc. Chem. Res 2002, 35, 57–69. [DOI] [PubMed] [Google Scholar]
  • (116).Fleming GR; van Grondelle R Femtosecond spectroscopy of photosynthetic light-harvesting systems. Curr. Opin. Struct. Biol 1997, 7, 738–748. [DOI] [PubMed] [Google Scholar]
  • (117).Shi B; Gao F; Liang W A simplified approach for the coupling of excitation energy transfer. Chem. Phys 2012, 394, 56–63. [Google Scholar]
  • (118).Harcourt RD; Scholes GD; Ghiggino KP Rate Expressions for Excitation Transfer .2. Electronic Considerations of Direct and through-Configuration Exciton Resonance Interactions. J. Chem. Phys 1994, 101, 10521–10525. [Google Scholar]
  • (119).Scholes GD; Harcourt RD; Ghiggino KP Rate Expressions for Excitation Transfer .3. An Ab-Initio Study of Electronic Factors in Excitation Transfer and Exciton Resonance Interactions. J. Chem. Phys 1995, 102, 9574–9581. [Google Scholar]
  • (120).You Z-Q; Hsu CP The fragment spin difference scheme for triplet-triplet energy transfer coupling. J. Chem. Phys 2010, 133, 074105. [DOI] [PubMed] [Google Scholar]
  • (121).Gajdos F; Valner S; Hoffmann F; Spencer J; Breuer M; Kubas A; Dupuis M; Blumberger J Ultrafast Estimation of Electronic Couplings for Electron Transfer between pi-Conjugated Organic Molecules. J. Chem. Theory Comput 2014, 10, 4653–4660. [DOI] [PubMed] [Google Scholar]
  • (122).Kubas A; Gajdos F; Heck A; Oberhofer H; Elstner M; Blumberger J Electronic couplings for molecular charge transfer: benchmarking CDFT, FODFT and FODFTB against high-level ab initio calculations. II. Phys. Chem. Chem. Phys 2015, 17, 14342–14354. [DOI] [PubMed] [Google Scholar]
  • (123).Bernstein DS Matrix Mathematics: Theory, Facts, and Formulas, 2nd ed.; Princeton University Press, July 26, 2009; p 184, ISBN-10: 0691140391, ISBN-13: 978-0691140391. [Google Scholar]

Associated Data

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

Supplementary Materials

Supporting Information of BLE

RESOURCES