14
THE JOURNAL OF CHEMICAL PHYSICS 144, 204103 (2016) First time combination of frozen density embedding theory with the algebraic diagrammatic construction scheme for the polarization propagator of second order Stefan Prager, 1,a) Alexander Zech, 2,b) Francesco Aquilante, 3,c) Andreas Dreuw, 1,d) and Tomasz A. Wesolowski 2,e) 1 Interdisciplinary Center for Scientific Computing, Ruprecht-Karls University, Im Neuenheimer Feld 368, 69120 Heidelberg, Germany 2 Dèpartement de Chimie Physique, Université de Genève, 30 Quai Ernest-Ansermet, 1211 Genève 4, Switzerland 3 Dipartimento di Chimica “G. Ciamician,” Università di Bologna, Via Selmi 2, 40126 Bologna, Italy (Received 24 February 2016; accepted 25 April 2016; published online 24 May 2016) The combination of Frozen Density Embedding Theory (FDET) and the Algebraic Diagrammatic Construction (ADC) scheme for the polarization propagator for describing environmental eects on electronically excited states is presented. Two dierent ways of interfacing and expressing the so-called embedding operator are introduced. The resulting excited states are compared with supermolecular calculations of the total system at the ADC(2) level of theory. Molecular test systems were chosen to investigate molecule–environment interactions of varying strength from dispersion interaction up to multiple hydrogen bonds. The overall dierence between the supermolecular and the FDE-ADC calculations in excitation energies is lower than 0.09 eV (max) and 0.032 eV in average, which is well below the intrinsic error of the ADC(2) method itself. Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4948741] I. INTRODUCTION Almost all chemical reactions occur in liquid phase and also almost all chemical properties are measured in condensed phases. In these cases, the investigated molecule interacts with its environment, leading to a change of its chemical properties. Electronically excited states, especially, can be strongly influenced by the environment. A large variety of molecules and solvents exhibit this eect of solvatochromism. 1 Hence, including the environment is crucial for comparable calculations and reliable predictions using quantum chemical methods. However, including all environment molecules in quantum chemical calculations exceeds nowadays computational limits. Dierent methods for considering the environment have been proposed, ranging from an implicit treatment as in polarizable continuum models (PCM) 2 to explicit models treating the environment at a molecular mechanics level of theory (QM/MM). 3 It is known that implicit approaches can only model dispersion and Coulombic interaction but may fail in describing specific solvent interactions, for instance hydrogen bonds. 46 In these cases, an explicit model for the environment is needed. Density embedding methods 7,8 constitute such an explicit model, and have received increasing attention in recent years (for the latest reviews see Refs. 9 and 10). Frozen-Density Embedding Theory (FDET) 7,11 in particular provides a formal a) Electronic mail: [email protected] b) Electronic mail: [email protected] c) Electronic mail: [email protected] d) Electronic mail: [email protected] e) Electronic mail: [email protected] framework in which the whole system is described by means of two independent quantities: the embedded wavefunction (interacting or not) and the density associated with the environment. The latter can in principle be taken from lower level quantum mechanical methods, statistical theories for ensembles, or even from experiment. While the original work used DFT 1214 methods to describe the embedded species, also FDET variants based on wavefunction methods have been developed. 11,1517 Carter and collaborators were the first to combine the FDET embedding potential with wavefunction based methods in their approach in which the independent variables are the total density and the density of the embedded species. 8,18 This combination is especially useful for the calculation of excited states where DFT methods have known limitations. 1921 The embedding is accomplished by means of an embedding potential which depends on the density of the embedded species ρ A ( r ). In the canonical form of FDET, herein referred to as conventional FDET, an iterative scheme is applied to obtain a wavefunction which is self- consistent with respect to the embedding potential. The iterative scheme requires reconstruction of the embedding potential with an updated density taken from the previous iteration and solving the many-body problem with the newly generated embedding potential. 16,17,22,23 This circumstance makes high-level ab initio methods unfavorable. Furthermore, the ρ A -dependency of the embedding potential leads to non- orthogonal wavefunctions within conventional FDET. In order to overcome ρ A -dependency and reduce the computational cost of conventional FDET, an approximation is usually made in which the embedding potential is evaluated at some 0021-9606/2016/144(20)/204103/14/$30.00 144, 204103-1 Published by AIP Publishing.

First time combination of frozen density embedding theory with … · 2016. 5. 26. · THE JOURNAL OF CHEMICAL PHYSICS 144, 204103 (2016) First time combination of frozen density

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • THE JOURNAL OF CHEMICAL PHYSICS 144, 204103 (2016)

    First time combination of frozen density embedding theorywith the algebraic diagrammatic construction schemefor the polarization propagator of second order

    Stefan Prager,1,a) Alexander Zech,2,b) Francesco Aquilante,3,c) Andreas Dreuw,1,d)and Tomasz A. Wesolowski2,e)1Interdisciplinary Center for Scientific Computing, Ruprecht-Karls University, Im Neuenheimer Feld 368,69120 Heidelberg, Germany2Dèpartement de Chimie Physique, Université de Genève, 30 Quai Ernest-Ansermet,1211 Genève 4, Switzerland3Dipartimento di Chimica “G. Ciamician,” Università di Bologna, Via Selmi 2, 40126 Bologna, Italy

    (Received 24 February 2016; accepted 25 April 2016; published online 24 May 2016)

    The combination of Frozen Density Embedding Theory (FDET) and the Algebraic DiagrammaticConstruction (ADC) scheme for the polarization propagator for describing environmental effectson electronically excited states is presented. Two different ways of interfacing and expressingthe so-called embedding operator are introduced. The resulting excited states are compared withsupermolecular calculations of the total system at the ADC(2) level of theory. Molecular test systemswere chosen to investigate molecule–environment interactions of varying strength from dispersioninteraction up to multiple hydrogen bonds. The overall difference between the supermolecular andthe FDE-ADC calculations in excitation energies is lower than 0.09 eV (max) and 0.032 eV inaverage, which is well below the intrinsic error of the ADC(2) method itself. Published by AIPPublishing. [http://dx.doi.org/10.1063/1.4948741]

    I. INTRODUCTION

    Almost all chemical reactions occur in liquid phaseand also almost all chemical properties are measured incondensed phases. In these cases, the investigated moleculeinteracts with its environment, leading to a change of itschemical properties. Electronically excited states, especially,can be strongly influenced by the environment. A largevariety of molecules and solvents exhibit this effect ofsolvatochromism.1 Hence, including the environment iscrucial for comparable calculations and reliable predictionsusing quantum chemical methods. However, including allenvironment molecules in quantum chemical calculationsexceeds nowadays computational limits. Different methodsfor considering the environment have been proposed, rangingfrom an implicit treatment as in polarizable continuum models(PCM)2 to explicit models treating the environment at amolecular mechanics level of theory (QM/MM).3 It is knownthat implicit approaches can only model dispersion andCoulombic interaction but may fail in describing specificsolvent interactions, for instance hydrogen bonds.4–6 In thesecases, an explicit model for the environment is needed.

    Density embedding methods7,8 constitute such an explicitmodel, and have received increasing attention in recent years(for the latest reviews see Refs. 9 and 10). Frozen-DensityEmbedding Theory (FDET)7,11 in particular provides a formal

    a)Electronic mail: [email protected])Electronic mail: [email protected])Electronic mail: [email protected])Electronic mail: [email protected])Electronic mail: [email protected]

    framework in which the whole system is described by meansof two independent quantities: the embedded wavefunction(interacting or not) and the density associated with theenvironment. The latter can in principle be taken from lowerlevel quantum mechanical methods, statistical theories forensembles, or even from experiment.

    While the original work used DFT12–14 methods todescribe the embedded species, also FDET variants based onwavefunction methods have been developed.11,15–17 Carter andcollaborators were the first to combine the FDET embeddingpotential with wavefunction based methods in their approachin which the independent variables are the total density andthe density of the embedded species.8,18 This combination isespecially useful for the calculation of excited states whereDFT methods have known limitations.19–21

    The embedding is accomplished by means of anembedding potential which depends on the density ofthe embedded species ρA(r⃗). In the canonical form ofFDET, herein referred to as conventional FDET, an iterativescheme is applied to obtain a wavefunction which is self-consistent with respect to the embedding potential. Theiterative scheme requires reconstruction of the embeddingpotential with an updated density taken from the previousiteration and solving the many-body problem with the newlygenerated embedding potential.16,17,22,23 This circumstancemakes high-level ab initio methods unfavorable. Furthermore,the ρA-dependency of the embedding potential leads to non-orthogonal wavefunctions within conventional FDET. In orderto overcome ρA-dependency and reduce the computationalcost of conventional FDET, an approximation is usuallymade in which the embedding potential is evaluated at some

    0021-9606/2016/144(20)/204103/14/$30.00 144, 204103-1 Published by AIP Publishing.

    http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741http://dx.doi.org/10.1063/1.4948741mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]

  • 204103-2 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    fixed density ρrefA(r⃗).16,23–26 Recently an alternative method,

    linearized FDET, has been proposed, which is based on a fixeddensity ρref

    A(r⃗) for the embedding potential, but still provides

    consistency between wavefunction and energy.27,28

    The algebraic diagrammatic construction (ADC) for thepolarization propagator is a robust and accurate method for thecalculation of excited states.29 Additionally, its mathematicalstructure is well suited to include external one-particleoperators since it is based on a non-iterative ground statemethod, making it straight forward to include environmentalmodels in ADC. ADC is size-consistent, Hermitian andsystematically improvable.30 So far, including environmentaleffects into an ADC calculation has been accomplished bymeans of a polarizable continuum model (PCM)6,31 andthrough QM/MM calculations.32

    In this article, we report a multiscale variant using FDETin combination with ADC. After giving a brief introductionto ADC and FDET in Sec. II, we outline the computationaldetails in Sec. III. The results of our test calculations arepresented in Sec. IV. A discussion and conclusion follow inSec. V.

    II. THEORETICAL METHODOLOGY

    A. Algebraic diagrammatic construction schemefor the polarization propagator of secondorder (ADC(2))

    A full derivation of the algebraic diagrammatic construc-tion scheme for the polarization propagator (ADC) isgiven in the literature.30,33,34 ADC can be derived usingthe intermediate state (IS) formalism. In this formalism, acomplete set of orthonormalized intermediate states

    �Ψ̃I

    �is

    constructed by applying creation and annihilation operatorsĈI =∈

    ĉ†aĉi; ĉ

    †aĉ†bĉiĉj; . . .

    onto the exact ground state

    wavefunction followed by Gram-Schmidt orthogonalization.In this intermediate state representation (ISR), the excited statewavefunction can be expressed in the basis of the intermediatestates according to

    Ψn =J

    XnJ�Ψ̃J

    �, (1)

    with XnJ as the eigenvectors of the ADC matrix, which isbuilt as a matrix representation of the Hamiltonian shifted bythe exact ground state energy E0,

    MI J =

    Ψ̃I

    �Ĥ − E0

    �Ψ̃J

    �. (2)

    This leads to the corresponding Hermitian eigenvalue problem,

    MX = ΩX, X†X = 1, (3)

    with X as the eigenvector matrix andΩ as the diagonal matrixof the eigenvalues which correspond to the exact excitationenergies,

    Ωn = En − E0. (4)

    This eigenvalue problem can be solved by diagonalizationof the ADC matrix M using typical iterative diagonalizationschemes like the Davidson algorithm.35 Since neither theexact ground state wavefunction nor the exact ground state

    energy are known, they are replaced by the correlated Møller-Plesset wavefunction and energy, respectively. Applying the ISformalism to 2nd order perturbation theory (MP(2)) generatesthe 2nd order ADC equations (ADC(2)).

    B. Frozen density embedding theory (FDET)

    This section will deal with the key features ofFrozen Density Embedding Theory (FDET) with respect towavefunction embedding, while a more detailed view on theFDET formalism used in this study is given in Refs. 11 and 28.

    In FDET, the supersystem is divided into the embeddedspecies (A) and the environment (B). Although FDET holds forall possible choices for ρA and ρB, the practical FDET basedmethods, which hinge on approximate density functionalfor the embedding potential, are designed for systems withnon-covalently bound environments like systems amenablefor QM/MM type of embedding methods. If charge transferbetween subsystems cannot be excluded then modeling suchshould be made by means of neither FDET nor QM/MM typeof methods. In this particular variant of FDET, a wavefunctionis used to describe the electronic state of the embeddedspecies while the electron density ρB(r⃗) is the descriptor forthe environment. Thus, the total energy of the supersystem(AB) takes the form of a functional EEWF

    AB[ΨA, ρB] depending

    on the embedded wavefunction (EWF) and the environmentdensity. Stationary wavefunctions are determined through theEuler-Lagrange equation

    δEEWFAB

    [ΨA, ρB]δΨ

    A

    − λΨA = 0, (5)

    where λ is the Lagrange multiplier referring to thenormalization. While Eq. (5) holds rigorously for theelectronic ground state, other than ground-state solutionscan be associated with excited states on the basis of theLevy-Perdew theorem36 as noticed by Khait and Hoffmann.37

    The total energy of the supersystem follows as

    EEWFAB�ΨA, ρB

    �=

    ΨA

    �ĤA

    �ΨA

    �+ V nucB

    �ρA

    + Jint�ρA, ρB

    �+ Enadxc,T

    �ρA, ρB

    + EHKvB�ρB

    �+ V nucA

    �ρB

    �, (6)

    where the following terms arise from the interaction of thetwo subsystems:

    Jint�ρA, ρB

    �=

    ρA(r⃗)ρB(r⃗ ′)�

    r⃗ − r⃗ ′� dr⃗dr⃗ ′, (7)

    V nucA�ρB

    �=

    ρB(r⃗)vA(r⃗)dr⃗ , (8)

    V nucB�ρA

    �=

    ρA(r⃗)vB(r⃗)dr⃗ , (9)

    Enadxc,T�ρA, ρB

    �= Enadxc

    �ρA, ρB

    �+ Tnads

    �ρA, ρB

    +∆FMD�ρA

    �. (10)

    The first three terms describe classical electrostatic interac-tions while Enadxc,T

    �ρA, ρB

    �comprises terms representing non-

    additivity of the exchange-correlation and kinetic densityfunctionals and a ρA dependent functional ∆FMD

    �ρA

  • 204103-3 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    (usually neglected in practice) depending on the form ofthe wavefunction used to solve Eq. (5). The non-additiveenergy bifunctional is generally defined as

    Enad�ρA, ρB

    �= E

    �ρA + ρB

    �− E

    �ρA

    �− E

    �ρB

    �. (11)

    The necessary condition for the stationary wavefunction tosatisfy is the following Schrödinger-like equation:

    �ĤA + v̂emb

    �ΨA = ϵΨA, (12)

    where v̂emb is the embedding operator describing the effectsof the environment. This operator is determined uniquely bythree charge densities (ρA(r⃗), ρB(r⃗), and that generated bythe nuclei of the environment (equivalently represented by theelectrostatic potential vB(r⃗))). It is thus the functional of thesethree functions. It reads:

    vemb[ρA, ρB, vB](r⃗)

    = vB(r⃗) +

    ρB(r⃗ ′)�r⃗ − r⃗ ′

    �dr⃗ ′ +δEnadxc,T[ρA, ρB]

    δρA

    . (13)

    This potential is obtained as the functional derivative of thecorresponding terms in the total energy functional with respectto ρA(r⃗).

    In this study, a particular version of FDET is usedwhere the non-additive energy functional Enadxc,T[ρA, ρB] isapproximated by a functional which is linear in ρA(r⃗). Theapproximation is constructed as a Taylor-expansion of thenon-additive energy functional at a reference density ρref

    A(r⃗)

    with the series being truncated after the linear term,

    Enadxc,T[ρA, ρB] ≈ Enadxc,T[ρrefA , ρB]

    +

    (ρA(r⃗) − ρrefA (r⃗)

    ) δEnadxc,T[ρrefA , ρB]δρ

    refA(r⃗) dr⃗

    = Enadxc,T[ρrefA , ρB] + ∆lin[ρA, ρrefA , ρB]. (14)From here on, we will refer to this variant of FDET

    as linearized FDET. The expression for the total energy inlinearized FDET is

    EL−FDETAB [ΨA, ρB, ρrefA ] =

    ΨA

    �ĤA

    �ΨA

    �+ V nucB

    �ρA

    + Jint�ρA, ρB

    �+ Enadxc,T[ρrefA , ρB]

    +∆lin[ρA, ρrefA , ρB]+ EHKvB

    �ρB

    �+ V nucA

    �ρB

    �. (15)

    The embedding potential obtained as functional derivative ofthe corresponding terms in the total energy functional givenin Eq. (15) reads

    vemb[ρrefA , ρB, vB](r⃗)

    = vB(r⃗) +

    ρB(r⃗ ′)�r⃗ − r⃗ ′

    �dr⃗ ′ +δEnadxc,T[ρrefA , ρB]

    δρrefA

    . (16)

    The potential will be denoted as v linemb(r⃗) = vemb[ρref

    A, ρB, vB](r⃗). Note that although both the potentials that

    are given in Eqs. (13) and (16) are functionals of three chargedensities, the latter one does not depend on the electronic stateof the embedded system.

    Since the embedding potential is added to the 1-particleHamiltonian, all electrostatic and part of the non-electrostatic

    interaction terms are already included when evaluating theexpectation value of the Hamiltonian. In order to arrive at thesame expression as Eq. (15), only constant, state-independentterms and the DFT energy of system B need to be added,

    ⟨ΨA|v̂emb|ΨA⟩ = V nucB�ρA

    �+ Jint

    �ρA, ρB

    +

    ρA(r⃗)

    δEnadxc,T[ρrefA , ρB]δρ

    refA(r⃗) dr⃗ . (17)

    Therefore, the energy difference between two states I and J isevaluated as the difference of two self-consistent expressionsfor the total energy given in Eq. (15) and reads (for the fullderivation see Zech et al.):28

    ∆EI J = ⟨ΨJA|ĤA + v̂ linemb|ΨJA⟩ − ⟨ΨIA|ĤA + v̂ linemb|ΨIA⟩. (18)Benchmark calculations on linearized FDET show that thisapproximation yields only negligible errors in excitationenergies compared to conventional FDET. Even in states withlarge differences in electron densities between ground andexcited state as in charge-transfer states, the approximation isvalid.28

    III. COMPUTATIONAL METHODOLOGY

    In general, the implementation of the FDE-ADC methodused in this work consists of four separate steps: the individualcalculations of the density matrices of A and B, the building ofthe embedding potential, and the final calculation applying theembedding potential on system A. In our implementation, thecore system (A) is calculated with Q-Chem38 using Møller-Plesset perturbation theory39,40 of second order (MP(2)) whilethe environment (B) is calculated either in Q-Chem or inMolcas41,42 using Hartree-Fock (HF)43 or density-functional-theory (DFT).12–14 Both density matrices of A and B areused to construct the embedding potential within Molcas.The embedding potential subsequently serves as input forQ-Chem and is added to the 1-particle Hamiltonian used inthe HF calculation,

    Õ1 = O1 + vemb[ρrefA , ρB]. (19)The orbitals and orbital energies obtained from the modifiedHF calculation include the influence of the environment andserve as input for a subsequent ADC(2) calculation. Theresulting excited states are influenced by the environmentby virtue of the modified orbitals and orbital energies.For applying this method, no modifications to the originalADC equations need to be made. Furthermore, the excitationenergies are a direct result (cf. Eq. (18)), i.e., no furthercorrections have to be applied. The combination of thesetwo steps, the HF calculation including the embeddingpotential and the following ADC(2) calculation, are furtherreferred to as FDE-ADC(2) calculation. Currently, twodifferent approaches are implemented to perform FDE-ADC calculations, which we call supermolecular expansion(SE) and reassembling of density matrix (RADM), whichdistinguish the different numerical procedures to perform thecontraction of the required density matrices.

  • 204103-4 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    In SE, the calculations of both systems A and B andthe final FDE-ADC(2) calculation are performed in the basisfunctions of A + B. Although this is not very efficient, itprovides an easy way to analyze this method in detail as aproof of principle, since there is no basis-set superpositionerror (BSSE) included.

    In our newly developed approach called RADM, thecalculation of A is split. A is calculated in the monomerbasis on MP(2) level of theory and additionally in thesupermolecular basis on HF level of theory, obtaining twoseparate density matrices. A new density matrix in the basisof A and B is generated by adding the HF-to-MP(2) differencedensity matrix in the basis of A to the AA block of theHF density matrix in the basis of A + B. The embeddingpotential is preliminary calculated in the basis of A + B andtruncated afterwards to the basis of A. This approximation canbe made because the off-diagonal blocks AB and BA in theembedding potential matrix are almost zero and the BB blockhardly contributes to A since the density of A is vanishinglysmall on the basis of B. The final FDE-ADC(2) calculationincluding the embedding potential, i.e., the computationallymost demanding step, is performed in the basis of A only(Fig. 1).

    For the FDE-ADC calculations, a development versionof Q-Chem based on version 4.3 and a developmentversion of Molcas based on version 8.0 have been used.Molecular pictures were captured using Avogadro 1.1.0.44

    Unless otherwise indicated, all calculations have been carriedout at MP(2)/cc-pVDZ45 and ADC(2)/cc-pVDZ level oftheory for system A and HF/cc-pVDZ level of theory forsystem B.

    IV. RESULTS

    FDE-ADC(2) was tested for four different systems withvarious environments of different interaction strengths. Inall benchmarks, the supersystem is optimized at MP(2)/cc-pVDZ level of theory and the five lowest electronicallyexcited states are calculated at ADC(2)/cc-pVDZ level of

    theory. Additionally, the five lowest excited states of theisolated system A were calculated at ADC(2)/cc-pVDZ levelof theory without re-optimizing the geometry. In this way,the electronic interactions between the core system andenvironment can be analyzed directly without the influenceof geometry changes due to environment interactions. TheFDE-ADC(2) calculations were carried out using the sameoptimized geometries for system A and B as in the supersystemcalculation. For all investigations, the singlet multiplicityand neutral charge were conserved. The test systems werechosen with respect to various kinds of system-environmentinteractions. Benzene with hydrogen fluoride (Sec. IV A 1)in two different orientations was chosen as a system withonly weak dispersion interactions and polarization of anaromatic π-system. Benzaldehyde with two water molecules(Sec. IV A 2) forming a water dimer is an example forsystems with a hydrogen bond between embedded speciesand environment in addition to a polarization of the π-system.The last system, uracil with five water molecules (Sec. IV A 3),shows the strongest interactions because all water moleculesare involved in hydrogen bonds with the embedded species.Transitions involving the electron lone pairs are expected tobe more strongly shifted than typical π → π∗ states due to theenvironment interactions.

    A. Results of FDE-ADC(2) usingthe supermolecular expansion

    In the first test, the supermolecular expansion was used.This approach is the mathematically exact implementation ofFDE-ADC without any further approximation and withoutrestrictions to the basis set. Although no benefit incomputational cost with respect to the supersystem calculationcan be achieved, this serves as a benchmark for furtherapproximate FDE-ADC approaches.

    1. Benzene with one hydrogen fluoride molecule

    As a first test system, a hydrogen fluoride molecule isplaced apical to the benzene ring, almost along the C6-axis

    FIG. 1. Schematic representation ofthe reassembling of density matrix(RADM) method. In the upper line, theisolated system A is calculated and inthe lower line, the isolated system B.In system A, the HF (grey) density ma-trix on the basis functions of A and Bis combined with the HF/MP(2) (red)difference density matrix on the basisfunctions of A to build the RADM ma-trix. In combination, the embedding po-tential (yellow) is created and truncatedto the elements in the basis of A. Fi-nally, FDE-ADC(2) (blue) calculationsare performed in the basis of A only.

  • 204103-5 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    FIG. 2. Molecular structures of the two benzene-hydrogen fluoride supersys-tems. Benzene with apical hydrogen fluoride (left) and side-on (right).

    with the hydrogen pointing towards the benzene (Fig. 2 left).The distance between the plane of the benzene ring and thehydrogen atom is 2.161 Å.

    The frontier orbitals of benzene, i.e., the highest occupiedmolecular orbital (HOMO), HOMO-1, the lowest unoccupiedmolecular orbital (LUMO) and LUMO+1 are the typical πorbitals. The lower lying occupied orbitals (HOMO-2 andHOMO-3) show σ-character while the higher unoccupiedorbitals (LUMO+2 and LUMO+3) can be described asRydberg-orbitals. The analysis of the five lowest electronicallyexcited states of isolated benzene shows four locally excitedstates characterized by π → π∗ transitions and an energeticallyhigher lying Rydberg state. The five lowest excited states andtheir character are given in Table I.

    In the supermolecule calculations of benzene togetherwith hydrogen fluoride, the energetic order of the frontierorbitals as well as the energetic order and the characterof the excited states are changed due to the influence of

    the environment and as the fifth excited state a σ → π∗transition arises. The HOMO and HOMO-1 switch theirenergetic order but since they are almost degenerate, it haspractically no influence. However, this change leads to adifferent orbital transition pattern in the characterization ofthe excited states. A HOMO → LUMO transition in theisolated benzene corresponds now to a HOMO-1 → LUMOtransition in the C6H6-HF supermolecular complex. Thefive lowest excited states are listed and characterized inTable I.

    In the FDE-ADC calculation, the results of thesupersystem calculation are nicely reproduced. The lowestfour excited states are in almost perfect agreement with thecalculation of the supersystem regarding excitation energies,oscillator strengths and orbital transitions. The 3rd excitedstate is marginally shifted to higher excitation energies insteadof slightly lower excitation energies, as the supersystemcalculation would infer. The 5th and 6th excited state changedthe energetic order in FDE-ADC compared to the supersystemcalculation because of the small energy difference betweenthese excited states of only 0.02 eV (Table I). For the sakeof comparison, the S6 is used and labeled as S5 according tothe energetic order of this specific state in the supersystemcalculation.

    The difference of the excitation energies and the oscillatorstrengths of the isolated system and the FDE-ADC calculationto the supersystem calculation is shown in Fig. 3. With amaximum deviation of 0.03 eV in excitation energies for thissystem, the polarization of the π-system is reproduced almostquantitatively.

    TABLE I. Excitation energies, oscillator strengths, and orbital transitions for the five energetically lowest electronically excited singlet states of isolatedbenzene, the supersystem benzene with apical hydrogen fluoride, and the FDE-ADC calculations in the supermolecular expansion of benzene (A) and hydrogenfluoride (B).

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    Weight (%)

    State Isol. Supersys. FDE-ADC Isol. Supersys. FDE-ADC H+x → L+ya Isol. Supersys. FDE-ADC

    S1 5.287 5.323 5.328 0.0000 0.0001 0.0001 0 → 0 (π → π∗) 8.3 45.2 45.10 → 1 (π → π∗) 36.6 . . . . . .−1 → 0 (π → π∗) 36.6 . . . . . .−1 → 1 (π → π∗) 8.2 43.6 43.5

    S2 6.677 6.692 6.704 0.0000 0.0000 0.0000 0 → 0 (π → π∗) 37.9 . . . . . .0 → 1 (π → π∗) 8.5 45.0 45.6−1 → 0 (π → π∗) 8.2 46.2 45.4−1 → 1 (π → π∗) 37.2 . . . . . .

    S3 7.520 7.500 7.530 0.7155 0.6810 0.7000 0 → 0 (π → π∗) 18.3 . . . . . .0 → 1 (π → π∗) 25.8 44.1 43.4−1 → 0 (π → π∗) 25.7 43.0 43.6−1 → 1 (π → π∗) 18.5 . . . . . .

    S4 7.522 7.527 7.548 0.7162 0.6867 0.7057 0 → 0 (π → π∗) 25.5 42.6 42.60 → 1 (π → π∗) 18.3 . . . . . .−1 → 0 (π → π∗) 18.4 . . . . . .−1 → 1 (π → π∗) 26.0 44.3 44.3

    S5 8.131 8.298 8.317 0.0000 0.0005 0.0003 0 → 2 (π → R) 90.2 . . . . . .−2 → 0 (σ → π∗) . . . 69.4 64.7b−3 → 1 (σ → π∗) . . . 14.3 18.3b

    aH=HOMO, L=LUMO, R=Rydberg-type orbital.bOccurs as S6 in the original FDE-ADC(2) calculation.

  • 204103-6 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    FIG. 3. Excitation energy difference (left) and oscillator strength difference(right) of the isolated benzene (blue) to the supersystem C6H6–HFapical andthe FDE-ADC(2) calculation (red) to the supersystem. The blue bars cor-respond to the influence of the environment on the excitation energies andoscillator strengths (solvatochromic shift) while the red bars indicate theaccuracy of the FDE-ADC(2) calculation in comparison with the ADC(2)calculation of the full system.

    In a second geometrical arrangement, the hydrogenfluoride is placed side-on and in plane with the benzene(Fig. 2 right). The geometry of benzene is slightly distortedby the hydrogen fluoride and differs therefore marginallyfrom the geometry used in the system above. The Hbenzene–Fdistance is 2.074 Å. However, the frontier orbitals are visuallyindistinguishable to the orbitals of the unperturbed benzeneor the orbitals of benzene of the previous system due tothe weak system-environment interaction. The five lowest

    electronically excited states consist of four π → π∗ transitionsand one Rydberg state as the S5 state.

    Calculating the supersystem, the frontier orbitals ofbenzene are almost identical to the frontier orbitals of theisolated benzene and only slightly distorted by the hydrogenfluoride. Only the LUMO+2, which is a Rydberg orbital in thecase of the isolated benzene, is now localized on the hydrogenfluoride and the original Rydberg orbital located at the benzeneis the LUMO+3. The excited states are more influenced bythe hydrogen fluoride in plane with the benzene ring thanperpendicular to it. The S1 to S4 states are still characterizedas local π → π∗ transitions on the benzene while the S5 state isa mixed Rydberg state delocalized over benzene and hydrogenfluoride.

    In the FDE-ADC calculation, a higher lying virtualRydberg orbital is lowered and becomes the LUMO, but thisorbital does not contribute to any local π → π∗ transitions.Since the energy of the higher lying unoccupied virtualorbitals (LUMO+1 to LUMO+10) is practically identicalto the supersystem calculation, the character of the excitedstates remains the same even if the number of the electronaccepting orbital is increased by 1. The excited states S3 andS4 are, as in the supersystem calculation, degenerate. Theexcited states of the supersystem and the isolated benzene aswell as the FDE-ADC calculations are characterized in detailin Table II.

    The influence of the hydrogen fluoride molecule onto thebenzene while it is located in-plane with the benzene-ringis reproduced almost exactly by the FDE-ADC calculations.The largest deviation in the excitation energies is lower than0.02 eV for the Rydberg state and lower than 0.005 for theπ → π∗ transitions. The differences in the excitation energiesand oscillator strengths between isolated benzene and the

    TABLE II. Excitation energies, oscillator strengths, and orbital transitions for the five energetically lowest electronically excited singlet states of isolatedbenzene, the supersystem benzene with side-on hydrogen fluoride, and the FDE-ADC calculations in the supermolecular expansion of benzene (A) and hydrogenfluoride (B).

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    Weight (%)

    State Isol. Supersys. FDE-ADC Isol. Supersys. FDE-ADC H+x → L+ya Isol. Supersys. FDE-ADC

    S1 5.287 5.435 5.435 0.0000 0.0001 0.0001 0 → 0 (π → π∗) 44.8 45.9 45.9b−1 → 1 (π → π∗) 44.8 43.8 43.8b

    S2 6.677 6.791 6.792 0.0000 0.0000 0.0000 0 → 1 (π → π∗) 46.0 46.3 46.1b−1 → 0 (π → π∗) 46.0 45.9 46.2b

    S3 7.521 7.659 7.660 0.7155 0.7447 0.7295 0 → 0 (π → π∗) 7.5 . . . 43.2b0 → 1 (π → π∗) 36.7 44.0 . . .−1 → 0 (π → π∗) 36.7 44.3 . . .−1 → 1 (π → π∗) 7.5 . . . 45.3b

    S4 7.522 7.659 7.664 0.7162 0.7262 0.7200 0 → 0 (π → π∗) 36.7 43.1 . . .0 → 1 (π → π∗) 7.5 . . . 44.3b−1 → 0 (π → π∗) 7.5 . . . 44.2b−1 → 1 (π → π∗) 36.8 45.2 . . .

    S5 8.131 8.177 8.163 0.0000 0.0004 0.0003 0 → 2 (π → R) 90.7 26.5 82.8b0 → 3 (π → R) . . . 63.6 . . .0 → 0 (π → π∗) . . . . . . 7.5

    aH=HOMO, L=LUMO, R=Rydberg-type orbital.bThe low-lying Rydberg orbital is neglected in counting, i.e., LUMO indices reduced by 1.

  • 204103-7 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    FIG. 4. Excitation energy difference (left) and oscillator strength difference(right) of the isolated benzene (blue) to the supersystem C6H6–HFside-onand the FDE-ADC(2) calculation (red) to the supersystem. The blue barscorrespond to the influence of the environment on the excitation energiesand oscillator strengths (solvatochromic shift) while the red bars indicate theaccuracy of the FDE-ADC(2) calculation in comparison with the ADC(2)calculation of the full system.

    FDE-ADC calculation to the supersystem calculation areshown in Fig. 4.

    2. Benzaldehyde with two water molecules

    Benzaldehyde is able to form hydrogen bonds with a polarprotic solvent. Because of this ability to act as a hydrogenacceptor, it was chosen together with two water molecules asa model system (Fig. 5). Because of the hydrogen bond withinthe water dimer, the strength of the hydrogen bond from thewater dimer to benzaldehyde is increased.

    FIG. 5. Molecular structure of the test system of benzaldehyde with twowater molecules.

    The frontier orbitals of isolated benzaldehyde optimizedas supersystem are characterized as follows: HOMO-3: π,HOMO-2: n, HOMO-1: π, HOMO: π, LUMO: π∗, LUMO+1:π∗, LUMO+2: Rydberg, LUMO+3: Rydberg, LUMO+4: π∗.Analysis of the five energetically lowest electronically excitedstates shows two states with n → π∗ character (S1 and S4) and3 π → π∗ transitions. A detailed characterization is given inTable III.

    Going to the supersystem, the frontier orbitals are hardlychanged but the HOMO and HOMO-1 change their energeticorder. Since these two orbitals are almost degenerate it canbe considered as negligible. However, this change has to beconsidered in the characterization of the excited states. Incontrast, the excited states are influenced significantly by thewater environment. As expected, the hydrogen-bonded watermolecule stabilizes the n-orbitals leading to a large increase

    TABLE III. Excitation energies, oscillator strengths, and orbital transitions for the five energetically lowest electronically excited singlet states of isolatedbenzaldehyde, the supersystem benzaldehyde with a hydrogen-bonded water dimer, and the FDE-ADC calculations in the supermolecular expansion ofbenzaldehyde (A) and the water dimer (B).

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    Weight (%)

    State Isol. Supersys. FDE-ADC Isol. Supersys. FDE-ADC H+x → L+ya Isol. Supersys. FDE-ADC

    S1 3.692 3.897 3.891 0.0001 0.0001 0.0001 −2 → 0 (n → π∗) 73.9 74.3 73.9−2 → 4 (n → π∗) 13.2 9.1 . . .−2 → 5 (n → R) . . . . . . 6.9

    S2 4.933 4.813 4.837 0.0104 0.0197 0.0174 0 → 0 (π → π∗) 18.5 61.9 59.10 → 1 (π → π∗) 20.5 . . . . . .−1 → 0 (π → π∗) 41.0 4.3 6.0−1 → 1 (π → π∗) 7.9 19.8 19.1

    S3 5.771 5.554 5.595 0.3258 0.3444 0.3519 0 → 0 (π → π∗) 60.9 7.3 9.8−1 → 0 (π → π∗) 22.2 76.1 73.5

    S4 6.272 6.751 6.742 0.0000 0.0059 0.0001 −2 → 1 (n → π∗) 83.6 82.1 81.4S5 6.884 6.753 6.784 0.3684 0.2942 0.2908 0 → 0 (π → π∗) 5.3 17.7 18.4

    0 → 1 (π → π∗) 41.9 5.6 7.1−1 → 0 (π → π∗) 23.5 . . . . . .−1 → 1 (π → π∗) 12.4 50.3 47.7

    0 → 4 (π → π∗) . . . 6.0 . . .aH=HOMO, L=LUMO, R=Rydberg-type orbital.

  • 204103-8 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    of the excitation energies of the corresponding n → π∗ states.Simultaneously, the excitation energies of the π → π∗ statesare reduced. However, the energetic order and the characterof the states are not changed by the environment (Table III).

    In the FDE-ADC calculation, the interactions of thehydrogen bonds are simulated almost quantitatively. Thedegeneracy of the HOMO and HOMO-1 is retained andalso the n-orbitals are correctly described. The excited statesin the FDE-ADC calculation show the same influence of theenvironment on the n → π∗ as well as on the π → π∗ states asseen in the supersystem calculation. The full characterizationis given in Table III.

    In this example, the largest deviation in the excitationenergies is lower than 0.01 eV for the n → π∗ states andlower than 0.05 eV for the π → π∗ transitions. Especiallythe reproduction of the influence of the hydrogen bonds onthe excited states is remarkable, since hydrogen bonds havelarger orbital interactions compared to dispersion interactionor polarization. The difference in excitation energies andoscillator strengths of the isolated benzaldehyde and theFDE-ADC(2) calculation to the supersystem calculation aresummarized in Fig. 6.

    Additionally, the benzaldehyde system was used as anexample system for testing the basis set dependence ofFDE-ADC(2). The same calculations as described above havebeen carried out using the cc-pVTZ basis set. Generally,the excitation energies are lowered due to the larger basis set.This is observed consistently in the supersystem calculation aswell as in the FDE-ADC calculation. Both in the supersystemcalculation and the FDE-ADC calculation, the S4 and S5state change their energetic order with respect to the isolatedbenzaldehyde. This change is reproduced very nicely in theFDE approach (see Table IV).

    Calculating the difference of the deviation in excitationenergies obtained with FDE-ADC and the supersystem

    FIG. 6. Excitation energy difference (left) and oscillator strength difference(right) of the isolated benzaldehyde (blue) to the supersystem of benzalde-hyde with a water dimer and the FDE-ADC(2) calculation (red) to the super-system. The blue bars correspond to the influence of the environment on theexcitation energies and oscillator strengths (solvatochromic shift) while thered bars indicate the accuracy of the FDE-ADC(2) calculation in comparisonwith the ADC(2) calculation of the full system.

    TABLE IV. Excitation energies and oscillator strengths for the FDE-ADC(2)/cc-pVTZ calculation in the SE approach using benzaldehyde assystem A and two water molecules as system B.

    Exc. energies (eV) Osc. strength

    State Isol. Supersys. FDE-ADC Isol. Supersys. FDE-ADC

    S1 3.623 3.827 3.820 0.0001 0.0002 0.0001S2 4.810 4.685 4.715 0.0127 0.0220 0.0188S3 5.549 5.350 5.399 0.3260 0.3378 0.3474S4 6.129 6.530 6.565 0.0000 0.2975 0.2876S5 6.645 6.599 6.587 0.3583 0.0001 0.0001

    calculation in the two basis sets according to

    ∆Ωbasis =���(Ω

    cc-pVDZFDE-ADC −Ω

    cc-pVDZsupersystem

    )−(Ω

    cc-pVTZFDE-ADC −Ω

    cc-pVTZsupersystem

    ) ��� , (20)with ٠as the excitation energy exhibits an error smallerthan 0.008 eV in all investigated electronically excited states.Hence, the deviation of FDE-ADC(2) from the supersystemcalculation does not vary with the larger basis set. It occursthat the error of the FDE-ADC(2) method compared to thesupersystem calculation is largely independent of the basis setsize. Certainly, the results of the ADC(2) and FDE-ADC(2)calculations in general are affected using a larger basis set.

    3. Uracil with five water molecules

    Uracil was chosen as a test system, because it has anaromatic π system and in addition two hydrogen bond donors(B and D in Fig. 7) and two oxygen atoms as hydrogenbond acceptors (A and C), thus providing a large variety ofenvironment interactions. However, uracil has also a moderate

    FIG. 7. Geometrical arrangement of the supersystem of uracil and five watermolecules.

  • 204103-9 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    size to allow benchmark calculations of the supersystemin reasonable time. The geometrical arrangement of thesupersystem is shown in Fig. 7.

    The frontier orbitals of uracil, optimized in thesupersystem, show π character for the HOMO-1 and HOMO,while the HOMO-2 and HOMO-3 are n-orbitals and localizedmainly at the oxygen atom OA and OC, respectively. TheLUMO and LUMO+2 exhibit π∗ character, the LUMO+1 isa Rydberg orbital. Analyzing the five lowest electronicallyexcited states of the isolated uracil, optimized in thesupersystem structure, n → π∗ transitions are observed asthe S1, S3 and S5 states. The S2 and S4 states are local π → π∗transitions. The detailed characterization of the five lowestexcited states is given in Table V.

    In the supersystem structure, the water molecules interactstrongly with the uracil forming hydrogen bonds. Two watermolecules act as hydrogen bond donors for the OA, one asan acceptor for NB, forming a hydrogen bond chain overone of the previously mentioned water molecules to OA.Another hydrogen bond chain is formed from ND over twowater molecules to OC. These hydrogen bonds will influencethe n-orbitals and hence also the n → π∗ excitations butwill do so differently for each state since the interactionis more pronounced at the OA than at the OC. The rest ofthe frontier orbitals are qualitatively unchanged except forthe LUMO+1 and LUMO+2, which change their energeticorder. Analyzing the five lowest electronically excited statesof the supersystem, the excitation energy of the S1 state is, asexpected, largely increased due to the stabilizing effects of theenvironment on the n-orbitals. Accordingly, the influenceon the two other n → π∗ transitions is weaker but still

    significant. However, some orbitals delocalized over uraciland one or more water molecules contribute only slightly tothe excitation. A detailed description of the excited states isgiven in Table V.

    In the FDE-ADC calculations, the uracil was consideredas system A while all 5 water molecules are used asthe environment B. As in the supersystem calculation, theenergetic order of the LUMO+1 and LUMO+2 is changedcompared to the isolated uracil calculation. The remainingfrontier orbitals are essentially identical to the orbitals ofisolated uracil, but higher lying virtual orbitals are largelydistorted accounting for the effect of the environment. Theresults of the FDE-ADC calculations are in very goodagreement with the supersystem benchmark. The largest shiftin excitation energy (S1) and the largest shift in oscillatorstrength (S4) are well reproduced. Also, the different influenceof the hydrogen bonds on the excited states is included inthe FDE-ADC calculation. The characterization of the excitedstates is given in Table V.

    For this system, the largest error is 0.09 eV for theexcitation energies and 0.02 for the oscillator strength,which is slightly higher than in the previous systems. Still,considering the strong interaction between uracil and the waterenvironment, this error is acceptable for this approximatetreatment of the environment. The slightly larger error arisesbecause some orbitals, localized on the water molecules ofthe environment, contribute slightly to the excitations inthe supersystem but are not considered in the FDE-ADCcalculation. Comparison of the FDE-ADC calculation and theisolated uracil calculation to the supersystem is presented inFig. 8.

    TABLE V. Excitation energies, oscillator strengths, and orbital transitions for the five energetically lowest electronically excited singlet states of isolated uracil,the supersystem uracil with five hydrogen-bonded water molecules, and the FDE-ADC calculations in the supermolecular expansion of uracil (A) and the fivewater molecules (B).

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    Weight (%)

    State Isol. Supersys. FDE-ADC Isol. Supersys. FDE-ADC H+x → L+ya Isol. Supersys. FDE-ADC

    S1 4.440 5.143 5.108 0.0001 0.0016 0.0007 −2 → 0 (n → π∗) 77.2 72.2 72.9−3 → 0 (n → π∗) . . . . . . 5.8−4 → 0 (n → π∗) . . . 4.6 . . .

    S2 5.439 5.215 5.291 0.2261 0.2371 0.2162 0 → 0 (π → π∗) 81.3 80.7 80.1S3 5.866 6.211 6.221 0.0000 0.0001 0.0002 −2 → 0 (n → π∗) 4.6 . . . 5.8

    −2 → 1 (n → π∗) 35.7b 20.0 . . .−3 → 0 (n → π∗) 28.3 17.4 31.0−3 → 1 (n → π∗) 16.4b 10.8 . . .−4 → 0 (n → π∗) . . . 20.2 . . .−4 → 1 (n → π∗) . . . 5.6 . . .−2 → 6 (n → R) . . . . . . 24.8−3 → 6 (n → R) . . . . . . 13.6

    S4 6.194 6.292 6.320 0.0311 0.0963 0.0850 −1 → 0 (π → π∗) 85.9 84.4 84.1S5 6.468 6.657 6.749 0.0001 0.0005 0.0021 −2 → 1 (n → π∗) 38.0b 22.8 . . .

    −3 → 0 (n → π∗) 44.6 15.3 45.9−3 → 1 (n → π∗) . . . 12.4 . . .−4 → 0 (n → π∗) . . . 20.4 . . .−4 → 1 (n → π∗) . . . 6.6 . . .−2 → 6 (n → R) . . . . . . 19.1−3 → 6 (n → R) . . . . . . 9.2

    aH=HOMO, L=LUMO, R=Rydberg-type orbital.bOrbital index for LUMO+1 and LUMO+2 interchanged.

  • 204103-10 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    FIG. 8. Excitation energy difference (left) and oscillator strength difference(right) of the isolated uracil (blue) to the supersystem of uracil with five watermolecules and the FDE-ADC(2) calculation (red) to the supersystem. Theblue bars correspond to the influence of the environment on the excitationenergies and oscillator strengths (solvatochromic shift) while the red barsindicate the accuracy of the FDE-ADC(2) calculation in comparison with theADC(2) calculation of the full system.

    B. Results of FDE-ADC(2) using the reassemblingof density matrix approach

    In this chapter, the newly developed reassembling ofdensity matrix (RADM) approach was used within the FDE-ADC calculations. Using RADM, the density matrix forthe FDE-ADC(2) calculation is conducted within the basisfunctions of A only leading to a substantial decrease of thecomputational costs. The neglect of the embedding potentialon the basis functions on B induces, however, an additionalerror, which is also discussed in this section.

    As in the previous chapters, the FDE-ADC calculationis tested against the supersystem calculation and the isolatedsystem A. The systems benzene with one hydrogen fluorideside-on, benzaldehyde with two water molecules and uracilwith five water molecules were chosen as test systems.

    1. Benzene with one hydrogen fluoridemolecule side-on

    The FDE-ADC calculations on benzene with onehydrogen fluoride show almost identical orbitals comparedto the isolated benzene. Only the LUMO+2 Rydberg orbitalshows a distorted surface in the region pointing towardsthe hydogenfluoride molecule. In contrast to using thesupermolecular expansion for FDE-ADC (SE-FDE-ADC) nohigher lying virtual orbital is lowered to become the LUMO.HOMO-1, HOMO, LUMO and LUMO+1 exhibit π andπ∗-character, respectively. The character of the five lowestelectronically excited states is the same as in the supersystemcalculation. The S1 to S4 are local π → π∗ transitions, the S5 isa Rydberg state. As in the supersystem, the S3 and S4 states aredegenerate. In this case, the state corresponding to the S4 ofthe supersystem calculation shows a slightly lower excitationenergy and becomes the S3 state (Table VI).

    Using RADM, the excited states are in very goodagreement with the supersystem calculation and almostidentical to the SE-results. The differences of RADM-FDE-ADC to the supersystem are below 0.006 eV for the localπ → π∗ transitions and below 0.03 eV for the Rydberg state.In comparison to the SE-FDE-ADC(2) results, the error ofthe RADM approximation alone is smaller than 0.04 eVfor the excitation energy of the Rydberg state and almostzero (0.0001 eV) for the local π → π∗ excitation energies.In summary, the differences of SE-FDE-ADC, RADM-FDE-ADC and the isolated benzene calculations to the supersystemcalculation are collected in Fig. 9.

    2. Benzaldehyde with two water molecules

    Analogous to the SE calculations, benzaldehyde wasused as system A. The orbitals obtained in the RADM-FDE-ADC calculation are very similar to the orbitals ofthe isolated benzaldehyde and only slightly distorted. Thefrontier orbitals HOMO-1 to LUMO+1 and LUMO+4 showπ and π∗-symmetry, respectively. The HOMO-2 is an n-orbitallocalized at the oxygen and the LUMO+2 and LUMO+3 are

    TABLE VI. Excitation energies, oscillator strengths, and orbital transitions for the FDE-ADC calculation usingthe RADM approach for benzene as system A and one hydrogen fluoride molecule in plane as system B.

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    RADM ∆Ωa ∆Ωb f ∆ f a ∆ f b H+x → L+y Weight (%)

    S1 5.436 0.001 0.001 0.0001 0.0000 0.0000 0 → 0 (π → π∗) 46.0−1 → 1 (π → π∗) 43.8

    S2 6.792 0.001 0.000 0.0000 0.0000 0.0000 −1 → 0 (π → π∗) 46.20 → 1 (π → π∗) 46.0

    S3 7.661 0.002 0.001 0.7298 −0.0149 0.0003 0 → 0 (π → π∗) 43.2−1 → 1 (π → π∗) 45.3

    S4 7.665 0.006 0.001 0.7184 −0.0079 −0.0016 0 → 1 (π → π∗) 44.4−1 → 0 (π → π∗) 44.1

    S5 8.199 0.022 0.036 0.0006 0.0001 0.0003 0 → 2 (π → R) 90.3aReferenced to the supersystem calculation.bReferenced to the SE-FDE-ADC(2) calculation.

  • 204103-11 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    FIG. 9. Excitation energy difference (left) and oscillator strength difference(right) of the isolated benzene (blue) to the supersystem C6H6–HFside-on, theSE-FDE-ADC(2) calculation (red) to the supersystem, and the RADM-FDE-ADC(2) calculation (green) to the supersystem. The blue bars correspond tothe influence of the environment on the excitation energies and oscillatorstrengths (solvatochromic shift) while the red and green bars indicate theaccuracy of the SE-FDE-ADC(2) and RADM-FDE-ADC(2) calculations incomparison with the ADC(2) calculation of the full system. The SE-FDE-ADC(2) results are taken from Sec. IV A 1.

    Rydberg orbitals. The five lowest electronically excited statesare closely related to the excited states of the supersystemexhibiting a local π → π∗ transition in the excited states S2,S3, and S5 and an n → π∗ transition in the excited states S1and S4.

    Also for this test, the agreement of the RADM-FDE-ADC calculation with the supersystem calculation is verygood, but shows, as expected, slightly larger differences thanthe SE calculation. The difference in the excitation energiescompared to the supersystem calculation is below 0.08 eV forthe calculated excited states. The error induced by the RADM

    FIG. 10. Excitation energy difference (left) and oscillator strength difference(right) of the isolated benzaldehyde (blue) to the supersystem of benzalde-hyde with a water dimer, the SE-FDE-ADC(2) calculation (red) to the super-system, and the RADM-FDE-ADC(2) calculation (green) to the supersystem.The blue bars correspond to the influence of the environment on the excitationenergies and oscillator strengths (solvatochromic shift) while the red andgreen bars indicate the accuracy of the SE-FDE-ADC(2) and RADM-FDE-ADC(2) calculations in comparison with the ADC(2) calculation of the fullsystem. The SE-FDE-ADC(2) results are taken from Sec. IV A 2.

    approximation alone compared to the SE-FDE-ADC(2) resultsis smaller than 0.04 eV in excitation energies. A graphicalcomparison is presented in Fig. 10.

    Additionally, this system was benchmarked using thediffuse basis set aug-cc-pVDZ in combination with theRADM-FDE-ADC(2) method. Diffuse basis functions canbetter describe the peripheral regions of the systems, in whichthe densities overlap. This results in a better descriptionof the embedding potential and therefore in even smallererrors in excitation energies. Since diffuse basis functions areincluded, the orbitals as well as the orbital transitions of theexcited states differ both for the supersystem calculation and

    TABLE VII. Excitation energies, oscillator strengths and orbital transitions for the FDE-ADC calculation in theRADM approach using benzaldehyde as system A and two water molecules as system B.

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    RADM ∆Ωa ∆Ωb f ∆ f a ∆ f b H+x → L+y Weight (%)

    S1 3.877 −0.020 0.014 0.0001 0.0000 0.0000 −2 → 0 (n → π∗) 74.1−2 → 4 (n → π∗) 12.2

    S2 4.851 0.038 0.014 0.0164 −0.0033 −0.0010 0 → 0 (π → π∗) 58.0−1 → 0 (π → π∗) 6.6−1 → 1 (π → π∗) 20.4

    S3 5.628 0.075 0.033 0.3490 0.0046 −0.0029 0 → 0 (π → π∗) 10.8−1 → 0 (π → π∗) 72.6

    S4 6.731 −0.020 −0.011 0.0001 −0.0058 0.0000 −2 → 1 (n → π∗) 83.9S5 6.803 0.050 0.019 0.2979 0.0037 0.0071 0 → 0 (π → π∗) 18.6

    0 → 1 (π → π∗) 8.1−1 → 1 (π → π∗) 48.4

    aReferenced to the supersystem calculation.bReferenced to the SE-FDE-ADC(2) calculation.

  • 204103-12 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    TABLE VIII. Excitation energies and oscillator strengths for the FDE-ADC(2)/aug-cc-pVDZ calculation in the RADM approach using benzalde-hyde as system A and two water molecules as system B.

    Exc. energies (eV) Osc. strength

    State Isol. Supersys. FDE-ADC Isol. Supersys. FDE-ADC

    S1 3.562 3.765 3.761 0.0001 0.0002 0.0002S2 4.812 4.682 4.716 0.0132 0.0234 0.0194S3 5.486 5.282 5.338 0.3308 0.3393 0.3534S4 5.943 6.351 6.350 0.0001 0.0154 0.0019S5 6.042 6.403 6.397 0.0115 0.0002 0.0001

    the RADM-FDE-ADC(2) calculation compared to the resultsobtained with the cc-pVDZ basis set. Applying the diffusebasis set aug-cc-pVDZ, the error of the RADM-FDE-ADC(2)calculation compared to the supersystem calculation is smallerfor all excitation energies in this system compared to theresults obtained with the cc-pVDZ basis set. (See Tables VIIand VIII.)

    3. Uracil with five water molecules

    In the strongly interacting system of uracil with five watermolecules, the orbitals are hardly perturbed using RADM-FDE-ADC. HOMO-2 and HOMO-3 exhibit n-type symmetrywhile the frontier orbitals from HOMO-1 to LUMO+1 areπ and π∗ type orbitals, respectively. The five energeticallylowest excited states are investigated. Their character isconserved compared to the supersystem calculation. Localπ → π∗ transitions can be observed in the S2 and S4 whilethe remaining excited states show n → π∗ transition character.(See Table IX.)

    As in the SE-FDE-ADC calculation, the difference to thesupersystem calculation is slightly larger than in the previousexamples due to the neglected contributions localized onthe environment, i.e., the five water molecules. The largestdifference in excitation energies is about 0.1 eV. The errorinduced by the RADM approximation alone compared to the

    FIG. 11. Excitation energy difference (left) and oscillator strength difference(right) of the isolated uracil (blue) to the supersystem of uracil with five watermolecules, the SE-FDE-ADC(2) calculation (red) to the supersystem, and theRADM-FDE-ADC(2) calculation (green) to the supersystem. The blue barscorrespond to the influence of the environment on the excitation energiesand oscillator strengths (solvatochromic shift) while the red and green barsindicate the accuracy of the SE-FDE-ADC(2) and RADM-FDE-ADC(2)calculations in comparison with the ADC(2) calculation of the full system.The SE-FDE-ADC(2) results are taken from Sec. IV A 2.

    SE-FDE-ADC(2) results is smaller than 0.07 eV in excitationenergies. Considering the large shift of the excitation energieswith up to 0.7 eV due to the environment, this difference is inexcellent agreement with the reference calculation (Fig. 11).

    For this system, the computation time was alsoinvestigated. The calculation of the supersystem takes about73 h CPU time on one core keeping all data for the ADCcalculation in memory (Intel Xeon E7-4870v2 2.3 GHz,software compiled using Intel C++ and Fortran compilers46

    v15.0 in combination with the MKL library47). In contrast, thecomplete FDE-ADC calculations using the RADM approachwhich consist of the MP(2) calculation on system A in thebasis functions of A, the HF calculation of system A in thebasis functions of A and B, the HF calculation of system

    TABLE IX. Excitation energies, oscillator strengths, and orbital transitions for the FDE-ADC calculation usingthe RADM approach for uracil as system A and five water molecules as system B.

    Excitation energies (eV) Oscillator strengthsOrb. trans.

    RADM ∆Ωa ∆Ωb f ∆ f a ∆ f b H+x → L+y Weight (%)

    S1 5.043 −0.100 −0.065 0.0004 −0.0011 −0.0003 −2 → 0 (n → π∗) 73.8−3 → 0 (n → π∗) 5.9

    S2 5.319 0.104 0.028 0.2092 −0.0279 0.0070 0 → 0 (π → π∗) 81.4S3 6.198 −0.013 −0.023 0.0001 0.0000 −0.0001 −2 → 0 (n → π∗) 6.0

    −2 → 1 (n → π∗) 31.8−3 → 0 (n → π∗) 30.2−3 → 1 (n → π∗) 16.9

    S4 6.331 0.039 0.011 0.0720 −0.0243 0.0130 −1 → 0 (π → π∗) 85.7S5 6.763 0.106 0.014 0.0010 0.0005 −0.0011 −2 → 1 (n → π∗) 24.6

    −3 → 0 (n → π∗) 48.4−3 → 1 (n → π∗) 10.0

    aReferenced to the supersystem calculation.bReferenced to the SE-FDE-ADC(2) calculation.

  • 204103-13 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    FIG. 12. Comparison of the computational cost in CPU time of a supersys-tem calculation and the RADM-FDE-ADC method.

    B in the basis functions of A and B, the creation of theembedding potential, and the final FDE-ADC(2) calculationsin the basis function of A takes about 3 h on the samemachine. This amounts to a saving of 70 h or more than95% of the time needed for the supersystem calculation. Forlarger systems with even more environment molecules, thepercentage of time saved will be drastically larger due tothe formal O(N5)-scaling of the ADC(2) calculation for thesupersystem (with N being the number of basis functions).

    A comparison of the computational time is shown inFig. 12. The size of the environment is successively increasedfrom 1 to 5 water molecules. The computational cost of thesupersystem calculation increases dramatically already for asmall environment due to the scaling of ADC(2). Hence, thecomputational cost is dominated by the calculation of theembedded species for a small environment. Of course, thistrend holds only for environments up to the same size assystem A. The calculation of the environment scales formallywith O(N3), which will become the most time-consuming stepfor large environments (e.g., more than 100 water molecules).But for such large systems, a full ADC calculation is no longerfeasible.

    V. DISCUSSION AND CONCLUSION

    In this work we have introduced a wavefunctionembedding approach based on the combination of ADCand linearized FDET. Three molecular model systems werestudied using two different FDE-ADC techniques in whichthe environment consisted of up to five water molecules.

    In the case of the supermolecular expansion (SE-FDE-ADC), we also tested different spatial arrangements andanalyzed the basis set dependence. Although polarizationof the environment is neglected when using a frozen density,the results from this study show that FDE-ADC is accurateenough to be employed in practical calculations for larger

    systems. Even large effects on the excited state due to stronginteractions with the environment were described correctly.Also large changes in oscillator strength due to the effect ofthe environment were reproduced with satisfactory accuracyby both FDE-ADC methods. The error of FDE-ADC islarger when orbitals of the environment contribute to theexcitation. This cannot be captured by FDET. In thesecases, the influence of the environment can both be over-or underestimated depending on the effect of the missingenvironmental orbital.

    While SE-FDE-ADC yields excitation energies very closeto the supersystem results (MAE = 0.025 eV), it providesno computational advantage over an ADC calculation ofthe supersystem. On the other hand, SE-FDE-ADC is wellsuited for an analysis of the FDE-ADC method since apartfrom linearization of the non-additive energy functionalEnadxc,T[ρA, ρB] no further approximations are introduced.

    With the second variant, RADM, it is possible toapproach a system size of practical relevance due to thetruncation of the embedding potential to the basis functionsof system A. Although accompanied by the introduction ofnew approximations, the RADM variant performs comparablywell (MAE = 0.040 eV). Computation of the reassembleddensity matrix requires an additional Hartree-Fock calculationof the supersystem and therefore increases the scaling bythe term O(N3

    A+B). Using the current implementation, also

    investigations of chromophores in solution at FDE-ADC(3)level of theory should be feasible using the existing ADC(3)implementation in Q-Chem.48 Also, state-of-the-art tools forsystematic analysis of electronic excitations can be combinedwith the FDE-ADC approach for a detailed investigation ofthe influence of the environment on the character of excitedstates.49,50

    Note that the evaluation of the total energy was not neededin the present work because of the relation for the excitationenergy given in Eq. (18) holding for linearized FDET. Also,the evaluation of the total energy (see Eq. (15)) involves notonly expectation values of some quantum operators but alsoexplicit integration over the real space.

    A future implementation of FDE-ADC should includethe so called Monomer Expansion (ME). In this approach, thecalculation of the subsystem involves only the respective basisfunctions of one system, which reduces the computational costeven further without inducing an error due to a truncation asin the RADM approach.

    In total, the FDE-ADC method is a promising approachfor considering environmental effects on electronically excitedstates. The error of this method is lower than the intrinsicerror of the used ADC method and using the RADMapproximation explicit treatment of extended environments isalready feasible. In further studies, charge-transfer complexesare going to be investigated in polar solvents using the RADM-FDE-ADC method.

    ACKNOWLEDGMENTS

    This research was supported by grants from the SwissNational Science Foundation (No. 200021_152779). S.P.thanks the Heidelberg Graduate School “Mathematical and

  • 204103-14 Prager et al. J. Chem. Phys. 144, 204103 (2016)

    Computational Methods for the Sciences” (HGS MathComp)for financial support.

    1W. Liptay, Angew. Chem., Int. Ed. 8, 177 (1969).2J. Tomasi, B. Mennucci, and R. Cammi, Chem. Rev. 105, 2999 (2005).3A. Warshel and M. Levitt, J. Mol. Biol. 103, 227 (1976).4B. Mennucci, J. Tomasi, R. Cammi, J. R. Cheeseman, M. J. Frisch, F. J.Devlin, S. Gabriel, and P. J. Stephens, J. Phys. Chem. A 106, 6102 (2002).

    5R. A. Klein, B. Mennucci, and J. Tomasi, J. Phys. Chem. A 108, 5851(2004).

    6J.-M. Mewes, Z.-Q. You, M. Wormit, T. Kriesche, J. M. Herbert, and A.Dreuw, J. Phys. Chem. A 119, 5446 (2015).

    7T. A. Wesolowski and A. Warshel, J. Phys. Chem. 97, 8050 (1993).8N. Govind, Y. A. Wang, A. J. R. da Silva, and E. A. Carter, Chem. Phys. Lett.295, 129 (1998).

    9L. I. Bendavid and E. A. Carter, Status in Calculating Electronic ExcitedStates in Transition Metal Oxides from First Principles (Springer, 2014),pp. 1–52.

    10T. A. Wesolowski, S. Shedge, and X. Zhou, Chem. Rev. 115, 5891 (2015).11T. A. Wesolowski, Phys. Rev. A 77, 012504 (2008).12P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).13W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965).14R. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules

    (Oxford University Press, Clarendon Press, New York, Oxford, England,1989).

    15F. Aquilante and T. A. Wesołowski, J. Chem. Phys. 135, 084120 (2011).16C. Daday, C. König, O. Valsson, J. Neugebauer, and C. Filippi, J. Chem.

    Theory Comput. 9, 2355 (2013).17C. Daday, C. König, J. Neugebauer, and C. Filippi, ChemPhysChem 15, 3205

    (2014).18N. Govind, Y. A. Wang, and E. A. Carter, J. Chem. Phys. 110, 7677 (1999).19A. Dreuw, J. L. Weisman, and M. Head-Gordon, J. Chem. Phys. 119, 2943

    (2003).20A. Dreuw and M. Head-Gordon, J. Am. Chem. Soc. 126, 4007 (2004).21A. D. Laurent and D. Jacquemin, Int. J. Quantum Chem. 113, 2019 (2013).22M. Dulak, J. W. Kaminski, and T. A. Wesolowski, Int. J. Quantum Chem.

    109, 1886 (2009); e-print arXiv:0610288 [physics].23T. Dresselhaus, J. Neugebauer, S. Knecht, S. Keller, Y. Ma, and M. Reiher,

    J. Chem. Phys. 142, 044111 (2015).24A. S. P. Gomes, C. R. Jacob, and L. Visscher, Phys. Chem. Chem. Phys. 10,

    5353 (2008).25S. Hofener, A. S. P. Gomes, and L. Visscher, J. Chem. Phys. 139, 104106

    (2013).26D. K. Kanan, S. Sharifzadeh, and E. A. Carter, Chem. Phys. Lett. 519, 18

    (2012).27T. A. Wesolowski, J. Chem. Phys. 140, 18A530 (2014).28A. Zech, F. Aquilante, and T. A. Wesolowski, J. Chem. Phys. 143, 164106

    (2015).29M. Wormit, D. R. Rehn, P. H. Harbach, J. Wenzel, C. M. Krauter, E.

    Epifanovsky, and A. Dreuw, Mol. Phys. 112, 774 (2014).30A. Dreuw and M. Wormit, WIREs Comput. Mol. Sci. 5, 82 (2015).31B. Lunkenheimer and A. Khn, J. Chem. Theory Comput. 9, 977 (2013).32A. P. Gamiz-Hernandez, I. N. Angelova, R. Send, D. Sundholm, and V. R.

    I. Kaila, Angew. Chem., Int. Ed. 54, 11564 (2015).33J. Schirmer, Phys. Rev. A 26, 2395 (1982).34J. Schirmer and A. B. Trofimov, J. Chem. Phys. 120, 11449 (2004).

    35E. R. Davidson, J. Comput. Phys. 17, 87 (1975).36J. P. Perdew and M. Levy, Phys. Rev. B 31, 6264 (1985).37Y. G. Khait and M. R. Hoffmann, J. Chem. Phys. 133, 44107 (2010).38Y. Shao, Z. Gan, E. Epifanovsky, A. T. Gilbert, M. Wormit, J. Kussmann,

    A. W. Lange, A. Behn, J. Deng, X. Feng, D. Ghosh, M. Goldey, P. R. Horn,L. D. Jacobson, I. Kaliman, R. Z. Khaliullin, T. Ku, A. Landau, J. Liu,E. I. Proynov, Y. M. Rhee, R. M. Richard, M. A. Rohrdanz, R. P. Steele,E. J. Sundstrom, H. Lee Woodcock III, P. M. Zimmerman, D. Zuev, B.Albrecht, E. Alguire, B. Austin, G. J. O. Beran, Y. A. Bernard, E. Berquist,K. Brandhorst, K. B. Bravaya, S. T. Brown, D. Casanova, C.-M. Chang,Y. Chen, S. H. Chien, K. D. Closser, D. L. Crittenden, M. Diedenhofen,R. A. DiStasio, Jr., H. Do, A. D. Dutoi, R. G. Edgar, S. Fatehi, L. Fusti-Molnar, A. Ghysels, A. Golubeva-Zadorozhnaya, J. Gomes, M. W. Hanson-Heine, P. H. Harbach, A. W. Hauser, E. G. Hohenstein, Z. C. Holden,T.-C. Jagau, H. Ji, B. Kaduk, K. Khistyaev, J. Kim, J. Kim, R. A. King, P.Klunzinger, D. Kosenkov, T. Kowalczyk, C. M. Krauter, K. U. Lao, A. D.Laurent, K. V. Lawler, S. V. Levchenko, C. Y. Lin, F. Liu, E. Livshits, R. C.Lochan, A. Luenser, P. Manohar, S. F. Manzer, S.-P. Mao, N. Mardirossian,A. V. Marenich, S. A. Maurer, N. J. Mayhall, E. Neuscamman, C. M. Oana,R. Olivares-Amaya, D. P. ONeill, J. A. Parkhill, T. M. Perrine, R. Peverati, A.Prociuk, D. R. Rehn, E. Rosta, N. J. Russ, S. M. Sharada, S. Sharma, D. W.Small, A. Sodt, T. Stein, D. Stck, Y.-C. Su, A. J. Thom, T. Tsuchimochi, V.Vanovschi, L. Vogt, O. Vydrov, T. Wang, M. A. Watson, J. Wenzel, A. White,C. F. Williams, J. Yang, S. Yeganeh, S. R. Yost, Z.-Q. You, I. Y. Zhang, X.Zhang, Y. Zhao, B. R. Brooks, G. K. Chan, D. M. Chipman, C. J. Cramer,W. A. Goddard III, M. S. Gordon, W. J. Hehre, A. Klamt, H. F. Schaefer III,M. W. Schmidt, C. D. Sherrill, D. G. Truhlar, A. Warshel, X. Xu, A. Aspuru-Guzik, R. Baer, A. T. Bell, N. A. Besley, J.-D. Chai, A. Dreuw, B. D. Dunietz,T. R. Furlani, S. R. Gwaltney, C.-P. Hsu, Y. Jung, J. Kong, D. S. Lambrecht,W. Liang, C. Ochsenfeld, V. A. Rassolov, L. V. Slipchenko, J. E. Subotnik,T. V. Voorhis, J. M. Herbert, A. I. Krylov, P. M. Gill, and M. Head-Gordon,Mol. Phys. 113, 184 (2015).

    39C. Cramer, Essentials of Computational Chemistry: Theories and Models(Wiley, Chichester, West Sussex, England, Hoboken, NJ, 2004).

    40F. Jensen, Introduction to Computational Chemistry (John Wiley & Sons,Chichester, England, Hoboken, NJ, 2007).

    41G. Karlstroem, R. Lindh, P.-Å. Malmqvist, B. O. Roos, U. Ryde, V.Veryazov, P.-O. Widmark, M. Cossi, B. Schimmelpfennig, P. Neogrády,and L. Seijo, in Proceedings of the Symposium on Software Developmentfor Process and Materials Design [Comput. Mater. Sci. 28, 222 (2003)].

    42F. Aquilante, L. De Vico, N. Ferré, G. Ghigo, P.-Å. Malmqvist, P. Neogrády,T. B. Pedersen, M. Pitok, M. Reiher, B. O. Roos, L. Serrano-Andrs, M.Urban, V. Veryazov, and R. Lindh, J. Comput. Chem. 31, 224 (2010).

    43A. Szabo, Modern Quantum Chemistry: Introduction to Advanced Elec-tronic Structure Theory (Dover Publications, Mineola, NY, 1996).

    44M. Hanwell, D. Curtis, D. Lonie, T. Vandermeersch, E. Zurek, and G.Hutchison, J. Cheminf. 4, 17 (2012).

    45T. H. Dunning, J. Chem. Phys. 90, 1007 (1989).46See https://software.intel.com/en-us/intel-compilers for Intel C++ and

    Fortran Compilers, 2015.47See https://software.intel.com/en-us/intel-mkl for Intel Math Kernel Library

    (MKL), 2015.48P. H. P. Harbach, M. Wormit, and A. Dreuw, J. Chem. Phys. 141, 064113

    (2014).49F. Plasser, M. Wormit, and A. Dreuw, J. Chem. Phys. 141, 024106 (2014).50F. Plasser, S. A. Baeppler, M. Wormit, and A. Dreuw, J. Chem. Phys. 141,

    024107 (2014).

    http://dx.doi.org/10.1002/anie.196901771http://dx.doi.org/10.1021/cr9904009http://dx.doi.org/10.1016/0022-2836(76)90311-9http://dx.doi.org/10.1021/jp020124thttp://dx.doi.org/10.1021/jp0487408http://dx.doi.org/10.1021/jp511163yhttp://dx.doi.org/10.1021/j100132a040http://dx.doi.org/10.1016/S0009-2614(98)00939-7http://dx.doi.org/10.1021/cr500502vhttp://dx.doi.org/10.1103/PhysRevA.77.012504http://dx.doi.org/10.1103/PhysRev.136.B864http://dx.doi.org/10.1103/PhysRev.140.A1133http://dx.doi.org/10.1063/1.3624888http://dx.doi.org/10.1021/ct400086ahttp://dx.doi.org/10.1021/ct400086ahttp://dx.doi.org/10.1002/cphc.201402459http://dx.doi.org/10.1063/1.478679http://dx.doi.org/10.1063/1.1590951http://dx.doi.org/10.1021/ja039556nhttp://dx.doi.org/10.1002/qua.24438http://dx.doi.org/10.1002/qua.22011http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://arxiv.org/abs/0610288http://dx.doi.org/10.1063/1.4906152http://dx.doi.org/10.1039/b805739ghttp://dx.doi.org/10.1063/1.4820488http://dx.doi.org/10.1016/j.cplett.2011.11.003http://dx.doi.org/10.1063/1.4870014http://dx.doi.org/10.1063/1.4933372http://dx.doi.org/10.1080/00268976.2013.859313http://dx.doi.org/10.1002/wcms.1206http://dx.doi.org/10.1021/ct300763vhttp://dx.doi.org/10.1002/anie.201501609http://dx.doi.org/10.1103/PhysRevA.26.2395http://dx.doi.org/10.1063/1.1752875http://dx.doi.org/10.1016/0021-9991(75)90065-0http://dx.doi.org/10.1103/PhysRevB.31.6264http://dx.doi.org/10.1063/1.3460594http://dx.doi.org/10.1080/00268976.2014.952696http://dx.doi.org/10.1016/S0927-0256(03)00109-5http://dx.doi.org/10.1002/jcc.21318http://dx.doi.org/10.1186/1758-2946-4-17http://dx.doi.org/10.1063/1.456153https://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-compilershttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttps://software.intel.com/en-us/intel-mklhttp://dx.doi.org/10.1063/1.4892418http://dx.doi.org/10.1063/1.4885819http://dx.doi.org/10.1063/1.4885820