Skip to main content
The Journal of Chemical Physics logoLink to The Journal of Chemical Physics
. 2022 Jun 1;156(21):210901. doi: 10.1063/5.0088271

Evaluation of molecular photophysical and photochemical properties using linear response time-dependent density functional theory with classical embedding: Successes and challenges

WanZhen Liang 1,a), Zheng Pei 1, Yuezhi Mao 2,b), Yihan Shao 3,c)
PMCID: PMC9162785  PMID: 35676148

Abstract

Time-dependent density functional theory (TDDFT) based approaches have been developed in recent years to model the excited-state properties and transition processes of the molecules in the gas-phase and in a condensed medium, such as in a solution and protein microenvironment or near semiconductor and metal surfaces. In the latter case, usually, classical embedding models have been adopted to account for the molecular environmental effects, leading to the multi-scale approaches of TDDFT/polarizable continuum model (PCM) and TDDFT/molecular mechanics (MM), where a molecular system of interest is designated as the quantum mechanical region and treated with TDDFT, while the environment is usually described using either a PCM or (non-polarizable or polarizable) MM force fields. In this Perspective, we briefly review these TDDFT-related multi-scale models with a specific emphasis on the implementation of analytical energy derivatives, such as the energy gradient and Hessian, the nonadiabatic coupling, the spin–orbit coupling, and the transition dipole moment as well as their nuclear derivatives for various radiative and radiativeless transition processes among electronic states. Three variations of the TDDFT method, the Tamm–Dancoff approximation to TDDFT, spin–flip DFT, and spin-adiabatic TDDFT, are discussed. Moreover, using a model system (pyridine–Ag20 complex), we emphasize that caution is needed to properly account for system–environment interactions within the TDDFT/MM models. Specifically, one should appropriately damp the electrostatic embedding potential from MM atoms and carefully tune the van der Waals interaction potential between the system and the environment. We also highlight the lack of proper treatment of charge transfer between the quantum mechanics and MM regions as well as the need for accelerated TDDFT modelings and interpretability, which calls for new method developments.

I. INTRODUCTION

To study the systems at the atomic and molecular levels, one has to follow the principles of quantum mechanics (QM). Even though the first-principles electronic structure theory and the quantum dynamics methods have achieved great success in describing the ground state of molecular systems, computing excited states still faces many challenges because it implies not only to find the higher-energy solution of the electronic Schrödinger equation but also to solve many challenging conceptual and technical problems, such as the electron–nuclear coupling, the state mixing of the same spin and different spin multiplicities, or the entangled optical and dark processes, which usually require one to abandon the approximations such as the adiabatic Born–Oppenheimer (BO) and to calculate the accurate structure properties, such as potential energy surfaces, nuclear forces, harmonic and non-harmonic vibrational frequencies, nonadiabatic couplings (NAC), and spin–orbit couplings (SOC).1 Furthermore, most of the photophysical and photochemical processes of technological and biological interest take place in the condensed phase or in the presence of an environment surrounding the molecular system such as solvents, protein scaffolds, and near semiconductor or metal surfaces. In this case, one has to account for the system–environment interaction for properly describing these processes.2

For composite systems, such as solute–solvent systems, fluorescent proteins, and even nanostructures that integrate molecules or semiconductors with metal nanoparticles (MNPs), a small region of the system plays a predominant role in the excited-state properties and processes of the whole system. Thus, the total system can usually be partitioned into an active region ranging from a single molecule to a few molecules and an embedding environment. The former is mainly responsible for the observed process or property, and the latter does not take a direct part in the process but acts as a perturbation affecting the electronic structure and the dynamics of the core active system. The active region is then treated by a high-level QM method, whereas the rest is described by a lower-level approach, such as molecular mechanical (MM) force fields and polarizable continnum models (PCMs). The integration of QM methods with classical descriptions within multiscale models provides a natural way to focus the computation on a small area in a larger environment without significantly increasing the computational cost.3 Currently, scientists have developed a family of multiscale theoretical models, such as QM/(MM),4,5 QM/PCM,6,7 QM/MM/PCM,8 and QM/electromagnetism (EM) (the hybrid QM and electromagnetic mechanics or electrodynamics method).9–12

The QM methods for describing the molecular electronic systems can be divided into three categories: semi-empirical quantum chemistry methods, ab initio wavefunction-based methods, and density functional theory (DFT) methods.13 The post-Hartree–Fock (HF) methods have been very successful in describing the ground and excited-state electronic states of small molecules, especially when the molecules are near the equilibrium configuration. The semi-empirical methods are much faster than the ab initio ones. However, they require a couple of atom-specified parameters and are thus only suitable for parameterized systems. For a medium and even large sized molecular system, the most widely used method is DFT. The last two decades have witnessed an explosion of DFT-based calculations in the field of computational chemistry.14–16 The modeling of ground state processes, such as chemical,17 surface,18 and enzyme reactions,19 is mostly carried out using Kohn–Sham DFT (KS-DFT)20,21 calculations. Meanwhile, excited-state processes (such as UV-vis absorption,22–25 fluorescence,26,27 phosphorescence,28–30 chemiluminescence,31 bioluminescence,32–34 and energy transfer processes in natural and artificial photosynthetic systems and organic and dye-sensitized solar cells35–40) have been modeled extensively using time-dependent DFT (TDDFT),41,42 which is the excited-state counterpart of the KS-DFT.

The vast popularity of KS-DFT and TDDFT calculations arises from three factors. The first is their cost-effectiveness. They are simply orders of magnitudes cheaper than highly reliable quantum chemistry methods, such as coupled-cluster methods,43,44 equation-of-motion coupled-cluster methods,45–49 and multi-configurational and multi-reference wavefunction methods.50–52 Moreover, the high-order energy derivatives, transition couplings, or molecular properties from DFT and TDDFT are relatively simple to be implemented and cost-efficient, which are essential for studying the rich chemical and physical events we are interested in, e.g., the evaluation of spectroscopic intensities and rates and quantum yields of excited-state processes. The second is their reasonable accuracy if a suitable exchange–correlation (XC) functional is employed for the particular problem at hand. For example, a valence electronic excitation might be best described with a conventional hybrid functional (such as B3LYP and PBE0), whereas a charge-transfer or Rydberg excitation necessitates the use of range-separated hybrid functionals.53,54 Besides the above-mentioned two factors, the DFT/TDDFT methods have also been implemented in many efficient and user-friendly commercial or noncommercial software packages that were developed and made accessible to a wide community of researchers. This wide availability of DFT/TDDFT methods in modern computational chemistry programs combined with versatile and user-friendly visualization tools makes them accessible to researchers with no sophisticated background in theoretical chemistry.

Two versions of the TDDFT implementation have been familiar to the quantum chemistry community. One is to numerically solve the time-dependent Kohn–Sham equations directly in the real-time domain (RT-TDDFT).55 The RT-TDDFT approach has been implemented in the real space, in plane wave bases, and in atomic orbital basis sets and has been used to calculate the linear and nonlinear spectra and describe the photochemistry by combining the Ehrenfest or surface hopping nuclear dynamics. Many review articles have summarized the recent progress of RT-TDDFT approaches (e.g., Refs. 56–60 and references therein). The other is the implementation of TDDFT at the linear response level (LR-TDDFT),61–66 which looks like the standard implementation of LR-time-dependent Hartree–Fock (TDHF) already available in most quantum chemistry codes. A set of static Casida’s equations was derived and solved to obtain the molecular excitation energies and transition vectors as well as the oscillator strengths. Within the framework of LR-TDDFT, the nuclear gradient67–69 and Hessian70,71 of excited-state energies and the nonadiabatic coupling between electronic states72–74 have all been implemented. Consequently, this LR-TDDFT scheme has become the most widely used one to extract the energies of low-lying excited states, optimize excited-state geometries, calculate Stokes shifts, and explore the excited-state potential energy surfaces. Here, we only focus on reviewing approaches based on LR-TDDFT with a specific emphasis on their applications to the calculation of electronic structure quantities required by describing molecular excited-state properties and processes.

Most earlier TDDFT calculations were focused on gas-phase excited states. However, as mentioned earlier, there has been an ever-growing interest to extend these calculations to composite systems for the modeling of excited-state processes of a single or a few molecules embedded in a condensed-phase medium, such as solvents,75–77 macromolecules,78–80 and metal nanoparticles.10,81–84 For these system–environment complexes, multiscale modeling approaches have been proposed with different levels of theories for the different regions. There are generally two categories of embedding methods for TDDFT calculations within the environment: density embedding and classical embedding. Within the category of density-embedding schemes, recent efforts include the projection-based embedding theory,85,86 frozen-density embedding,87 and polarized many-body expansion scheme88. When the excited states of molecular aggregates are considered, the subsystem TDDFT methods have shown accuracy and efficiency for large systems89 and inter-region charge-transfer excitations.90,91 On the other hand, the hybrid QM/MM method has been more commonly used in TDDFT calculations with the aforementioned classical embedding, such as QM/MM, QM/PCM, QM/MM/PCM, and QM/EM. In the following, we mainly focus on TDDFT calculations with a classical embedding description of the environment using either PCM6,7 or polarizable MM force fields (MMpol).92

While tremendous progress has been made in TDDFT/PCM and TDDFT/MM methodologies, the application of these methods to model excited-state properties and processes in a condensed-phase medium still faces three great challenges (in addition to the choice of suitable DFT XC functionals):

  • System–environment interactions. For each system–environment complex, there lacks a standard protocol for tuning the QM/MM electrostatic and van der Waals interactions between the system and its environment, which is critical to avoid over-polarization and geometry distortions. In addition, missing are approaches for describing full charge transfer excitations that move system electrons into the environment or partial charge transfer excitations where the participating system orbitals extend into the environment.

  • Computational cost. Standard single-point TDDFT excitation energy calculations and geometry optimizations are currently limited to systems with up to a couple of 100 atoms. To theoretically study a variety of molecular excited-state properties and processes, one has to calculate the derivatives of excitation energies with respect to perturbation parameters, such as the nuclear coordinates and external fields, the NACs, and the nuclear derivatives of the transition dipole or SOC matrix element. These quantities require even much higher computational cost than the excitation energy calculation. Another bottleneck in TDDFT/MMpol calculations arises from the need to solve for polarizable charges or dipoles, which can become rather expensive with a large polarizable MM region.93,94

  • Interpretability. While numerous TDDFT calculations are routinely performed on new dye molecules and bioluminescence probes, few analysis tools beyond electron (difference and transition) densities, atomic partial charges, and excited-state and transition dipole moments are available for helping to directly interpret the observed photochemical properties.95 New analysis procedures and design principles only started to emerge recently for predicting the substituent, solvent, and macromolecular effects on the measured spectra and quantum yield.96–100

This Perspective is organized as follows. In Sec. II, we briefly introduce a variety of electronic structure quantities required by the evaluation of molecular properties, which are closely related to the calculations of the derivatives of the energy or wavefunction with respect to the nuclear coordinates or other perturbation parameters. The current development of hybrid TDDFT with MM and PCM is summarized in Sec. III, which is followed by a discussion on the choice of TDDFT methods in Sec. IV. We then elaborate in Sec. V on the three challenges listed above for the TDDFT modeling of excited-state processes in complex environments. A summary statement will be given in Sec. VI.

II. A VARIETY OF ELECTRONIC STRUCTURE QUANTITIES FOR EVALUATION OF MOLECULAR PHOTOPHYSICAL AND PHOTOCHEMICAL PROPERTIES

A. Quantities for describing properties of a specific electronic excited state

Molecular properties characterize molecules and their behaviors, and they represent the link between experimental observable quantities and theoretical calculations. There are a wide variety of molecular properties, which are related to numerous physical phenomena. At first, we address a variety of properties of molecules at a given electronic state, which are related with the calculations of energy derivatives with respect to the perturbation parameters.101,102 For example, the forces and the force constants are given as the first and second derivatives of the molecular energy with respect to the nuclear coordinates, respectively,67,103 and the infrared intensities can be computed as the cross second derivatives of the molecular energy with respect to the components of a static electric field and nuclear coordinates.104–106 As such, formulating energy derivatives with respect to the external perturbation variables yields a wealth of time-independent molecular properties.

1. The excited-state energy gradient and Hessian within TDDFT

The energy gradient and hessian with respect to nuclear coordinates have been widely used to explore molecular potential energy surfaces and to calculate vibrational spectra and thermochemical parameters. For a given molecular geometry, the energy of the Ith excited state can be expressed as the sum of the ground-state energy and the excitation energy,

EI=Eg+ωI, (1)

where Eg is the ground-state energy and ωI is the TDDFT excitation energy of the Ith excited state. Then, the energy derivatives of a given excited state I can be evaluated as the sum of the ground-state energy derivatives and the derivatives of the corresponding TDDFT excitation energy.

In the framework of LR-TDDFT, the excitation energy is defined as the pole of the LR function to an external perturbation. The excitation energies and the corresponding transition vectors can be obtained from an usual non-Hermitian eigenvalue equation,42

ABBAXY=ω1001XY. (2)

The matrix representations of A and B in the molecular orbital (MO) basis are written as

Aiaσ,jbσ=δabδijδσσ(εaσεiσ)+FiaσPjbσ, (3)
Baiσ,bjσ=FaiσPjbσ. (4)

Here, ɛ is the pth σ (σ = α or β) spin Kohn–Sham MO energy. The indices i, j, k, …; a, b, c, …; and p, q, r, … denote occupied, virtual, and generic MOs, respectively. The transition vectors |X⟩ and |Y⟩ are defined on the same Hilbert spaces of occ ⨂ vir, which satisfy the following biorthonormal condition:

XI,YI|XJ,YJ=δIJ. (5)

After these transition vectors are solved, the excitation energy of the Ith TDDFT excited state can be expressed as

ωI=12(XI+YI)(A+B)(XI+YI)+12(XIYI)(AB)(XIYI). (6)

Direct differentiation of ωI in Eq. (6) with respect to the nuclear coordinates (R) requires one to calculate first-order nuclear derivatives of molecular orbital coefficients (C) and transition vectors (X and Y). In order to avoid these calculations, one would define an auxiliary energy functional as68

Le[X,ωI,C,Z,W]=Eg+ωIλI(Tr(XI2YI2)1)+aiσZaiσFaiσ+pqσWpqσ(Sqpσδpq). (7)

With the construction of the Lagrangian Le, one has

dEIdRi=LeRi (8)

since the Le is stationary with respect to all the variational parameters, including C, X, and Y. The Lagrange multipliers Z and W are determined by the condition

LeCμpσ=0, (9)

and the equation derived thereof is called the z-vector equation.107 Once the energy gradient of an excited state is formulated, one can straightforwardly differentiate the energy gradient to derive the excited-state Hessian. The calculation of the excited-state Hessian requires one to calculate the first-order nuclear derivatives of molecular orbital coefficients (CR) and transition vectors (XR and YR) explicitly, which can be obtained by solving the coupled perturbed self-consistent field (CPSCF) equations103 and CP-TDDFT equations,71 respectively. A detailed formulation of the energy Hessian of TDDFT excited states can be found in Ref. 70 [with the Tamm–Dancoff approximation (TDA)] and Ref. 71 (full TDDFT). The reader is referred to Refs. 69–71 and 108–116 for additional details.

2. Electric dipole moments, polarizabilities, and infrared/Raman intensities

The molecular dipole moment and polarizability of a given electronic state are defined as

μx=EFx, (10)
αxy=μxFy=2EFxFy, (11)

where Fx/Fy are the x/y components of the perturbing electric field. The expectation value of the electric dipole moment of a specific electronic state I is usually calculated as

μ¯xI=ΨI|μ^x|ΨI=Tr(PIμx). (12)

For the Ith excited state, its one-electron reduced density matrix, PI, is composed of70,71,116,117

PI=P0+ΔPI+ZI. (13)

Here, P0 is the reduced one-electron density matrix of the ground state; ΔPI is the unrelaxed difference density matrix, ΔPI=12Cv(XI+YI)(XI+YI)Cv+12Cv(XIYI)(XIYI)Cv12Co(XI+YI)(XI+YI)Co12Co(XIYI)(XIYI)Co, where Co and Cv are the occupied and virtual MO coefficient matrices, respectively; and ZI corresponds to the z-vector, denoting the relaxed part of the excited-state density matrix.

Moreover, the Ith electronic state’s polarizability tensor can be calculated as

α¯xyI=Tr(PIFyμx), (14)

where PIFy denotes the derivative of the one-electron density matrix of a given electronic excited state with respect to the y component of an external field. This expression indicates that it is much easier and faster to calculate the polarizability than the nuclear Hessian because the former is merely related to the energy derivatives with respect to external electric fields so that the explicit derivatives of atomic orbitals in one- and two-electron integrals with respect to nuclear coordinates are not required.

One may be further interested in the nuclear derivatives of the above two quantities (μ and α¯) since they are closely related to the calculations of infrared (IR) and Raman intensities, respectively.118–120 In the harmonic approximation, the transition electric dipole moment, μfiI, and the polarizability, αfiI, between the initial and final vibronic states, which enter the expressions of IR absorption and Raman scattering intensities, can be expanded in a power series of the molecular normal coordinates Q as

μfiI=μ¯I(0)ΛIf|ΛIi+kμ¯IQk0ΛIf|Qk|ΛIi+, (15)
αfiI=α¯I(0)ΛIf|ΛIi+kα¯IQk0ΛIf|Qk|ΛIi+. (16)

Under the dipole approximation, the IR intensity corresponding to the transition from the ith vibrational state associated with the electronic state I to the fth vibrational state within the same electronic state is proportional to μ¯IQk02, and the Raman intensity is closely related to the calculation of nuclear derivatives of molecular polarizability.118,121 Here, we assume that the pure-spin BO (psBO) wavefunctions of initial and final vibronic states can be written as products of an electronic wavefunction Ψ and a vibrational wavefunction Λ, such as Φnσ(q,Q)=ΨNσ(q,Q)ΛNn(Q), with normal coordinates Qk=i=13Nlikqi. The qi’s denote the mass-weighted Cartesian coordinates, and the transformation coefficients lik are chosen such that the potential energy does not depend on cross products.

B. Quantities for describing transitions between electronic states

Transition couplings between different electronic states play central roles in many physical and chemical processes. To characterize a photophysical process, such as optical absorption and radiative emission (fluorescence and phosphorescence), or radiativeless transition, such as internal conversion (IC), intersystem crossing (ISC), and decay through a conical intersection (CI), one usually has to account for simultaneous changes in the vibrational and electronic states. Within the perturbation theory framework, one will need to combine the electronic structure theories and the quantum dynamics methods. The former is adopted to obtain the electronic structure parameters, and the latter is used to describe the nuclear dynamics.

According to the Fermi’s golden rule,122 the transition rate (k) from the initial state i to a dense manifold of final states f can be defined as

ki=2πfΦf|H^|Φi2δω(EfEi), (17)

where Φi and Φf are the wavefunctions of the initial and final vibronic states, respectively, and ω is the frequency of the external radiation. H^=H^H^el (the electronic Hamiltonian) denotes the perturbation Hamiltonian and can be written as

H^=μf+T^N+H^SO, (18)

where the first, second, and third operators correspond to the matter–field interaction, the nuclear kinetic energy, and spin–orbit interaction, respectively. The radiative transition, such as electronic absorption or emission, is usually governed by the matrix element of Φf|μ^|Φi while the nonradiative processes, such as IC and ISC or reversed ISC (RISC), are determined by the matrix elements of the latter two operators, respectively. For example, if we restrict ourselves up to second-order terms, the coupling matrix element of an ISC process can be expressed as

Hfi=Φf3H^SOΦi1+mi[Φf3H^SOΦm1Φm1T^NΦi1/(EiEm)]+nf[Φf3T^NΦn3Φn3H^SOΦi1/(EfEn)],=HSOFIQ=0ΛFf(Q)|ΛIi(Q)+k=13N6HSOFIQkQ=0ΛFf(Q)|Qk|ΛIi(Q)k=13N6MIHSOFMdMI,1k(EIEM)+NFdFN,3kHSONI(EFEN)Q=0×ΛFf(Q)|Qk|ΛIi(Q). (19)

Here, the summations extend over the complete sets of psBO vibronic states of the given multiplicity, and the psBO wavefunctions of the initial and final vibronic states have been written as products of an electronic wavefunction Ψ and a vibrational wavefunction Λ. The relation

ΦeσT^NΦrσ2k=13N6ΨEσkΨRσ×ΛEe(Q)|kΛRr(Q) (20)

has been applied, and

dER,σk=ΨEσ/QkΨRσ (21)

is defined as the NAC vector between two excited states of the given spin multiplicity σ. The spin–orbit coupling matrix element (SOCME) has been assumed to be normal mode-dependent, and it reads

HSOFI=ΨF|H^SO|ΨI=HSOFI(Q=0)+HSOFIQ|Q=0Q+, (22)

which thus includes the Franck–Condon (FC) and Herzberg–Teller (HT) contributions. Although Eq. (19) specifically refers to a S → T crossing, the corresponding expression for the RISC is readily written down.

The transition dipole moment between pure S0 and T1 states is zero. In the case of phosphorescence, one has to resort to perturbation theory by considering the SOC perturbation to give rise to admixtures of states with opposite spin.

It is clear that to quantitatively describe the radiative and radiationless processes, such as the molecular vibronic spectra123 and the transition rates, within the perturbation theory framework, one has to know the geometric and electronic structure parameters at first. More specifically, the energy gradients of ground and excited states are required to reach the minimum on each potential energy surface. Meanwhile, Hessians of the ground- and excited-state energies are indispensable for acquiring the normal modes for each electronic state, establishing the Duschinsky relation,124 as well as incorporating the FC/HT and HT contributions.

1. Non-adiabatic coupling

The nonadiabatic coupling (NAC) is needed as long as the psBO approximation breaks down, which is quite common for radiationless processes, such as IC,125 decay through CI,126,127 and potentially ISC and RISC.1 These nonadiabatic transitions are governed by T^N mentioned in Eq. (18), and the NAC between two vibronic states could be derived as128

HNAC=Φm(q,Q)22k2Qk2Φn(q,Q)2kdmnkΛm(Q)Λn(Q)Qk, (23)

where T^N is split into two momentum operators for electronic and vibrational states, respectively, and

dmnk=Ψm(q,Q)Ψn(q,Q)Qk (24)

is the derivative coupling (DC). The DC associated with TDDFT excited states can be formulated using the equation-of-motion (EOM) and response theory72,73,129 or pseudo-wavefunction approach.74,130 The DC implementations with different DFT methods are summarized in Table I. In short, it is not necessary to explicitly solve for either orbital responses or amplitude derivatives in the computation of dmnk; only one z-vector needs to be solved for instead. One may refer to Ref. 131 for a review about the derivative couplings in TDDFT.

TABLE I.

Development of analytical gradient, Hessian, and nonadiabatic coupling (NAC) of various TDDFT and TDDFT/Classical embedding methods.

Method Energy Gradient Hessian NAC
TDDFT Reference 61 References 67, 68, 108, References 70 and 71 References 72–74
110, and 150154 and 155–159
SF-DFT References 160–163 References 164 and 165 References 166 and 167
SA-TDDFT References 130 and 168 Reference 168 References 130 and 169
TDDFT/PCM References 22, 75, and 170174 References 69, 114, and 115 Reference 116
TDDFT/MM Many Reference 175 Reference 175 References 176 and 177
TDDFT/MMpol References 83, 92, and 178181 References 111113 and 182

2. Transition dipole moment and its nuclear derivatives

For studying radiative transitions, the transition dipole moment μ and frequently its nuclear derivative need to be calculated. Like SOCME in Eq. (22), μ can also be expanded as

ΛGg|μGK|ΛKk=μGK(Q=0)ΛGg|ΛKk+jμGKQjQ=0ΛGg|Qj|ΛKk+. (25)

FC approximation just keeps the first term above and takes μGK=ΨG|μ^|ΨK as a constant. The contribution of the second term could also be significant especially when μGK is small. In these situations, one must invoke the FCHT approximation to take the first two terms into account.

The nuclear derivatives of the transition dipole moments between the ground and excited states can be evaluated as μGEx=Tr(δρxμ)+Tr(δρμx). Here, δρ and μ are the transition density matrix and the dipole moment matrix, respectively, in the basis of molecular or atomic orbitals. Observably, μGEx is the geometry-dependent reference. δρx has already been evaluated during the analytical calculation of the excited-state Hessian, as showed in Ref. 70 (with TDA) and Ref. 71 (full TDDFT). Therefore, the computational time will be saved if one obtains μGEx as a by-product of the calculation of the excited-state Hessian.

3. SOC matrix element between different electronic states and its nuclear derivatives

The SOC operator acts on both the angular and spin components of electronic states, leading to its defining characteristics of mixing the orbital and spin degrees of freedom, thus allowing electronic states of different multiplicities to couple. It plays an essential role in photoinduced phenomena that involves change in the spin multiplicities of electronic states as it governs the rates of phosphorescence emission132 and ISC/RISC processes.133 Since the TDDFT has been routinely used in studying photoexcited molecular systems, it is desirable to calculate SOC at the TDDFT level of theory.

Given that SOC stems from relativistic Dirac theory, the SOC effects can be taken into account naturally in relativistic TDDFT,134 such as the four- and two-component relativistic theories. This type of treatment is rigorous and of high quality but computational demanding. For molecules in which the SOC is weak and therefore serves as a perturbation, the perturbative approaches are readily applicable.134,135 In this case, the pure-spin states are first obtained from scalar relativistic or non-relativistic LR-TDDFT;136–144 then, one could calculate SOC matrix elements among these excited states.

The perturbational treatment is not so accurate as the variational approaches are. However, the perturbative SOC approach is less computational demanding and allows one to identify the weights of singlet and triplet characters in a spin-adiabatic state.145 Beyond the linear-response regime, the calculation of SOCMEs has also been performed within quadratic-response TDDFT, which can be viewed as implicit sum-over-states compared to perturbation theory.146,147 It is worth noting that Valentine and Li proposed a novel procedure to calculate SOCMEs from variational relativistic electronic structure theory.148

As shown in Eq. (19), the calculation of ISC or RISC rate usually requires one to account for both the HT-type vibronic and the spin-vibronic effects.1,149 The former vibronic effect requires one to know the nuclear derivatives of SOCME, whose analytical implementation is currently under development.

III. ELECTRONIC EXCITED STATES WITHIN TDDFT/PCM AND TDDFT/MM

When a system of interest is embedded into a condensed-phase medium, its potential energy surfaces are perturbed, which can potentially cause a substantial change in the molecular properties, spectral profiles, and various transition rates and the corresponding quantum yield.

Due to the formidable computational cost of the realistic complexity, a full-QM modeling of the electronic structure of a system and its environment is usually infeasible. For instance, QM modeling, especially plane-wave density functional theory, has been employed to describe plasmonic metal surfaces and adsorbed reactants in studies of photon-driven catalysis.183–185 However, given its steeply increasing cost, routine ab initio QM modeling is typically limited to systems (or subsystems) smaller than 2 or 3 nm. This has inspired the development of methods for capturing the environment effects via lower-level theories, such as classical analytical models (specifically for metal nanoparticles),186–191 implicit solvation models, polarizable or nonpolarizable MM force fields, and coarse-grained models.192

When a chemical system is immersed in liquid solutions, its properties may change significantly, especially in polar solvents. The solvent effects are usually accounted for by a continuum solvent description, such as the family of PCM6,7 or a mixed treatment that integrates different descriptions of the solvent, for example, representing the first solvation shell with explicit solvent molecules while using a continuum model for the bulk. In the PCM method, the electrostatic polarization interaction between solutes and solvents is determined via the induced apparent surface charge method for which one needs to specify a surface that defines the continuum boundary. TDDFT coupled with PCM for excitation energy calculations has become a general routine in most quantum chemistry software packages. The implementation of the analytical gradient69,114,115 and Hessian116 of TDDFT/PCM excited states has also been realized by several groups. TDDFT/PCM has been shown to be successful in supporting the analysis of experimental data, providing useful insights for a better understanding of photophysical and photochemical pathways in a solution.

The PCM has also been applied to model nanoparticles (NPs), leading to the PCM-NP approach. In the quasi-static limit, the NP experiences an electrostatic potential at a given frequency, which induces polarization charges on the nanoparticle surface. The QM/PCM-NP approach, which explicitly includes mutual polarization effects between the molecule and NP, has been applied to calculate the absorption factor as well as the radiative and nonradiative decay rates.3

The all-atom polarizable molecular mechanical models have emerged in the last few decades. Through introducing induced charges and/or dipoles at each atom site of the MM region, these MMpol models can efficiently capture the electronic motion in solvents, proteins, and large MNPs (>10 nm).193,194 Among the MMpol models for MNPs, of particular interest to us are the discrete interaction model (DIM)83,195,196 and its coordination-dependent variant (cd-DIM)193 with the induced charges and dipoles representing polarization effects. Several similar all-atom MMpol models have also been developed, such as the point–dipole interaction model197,198 (including its combination with either electronegativity equalization199 or charge-transfer200), charge–dipole interaction model,201,202 and atomic dipole approximation model.194 Using a polarizable force field in QM/MM calculations is advantageous in light of the fact that the electronic polarization of the MM region can be described, especially on the occasion when electronic excitation is involved in the QM region.

Comparing with the QM/PCM methods, QM/MM or QM/MMpol approaches, which allow us to incorporate the explicit atomic or molecular details of surrounding molecules, can be applied to more complex environments. There are three embedding schemes for QM/MM, namely, mechanical,203 electrostatic, and polarizable.204 It is essential to combine TDDFT with MM to explore the molecular excited-state structures and dynamical properties in complex condensed-phase systems. Table I summarizes the current status of TDDFT-related approaches. Currently, the analytic nuclear gradients of TDDFT/MM and TDDFT/MMpol excited states are made available by a few groups. The implementation of analytic second-order energy derivatives175 for TDDFT/MM with a non-polarizable force field has also been realized. The newly proposed hybrid methods, TDDFT/DIM (for metal nanoparticles)83,178 and TDDFT/MM with fluctuatingcharges and fluctuating dipoles (for solvents),181 have only been made available for excitation energy calculations. Note that within the hybrid schemes, NAC calculations have only been carried out using non-polarizable AMEBA,176,177 and the incorporation of the environment effect in this model needs to be reviewed carefully.

IV. CHOICE OF TDDFT MODELS

In the TDDFT modeling of electronic transitions, one can choose among several variations of the TDDFT methodology for the problem in hand.

A. TDA vs full-TDDFT

Tamm–Dancoff approximation (TDA) amounts to the negligence of the Casida B matrix in the TDDFT working equations. As a result, all Y amplitudes become zero. Using perturbation theory, it is easy to show that TDA excitation energies would always be larger than corresponding TDDFT values with the leading difference being second-order: ωmTDAωmTDDFTnXmTDABXnTDA2/(ωmTDA+ωnTDA). In general, B matrix elements are relatively small, so TDA produces results that closely resemble TDDFT values. When using pure functionals [such as Becke88-Lee-Yang-Parr (BLYP) and Perdew–Burke–Ernzerhof (PBE)] or conventional hybrid functionals (such as B3LYP and PBE0), however, full TDDFT tends to systematically underestimate the energy of excited states with a small charge-transfer or Rydberg character. In those cases, as shown in a recent TDDFT benchmarking study,205 TDA can produce slightly more accurate excitation energies than full TDDFT as shown in Fig. 1. On the other hand, for range-separated hybrid functionals, full TDDFT can produce more accurate results (compared to CC2 values). Full TDDFT usually encounters problems with triplet states because the DFT ground state is used as the reference, which in many cases even leads to triplet instabilities. The TDA can be used to circumvent the triplet instabilities. Furthermore, TDDFT obeys the Thomas–Reiche–Kuhn sum rule of the oscillator strengths206 and therefore yields more precise results for the oscillator strength and other related physical quantities. The TDA violates the sum rule, which leads to poor results in the calculated oscillator strengths. Interestingly, it is found that the TDA performs better in the calculation of NACs than the full TDDFT, contrary to the conjecture that the TDA might cause the NAC results to deteriorate and violate the sum rule.207 For the excited-state harmonic vibrational frequencies, however, full TDDFT does not seem to be advantageous since the numerical tests demonstrate that the accuracy of TDDFT with and without TDA is comparable to each other.71

FIG. 1.

FIG. 1.

Root-mean-square (RMS) and mean-signed-average (MSA) differences (in eV) between vertical excitation energies calculated at the TDDFT, TDA, and CC2208 levels using the def2-TZVP209 basis set. Excitation energies of the five lowest excited states of the 11 chromophores are included. Reproduced with permission from Shao et al., J. Chem. Theory Comput. 16, 587–600 (2019). Copyright 2019 American Chemical Society.

B. Spin–flip DFT

Spin–flip (SF) DFT, which can be viewed as a special version of TDA,160–162 handles up to four lowest electronic states of a system with an even number of electrons (two closed-shell singlets, an open-shell singlet, and a mz = 0 triplet) on the same footing, thus offering a balanced description for these states. These states are obtained by using a high-spin (mz = 1 triplet) reference state and solving the eigenproblem within spin–flip excitations, where an electron is excited from an α occupied molecular orbital to a β virtual orbital. In general, for a system whose ground state has a total spin quantum number S (multiplicity 2S + 1), the reference needs to be in an S + 1 state in order to target the lowest-energy states with total spin quantum number S and S + 1 via spin–flip excitations. Compared to standard TDA and TDDFT, the spin–flip TDA model usually provides more accurate results for diradicals and other systems with multi-reference characters. After the proposition of the original SF-DFT,160 there have been various improved variations, such as the spin-adapted SF-DFT method aiming to eliminate the spin-contamination issue.163,210,211

When CI appears between the electronic ground state and an excited state, it is an advantage to use the SF method to provide a balanced description of the two electronic states. Gordon and Minezawa180 used the SF-DFT combined with the effective fragment potential (EFP) and illustrated that polar solvents could change the CI geometry dramatically and strongly stabilize its energy. Herbert and Zhang166 studied the CI point of the D0 and D1 surfaces of the H3 molecule. They have demonstrated that the SF-configuration interaction singles (SF-CIS) and SF-BH&HLYP (the BH&HLYP functional, in which LYP correlation is combined with Becke’s “half and half” (BH&H) exchange functional, consisting of 50%Hartree-Fock exchange and 50% Becke88 exchange) methods could locate CI points that involve the reference (ground) state correctly, which is problematic for the spin-conserved counterparts of these methods.212 What’s more, equipped with analytic derivative couplings, their method could substantially reduce the cost of locating minimum-energy crossing points.166

Analytic gradient164,165 and nonadiabatic coupling166,167 have been developed for gas-phase SF-DFT calculations, while the analytical Hessian is still under development. As far as we know, there has been no much discussion on SF-DFT/PCM or SF-DFT/MMpol models. Within a linear-response framework, however, a PCM or MMpol modeling of the complex environment is expected to yield no additional contribution to the A matrix (due to a transition density in the αβ block). Therefore, theoretically, it reduces to a “zeroth-order” model, where PCM surface charges or MMpol permanent/induced moments affect the results only by polarizing the KS orbitals.

C. Spin-adiabatic DFT/TDDFT states

Spin-adiabatic (SA) TDDFT has gained attention only in recent years.130,168 Compared with the spin-diabatic states, the spin-adiabatic counterparts have shown advantages130 in Tully’s fewest switches surface hopping (FSSH) dynamics213 to evaluate the rates and branching ratios of ISC processes where spin multiplicity changes. For a system with an even number of electrons, there are 4n + 1 spin-diabatic states: the ground state, the lowest n open-shell singlet excited-states, and the lowest 3n triplet excited-states (with ms = −1, 0, 1). They could be coupled together via the SOC operator,

H(0)+VSO=ES00VS0T1SO0VS0TnSO0ES1VS1T1SO0VS1TnSOVT1S0SOVT1S1SOET1VT1SnSOVT1TnSO00VSnT1SOESnVSnTnSOVTnS0SOVTnS1SOVTnT1SOVTnSnSOETn. (26)

Note that the couplings between different triplet states could be nonzeros. Formally, all the spin-adiabatic states can be obtained by diagonalizing this (4n + 1) × (4n + 1) matrix.

Bellonzi et al.130,168 implemented an efficient iterative procedure for diagonalizing the above coupling matrix between TDA singlets and triplets but without including the ground state. Furthermore, they formulated the analytical gradient of the obtained spin-adiabatic TDA states168 and developed the nonadiabatic couplings between the spin-adiabatic TDA states.130 This new and substantial advance is expected to facilitate the study of ISC and allow for a balanced treatment between ISC and IC events.

On the other hand, a future extension of this methodology to include the ground state as well in the coupling [Eq. (26)] would be desirable, which will provide a general framework for, for instance, the study of phosphorescence decay and the construction of spin-adiabatic potential energy surfaces for modeling spin-crossing reactions.214 Furthermore, in the study of spin-crossing reactions, it was found to be a poor approximation to describe the lowest singlet and triplet states using the same set of molecular orbitals.214 Therefore, it should be more suitable to have one set of optimized orbitals for the singlet state and another for the triplet state.

V. CHALLENGES WITH TDDFT MODELING OF EXCITED STATES IN THE CONDENSED PHASE

A. Challenges with the modeling of system–environment interactions

To illustrate the challenges in describing several QM–MM interactions of an actual system, a pyridine–Ag20 complex in Fig. 2 is used as a showcase. As shown in the following, this is a highly challenging system for classical embedding due to the strong interaction between the pyridine molecule and the metal cluster. Our primary focus is on the effect of the metal cluster on electronic excitations within the pyridine molecule. This complex is handled using a QM/DIM model adopted from Jensen et al.,83,84,178,215 which combines a DIM description for the Ag20 cluster and a DFT or TDDFT description for pyridine. All calculations are carried out using a preliminary implementation216 of the QM/DIM model within the Q-Chem software package.217

FIG. 2.

FIG. 2.

Scheme of pyridine–silver20 (Py–Ag20) configuration.

1. Charge transfer between QM and MM regions

As indicated by charge populations in Table II, for the electronic ground state of Py–Ag20, there is a charge migration (0.16–0.30 e) from the pyridine molecule to the Ag20 cluster. Energy decomposition analysis (EDA) based on absolutely localized molecular orbitals (ALMO)218 for this complex also shows a substantial charge-transfer stabilization (−49.09 kJ/mol) together with a similar amount of polarization energy (−38.83 kJ/mol).

TABLE II.

Net charges on the pyridine molecule within Merz–Kollman ESP,219 ChElPG,220 and fragment-based Hirshfeld (FBH)221 charge partitioning schemes and energy decomposition analysis of the ground-state energy of the pyridine–silver20 complex calculated with the ωB97X-D222 functional and 6-31G(d) basis set and SRSC ECP.

QPy (a.u.) ΔE (kJ/mol)
QESP QChElPG QFBH ΔEFrozen ΔEPol ΔECT ΔETotal
0.30 0.22 0.16 7.68 −38.83 −49.09 −80.23

Starting from this ground state with partial charge transfer, the lowest 200 excited states from a full-TDDFT calculation are mostly local excitations on the metal cluster with a few states featuring charge transfer between the molecule and the metal cluster. None of these low-lying states correspond to local excitations on the pyridine molecule. This poses a serious challenge to the modeling of electronic excitations in these complexes since the full or partial charge-transfer excitations are not incorporated into the existing hybrid QM/MM approaches, such as TDDFT/DIM. Here, state-selective optimization schemes223–225 might help capture the local excited states in large systems.

2. Proper damping of QM/MM electrostatics

To show the effect of the damped QM/MM electrostatics in the QM/DIM modeling of the Py–Ag20 complex, several frontier molecular orbital energies are plotted in Fig. 3. The orbital energies for the gas-phase pyridine and the Py–Ag20 complex are obtained from DFT/DIM and ALMO-DFT226 calculations, respectively.

FIG. 3.

FIG. 3.

Frontier orbital energies (three occupied orbitals and two unoccupied orbitals) from DFT calculations of the gas-phase pyridine, DFT/DIM calculations (with Gaussian exponents of 0.4, 0.6, 0.8, and 1.0 a.u.), and ALMO-DFT calculations for the Py–Ag20 complex. All calculations used the ωB97X-D functional, 6-31G(d) basis set, and SRSC ECP.

As shown in Fig. 3, all frontier MO energies of the pyridine molecule are shifted down upon its adsorption onto the Ag20 cluster. The highest occupied molecule orbital (HOMO), which is the only σ-orbital out of the five orbitals, is subjected to the largest decrease in its energy, thus becoming HOMO − 1 or HOMO − 2 in DFT/DIM and ALMO-DFT calculations. Overall, we find that using a damping factor (αm) of 0.6 or 0.8 a.u. for the induced dipoles in the DIM model best reproduces the relative energies between these frontier orbitals obtained from the ALMO-DFT reference. The five lowest excitation energies show a similar trend as the frontier orbitals with the increasing damping factor (see Fig. 4). In fact, TDDFT-TDA/DIM calculations with a damping factor αm of 1.0 a.u. would best reproduce the results of the ALMO-TDDFT-TDA calculations,97,227 which can be used as the benchmark of the local excited states.

FIG. 4.

FIG. 4.

First five excitation energies from TDDFT-TDA calculations of the gas-phase pyridine, TDDFT-TDA/DIM calculations (with Gaussian exponents of 0.4, 0.6, 0.8, and 1.0 a.u.), and ALMO-TDDFT-TDA calculations for the Py–Ag20 complex. All calculations used the ωB97X-D functional, 6-31G(d) basis, and SRSC ECP.

3. Description of QM/MM van der Waals interactions

The van der Waals (vdW) interactions between the QM and MM regions are crucial for obtaining accurate potential energy profiles. In our implementation of the QM/DIM method, a distance-dependent Lennard-Jones (i.e., 12–6) potential proposed by Payton et al.215 is adopted with a small modification. Without the repulsive interaction, the optimized Py–Ag bond would become very short due to a strongly attractive QM/MM electrostatic interaction.

To avoid a system-dependent tuning of QM/MM vdW parameters as we have performed with the ɛ0 and ɛ1 parameters for the nitrogen atom, one should make the vdW parameters of the QM atoms dependent on the electronic structure (densities, molecular orbitals, or atomic charges) of the QM region.228–230 Within the ab initio QM/MM framework, one can tap into the Tkatchenko–Scheffler dispersion model,231 where the C6 parameter of an atom is computed from its Hirshfeld-weight-based atomic volume in the molecule. This has led to two density-dependent QM/MM vdW models by Giovannini et al.229 and by Curutchet et al.,230 respectively. We expect that these models be further tested and developed before becoming widely adopted in condensed-phase QM/MM simulations. Alternatively, one can predefine localized molecular orbitals on the MM atoms and interact them with orbitals from the QM region to acquire dispersion and repulsion interactions. This has been implemented in DFT/EFP models.232–234

B. Challenges with the computational costs

TDDFT modeling of condensed-phase excited-state processes can become very expensive, especially with a large QM region or a long simulation (1 ps or longer). The first bottleneck of these modelings come from the QM calculation itself, namely, the computation of DFT and TDDFT energy gradient/Hessian and state-to-state derivative couplings. It is time-consuming to evaluate many Fock-like or Fock-derivative-like matrices with some of them involving the third and fourth derivatives of the exchange–correlation functional. Two ways can be pursued to accelerate these calculations. First, following Grimme and co-workers, one can approximate the two-electron integrals and transition density, which has led to the simplified TDA scheme for a large system.235–237 Second, one should be able to extrapolate SCF orbitals, TDDFT amplitudes, and z-vectors from previous geometries during a geometry optimization or molecular dynamics simulation. To extrapolate SCF orbitals, the Fock matrix extrapolation scheme from Pulay, Herbert, and others238,239 and extended Lagrangian from Niklasson240,241 can already reduce the SCF cycles. For excited state simulations with TDDFT energies/forces, one should also be able to extrapolate amplitudes and z-vectors in a way similar to the handling of MP2 z-vectors by Steele and Tully.242

For TDDFT/MMpol with an extended (solvent, protein, or metal-cluster) environment, a second bottleneck comes from the solution of a large number of induced charges/dipoles during each SCF iteration.3,76,243,244 Several efforts have been devoted to the reduction of this computational cost.93,94 First, in order to avoid the O(N3) complexity associated with directly solving linear equations for induced charges or dipoles, iterative solvers can be employed, such as Jacobi iterations (JI)245–247 and preconditioned conjugate gradient (PCG)248 and the truncated version of JI249 and PCG.250,251 For large systems, Nocito and Beran proposed a divide-and-conquer Jacobi iteration (DA-JI) method252,253 and also modified an always stable predictor-corrector (ASPC) algorithm.254 Second, one can truncate the mutual polarization among MM atoms such as in the 3-AMOEBA scheme.255 Finally, the extended Lagrangian method can be employed to remove the need to fully converge MM induced dipoles at each geometry.256–258

C. Challenges with the interpretability of computational results

Fundamentally, we are interested in the effects of various internal and external perturbations on excited-state processes, namely, how they shift absorption/emission wavelengths and how they modulate the radiative/nonradiative transition rates and thus change the overall quantum yield. Internally, a system can be “perturbed” by chemical modifications, such as substitutions and combination of different donors/bridges/acceptors. Externally, a system can be placed in different solvents, protein mutants, or metal clusters of various shapes, sizes, and compositions.

TDDFT modeling has proven to be extremely valuable in these studies. In addition to the TDDFT calculations themselves, many efforts have been devoted to the challenging task of analyzing the results of these calculations. Given a computed TDDFT excited state, one can now readily (a) visualize the density change from the ground state and compute the corresponding charge populations; (b) visualize the canonical orbitals and natural transition orbitals associated with the excitation; and (c) characterize it as being a local excitation, a charge-transfer excitation, or a hybrid.

The characterization of the lowest excited states of a system is especially important in many design problems, such as new thermally activated delayed fluorescence (TADF) molecules for which a hybridized local and intramolecular charge-transfer (HLCT) excitation might optimally enhance the reverse internal system crossing while retaining a decent fluorescence intensity. The Λ values from Peach et al.259 and charge transfer numbers (CTN) from Plasser et al.260,261 have been widely used for such characterizations.

While the electronic density change upon an excitation is commonly visualized, it was only until recently when one can start to analyze the excitation energy distribution, which can also be of interest. For several example systems, we showed that it is feasible to compute and visualize the LR-TDDFT excitation energy density.262 It has also provided a way for one to formulate the effective energies of the hole and particle for a charge-transfer excitation.

In the design of new fluorophores/dyes, guiding principles or inexpensive tools are needed to help quickly predict how a chemical modification or environmental change (as mentioned above) might affect the photophysical properties of a fluorophore (dye) molecule. As for many of these fluorophores, the HOMO → LUMO transition or other transitions between frontier orbitals play the dominant role; we expect the ALMO-based orbital interaction analysis96,100 to be increasingly used to analyze or predict substituent effects or donor–acceptor interactions. On the other hand, the substitution/mutation effects on the rate of radiative and non-radiative transitions are less well-studied and poorly understood, thus providing opportunities for the development of new analysis tools.

VI. SUMMARY

In this Perspective, we have provided an overview on the successes and challenges of TDDFT-related approaches for modeling the excited-state properties and processes. A particular emphasis is placed on which electronic structure properties are accessible with which methods rather than on one single topic, such as the accuracy of excitation energies.

First, we introduced the main electronic structure quantities required by the description of molecular excited-state properties and processes, such as the analytic energy derivatives of the TDDFT excited state, and the NAC and SOC matrix elements. Second, we gave a summary of the recent development of hybrid embedding methods for the description of molecular excited states in complex environments, noting that the evaluation of NAC and SOCME accounting for the environment effects, the analytic nuclear derivative of SOCME, and the analytic Hessian of TDDFT/MMpol have not been successfully implemented into the electronic structure packages. Third, we discussed the advantages of three variations of the TDDFT methods. Specifically, TDA might slightly outperform full TDDFT with pure or conventional hybrid functionals; spin–flip DFT is useful for systems with some multi-reference characters (such as regions around conical interactions); and spin-adiabatic TDDFT could be helpful for studying spin-crossing reactions or the intersystem crossing events. In the end, we elaborated on the challenges in accurately describing interactions between the QM and MM regions, reducing the computational costs associated with modeling excited states in complex environments, and making the results of excited-state calculations physically more transparent and easier to interpret. These challenges have afforded many new opportunities for the development of advanced computational and analysis techniques to treat excited-state-involved processes in complex environments in the future.

ACKNOWLEDGMENTS

W.L. acknowledges financial support from the National Natural Science Foundation of China (Grant Nos. 21833006 and 22173074). Y.S. was supported by the National Institutes of Health (Grant No. R01GM135392), the National Science Foundation (Grant No. CHE-2102071), and the Office of the Vice President of Research and the College of Art and Sciences at the University of Oklahoma (OU).

Contributor Information

WanZhen Liang, Email: mailto:liangwz@xmu.edu.cn.

Yuezhi Mao, Email: mailto:yuezhi.mao@stanford.edu.

Yihan Shao, Email: mailto:yihan.shao@ou.edu.

AUTHOR DECLARATIONS

Conflict of Interest

The authors have no conflicts to disclose.

DATA AVAILABILITY

The data that support the findings of this study are available within the article.

REFERENCES

  • 1.Penfold T. J., Gindensperger E., Daniel C., and Marian C. M., “Spin-vibronic mechanism for intersystem crossing,” Chem. Rev. 118, 6975–7025 (2018). 10.1021/acs.chemrev.7b00617 [DOI] [PubMed] [Google Scholar]
  • 2.Nottoli M., Cupellini L., Lipparini F., Granucci G., and Mennucci B., “Multiscale models for light-driven processes,” Annu. Rev. Phys. Chem. 72, 489–513 (2021). 10.1146/annurev-physchem-090419-104031 [DOI] [PubMed] [Google Scholar]
  • 3.Mennucci B. and Corni S., “Multiscale modelling of photoinduced processes in composite systems,” Nat. Rev. Chem. 3, 315–330 (2019). 10.1038/s41570-019-0092-4 [DOI] [Google Scholar]
  • 4.Warshel A. and Levitt M., “Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme,” J. Mol. Biol. 103, 227–249 (1976). 10.1016/0022-2836(76)90311-9 [DOI] [PubMed] [Google Scholar]
  • 5.Field M. J., Bash P. A., and Karplus M., “A combined quantum mechanical and molecular mechanical potential for molecular dynamics simulations,” J. Comput. Chem. 11, 700–733 (1990). 10.1002/jcc.540110605 [DOI] [Google Scholar]
  • 6.Tomasi J., Mennucci B., and Cammi R., “Quantum mechanical continuum solvation models,” Chem. Rev. 105, 2999–3093 (2005). 10.1021/cr9904009 [DOI] [PubMed] [Google Scholar]
  • 7.Mennucci B., “Polarizable continuum model,” Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2, 386–404 (2012). 10.1002/wcms.1086 [DOI] [Google Scholar]
  • 8.Cappelli C., “Integrated QM/polarizable MM/continuum approaches to model chiroptical properties of strongly interacting solute–solvent systems,” Int. J. Quantum Chem. 116, 1532–1542 (2016). 10.1002/qua.25199 [DOI] [Google Scholar]
  • 9.Gao Y. and Neuhauser D., “Dynamical quantum-electrodynamics embedding: Combining time-dependent density functional theory and the near-field method,” J. Chem. Phys. 137, 074113 (2012). 10.1063/1.4745847 [DOI] [PubMed] [Google Scholar]
  • 10.Chen H., McMahon J. M., Ratner M. A., and Schatz G. C., “Classical electrodynamics coupled to quantum mechanics for calculation of molecular optical properties: A RT-TDDFT/FDTD approach,” J. Phys. Chem. C 114, 14384–14392 (2010). 10.1021/jp1043392 [DOI] [Google Scholar]
  • 11.Yam C., Meng L., Chen G., Chen Q., and Wong N., “Multiscale quantum mechanics/electromagnetics simulation for electronic devices,” Phys. Chem. Chem. Phys. 13, 14365–14369 (2011). 10.1039/c1cp20766k [DOI] [PubMed] [Google Scholar]
  • 12.Sun J., Ding Z., Yu Y., and Liang W., “Nonlinear features of Fano resonance: A QM/EM study,” Phys. Chem. Chem. Phys. 23, 15994–16004 (2021). 10.1039/d1cp02459k [DOI] [PubMed] [Google Scholar]
  • 13.Akimov A. V. and Prezhdo O. V., “Large-scale computations in chemistry: A bird’s eye view of a vibrant field,” Chem. Rev. 115, 5797–5890 (2015). 10.1021/cr500524c [DOI] [PubMed] [Google Scholar]
  • 14.Pribram-Jones A., Gross D. A., and Burke K., “DFT: A theory full of holes?,” Annu. Rev. Phys. Chem. 66, 283–304 (2015). 10.1146/annurev-physchem-040214-121420 [DOI] [PubMed] [Google Scholar]
  • 15.Mardirossian N. and Head-Gordon M., “Thirty years of density functional theory in computational chemistry: An overview and extensive assessment of 200 density functionals,” Mol. Phys. 115, 2315–2372 (2017). 10.1080/00268976.2017.1333644 [DOI] [Google Scholar]
  • 16.Wasserman A., Nafziger J., Jiang K., Kim M.-C., Sim E., and Burke K., “The importance of being inconsistent,” Annu. Rev. Phys. Chem. 68, 555–581 (2017). 10.1146/annurev-physchem-052516-044957 [DOI] [PubMed] [Google Scholar]
  • 17.Hillier I. H., “Chemical reactivity studied by hybrid QM/MM methods,” J. Mol. Struct.: THEOCHEM 463, 45–52 (1999). 10.1016/s0166-1280(98)00391-1 [DOI] [Google Scholar]
  • 18.Sousa C., Tosoni S., and Illas F., “Theoretical approaches to excited-state-related phenomena in oxide surfaces,” Chem. Rev. 113, 4456–4495 (2012). 10.1021/cr300228z [DOI] [PubMed] [Google Scholar]
  • 19.Senn H. M. and Thiel W., “QM/MM methods for biomolecular systems,” Angew. Chem., Int. Ed. 48, 1198–1229 (2009). 10.1002/anie.200802019 [DOI] [PubMed] [Google Scholar]
  • 20.Hohenberg P. and Kohn W., “Inhomogeneous electron gas,” Phys. Rev. 136, B864–B871 (1964). 10.1103/physrev.136.b864 [DOI] [Google Scholar]
  • 21.Kohn W. and Sham L. J., “Self-consistent equations including exchange and correlation effects,” Phys. Rev. 140, A1133–A1138 (1965). 10.1103/physrev.140.a1133 [DOI] [Google Scholar]
  • 22.Adamo C. and Barone V., “A TDDFT study of the electronic spectrum of s-tetrazine in the gas-phase and in aqueous solution,” Chem. Phys. Lett. 330, 152–160 (2000). 10.1016/s0009-2614(00)01082-4 [DOI] [Google Scholar]
  • 23.Aikens C. M., Li S., and Schatz G. C., “From discrete electronic states to plasmons: TDDFT optical absorption properties of Agn (n = 10, 20, 35, 56, 84, 120) tetrahedral clusters,” J. Phys. Chem. C 112, 11272–11279 (2008). 10.1021/jp802707r [DOI] [Google Scholar]
  • 24.Baseggio O., De Vetta M., Fronzoni G., Stener M., Sementa L., Fortunelli A., and Calzolari A., “Photoabsorption of icosahedral noble metal clusters: An efficient TDDFT approach to large-scale systems,” J. Phys. Chem. C 120, 12773–12782 (2016). 10.1021/acs.jpcc.6b04709 [DOI] [Google Scholar]
  • 25.Grabarek D. and Andruniów T., “Assessment of functionals for TDDFT calculations of one- and two-photon absorption properties of neutral and anionic fluorescent proteins chromophores,” J. Chem. Theory Comput. 15, 490–508 (2018). 10.1021/acs.jctc.8b00769 [DOI] [PubMed] [Google Scholar]
  • 26.Amat A., Miliani C., Romani A., and Fantacci S., “DFT/TDDFT investigation on the UV-vis absorption and fluorescence properties of alizarin dye,” Phys. Chem. Chem. Phys. 17, 6374–6382 (2015). 10.1039/c4cp04728a [DOI] [PubMed] [Google Scholar]
  • 27.Wang B., Wang X., Wang W., and Liu F., “Exploring the mechanism of fluorescence quenching and aggregation-induced emission of a phenylethylene derivative by QM (CASSCF and TDDFT) and ONIOM(QM:MM) calculations,” J. Phys. Chem. C 120, 21850–21857 (2016). 10.1021/acs.jpcc.6b07963 [DOI] [Google Scholar]
  • 28.Wan Q., Xia J., Lu W., Yang J., and Che C.-M., “Kinetically controlled self-assembly of phosphorescent AuIII aggregates and ligand-to-metal–metal charge transfer excited state: A combined spectroscopic and DFT/TDDFT study,” J. Am. Chem. Soc. 141, 11572–11582 (2019). 10.1021/jacs.9b04007 [DOI] [PubMed] [Google Scholar]
  • 29.Babazadeh M., Burn P. L., and Huang D. M., “Calculating transition dipole moments of phosphorescent emitters for efficient organic light-emitting diodes,” Phys. Chem. Chem. Phys. 21, 9740–9746 (2019). 10.1039/c9cp01045a [DOI] [PubMed] [Google Scholar]
  • 30.Grotjahn R. and Kaupp M., “Validation of local hybrid functionals for excited states: Structures, fluorescence, phosphorescence, and vibronic spectra,” J. Chem. Theory Comput. 16, 5821–5834 (2020). 10.1021/acs.jctc.0c00520 [DOI] [PubMed] [Google Scholar]
  • 31.Tang Y.-Q. and Liu Y.-J., “Theoretical study on chemiluminescence of H2O2-dependent tetrachloro-1,4-benzoquinone,” J. Org. Chem. 85, 9042–9050 (2020). 10.1021/acs.joc.0c00955 [DOI] [PubMed] [Google Scholar]
  • 32.Ding B.-W. and Liu Y.-J., “Bioluminescence of firefly squid via mechanism of single electron-transfer oxygenation and charge-transfer-induced luminescence,” J. Am. Chem. Soc. 139, 1106–1119 (2017). 10.1021/jacs.6b09119 [DOI] [PubMed] [Google Scholar]
  • 33.Min C.-G., Liu Q.-B., Leng Y., Huang S.-J., Liu C.-X., Yang X.-K., Ren A.-M., and Pinto da Silva L., “Development of firefly oxyluciferin derivatives as pH sensitive fluorescent Probes: A DFT/TDDFT study,” Comput. Theor. Chem. 1133, 18–24 (2018). 10.1016/j.comptc.2018.04.008 [DOI] [Google Scholar]
  • 34.Zemmouche M., García-Iriepa C., and Navizet I., “Light emission colour modulation study of oxyluciferin synthetic analogues via QM and QM/MM approaches,” Phys. Chem. Chem. Phys. 22, 82–91 (2020). 10.1039/c9cp04687a [DOI] [PubMed] [Google Scholar]
  • 35.Iozzi M. F., Mennucci B., Tomasi J., and Cammi R., “Excitation energy transfer (EET) between molecules in condensed matter: A novel application of the polarizable continuum model (PCM),” J. Chem. Phys. 120, 7029–7040 (2004). 10.1063/1.1669389 [DOI] [PubMed] [Google Scholar]
  • 36.Curutchet C., Muñoz-Losa A., Monti S., Kongsted J., Scholes G. D., and Mennucci B., “Electronic energy transfer in condensed phase studied by a polarizable QM/MM model,” J. Chem. Theory Comput. 5, 1838–1848 (2009). 10.1021/ct9001366 [DOI] [PubMed] [Google Scholar]
  • 37.Difley S. and Van Voorhis T., “Exciton/charge-transfer electronic couplings in organic semiconductors,” J. Chem. Theory Comput. 7, 594–601 (2011). 10.1021/ct100508y [DOI] [PubMed] [Google Scholar]
  • 38.Saikin S. K., Eisfeld A., Valleau S., and Aspuru-Guzik A., “Photonics meets excitonics: Natural and artificial molecular aggregates,” Nanophotonics 2, 21–38 (2013). 10.1515/nanoph-2012-0025 [DOI] [Google Scholar]
  • 39.Collini E., “Spectroscopic signatures of quantum-coherent energy transfer,” Chem. Soc. Rev. 42, 4932 (2013). 10.1039/c3cs35444j [DOI] [PubMed] [Google Scholar]
  • 40.Mirkovic T., Ostroumov E. E., Anna J. M., van Grondelle R., Govindjee, and Scholes G. D., “Light absorption and energy transfer in the antenna complexes of photosynthetic organisms,” Chem. Rev. 117, 249–293 (2016). 10.1021/acs.chemrev.6b00002 [DOI] [PubMed] [Google Scholar]
  • 41.Runge E. and Gross E. K. U., “Density-functional theory for time-dependent systems,” Phys. Rev. Lett. 52, 997–1000 (1984). 10.1103/physrevlett.52.997 [DOI] [Google Scholar]
  • 42.Casida M. E. and Huix-Rotllant M., “Progress in time-dependent density-functional theory,” Annu. Rev. Phys. Chem. 63, 287–323 (2012). 10.1146/annurev-physchem-032511-143803 [DOI] [PubMed] [Google Scholar]
  • 43.Jeziorski B. and Monkhorst H. J., “Coupled-cluster method for multideterminantal reference states,” Phys. Rev. A 24, 1668–1681 (1981). 10.1103/physreva.24.1668 [DOI] [Google Scholar]
  • 44.Watts J. D., Gauss J., and Bartlett R. J., “Coupled-cluster methods with noniterative triple excitations for restricted open-shell Hartree–Fock and other general single determinant reference functions. Energies and analytical gradients,” J. Chem. Phys. 98, 8718–8733 (1993). 10.1063/1.464480 [DOI] [Google Scholar]
  • 45.Stanton J. F. and Bartlett R. J., “The equation of motion coupled-cluster method. A systematic biorthogonal approach to molecular excitation energies, transition probabilities, and excited state properties,” J. Chem. Phys. 98, 7029–7039 (1993). 10.1063/1.464746 [DOI] [Google Scholar]
  • 46.Watts J. D. and Bartlett R. J., “Economical triple excitation equation-of-motion coupled-cluster methods for excitation energies,” Chem. Phys. Lett. 233, 81–87 (1995). 10.1016/0009-2614(94)01434-w [DOI] [Google Scholar]
  • 47.Hirata S., “Higher-order equation-of-motion coupled-cluster methods,” J. Chem. Phys. 121, 51 (2004). 10.1063/1.1753556 [DOI] [PubMed] [Google Scholar]
  • 48.Krylov A. I., “Equation-of-motion coupled-cluster methods for open-shell and electronically excited species: The Hitchhiker’s guide to Fock space,” Annu. Rev. Phys. Chem. 59, 433–462 (2008). 10.1146/annurev.physchem.59.032607.093602 [DOI] [PubMed] [Google Scholar]
  • 49.Oana C. M. and Krylov A. I., “Cross sections and photoelectron angular distributions in photodetachment from negative ions using equation-of-motion coupled-cluster Dyson orbitals,” J. Chem. Phys. 131, 124114 (2009). 10.1063/1.3231143 [DOI] [PubMed] [Google Scholar]
  • 50.Klooster R., Broer R., and Filatov M., “Calculation of X-ray photoelectron spectra with the use of the normalized elimination of the small component method,” Chem. Phys. 395, 122–127 (2012). 10.1016/j.chemphys.2011.05.009 [DOI] [Google Scholar]
  • 51.Josefsson I., Eriksson S. K., Ottosson N., Öhrwall G., Siegbahn H., Hagfeldt A., Rensmo H., Björneholm O., and Odelius M., “Collective hydrogen-bond dynamics dictates the electronic structure of aqueous I3,” Phys. Chem. Chem. Phys. 15, 20189 (2013). 10.1039/c3cp52866a [DOI] [PubMed] [Google Scholar]
  • 52.Ponzi A., Angeli C., Cimiraglia R., Coriani S., and Decleva P., “Dynamical photoionization observables of the CS molecule: The role of electron correlation,” J. Chem. Phys. 140, 204304 (2014). 10.1063/1.4876495 [DOI] [PubMed] [Google Scholar]
  • 53.Magyar R. J. and Tretiak S., “Dependence of spurious charge-transfer excited states on orbital exchange in TDDFT: Large molecules and clusters,” J. Chem. Theory Comput. 3, 976–987 (2007). 10.1021/ct600282k [DOI] [PubMed] [Google Scholar]
  • 54.Isegawa M., Peverati R., and Truhlar D. G., “Performance of recent and high-performance approximate density functionals for time-dependent density functional theory calculations of valence and Rydberg electronic transition energies,” J. Chem. Phys. 137, 244104 (2012). 10.1063/1.4769078 [DOI] [PubMed] [Google Scholar]
  • 55.Yabana K. and Bertsch G. F., “Time-dependent local-density approximation in real time,” Phys. Rev. B 54, 4484 (1996). 10.1103/physrevb.54.4484 [DOI] [PubMed] [Google Scholar]
  • 56.Liu J., Guo Z., Sun J., and Liang W., “Theoretical studies on electronic spectroscopy and dynamics with the real-time time-dependent density functional theory,” Front. Chem. China 5, 11–28 (2010). 10.1007/s11458-009-0205-6 [DOI] [Google Scholar]
  • 57.Chen S., Kwok Y., and Chen G., “Time-dependent density functional theory for open systems and its applications,” Acc. Chem. Res. 51, 383–393 (2018). 10.1021/acs.accounts.7b00382 [DOI] [PubMed] [Google Scholar]
  • 58.Li X., Govind N., Isborn C., DePrince A. E., and Lopata K., “Real-time time-dependent electronic structure theory,” Chem. Rev. 120, 9951–9993 (2020). 10.1021/acs.chemrev.0c00223 [DOI] [PubMed] [Google Scholar]
  • 59.Nakatsukasa T., Matsuyanagi K., Matsuo M., and Yabana K., “Time-dependent density-functional description of nuclear dynamics,” Rev. Mod. Phys. 88, 045004 (2016). 10.1103/revmodphys.88.045004 [DOI] [Google Scholar]
  • 60.Marques M., Castro A., Bertsch G. F., and Rubio A., “Octopus: A first-principles tool for excited electron–ion dynamics,” Comput. Phys. Commun. 151, 60–78 (2003). 10.1016/s0010-4655(02)00686-0 [DOI] [Google Scholar]
  • 61.Casida M. E., “Time-dependent density functional response theory for molecules,” in Recent Advances in Density Functional Methods (World Scientific, 1995), pp. 155–192. [Google Scholar]
  • 62.Bauernschmitt R. and Ahlrichs R., “Stability analysis for solutions of the closed shell Kohn–Sham equation,” J. Chem. Phys. 104, 9047–9052 (1996). 10.1063/1.471637 [DOI] [Google Scholar]
  • 63.Rajagopal A. K., “Time-dependent variational principle and the effective action in density-functional theory and Berry’s phase,” Phys. Rev. A 54, 3916–3922 (1996). 10.1103/physreva.54.3916 [DOI] [PubMed] [Google Scholar]
  • 64.Görling A., “Exact exchange kernel for time-dependent density-functional theory,” Int. J. Quantum Chem. 69, 265–277 (1998). [DOI] [Google Scholar]
  • 65.Casida M. E., Jamorski C., Casida K. C., and Salahub D. R., “Molecular excitation energies to high-lying bound states from time-dependent density-functional response theory: Characterization and correction of the time-dependent local density approximation ionization threshold,” J. Chem. Phys. 108, 4439–4449 (1998). 10.1063/1.475855 [DOI] [Google Scholar]
  • 66.Champagne B., Perpète E. A., van Gisbergen S. J. A., Baerends E.-J., Snijders J. G., Soubra-Ghaoui C., Robins K. A., and Kirtman B., “Assessment of conventional density functional schemes for computing the polarizabilities and hyperpolarizabilities of conjugated oligomers: An ab initio investigation of polyacetylene chains,” J. Chem. Phys. 109, 10489–10498 (1998). 10.1063/1.477731 [DOI] [Google Scholar]
  • 67.Van Caillie C. and Amos R. D., “Geometric derivatives of excitation energies using SCF and DFT,” Chem. Phys. Lett. 308, 249–255 (1999). 10.1016/s0009-2614(99)00646-6 [DOI] [Google Scholar]
  • 68.Furche F. and Ahlrichs R., “Adiabatic time-dependent density functional methods for excited state properties,” J. Chem. Phys. 117, 7433–7447 (2002). 10.1063/1.1508368 [DOI] [Google Scholar]
  • 69.Scalmani G., Frisch M. J., Mennucci B., Tomasi J., Cammi R., and Barone V., “Geometries and properties of excited states in the gas phase and in solution: Theory and application of a time-dependent density functional theory polarizable continuum model,” J. Chem. Phys. 124, 094107 (2006). 10.1063/1.2173258 [DOI] [PubMed] [Google Scholar]
  • 70.Liu J. and Liang W., “Analytical Hessian of electronic excited states in time-dependent density functional theory with Tamm-Dancoff approximation,” J. Chem. Phys. 135, 014113 (2011). 10.1063/1.3605504 [DOI] [PubMed] [Google Scholar]
  • 71.Liu J. and Liang W., “Analytical approach for the excited-state Hessian in time-dependent density functional theory: Formalism, implementation, and performance,” J. Chem. Phys. 135, 184111 (2011). 10.1063/1.3659312 [DOI] [PubMed] [Google Scholar]
  • 72.Li Z. and Liu W., “First-order nonadiabatic coupling matrix elements between excited states: A Lagrangian formulation at the CIS, RPA, TD-HF, and TD-DFT levels,” J. Chem. Phys. 141, 014110 (2014). 10.1063/1.4885817 [DOI] [PubMed] [Google Scholar]
  • 73.Li Z., Suo B., and Liu W., “First order nonadiabatic coupling matrix elements between excited states: Implementation and application at the TD-DFT and pp-TDA levels,” J. Chem. Phys. 141, 244105 (2014). 10.1063/1.4903986 [DOI] [PubMed] [Google Scholar]
  • 74.Ou Q., Bellchambers G. D., Furche F., and Subotnik J. E., “First-order derivative couplings between excited states from adiabatic TDDFT response theory,” J. Chem. Phys. 142, 064114 (2015). 10.1063/1.4906941 [DOI] [PubMed] [Google Scholar]
  • 75.Cammi R. and Tomasi J., “Nonequilibrium solvation theory for the polarizable continuum model: A new formulation at the SCF level with application to the case of the frequency-dependent linear electric response function,” Int. J. Quantum Chem. 56, 465–474 (1995). 10.1002/qua.560560850 [DOI] [Google Scholar]
  • 76.Morzan U. N., Alonso de Armiño D. J., Foglia N. O., Ramírez F., González Lebrero M. C., Scherlis D. A., and Estrin D. A., “Spectroscopy in complex environments from QM–MM simulations,” Chem. Rev. 118, 4071–4113 (2018). 10.1021/acs.chemrev.8b00026 [DOI] [PubMed] [Google Scholar]
  • 77.Yang L., Nian S., Zhang G., Sharman E., Miao H., Zhang X., Chen X., Luo Y., and Jiang J., “Role of hydrogen bonding in green fluorescent protein-like chromophore emission,” Sci. Rep. 9, 11640 (2019). 10.1038/s41598-019-47660-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Curutchet C. and Mennucci B., “Quantum chemical studies of light harvesting,” Chem. Rev. 117, 294–343 (2016). 10.1021/acs.chemrev.5b00700 [DOI] [PubMed] [Google Scholar]
  • 79.Cupellini L., Caprasecca S., Guido C. A., Müh F., Renger T., and Mennucci B., “Coupling to charge transfer states is the key to modulate the optical bands for efficient light harvesting in purple bacteria,” J. Phys. Chem. Lett. 9, 6892–6899 (2018). 10.1021/acs.jpclett.8b03233 [DOI] [PubMed] [Google Scholar]
  • 80.Loco D., Protti S., Mennucci B., and Mezzetti A., “Critical assessment of solvent effects on absorption and fluorescence of 3HF in acetonitrile in the QM/PCM framework: A synergic computational and experimental study,” J. Mol. Struct. 1182, 283–291 (2019). 10.1016/j.molstruc.2018.12.085 [DOI] [Google Scholar]
  • 81.Vukovic S., Corni S., and Mennucci B., “Fluorescence enhancement of chromophores close to metal nanoparticles. Optimal setup revealed by the polarizable continuum model,” J. Phys. Chem. C 113, 121–133 (2008). 10.1021/jp808116y [DOI] [Google Scholar]
  • 82.Aroca R. F., “Plasmon enhanced spectroscopy,” Phys. Chem. Chem. Phys. 15, 5355 (2013). 10.1039/c3cp44103b [DOI] [PubMed] [Google Scholar]
  • 83.Morton S. M. and Jensen L., “A discrete interaction model/quantum mechanical method to describe the interaction of metal nanoparticles and molecular absorption,” J. Chem. Phys. 135, 134103 (2011). 10.1063/1.3643381 [DOI] [PubMed] [Google Scholar]
  • 84.Hu Z. and Jensen L., “A discrete interaction model/quantum mechanical method for simulating plasmon-enhanced two-photon absorption,” J. Chem. Theory Comput. 14, 5896–5903 (2018). 10.1021/acs.jctc.8b00893 [DOI] [PubMed] [Google Scholar]
  • 85.Manby F. R., Stella M., Goodpaster J. D., and Miller T. F., “A simple, exact density-functional-theory embedding scheme,” J. Chem. Theory Comput. 8, 2564–2568 (2012). 10.1021/ct300544e [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Lee S. J. R., Welborn M., Manby F. R., and Miller T. F. III, “Projection-based wavefunction-in-DFT embedding,” Acc. Chem. Res. 52, 1359–1368 (2019). 10.1021/acs.accounts.8b00672 [DOI] [PubMed] [Google Scholar]
  • 87.Wesolowski T. A., Shedge S., and Zhou X., “Frozen-density embedding strategy for multilevel simulations of electronic structure,” Chem. Rev. 115, 5891–5928 (2015). 10.1021/cr500502v [DOI] [PubMed] [Google Scholar]
  • 88.Veccham S. P., Lee J., and Head-Gordon M., “Making many-body interactions nearly pairwise additive: The polarized many-body expansion approach,” J. Chem. Phys. 151, 194101 (2019). 10.1063/1.5125802 [DOI] [PubMed] [Google Scholar]
  • 89.Nakai H. and Yoshikawa T., “Development of an excited-state calculation method for large systems using dynamical polarizability: A divide-and-conquer approach at the time-dependent density functional level,” J. Chem. Phys. 146, 124123 (2017). 10.1063/1.4978952 [DOI] [PubMed] [Google Scholar]
  • 90.Tölle J., Böckers M., Niemeyer N., and Neugebauer J., “Inter-subsystem charge-transfer excitations in exact subsystem time-dependent density-functional theory,” J. Chem. Phys. 151, 174109 (2019). 10.1063/1.5121908 [DOI] [PubMed] [Google Scholar]
  • 91.Tölle J. and Neugebauer J., “The seamless connection of local and collective excited states in subsystem time-dependent density functional theory,” J. Phys. Chem. Lett. 13, 1003–1018 (2022). 10.1021/acs.jpclett.1c04023 [DOI] [PubMed] [Google Scholar]
  • 92.Loco D., Polack É., Caprasecca S., Lagardère L., Lipparini F., Piquemal J.-P., and Mennucci B., “A QM/MM approach using the AMOEBA polarizable embedding: From ground state energies to electronic excitations,” J. Chem. Theory Comput. 12, 3654–3661 (2016). 10.1021/acs.jctc.6b00385 [DOI] [PubMed] [Google Scholar]
  • 93.Lipparini F., “General linear scaling implementation of polarizable embedding schemes,” J. Chem. Theory Comput. 15, 4312–4317 (2019). 10.1021/acs.jctc.9b00585 [DOI] [PubMed] [Google Scholar]
  • 94.Bondanza M., Nottoli M., Cupellini L., Lipparini F., and Mennucci B., “Polarizable embedding QM/MM: The future gold standard for complex (bio)systems?,” Phys. Chem. Chem. Phys. 22, 14433–14448 (2020). 10.1039/d0cp02119a [DOI] [PubMed] [Google Scholar]
  • 95.Plasser F., “TheoDORE: A toolbox for a detailed and automated analysis of electronic excited state computations,” J. Chem. Phys. 152, 084108 (2020). 10.1063/1.5143076 [DOI] [PubMed] [Google Scholar]
  • 96.Mao Y., Head-Gordon M., and Shao Y., “Unraveling substituent effects on Frontier orbitals of conjugated molecules using an absolutely localized molecular orbital based analysis,” Chem. Sci. 9, 8598–8607 (2018). 10.1039/c8sc02990c [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Ge Q., Mao Y., and Head-Gordon M., “Energy decomposition analysis for exciplexes using absolutely localized molecular orbitals,” J. Chem. Phys. 148, 064105 (2018). 10.1063/1.5017510 [DOI] [PubMed] [Google Scholar]
  • 98.Ge Q. and Head-Gordon M., “Energy decomposition analysis for excimers using absolutely localized molecular orbitals within time-dependent density functional theory and configuration interaction with single excitations,” J. Chem. Theory Comput. 14, 5156–5168 (2018). 10.1021/acs.jctc.8b00537 [DOI] [PubMed] [Google Scholar]
  • 99.Orozco-Gonzalez Y., Kabir M. P., and Gozem S., “Electrostatic spectral tuning maps for biological chromophores,” J. Phys. Chem. B 123, 4813–4824 (2019). 10.1021/acs.jpcb.9b00489 [DOI] [PubMed] [Google Scholar]
  • 100.Pei Z., Ou Q., Mao Y., Yang J., de la Lande A., Plasser F., Liang W., Shuai Z., and Shao Y., “Elucidating the electronic structure of a delayed fluorescence emitter via orbital interactions, excitation energy components, charge-transfer numbers, and vibrational reorganization energies,” J. Phys. Chem. Lett. 12, 2712–2720 (2021). 10.1021/acs.jpclett.1c00094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Head-Gordon M., “Quantum chemistry and molecular processes,” J. Phys. Chem. 100, 13213–13225 (1996). 10.1021/jp953665+ [DOI] [Google Scholar]
  • 102.Yamaguchi Y., Schaefer H. F. III, Osamura Y., Goddard J. D. et al. , A New Dimension to Quantum Chemistry: Analytic Derivative Methods in Ab Initio Molecular Electronic Structure Theory (Oxford University Press, New York, 1994). [Google Scholar]
  • 103.Pople J. A., Krishnan R., Schlegel H. B., and Binkley J. S., “Derivative studies in Hartree-Fock and Møller-Plesset theories,” Int. J. Quantum Chem. 16, 225–241 (2009). 10.1002/qua.560160825 [DOI] [Google Scholar]
  • 104.Cammi R., Cappelli C., Corni S., and Tomasi J., “On the calculation of infrared intensities in solution within the polarizable continuum model,” J. Phys. Chem. A 104, 9874–9879 (2000). 10.1021/jp001749z [DOI] [Google Scholar]
  • 105.Califano S., Vibrational States (Wiley, New York, 1976). [Google Scholar]
  • 106. Wilson E. B., Decius J. C. , and Cross P. C., Molecular Vibrations: The Theory of Infrared and Raman Vibrational Spectra (Dover Publications, New York, 1980). [Google Scholar]
  • 107.Handy N. C. and Schaefer H. F., “On the evaluation of analytic energy derivatives for correlated wave functions,” J. Chem. Phys. 81, 5031–5033 (1984). 10.1063/1.447489 [DOI] [Google Scholar]
  • 108.Liu F., Gan Z., Shao Y., Hsu C.-P., Dreuw A., Head-Gordon M., Miller B. T., Brooks B. R., Yu J.-G., Furlani T. R., and Kong J., “A parallel implementation of the analytic nuclear gradient for time-dependent density functional theory within the Tamm–Dancoff approximation,” Mol. Phys. 108, 2791–2800 (2010). 10.1080/00268976.2010.526642 [DOI] [Google Scholar]
  • 109.Furche F. and Ahlrichs R., “Erratum: ‘Time-dependent density functional methods for excited state properties’ [J. Chem. Phys. 117, 7433 (2002)],” J. Chem. Phys. 121, 12772 (2004). 10.1063/1.1824903 [DOI] [PubMed] [Google Scholar]
  • 110.Rappoport D. and Furche F., “Analytical time-dependent density functional derivative methods within the RI-J approximation, an approach to excited states of large molecules,” J. Chem. Phys. 122, 064105 (2005). 10.1063/1.1844492 [DOI] [PubMed] [Google Scholar]
  • 111.Si D. and Li H., “Analytic energy gradient in combined time-dependent density functional theory and polarizable force field calculation,” J. Chem. Phys. 133, 144112 (2010). 10.1063/1.3491814 [DOI] [PubMed] [Google Scholar]
  • 112.Zeng Q. and Liang W., “Analytic energy gradient of excited electronic state within TDDFT/MMpol framework: Benchmark tests and parallel implementation,” J. Chem. Phys. 143, 134104 (2015). 10.1063/1.4931734 [DOI] [PubMed] [Google Scholar]
  • 113.Menger M. F. S. J., Caprasecca S., and Mennucci B., “Excited-state gradients in polarizable QM/MM models: An induced dipole formulation,” J. Chem. Theory Comput. 13, 3778–3786 (2017). 10.1021/acs.jctc.7b00475 [DOI] [PubMed] [Google Scholar]
  • 114.Wang Y. and Li H., “Excited state geometry of photoactive yellow protein chromophore: A combined conductorlike polarizable continuum model and time-dependent density functional study,” J. Chem. Phys. 133, 034108 (2010). 10.1063/1.3462248 [DOI] [PubMed] [Google Scholar]
  • 115.Guido C. A., Scalmani G., Mennucci B., and Jacquemin D., “Excited state gradients for a state-specific continuum solvation approach: The vertical excitation model within a Lagrangian TDDFT formulation,” J. Chem. Phys. 146, 204106 (2017). 10.1063/1.4983696 [DOI] [PubMed] [Google Scholar]
  • 116.Liu J. and Liang W., “Analytical second derivatives of excited-state energy within the time-dependent density functional theory coupled with a conductor-like polarizable continuum model,” J. Chem. Phys. 138, 024101 (2013). 10.1063/1.4773397 [DOI] [PubMed] [Google Scholar]
  • 117.Liu J. and Liang W. Z., “Molecular-orbital-free algorithm for the excited-state force in time-dependent density functional theory,” J. Chem. Phys. 134, 044114 (2011). 10.1063/1.3548063 [DOI] [PubMed] [Google Scholar]
  • 118.Neugebauer J., Reiher M., Kind C., and Hess B. A., “Quantum chemical calculation of vibrational spectra of large molecules? Raman and IR spectra for buckminsterfullerene,” J. Comput. Chem. 23, 895–910 (2002). 10.1002/jcc.10089 [DOI] [PubMed] [Google Scholar]
  • 119.Walter M. and Moseler M., “Ab initio wavelength-dependent Raman spectra: Placzek approximation and beyond,” J. Chem. Theory Comput. 16, 576–586 (2019). 10.1021/acs.jctc.9b00584 [DOI] [PubMed] [Google Scholar]
  • 120.Porezag D. and Pederson M. R., “Infrared intensities and Raman-scattering activities within density-functional theory,” Phys. Rev. B 54, 7830–7836 (1996). 10.1103/physrevb.54.7830 [DOI] [PubMed] [Google Scholar]
  • 121.Quinet O. and Champagne B., “Time-dependent Hartree–Fock schemes for analytical evaluation of the Raman intensities,” J. Chem. Phys. 115, 6293 (2001). 10.1063/1.1398310 [DOI] [Google Scholar]
  • 122.Feshbach H., Quantum Mechanics (American Association for the Advancement of Science, New York, 1962). [Google Scholar]
  • 123.Liang W., Ma H., Zang H., and Ye C., “Generalized time-dependent approaches to vibrationally resolved electronic and Raman spectra: Theory and applications,” Int. J. Quantum Chem. 115, 550–563 (2015). 10.1002/qua.24824 [DOI] [Google Scholar]
  • 124.Duschinsky F., “On the interpretation of electronic spectra of polyatomic molecules. I: The Franck-Condon principle,” Acta Physicochim. URSS 7, 551 (1937). [Google Scholar]
  • 125.Ou Q., Peng Q., and Shuai Z., “Toward quantitative prediction of fluorescence quantum efficiency through combining direct vibrational conversion and surface crossing: BODIPYs as example,” J. Phys. Chem. Lett. 11, 7790–7797 (2020). 10.1021/acs.jpclett.0c02054 [DOI] [PubMed] [Google Scholar]
  • 126.Domcke W. and Yarkony D. R., “Role of conical intersections in molecular spectroscopy and photoinduced chemical dynamics,” Annu. Rev. Phys. Chem. 63, 325–352 (2012). 10.1146/annurev-physchem-032210-103522 [DOI] [PubMed] [Google Scholar]
  • 127.Malhado J. P., Bearpark M. J., and Hynes J. T., “Non-adiabatic dynamics close to conical intersections and the surface hopping perspective,” Front. Chem. 2, 97 (2014). 10.3389/fchem.2014.00097 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Hayashi M., Mebel A. M., Liang K. K., and Lin S. H., “Ab initio calculations of radiationless transitions between excited and ground singlet electronic states of ethylene,” J. Chem. Phys. 108, 2044–2055 (1998). 10.1063/1.475584 [DOI] [Google Scholar]
  • 129.Rowe D. J., “Equations-of-motion method and the extended shell model,” Rev. Mod. Phys. 40, 153–166 (1968). 10.1103/revmodphys.40.153 [DOI] [Google Scholar]
  • 130.Bellonzi N., Alguire E., Fatehi S., Shao Y., and Subotnik J. E., “TD-DFT spin-adiabats with analytic nonadiabatic derivative couplings,” J. Chem. Phys. 152, 044112 (2020). 10.1063/1.5126440 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Matsika S., “Electronic structure methods for the description of nonadiabatic effects and conical intersections,” Chem. Rev. 121, 9407–9449 (2021). 10.1021/acs.chemrev.1c00074 [DOI] [PubMed] [Google Scholar]
  • 132.Baryshnikov G., Minaev B., and Ågren H., “Theory and calculation of the phosphorescence phenomenon,” Chem. Rev. 117, 6500–6537 (2017). 10.1021/acs.chemrev.7b00060 [DOI] [PubMed] [Google Scholar]
  • 133.Marian C. M., “Spin–orbit coupling and intersystem crossing in molecules,” Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2, 187–203 (2012). 10.1002/wcms.83 [DOI] [Google Scholar]
  • 134.Liu W. and Xiao Y., “Relativistic time-dependent density functional theories,” Chem. Soc. Rev. 47, 4481–4509 (2018). 10.1039/c8cs00175h [DOI] [PubMed] [Google Scholar]
  • 135.Kühn M. and Weigend F., “Phosphorescence lifetimes of organic light-emitting diodes from two-component time-dependent density functional theory,” J. Chem. Phys. 141, 224302 (2014). 10.1063/1.4902013 [DOI] [PubMed] [Google Scholar]
  • 136.Chiodo S. G. and Russo N., “DFT spin–orbit coupling between singlet and triplet excited states: A case of psoralen compounds,” Chem. Phys. Lett. 490, 90–96 (2010). 10.1016/j.cplett.2010.03.002 [DOI] [Google Scholar]
  • 137.Li Z., Suo B., Zhang Y., Xiao Y., and Liu W., “Combining spin-adapted open-shell TD-DFT with spin-orbit coupling,” Mol. Phys. 111, 3741–3755 (2013). 10.1080/00268976.2013.785611 [DOI] [Google Scholar]
  • 138.Ou Q. and Subotnik J. E., “Electronic relaxation in benzaldehyde evaluated via TD-DFT and localized diabatization: Intersystem crossings, conical intersections, and phosphorescence,” J. Phys. Chem. C 117, 19839–19849 (2013). 10.1021/jp405574q [DOI] [Google Scholar]
  • 139.Franco de Carvalho F., Curchod B. F. E., Penfold T. J., and Tavernelli I., “Derivation of spin-orbit couplings in collinear linear-response TDDFT: A rigorous formulation,” J. Chem. Phys. 140, 144103 (2014). 10.1063/1.4870010 [DOI] [PubMed] [Google Scholar]
  • 140.Gao X., Bai S., Fazzi D., Niehaus T., Barbatti M., and Thiel W., “Evaluation of spin-orbit couplings with linear-response time-dependent density functional methods,” J. Chem. Theory Comput. 13, 515–524 (2017). 10.1021/acs.jctc.6b00915 [DOI] [PubMed] [Google Scholar]
  • 141.Dinkelbach F., Kleinschmidt M., and Marian C. M., “Assessment of interstate spin–orbit couplings from linear response amplitudes,” J. Chem. Theory Comput. 13, 749–766 (2017). 10.1021/acs.jctc.6b01122 [DOI] [PubMed] [Google Scholar]
  • 142.Roemelt M., Maganas D., DeBeer S., and Neese F., “A combined DFT and restricted open-shell configuration interaction method including spin-orbit coupling: Application to transition metal L-edge X-ray absorption spectroscopy,” J. Chem. Phys. 138, 204101 (2013). 10.1063/1.4804607 [DOI] [PubMed] [Google Scholar]
  • 143.de Souza B., Farias G., Neese F., and Izsák R., “Predicting phosphorescence rates of light organic molecules using time-dependent density functional theory and the path integral approach to dynamics,” J. Chem. Theory Comput. 15, 1896–1904 (2019). 10.1021/acs.jctc.8b00841 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Duan J.-X., Zhou Y., Xie Z.-Z., Sun T.-L., and Cao J., “Incorporating spin–orbit effects into surface hopping dynamics using the diagonal representation: A linear-response time-dependent density functional theory implementation with applications to 2-thiouracil,” Phys. Chem. Chem. Phys. 20, 15445–15454 (2018). 10.1039/c8cp01852a [DOI] [PubMed] [Google Scholar]
  • 145.Mori K., Goumans T. P. M., van Lenthe E., and Wang F., “Predicting phosphorescent lifetimes and zero-field splitting of organometallic complexes with time-dependent density functional theory including spin–orbit coupling,” Phys. Chem. Chem. Phys. 16, 14523–14530 (2014). 10.1039/c3cp55438d [DOI] [PubMed] [Google Scholar]
  • 146.Tunell I., Rinkevicius Z., Vahtras O., Sałek P., Helgaker T., and Ågren H., “Density functional theory of nonlinear triplet response properties with applications to phosphorescence,” J. Chem. Phys. 119, 11024–11034 (2003). 10.1063/1.1622926 [DOI] [Google Scholar]
  • 147.Perumal S., Minaev B., and Ågren H., “Spin-spin and spin-orbit interactions in nanographene fragments: A quantum chemistry approach,” J. Chem. Phys. 136, 104702 (2012). 10.1063/1.3687002 [DOI] [PubMed] [Google Scholar]
  • 148.Valentine A. J. S. and Li X., “Toward the evaluation of intersystem crossing rates with variational relativistic methods,” J. Chem. Phys. 151, 084107 (2019). 10.1063/1.5113815 [DOI] [PubMed] [Google Scholar]
  • 149.Lin S., Pei Z., Zhang B., Ma H., and Liang W., “Vibronic coupling effect on the vibrationally resolved electronic spectra and intersystem crossing rates of a TADF emitter: 7-PhQAD,” J. Phys. Chem. A 126, 239–248 (2022). 10.1021/acs.jpca.1c08456 [DOI] [PubMed] [Google Scholar]
  • 150.Chiba M., Tsuneda T., and Hirao K., “Excited state geometry optimizations by analytical energy gradient of long-range corrected time-dependent density functional theory,” J. Chem. Phys. 124, 144106 (2006). 10.1063/1.2186995 [DOI] [PubMed] [Google Scholar]
  • 151.Nguyen K. A., Day P. N., and Pachter R., “Analytical energy gradients of Coulomb-attenuated time-dependent density functional methods for excited states,” Int. J. Quantum Chem. 110, 2247–2255 (2010). 10.1002/qua.22599 [DOI] [Google Scholar]
  • 152.Silverstein D. W., Govind N., van Dam H. J. J., and Jensen L., “Simulating one-photon absorption and resonance Raman scattering spectra using analytical excited state energy gradients within time-dependent density functional theory,” J. Chem. Theory Comput. 9, 5490–5503 (2013). 10.1021/ct4007772 [DOI] [PubMed] [Google Scholar]
  • 153.Burow A. M., Bates J. E., Furche F., and Eshuis H., “Analytical first-order molecular properties and forces within the adiabatic connection random phase approximation,” J. Chem. Theory Comput. 10, 180–194 (2013). 10.1021/ct4008553 [DOI] [PubMed] [Google Scholar]
  • 154.Grotjahn R., Furche F., and Kaupp M., “Development and implementation of excited-state gradients for local hybrid functionals,” J. Chem. Theory Comput. 15, 5508–5522 (2019). 10.1021/acs.jctc.9b00659 [DOI] [PubMed] [Google Scholar]
  • 155.Chernyak V. and Mukamel S., “Density-matrix representation of nonadiabatic couplings in time-dependent density functional (TDDFT) theories,” J. Chem. Phys. 112, 3572–3579 (2000). 10.1063/1.480511 [DOI] [Google Scholar]
  • 156.Hu C., Hirai H., and Sugino O., “Nonadiabatic couplings from time-dependent density functional theory: Formulation in the Casida formalism and practical scheme within modified linear response,” J. Chem. Phys. 127, 064103 (2007). 10.1063/1.2755665 [DOI] [PubMed] [Google Scholar]
  • 157.Tavernelli I., Tapavicza E., and Rothlisberger U., “Nonadiabatic coupling vectors within linear response time-dependent density functional theory,” J. Chem. Phys. 130, 124107 (2009). 10.1063/1.3097192 [DOI] [PubMed] [Google Scholar]
  • 158.Send R. and Furche F., “First-order nonadiabatic couplings from time-dependent hybrid density functional response theory: Consistent formalism, implementation, and performance,” J. Chem. Phys. 132, 044107 (2010). 10.1063/1.3292571 [DOI] [PubMed] [Google Scholar]
  • 159.Tavernelli I., Curchod B. F. E., Laktionov A., and Rothlisberger U., “Nonadiabatic coupling vectors for excited states within time-dependent density functional theory in the Tamm–Dancoff approximation and beyond,” J. Chem. Phys. 133, 194104 (2010). 10.1063/1.3503765 [DOI] [PubMed] [Google Scholar]
  • 160.Shao Y., Head-Gordon M., and Krylov A. I., “The spin–flip approach within time-dependent density functional theory: Theory and applications to diradicals,” J. Chem. Phys. 118, 4807–4818 (2003). 10.1063/1.1545679 [DOI] [Google Scholar]
  • 161.Wang F. and Ziegler T., “Time-dependent density functional theory based on a noncollinear formulation of the exchange-correlation potential,” J. Chem. Phys. 121, 12191 (2004). 10.1063/1.1821494 [DOI] [PubMed] [Google Scholar]
  • 162.Bernard Y. A., Shao Y., and Krylov A. I., “General formulation of spin-flip time-dependent density functional theory using non-collinear kernels: Theory, implementation, and benchmarks,” J. Chem. Phys. 136, 204103 (2012). 10.1063/1.4714499 [DOI] [PubMed] [Google Scholar]
  • 163.Zhang X. and Herbert J. M., “Spin-flip, tensor equation-of-motion configuration interaction with a density-functional correction: A spin-complete method for exploring excited-state potential energy surfaces,” J. Chem. Phys. 143, 234107 (2015). 10.1063/1.4937571 [DOI] [PubMed] [Google Scholar]
  • 164.Seth M., Mazur G., and Ziegler T., “Time-dependent density functional theory gradients in the Amsterdam density functional package: Geometry optimizations of spin-flip excitations,” Theor. Chem. Acc. 129, 331–342 (2010). 10.1007/s00214-010-0819-2 [DOI] [Google Scholar]
  • 165.Lee S., Kim E. E., Nakata H., Lee S., and Choi C. H., “Efficient implementations of analytic energy gradient for mixed-reference spin-flip time-dependent density functional theory (MRSF-TDDFT),” J. Chem. Phys. 150, 184111 (2019). 10.1063/1.5086895 [DOI] [PubMed] [Google Scholar]
  • 166.Zhang X. and Herbert J. M., “Analytic derivative couplings for spin-flip configuration interaction singles and spin-flip time-dependent density functional theory,” J. Chem. Phys. 141, 064104 (2014). 10.1063/1.4891984 [DOI] [PubMed] [Google Scholar]
  • 167.Zhang X. and Herbert J. M., “Analytic derivative couplings in time-dependent density functional theory: Quadratic response theory versus pseudo-wavefunction approach,” J. Chem. Phys. 142, 064109 (2015). 10.1063/1.4907376 [DOI] [PubMed] [Google Scholar]
  • 168.Bellonzi N., Medders G. R., Epifanovsky E., and Subotnik J. E., “Configuration interaction singles with spin-orbit coupling: Constructing spin-adiabatic states and their analytical nuclear gradients,” J. Chem. Phys. 150, 014106 (2019). 10.1063/1.5045484 [DOI] [PubMed] [Google Scholar]
  • 169.Minezawa N. and Nakajima T., “Trajectory surface hopping molecular dynamics simulation by spin-flip time-dependent density functional theory,” J. Chem. Phys. 150, 204120 (2019). 10.1063/1.5096217 [DOI] [PubMed] [Google Scholar]
  • 170.Cammi R. and Mennucci B., “Linear response theory for the polarizable continuum model,” J. Chem. Phys. 110, 9877–9886 (1999). 10.1063/1.478861 [DOI] [Google Scholar]
  • 171.Improta R., Barone V., Scalmani G., and Frisch M. J., “A state-specific polarizable continuum model time dependent density functional theory method for excited state calculations in solution,” J. Chem. Phys. 125, 054103 (2006). 10.1063/1.2222364 [DOI] [PubMed] [Google Scholar]
  • 172.You Z.-Q., Mewes J.-M., Dreuw A., and Herbert J. M., “Comparison of the Marcus and Pekar partitions in the context of non-equilibrium, polarizable-continuum solvation models,” J. Chem. Phys. 143, 204104 (2015). 10.1063/1.4936357 [DOI] [PubMed] [Google Scholar]
  • 173.Bi T.-J., Xu L.-K., Wang F., and Li X.-Y., “Solvent effects for vertical absorption and emission processes in solution using a self-consistent state specific method based on constrained equilibrium thermodynamics,” Phys. Chem. Chem. Phys. 20, 13178–13190 (2018). 10.1039/c8cp00930a [DOI] [PubMed] [Google Scholar]
  • 174.Guido C. A., Rosa M., Cammi R., and Corni S., “An open quantum system theory for polarizable continuum models,” J. Chem. Phys. 152, 174114 (2020). 10.1063/5.0003523 [DOI] [PubMed] [Google Scholar]
  • 175.Zeng Q., Liu J., and Liang W., “Molecular properties of excited electronic state: Formalism, implementation, and applications of analytical second energy derivatives within the framework of the time-dependent density functional theory/molecular mechanics,” J. Chem. Phys. 140, 18A506 (2014). 10.1063/1.4863563 [DOI] [PubMed] [Google Scholar]
  • 176.Wohlgemuth M., Bonačić-Koutecký V., and Mitrić R., “Time-dependent density functional theory excited state nonadiabatic dynamics combined with quantum mechanical/molecular mechanical approach: Photodynamics of indole in water,” J. Chem. Phys. 135, 054105 (2011). 10.1063/1.3622563 [DOI] [PubMed] [Google Scholar]
  • 177.Minezawa N. and Nakajima T., “Quantum mechanical/molecular mechanical trajectory surface hopping molecular dynamics simulation by spin-flip time-dependent density functional theory,” J. Chem. Phys. 152, 024119 (2020). 10.1063/1.5132879 [DOI] [PubMed] [Google Scholar]
  • 178.Morton S. M. and Jensen L., “A discrete interaction model/quantum mechanical method for describing response properties of molecules adsorbed on metal nanoparticles,” J. Chem. Phys. 133, 074103 (2010). 10.1063/1.3457365 [DOI] [PubMed] [Google Scholar]
  • 179.Kosenkov D. and Slipchenko L. V., “Solvent effects on the electronic transitions of p-nitroaniline: A QM/EFP study,” J. Phys. Chem. A 115, 392–401 (2011). 10.1021/jp110026c [DOI] [PubMed] [Google Scholar]
  • 180.Minezawa N. and Gordon M. S., “Optimizing conical intersections of solvated molecules: The combined spin-flip density functional theory/effective fragment potential method,” J. Chem. Phys. 137, 034116 (2012). 10.1063/1.4734314 [DOI] [PubMed] [Google Scholar]
  • 181.Giovannini T., Riso R. R., Ambrosetti M., Puglisi A., and Cappelli C., “Electronic transitions for a fully polarizable QM/MM approach based on fluctuating charges and fluctuating dipoles: Linear and corrected linear response regimes,” J. Chem. Phys. 151, 174104 (2019). 10.1063/1.5121396 [DOI] [PubMed] [Google Scholar]
  • 182.Minezawa N., De Silva N., Zahariev F., and Gordon M. S., “Implementation of the analytic energy gradient for the combined time-dependent density functional theory/effective fragment potential method: Application to excited-state molecular dynamics simulations,” J. Chem. Phys. 134, 054111 (2011). 10.1063/1.3523578 [DOI] [PubMed] [Google Scholar]
  • 183.Martirez J. M. P. and Carter E. A., “Excited-state N2 dissociation pathway on Fe-functionalized Au,” J. Am. Chem. Soc. 139, 4390–4398 (2017). 10.1021/jacs.6b12301 [DOI] [PubMed] [Google Scholar]
  • 184.Martirez J. M. P. and Carter E. A., “Prediction of a low-temperature N2 dissociation catalyst exploiting near-IR–to–visible light nanoplasmonics,” Sci. Adv. 3, eaao4710 (2017). 10.1126/sciadv.aao4710 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Mukherjee S., Libisch F., Large N., Neumann O., Brown L. V., Cheng J., Lassiter J. B., Carter E. A., Nordlander P., and Halas N. J., “Hot electrons do the impossible: Plasmon-induced dissociation of H2 on Au,” Nano Lett. 13, 240–247 (2013). 10.1021/nl303940z [DOI] [PubMed] [Google Scholar]
  • 186.Mie G., “Beiträge zur optik trüber medien, speziell kolloidaler metallösungen,” Ann. Phys. 330, 377–445 (1908). 10.1002/andp.19083300302 [DOI] [Google Scholar]
  • 187.Corni S. and Tomasi J., “Enhanced response properties of a chromophore physisorbed on a metal particle,” J. Chem. Phys. 114, 3739 (2001). 10.1063/1.1342241 [DOI] [Google Scholar]
  • 188.Draine B. T. and Flatau P. J., “Discrete-dipole approximation for scattering calculations,” J. Opt. Soc. Am. A 11, 1491 (1994). 10.1364/josaa.11.001491 [DOI] [Google Scholar]
  • 189.Yang W. H., Schatz G. C., and Van Duyne R. P., “Discrete dipole approximation for calculating extinction and Raman intensities for small particles with arbitrary shapes,” J. Chem. Phys. 103, 869 (1995). 10.1063/1.469787 [DOI] [Google Scholar]
  • 190.García de Abajo F. J. and Howie A., “Relativistic electron energy loss and electron-induced photon emission in inhomogeneous dielectrics,” Phys. Rev. Lett. 80, 5180 (1998). 10.1103/PhysRevLett.80.5180 [DOI] [Google Scholar]
  • 191.Hao F., Nehl C. L., Hafner J. H., and Nordlander P., “Plasmon resonances of a gold nanostar,” Nano Lett. 7, 729–732 (2007). 10.1021/nl062969c [DOI] [PubMed] [Google Scholar]
  • 192.Kmiecik S., Gront D., Kolinski M., Wieteska L., Dawid A. E., and Kolinski A., “Coarse-grained protein models and their applications,” Chem. Rev. 116, 7898–7936 (2016). 10.1021/acs.chemrev.6b00163 [DOI] [PubMed] [Google Scholar]
  • 193.Chen X., Moore J. E., Zekarias M., and Jensen L., “Atomistic electrodynamics simulations of bare and ligand-coated nanoparticles in the quantum size regime,” Nat. Commun. 6, 8921 (2015). 10.1038/ncomms9921 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Lim J., Kang S., Kim J., Kim W. Y., and Ryu S., “Non-empirical atomistic dipole-interaction-model for quantum plasmon simulation of nanoparticles,” Sci. Rep. 7, 15775 (2017). 10.1038/s41598-017-16053-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Jensen L. L. and Jensen L., “Electrostatic interaction model for the calculation of the polarizability of large noble metal nanoclusters,” J. Phys. Chem. C 112, 15697–15703 (2008). 10.1021/jp804116z [DOI] [Google Scholar]
  • 196.Zakomirnyi V. I., Rinkevicius Z., Baryshnikov G. V., Sørensen L. K., and Ågren H., “Extended discrete interaction model: Plasmonic excitations of silver nanoparticles,” J. Phys. Chem. C 123, 28867–28880 (2019). 10.1021/acs.jpcc.9b07410 [DOI] [Google Scholar]
  • 197.Applequist J., “An atom dipole interaction model for molecular optical properties,” Acc. Chem. Res. 10, 79–85 (1977). 10.1021/ar50111a002 [DOI] [Google Scholar]
  • 198.Jensen L., Åstrand P.-O., Osted A., Kongsted J., and Mikkelsen K. V., “Polarizability of molecular clusters as calculated by a dipole interaction model,” J. Chem. Phys. 116, 4001 (2002). 10.1063/1.1433747 [DOI] [Google Scholar]
  • 199.Smalo H. S., Åstrand P.-O., and Mayer A., “Combined nonmetallic electronegativity equalisation and point-dipole interaction model for the frequency-dependent polarisability,” Mol. Phys. 111, 1470–1481 (2013). 10.1080/00268976.2013.797116 [DOI] [Google Scholar]
  • 200.Davari N., Haghdani S., Åstrand P.-O., and Schatz G. C., “Local electric field factors by a combined charge-transfer and point-dipole interaction model,” RSC Adv. 5, 31594–31605 (2015). 10.1039/c5ra04183j [DOI] [Google Scholar]
  • 201.Mayer A., “Formulation in terms of normalized propagators of a charge-dipole model enabling the calculation of the polarization properties of fullerenes and carbon nanotubes,” Phys. Rev. B 75, 045407 (2007). 10.1103/physrevb.75.045407 [DOI] [Google Scholar]
  • 202.Mayer A., González A. L., Aikens C. M., and Schatz G. C., “A charge–dipole interaction model for the frequency-dependent polarizability of silver clusters,” Nanotechnology 20, 195204 (2009). 10.1088/0957-4484/20/19/195204 [DOI] [PubMed] [Google Scholar]
  • 203.Vreven T. and Morokuma K., “Chapter 3 hybrid methods: ONIOM(QM:MM) and QM/MM,” Annu. Rep. Comput. Chem. 2, 35–51 (2006). 10.1016/s1574-1400(06)02003-2 [DOI] [Google Scholar]
  • 204.Lin H. and Truhlar D. G., “QM/MM: What have we learned, where are we, and where do we go from here?,” Theor. Chem. Acc. 117, 185–199 (2006). 10.1007/s00214-006-0143-z [DOI] [Google Scholar]
  • 205.Shao Y., Mei Y., Sundholm D., and Kaila V. R. I., “Benchmarking the performance of time-dependent density functional theory methods on biochromophores,” J. Chem. Theory Comput. 16, 587–600 (2019). 10.1021/acs.jctc.9b00823 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.McLachlan A. D. and Ball M. A., “Time-dependent Hartree–Fock theory for molecules,” Rev. Mod. Phys. 36, 844–855 (1964). 10.1103/revmodphys.36.844 [DOI] [Google Scholar]
  • 207.Hu C., Sugino O., and Watanabe K., “Performance of Tamm-Dancoff approximation on nonadiabatic couplings by time-dependent density functional theory,” J. Chem. Phys. 140, 054106 (2014). 10.1063/1.4862904 [DOI] [PubMed] [Google Scholar]
  • 208.Christiansen O., Koch H., and Jørgensen P., “The second-order approximate coupled cluster singles and doubles model CC2,” Chem. Phys. Lett. 243, 409–418 (1995). 10.1016/0009-2614(95)00841-q [DOI] [Google Scholar]
  • 209.Weigend F. and Ahlrichs R., “Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy,” Phys. Chem. Chem. Phys. 7, 3297–3305 (2005). 10.1039/b508541a [DOI] [PubMed] [Google Scholar]
  • 210.Li Z. and Liu W., “Spin-adapted open-shell random phase approximation and time-dependent density functional theory. I. Theory,” J. Chem. Phys. 133, 064106 (2010). 10.1063/1.3463799 [DOI] [PubMed] [Google Scholar]
  • 211.Li Z., Liu W., Zhang Y., and Suo B., “Spin-adapted open-shell time-dependent density functional theory. II. Theory and pilot application,” J. Chem. Phys. 134, 134101 (2011). 10.1063/1.3573374 [DOI] [PubMed] [Google Scholar]
  • 212.Levine B. G., Ko C., Quenneville J., and MartÍnez T. J., “Conical intersections and double excitations in time-dependent density functional theory,” Mol. Phys. 104, 1039–1051 (2006). 10.1080/00268970500417762 [DOI] [Google Scholar]
  • 213.Tully J. C., “Molecular dynamics with electronic transitions,” J. Chem. Phys. 93, 1061–1071 (1990). 10.1063/1.45917z [DOI] [Google Scholar]
  • 214.Tao Y., Pei Z., Bellonzi N., Mao Y., Zou Z., Liang W., Yang Z., and Shao Y., “Constructing spin-adiabatic states for the modeling of spin-crossing reactions. I. A shared-orbital implementation,” Int. J. Quantum Chem. 120, e26123 (2020). 10.1002/qua.26123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Payton J. L., Morton S. M., Moore J. E., and Jensen L., “A discrete interaction model/quantum mechanical method for simulating surface-enhanced Raman spectroscopy,” J. Chem. Phys. 136, 214103 (2012). 10.1063/1.4722755 [DOI] [PubMed] [Google Scholar]
  • 216.Pei Z., Mao Y., Shao Y., and Liang W. Z., “Analytic high-order energy derivatives for metal nanoparticle-mediated infrared and Raman scattering spectra within the framework of quantum mechanics/molecular mechanics model with induced charges and dipoles,” (published online) (2022). [DOI] [PMC free article] [PubMed]
  • 217.Epifanovsky E., Gilbert A. T. B., Feng X., Lee J., Mao Y., Mardirossian N., Pokhilko P., White A. F., Coons M. P., Dempwolff A. L., Gan Z., Hait D., Horn P. R., Jacobson L. D., Kaliman I., Kussmann J., Lange A. W., Lao K. U., Levine D. S., Liu J., McKenzie S. C., Morrison A. F., Nanda K. D., Plasser F., Rehn D. R., Vidal M. L., You Z.-Q., Zhu Y., Alam B., Albrecht B. J., Aldossary A., Alguire E., Andersen J. H., Athavale V., Barton D., Begam K., Behn A., Bellonzi N., Bernard Y. A., Berquist E. J., Burton H. G. A., Carreras A., Carter-Fenk K., Chakraborty R., Chien A. D., Closser K. D., Cofer-Shabica V., Dasgupta S., de Wergifosse M., Deng J., Diedenhofen M., Do H., Ehlert S., Fang P.-T., Fatehi S., Feng Q., Friedhoff T., Gayvert J., Ge Q., Gidofalvi G., Goldey M., Gomes J., González-Espinoza C. E., Gulania S., Gunina A. O., Hanson-Heine M. W. D., Harbach P. H. P., Hauser A., Herbst M. F., Hernández Vera M., Hodecker M., Holden Z. C., Houck S., Huang X., Hui K., Huynh B. C., Ivanov M., Jász Á., Ji H., Jiang H., Kaduk B., Köhler S., Khistyaev K., Kim J., Kis G., Klunzinger P., Koczor-Benda Z., Koh J. H., Kosenkov D., Koulias L., Kowalczyk T., Krauter C. M., Kue K., Kunitsa A., Kus T., Ladjánszki I., Landau A., Lawler K. V., Lefrancois D., Lehtola S., Li R. R., Li Y.-P., Liang J., Liebenthal M., Lin H.-H., Lin Y.-S., Liu F., Liu K.-Y., Loipersberger M., Luenser A., Manjanath A., Manohar P., Mansoor E., Manzer S. F., Mao S.-P., Marenich A. V., Markovich T., Mason S., Maurer S. A., McLaughlin P. F., Menger M. F. S. J., Mewes J.-M., Mewes S. A., Morgante P., Mullinax J. W., Oosterbaan K. J., Paran G., Paul A. C., Paul S. K., Pavošević F., Pei Z., Prager S., Proynov E. I., Rák Á., Ramos-Cordoba E., Rana B., Rask A. E., Rettig A., Richard R. M., Rob F., Rossomme E., Scheele T., Scheurer M., Schneider M., Sergueev N., Sharada S. M., Skomorowski W., Small D. W., Stein C. J., Su Y.-C., Sundstrom E. J., Tao Z., Thirman J., Tornai G. J., Tsuchimochi T., Tubman N. M., Veccham S. P., Vydrov O., Wenzel J., Witte J., Yamada A., Yao K., Yeganeh S., Yost S. R., Zech A., Zhang I. Y., Zhang X., Zhang Y., Zuev D., Aspuru-Guzik A., Bell A. T., Besley N. A., Bravaya K. B., Brooks B. R., Casanova D., Chai J.-D., Coriani S., Cramer C. J., Cserey G., DePrince A. E., DiStasio R. A., Dreuw A., Dunietz B. D., Furlani T. R., Goddard W. A., Hammes-Schiffer S., Head-Gordon T., Hehre W. J., Hsu C.-P., Jagau T.-C., Jung Y., Klamt A., Kong J., Lambrecht D. S., Liang W., Mayhall N. J., McCurdy C. W., Neaton J. B., Ochsenfeld C., Parkhill J. A., Peverati R., Rassolov V. A., Shao Y., Slipchenko L. V., Stauch T., Steele R. P., Subotnik J. E., Thom A. J. W., Tkatchenko A., Truhlar D. G., Van Voorhis T., Wesolowski T. A., Whaley K. B., Woodcock H. L., Zimmerman P. M., Faraji S., Gill P. M. W., Head-Gordon M., Herbert J. M., and Krylov A. I., “Software for the frontiers of quantum chemistry: An overview of developments in the Q-Chem 5 package,” J. Chem. Phys. 155, 084801 (2021). 10.1063/5.0055522 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Khaliullin R. Z., Cobar E. A., Lochan R. C., Bell A. T., and Head-Gordon M., “Unravelling the origin of intermolecular interactions using absolutely localized molecular orbitals,” J. Phys. Chem. A 111, 8753–8765 (2007). 10.1021/jp073685z [DOI] [PubMed] [Google Scholar]
  • 219.Besler B. H., K. M. Merz, Jr., and Kollman P. A., “Atomic charges derived from semiempirical methods,” J. Comput. Chem. 11, 431–439 (1990). 10.1002/jcc.540110404 [DOI] [Google Scholar]
  • 220.Breneman C. M. and Wiberg K. B., “Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis,” J. Comput. Chem. 11, 361–373 (1990). 10.1002/jcc.540110311 [DOI] [Google Scholar]
  • 221.Rezac J. and de la Lande A., J. Chem. Theory Comput. 11, 528–537 (2015). 10.1021/ct501115m [DOI] [PubMed] [Google Scholar]
  • 222.Chai J.-D. and Head-Gordon M., “Long-range corrected hybrid density functionals with damped atom–atom dispersion corrections,” Phys. Chem. Chem. Phys. 10, 6615–6620 (2008). 10.1039/b810189b [DOI] [PubMed] [Google Scholar]
  • 223.Kovyrshin A. and Neugebauer J., “State-selective optimization of local excited electronic states in extended systems,” J. Chem. Phys. 133, 174114 (2010). 10.1063/1.3488230 [DOI] [PubMed] [Google Scholar]
  • 224.Kovyrshin A. and Neugebauer J., “Potential-energy surfaces of local excited states from subsystem- and selective Kohn–Sham-TDDFT,” Chem. Phys. 391, 147–156 (2011). 10.1016/j.chemphys.2011.02.007 [DOI] [Google Scholar]
  • 225.Kovyrshin A., Angelis F. D., and Neugebauer J., “Selective TDDFT with automatic removal of ghost transitions: Application to a perylene-dye-sensitized solar cell model,” Phys. Chem. Chem. Phys. 14, 8608 (2012). 10.1039/c2cp40840f [DOI] [PubMed] [Google Scholar]
  • 226.Khaliullin R. Z., Head-Gordon M., and Bell A. T., “An efficient self-consistent field method for large systems of weakly interacting components,” J. Chem. Phys. 124, 204105 (2006). 10.1063/1.2191500 [DOI] [PubMed] [Google Scholar]
  • 227.Closser K. D., Ge Q., Mao Y., Shao Y., and Head-Gordon M., “Superposition of fragment excitations for excited states of large clusters with application to helium clusters,” J. Chem. Theory Comput. 11, 5791–5803 (2015). 10.1021/acs.jctc.5b00703 [DOI] [PubMed] [Google Scholar]
  • 228.Kuechler E. R., Giese T. J., and York D. M., “Charge-dependent many-body exchange and dispersion interactions in combined QM/MM simulations,” J. Chem. Phys. 143, 234111 (2015). 10.1063/1.4937166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Giovannini T., Lafiosca P., and Cappelli C., “A general route to include Pauli repulsion and quantum dispersion effects in QM/MM approaches,” J. Chem. Theory Comput. 13, 4854–4870 (2017). 10.1021/acs.jctc.7b00776 [DOI] [PubMed] [Google Scholar]
  • 230.Curutchet C., Cupellini L., Kongsted J., Corni S., Frediani L., Steindal A. H., Guido C. A., Scalmani G., and Mennucci B., “Density-dependent formulation of dispersion–repulsion interactions in hybrid multiscale quantum/molecular mechanics (QM/MM) models,” J. Chem. Theory Comput. 14, 1671–1681 (2018). 10.1021/acs.jctc.7b00912 [DOI] [PubMed] [Google Scholar]
  • 231.Tkatchenko A. and Scheffler M., “Accurate molecular van der Waals interactions from ground-state electron density and free-atom reference data,” Phys. Rev. Lett. 102, 073005 (2009). 10.1103/PhysRevLett.102.073005 [DOI] [PubMed] [Google Scholar]
  • 232.Slipchenko L. V., Gordon M. S., and Ruedenberg K., “Dispersion interactions in QM/EFP,” J. Phys. Chem. A 121, 9495–9507 (2017). 10.1021/acs.jpca.7b05875 [DOI] [PubMed] [Google Scholar]
  • 233.Viquez Rojas C. I., Fine J., and Slipchenko L. V., “Exchange-repulsion energy in QM/EFP,” J. Chem. Phys. 149, 094103 (2018). 10.1063/1.5043107 [DOI] [PubMed] [Google Scholar]
  • 234.Viquez Rojas C. I. and Slipchenko L. V., “Exchange repulsion in quantum mechanical/effective fragment potential excitation energies: Beyond polarizable embedding,” J. Chem. Theory Comput. 16, 6408–6417 (2020). 10.1021/acs.jctc.9b01156 [DOI] [PubMed] [Google Scholar]
  • 235.Grimme S., “A simplified Tamm-Dancoff density functional approach for the electronic excitation spectra of very large molecules,” J. Chem. Phys. 138, 244104 (2013). 10.1063/1.4811331 [DOI] [PubMed] [Google Scholar]
  • 236.Bannwarth C. and Grimme S., “A simplified time-dependent density functional theory approach for electronic ultraviolet and circular dichroism spectra of very large molecules,” Comput. Theor. Chem. 1040–1041, 45–53 (2014). 10.1016/j.comptc.2014.02.023 [DOI] [Google Scholar]
  • 237.Risthaus T., Hansen A., and Grimme S., “Excited states using the simplified Tamm–Dancoff-approach for range-separated hybrid density functionals: Development and application,” Phys. Chem. Chem. Phys. 16, 14408–14419 (2014). 10.1039/c3cp54517b [DOI] [PubMed] [Google Scholar]
  • 238.Pulay P. and Fogarasi G., “Fock matrix dynamics,” Chem. Phys. Lett. 386, 272–278 (2004). 10.1016/j.cplett.2004.01.069 [DOI] [Google Scholar]
  • 239.Herbert J. M. and Head-Gordon M., “Accelerated, energy-conserving Born–Oppenheimer molecular dynamics via Fock matrix extrapolation,” Phys. Chem. Chem. Phys. 7, 3269 (2005). 10.1039/b509494a [DOI] [PubMed] [Google Scholar]
  • 240.Niklasson A. M. N., “Extended Born–Oppenheimer molecular dynamics,” Phys. Rev. Lett. 100, 123004 (2008). 10.1103/physrevlett.100.123004 [DOI] [PubMed] [Google Scholar]
  • 241.Niklasson A. M. N., “Density-matrix based extended Lagrangian Born–Oppenheimer molecular dynamics,” J. Chem. Theory Comput. 16, 3628–3640 (2020). 10.1021/acs.jctc.0c00264 [DOI] [PubMed] [Google Scholar]
  • 242.Steele R. P. and Tully J. C., “Accelerated ab initio molecular dynamics with response equation extrapolation,” Chem. Phys. Lett. 500, 167 (2010). 10.1016/j.cplett.2010.10.003 [DOI] [Google Scholar]
  • 243.Segatta F., Cupellini L., Garavelli M., and Mennucci B., “Quantum chemical modeling of the photoinduced activity of multichromophoric biosystems,” Chem. Rev. 119, 9361–9380 (2019). 10.1021/acs.chemrev.9b00135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Brunk E. and Rothlisberger U., “Mixed quantum mechanical/molecular mechanical molecular dynamics simulations of biological systems in ground and electronically excited states,” Chem. Rev. 115, 6217–6263 (2015). 10.1021/cr500628b [DOI] [PubMed] [Google Scholar]
  • 245.Ren P. and Ponder J. W., “Polarizable atomic multipole water model for molecular mechanics simulation,” J. Phys. Chem. B 107, 5933–5947 (2003). 10.1021/jp027815+ [DOI] [Google Scholar]
  • 246.Pulay P., “Convergence acceleration of iterative sequences. The case of scf iteration,” Chem. Phys. Lett. 73, 393–398 (1980). 10.1016/0009-2614(80)80396-4 [DOI] [Google Scholar]
  • 247.Pulay P., “Improved SCF convergence acceleration,” J. Comput. Chem. 3, 556–560 (1982). 10.1002/jcc.540030413 [DOI] [Google Scholar]
  • 248.Wang W. and Skeel R. D., “Fast evaluation of polarizable forces,” J. Chem. Phys. 123, 164107 (2005). 10.1063/1.2056544 [DOI] [PubMed] [Google Scholar]
  • 249.Simmonett A. C., Pickard F. C., Shao Y., Cheatham T. E., and Brooks B. R., “Efficient treatment of induced dipoles,” J. Chem. Phys. 143, 074115 (2015). 10.1063/1.4928530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Aviat F., Levitt A., Stamm B., Maday Y., Ren P., Ponder J. W., Lagardère L., and Piquemal J.-P., “Truncated conjugate gradient: An optimal strategy for the analytical evaluation of the many-body polarization energy and forces in molecular simulations,” J. Chem. Theory Comput. 13, 180–190 (2016). 10.1021/acs.jctc.6b00981 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Aviat F., Lagardère L., and Piquemal J.-P., “The truncated conjugate gradient (TCG), a non-iterative/fixed-cost strategy for computing polarization in molecular dynamics: Fast evaluation of analytical forces,” J. Chem. Phys. 147, 161724 (2017). 10.1063/1.4985911 [DOI] [PubMed] [Google Scholar]
  • 252.Nocito D. and Beran G. J. O., “Fast divide-and-conquer algorithm for evaluating polarization in classical force fields,” J. Chem. Phys. 146, 114103 (2017). 10.1063/1.4977981 [DOI] [PubMed] [Google Scholar]
  • 253.Nocito D. and Beran G. J. O., “Massively parallel implementation of divide-and-conquer Jacobi iterations using particle-mesh Ewald for force field polarization,” J. Chem. Theory Comput. 14, 3633–3642 (2018). 10.1021/acs.jctc.8b00328 [DOI] [PubMed] [Google Scholar]
  • 254.Nocito D. and Beran G. J. O., “Reduced computational cost of polarizable force fields by a modification of the always stable predictor-corrector,” J. Chem. Phys. 150, 151103 (2019). 10.1063/1.5092133 [DOI] [PubMed] [Google Scholar]
  • 255.Demerdash O. and Head-Gordon T., “Convergence of the many-body expansion for energy and forces for classical polarizable models in the condensed phase,” J. Chem. Theory Comput. 12, 3884–3893 (2016). 10.1021/acs.jctc.6b00335 [DOI] [PubMed] [Google Scholar]
  • 256.Albaugh A., Demerdash O., and Head-Gordon T., “An efficient and stable hybrid extended Lagrangian/self-consistent field scheme for solving classical mutual induction,” J. Chem. Phys. 143, 174104 (2015). 10.1063/1.4933375 [DOI] [PubMed] [Google Scholar]
  • 257.Albaugh A., Niklasson A. M. N., and Head-Gordon T., “Accurate classical polarization solution with no self-consistent field iterations,” J. Phys. Chem. Lett. 8, 1714–1723 (2017). 10.1021/acs.jpclett.7b00450 [DOI] [PubMed] [Google Scholar]
  • 258.Albaugh A., Head-Gordon T., and Niklasson A. M. N., “Higher-order extended Lagrangian Born–Oppenheimer molecular dynamics for classical polarizable models,” J. Chem. Theory Comput. 14, 499–511 (2018). 10.1021/acs.jctc.7b01041 [DOI] [PubMed] [Google Scholar]
  • 259.Peach M. J. G., Benfield P., Helgaker T., and Tozer D. J., “Excitation energies in density functional theory: An evaluation and a diagnostic test,” J. Chem. Phys. 128, 044118 (2008). 10.1063/1.2831900 [DOI] [PubMed] [Google Scholar]
  • 260.Plasser F., Wormit M., and Dreuw A., “New tools for the systematic analysis and visualization of electronic excitations. I. Formalism,” J. Chem. Phys. 141, 024106 (2014). 10.1063/1.4885819 [DOI] [PubMed] [Google Scholar]
  • 261.Plasser F., Bäppler S. A., Wormit M., and Dreuw A., “New tools for the systematic analysis and visualization of electronic excitations. II. Applications,” J. Chem. Phys. 141, 024107 (2014). 10.1063/1.4885820 [DOI] [PubMed] [Google Scholar]
  • 262.Pei Z., Yang J., Deng J., Mao Y., Wu Q., Yang Z., Wang B., Aikens C. M., Liang W., and Shao Y., “Analysis and visualization of energy densities. II. Insights from linear-response time-dependent density functional theory calculations,” Phys. Chem. Chem. Phys. 22, 26852–26864 (2020). 10.1039/d0cp04207b [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Data Availability Statement

The data that support the findings of this study are available within the article.


Articles from The Journal of Chemical Physics are provided here courtesy of American Institute of Physics

RESOURCES