16
Comparison between the BetheSalpeter Equation and Conguration Interaction Approaches for Solving a Quantum Chemistry Problem: Calculating the Excitation Energy for Finite 1D Hubbard Chains Qi Ou and Joseph E. Subotnik* Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United States ABSTRACT: We calculate the excitation energies of nite 1D Hubbard chains with a variety of dierent site energies from two perspectives: (i) the physics-based BetheSalpeter equation (BSE) method and (ii) the chemistry-based conguration interaction (CI) approach. Results obtained from all methods are compared against the exact values for three classes of systems: metallic, impurity-doped, and molecular (semiconducting/insulating) systems. While in a previous study we showed that the GW method holds comparative advantages versus traditional quantum chemistry approaches for calculating the ionization potentials and electron anities across a large range of Hamiltonians, we show now that the BSE method outperforms CI approaches only for metallic and semiconducting systems. For insulating molecular systems, CI approaches generate better results. 1. INTRODUCTION In the past few decades, great eorts have been made to explore the excited state properties of various systems via both experimental techniques and computational approaches. Accu- rate excited state energies are necessary for interpreting electronic spectra, constructing optically active molecules, and designing interfaces between optically active molecules and solid state materials. 113 Obviously, calculating exact excitation energies is impossible for large systems because the Coulomb interaction between electrons makes it exponentially dicult to diagonalize an N- body Hamiltonian. As such, we are usually left with approximate schemes. If one is treating isolated molecules, usually the best accuracy is obtained with multireference methods, especially multireference conguration interaction (MRCI). 1417 That being said, multireference methods are not terribly practical for solid state systems; for solids, single-reference methods are used almost exclusively. In the literature today, within the context of molecular chemistry, there is a large variety of single-reference methods for excited states. One approach is to use time-dependent density functional theory (TD-DFT) to capture dynamic correla- tion. 18, 19 A second approach is to extend conguration interaction singles (CIS). 20 Even though CIS is unreliable for ordering excited states, CIS is inexpensive and the CIS wave functions are often qualitatively correct. To go beyond CIS, several options exist, a few of which we list here: Time-dependent HartreeFock (TDHF): eectively a full version of CIS that includes electron correlation to the rst order. 21 CIS(D): a second-order perturbative correction to CIS that is applicable to nondegenerate cases 22,23 CC2: the approximate coupled-cluster singles and doubles model that iteratively determines the singles and doubles substitutions as the poles of a true linear response function 24,25 ADC(2): algebraic diagrammatic construction through second order 26,27 CIS(D n ): a family of quasi-degenerate second-order perturbative corrections to CIS 28 These single-reference approaches can very often yield qualitatively correct results. Now, in the context of solid state physics, the methods above have not been considered, either because the starting point for solids is always DFT (and not HF) 13,2932 or because the methods above were considered too expensive. That being said, in recent years, there has been an explosion of interest in correlated excited state methods based on a Greens function many-body perturbation theory (MBPT). 13,32 The basic idea is to treat correlated electrons (holes) as approximately independ- ent quasiparticles and address all electronelectron correlation eects through the self-energy of the quasiparticle. 3337 Within MBPT, one can calculate charged excitations with high precision using, for example, Hedins GW approximation 38 and neutral excitations through the BetheSalpeter equation (BSE). 39 In a recent study, for the case of electron attachment and detachment energies, we benchmarked GW calculations against conguration interaction (CI) approaches for a set of 1D Received: March 8, 2017 Published: November 28, 2017 Article pubs.acs.org/JCTC Cite This: J. Chem. Theory Comput. XXXX, XXX, XXX-XXX © XXXX American Chemical Society A DOI: 10.1021/acs.jctc.7b00246 J. Chem. Theory Comput. XXXX, XXX, XXXXXX

Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

Comparison between the Bethe−Salpeter Equation andConfiguration Interaction Approaches for Solving a QuantumChemistry Problem: Calculating the Excitation Energy for Finite 1DHubbard ChainsQi Ou† and Joseph E. Subotnik*

Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United States

ABSTRACT: We calculate the excitation energies of finite 1D Hubbard chains with a variety of differentsite energies from two perspectives: (i) the physics-based Bethe−Salpeter equation (BSE) method and (ii)the chemistry-based configuration interaction (CI) approach. Results obtained from all methods arecompared against the exact values for three classes of systems: metallic, impurity-doped, and molecular(semiconducting/insulating) systems. While in a previous study we showed that the GW method holdscomparative advantages versus traditional quantum chemistry approaches for calculating the ionizationpotentials and electron affinities across a large range of Hamiltonians, we show now that the BSE methodoutperforms CI approaches only for metallic and semiconducting systems. For insulating molecularsystems, CI approaches generate better results.

1. INTRODUCTION

In the past few decades, great efforts have been made to explorethe excited state properties of various systems via bothexperimental techniques and computational approaches. Accu-rate excited state energies are necessary for interpretingelectronic spectra, constructing optically active molecules, anddesigning interfaces between optically active molecules and solidstate materials.1−13

Obviously, calculating exact excitation energies is impossiblefor large systems because the Coulomb interaction betweenelectrons makes it exponentially difficult to diagonalize an N-body Hamiltonian. As such, we are usually left with approximateschemes. If one is treating isolated molecules, usually the bestaccuracy is obtained with multireference methods, especiallymultireference configuration interaction (MRCI).14−17 Thatbeing said, multireference methods are not terribly practical forsolid state systems; for solids, single-reference methods are usedalmost exclusively.In the literature today, within the context of molecular

chemistry, there is a large variety of single-reference methods forexcited states. One approach is to use time-dependent densityfunctional theory (TD-DFT) to capture dynamic correla-tion.18,19 A second approach is to extend configurationinteraction singles (CIS).20 Even though CIS is unreliable forordering excited states, CIS is inexpensive and the CIS wavefunctions are often qualitatively correct. To go beyond CIS,several options exist, a few of which we list here:

• Time-dependent Hartree−Fock (TDHF): effectively afull version of CIS that includes electron correlation to thefirst order.21

• CIS(D): a second-order perturbative correction to CISthat is applicable to nondegenerate cases22,23

• CC2: the approximate coupled-cluster singles and doublesmodel that iteratively determines the singles and doublessubstitutions as the poles of a true linear responsefunction24,25

• ADC(2): algebraic diagrammatic construction throughsecond order26,27

• CIS(Dn): a family of quasi-degenerate second-orderperturbative corrections to CIS28

These single-reference approaches can very often yieldqualitatively correct results.Now, in the context of solid state physics, the methods above

have not been considered, either because the starting point forsolids is always DFT (and not HF)13,29−32 or because themethods above were considered too expensive. That being said,in recent years, there has been an explosion of interest incorrelated excited state methods based on a Green’s functionmany-body perturbation theory (MBPT).13,32 The basic idea isto treat correlated electrons (holes) as approximately independ-ent quasiparticles and address all electron−electron correlationeffects through the self-energy of the quasiparticle.33−37 WithinMBPT, one can calculate charged excitations with high precisionusing, for example, Hedin’s GW approximation38 and neutralexcitations through the Bethe−Salpeter equation (BSE).39

In a recent study, for the case of electron attachment anddetachment energies, we benchmarked GW calculations againstconfiguration interaction (CI) approaches for a set of 1D

Received: March 8, 2017Published: November 28, 2017

Article

pubs.acs.org/JCTCCite This: J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

© XXXX American Chemical Society A DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

Page 2: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

Hubbard chains.40 In the present paper, we will now go one stepfurther. For the same model systems, we will benchmark neutralexcited state energies as calculated by both the BSE approach(which is popular in the physics community) and a set of CIapproaches (which is common in the chemistry community).Exact excitation energies are obtained using direct diagonaliza-tion of the full Hamiltonian so that the approximate results canbe evaluated for accuracy.An outline of this paper is as follows. In section 2, we review

BSE theory by deriving the relevant equations using standardquantum chemistry notation (section 2.A). Then we review therelevant CI theory (section 2.B) and our choice of model system(section 2.C). In section 3, we present the excitation energiespredicted by these different methods. In section 4, we discussvarious factors that influence the performance of the BSEmethods. In section 5, we conclude.Unless otherwise specified, we use lowercase latin letters to

denote spin molecular orbitals (MOs) (a, b, c, d for virtualorbitals, i, j, k, l, m for occupied orbitals, p, q, r, s, w for arbitraryorbitals) and Greek letters (α, β, γ, δ, λ, σ, μ, ν) to denote atomicorbitals (AOs). Position is denoted by r, and position/spintogether are denoted by x ≡ (r,σ). Thus, the notation ϕa(x)signifies an one-electron spin wave function indexed by “a”,where the spin of ϕa equals the spin of x. The electronic excitedstates obtained within the random-phase approximation (RPA)are denoted by ΨI or ΨJ (with uppercase latin indices I, J).

2. THEORYWe will now review both BSE theory and the relevant CI theory.Because BSE is less common in chemistry than is CI, we willfocus especially on deriving the relevant equations for BSE(which takes some time) using the standard quantum chemistrynotation. We will assume only basic familiarity with Green’sfunction formalisms, though we will assume some of the basicresults of ref 40 (which summarizes the GW approach in thelanguage of quantum chemistry). For BSE, we will followStrinati’s paper (ref 41) in detail and make a flowchart to makethe process easily repeatable. Note that Strinati’s paper derivesthe BSE working equations only in the Tamm−Dancoffapproximation, and thus, we will modify the derivation slightly.Below, we will also highlight the three key assumptions of BSEtheory.2.A. Bethe−Salpeter Equation. 2.A.1. Two-Particle

Correlation Function and Its BSE. BSE is derived using standardMBPT and the language of Green’s functions. We suppose that |N⟩ is the exact fully interacting ground state of the N-bodyelectronic system. Our final result agrees with Louie’s result in ref35. We define the one- and two-particle Green’s function as

′ ≡ − ℏ⟨ | Ψ Ψ ′ | ⟩†

G N T N(1, 1 ) i/ [ (1) (1 )]1 (1)

′ ′ ≡ − ℏ ⟨ | Ψ Ψ Ψ ′ Ψ ′ | ⟩† †

G N T N(12; 1 2 ) ( i/ ) [ (1) (2) (2 ) (1 )]22

(2)

Here we use the simplified notation for the indices (x1, t1) = (r1,σ1, t1)→ (1), (x2, t2) = (r2, σ2, t2)→ (2), i.e., (1) and (2) denoteboth space and time. We use the standard notation

∫ ∫= ∑σx rd d1 11. T denotes time-ordering. The two-particle

correlation function is defined as41

′ ′ ≡ ′ ′ − ′ ′L G G G(12; 1 2 ) (1, 1 ) (2, 2 ) (1, 2; 1 , 2 )1 1 2 (3)

Now, invoking a well-known trick in Green’s function theory,we imagine applying a one-particle potential u (x, x′; t) to theelectronic system. According to well-known functional derivativeidentities,41 one can show that (for t′2 = t2 + δ)

δδ

′ ′ = ′ ′

LG

u tx x(12; 1 2 )

(1, 1 )( , ; )2 2 2 (4a)

See Appendix A for a proof of this identity. Furthermore, if wedefine the quantity u(x1, t1; x2, t2) = u (x1, x2; t2)δ(t1 − t2), we canalso take a formal derivative of G with respect to u, and we find

δδ

′ ′ =′′

LGu

(12; 1 2 )(1, 1 )

(2, 2 )1

(4b)

See again Appendix A.Although eq 4b is applicable only when t2 = t2′, a diagrammatic

analysis shows that the BSE (as presented in eq 12) is in factapplicable for t2 ≠ t2′.41 That being said, for the present paper, wenote that we will eventually set t2′ = t2 + δ (after eq 26); we will useeq 4b only as a formal tool to manipulate L and derive the BSEmore quickly. Thus, applying the matrix identity

= −δδ

δδ

− −−A AA

xAx

1 11

, we can rewrite L(12; 1′2′) as

∬ δδ

′ ′ = − ′′

′′ ′

−L G

Gu

G(12; 1 2 ) d3 d3 (1, 3)(3, 3 )

(2 , 2)(3 , 1 )1

11

1

(5)

We must now evaluate eq 5. When using Green’s functions,one always distinguishes between the Green’s function for theinteracting Hamiltonian and the Green’s function for thenoninteracting Hamiltonian (which can be easily distinguished).The relationship between the noninteracting Green’s functionG1(0) and the exact Green’s function G1 is

= − − Σ− −G G u(1, 2) (1, 2) (1, 2) (1, 2)11

1(0) 1

(6)

where Σ(1, 2) denotes the self-energy (which is defined by eq 6).For BSE, the first approximation (Approximation #1) is to invokeGW theory for single particles, so that the self-energy in eq 6contains the contribution from two parts

Σ ≡ Σ + Σ(1, 2) (1, 2) (1, 2)H GW(7)

Here

∫δΣ = − ℏ +⎡⎣⎢

⎤⎦⎥v G(1, 2) (1, 2) i d3 (1, 3) (3, 3 )H

1 (8)

denotes the Hartree contribution (i.e., the mean-field Coulombiceffect of all electrons together), and

Σ = ℏ +G W(1, 2) i (1, 2) (1 , 2)GW1 (9)

captures the correlated motion of an individual electron in a seaof other electrons that results in the electronic screening of themean-field potential; the screening potential W is computedapproximately via a GW calculation. The notation A(1+, 2)signifies that time t1 should be replaced by t1 + δ. In eqs 6−9, wenote that for nonzero matrix elements, we require that (1) and(2) should have the same spin.At this point, we plug eq 6 into eq 5, and we find

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

B

Page 3: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

∬ δ δ δδ

δδ

δδ

δδ

′ ′ = ′ ′ ′ + Σ ′′

′ ′

= ′ ′ + ∬∬ ′ ′ ′ ′ Σ ′′

′′

= ′ ′ + ∬∬ ′ ′ ′ ′ Σ ′′

′ ′

⎡⎣⎢

⎤⎦⎥L G

uG

G G G GG

Gu

G G G GG

L

(12; 1 2 ) d3 d3 (1, 3) (2 , 3) (2, 3 )(3, 3 )(2 , 2)

(3 , 1 ) (10)

(1, 2 ) (2, 1 ) d3 d3 d4 d4 (1, 3) (3 , 1 )(3, 3 )(4, 4 )

(4, 4 )(2 , 2)

(11)

(1, 2 ) (2, 1 ) d3 d3 d4 d4 (1, 3) (3 , 1 )(3, 3 )(4, 4 )

(42; 4 2 ) (12)

1 1

1 1 1 11

1

1 1 1 11

Equation 12 is the BSE for L. If we define the electron−holeinteraction kernel K as41

δδ

′ ′ ≡ Σ ′′

KG

(34 ; 3 4)(3, 3 )(4, 4 )1 (13)

the BSE for L can then be written as

′ ′ = ′ ′

+ ∬∬ ′ ′ ′ ′ ′ ′ ′ ′

L G G

G G K L

(12; 1 2 ) (1, 2 ) (2, 1 )

d3 d3 d4 d4 (1, 3) (3 , 1 ) (34 ; 3 4) (42; 4 2 )1 1

1 1

(14)

If one plugs eq 7 into eq 13 (with the help of eqs 8 and 9) andneglects the derivative of the screened interactionWwith respectto G (which constitutes Approximation #2 for BSE), K can berewritten approximately as

δ δδ δ

′ ′ = − ℏ ′ ′+ ℏ ′ ′ ′

≡ ′ ′ + ′ ′

+

+K v

W

K K

(34 ; 3 4) i (3, 3 ) (4 , 4 ) (3, 4)i (3, 4) (3 , 4 ) (3 , 3 )

(15)

(34 ; 3 4) (34 ; 3 4) (16)x d

In a confusing twist of notation, the first term Kx, which resultsfrom the Coulomb potential in eq 13, is usually called theexchange term, while the second term Kd, which results from thescreened-exchange self-energy in eq 13, is usually called thedirect interaction term.2.A.2. Eigenvalue Problem for BSE. The derivation of the

effective eigenvalue problem for BSE starts from the analysis ofthe four time variables in the two-body Green’s function G(12;1′2′) (i.e. the second term in L(12; 1′2′) in eq 3). To begin thisanalysis, one considers a change of variables to

τ τ τ≡ − ≡ − ≡ − ′ ′t t t t t t1 1 1 2 2 2 1 2 (17)

where = + ′t t t( )112 1 1 and = + ′t t t( )2

12 2 2 . The particle−hole

Green’s function contains six classes as follows (with the order ofthe time variables for each class of the Green’s function shown inFigure 1)

θ τ τ τ

′ ′ = − ℏ ⟨ | Ψ Ψ ′ Ψ Ψ ′ | ⟩·

− | | − | |

† †

⎜ ⎟⎛⎝

⎞⎠

G N T T N(12; 1 2 ) ( i/ ) [ (1) (1 )] [ (2) (2 )]12

12

2I 2

1 2 (18)

θ τ τ τ

′ ′ = − ℏ ⟨ | Ψ Ψ ′ Ψ Ψ ′ | ⟩·

− − | | − | |

† †

⎜ ⎟⎛⎝

⎞⎠

G N T T N(12; 1 2 ) ( i/ ) [ (2) (2 )] [ (1) (1 )]12

12

2II 2

1 2 (19)

θτ τ

ττ τ

ττ τ

′ ′ = − − ℏ ⟨ | Ψ Ψ ′ Ψ Ψ ′ | ⟩·

− − − + + − + +

† †

⎛⎝⎜

⎞⎠⎟

G N T T N(12; 1 2 ) ( i/ ) [ (2) (1 )] [ (1) (2 )]

2 212 2 2

12 2 2

2III 2

2 1 1 2 1 2

(20)

θτ τ

ττ τ

ττ τ

′ ′ = − − ℏ ⟨ | Ψ Ψ ′ Ψ Ψ ′ | ⟩·

− − − + + − + +

† †

⎛⎝⎜

⎞⎠⎟

G N T T N(12; 1 2 ) ( i/ ) [ (1) (2 )] [ (2) (1 )]

2 212 2 2

12 2 2

2IV 2

1 2 2 1 2 1

(21)

θτ τ

ττ τ

ττ τ

′ ′ = − ℏ ⟨ | Ψ Ψ Ψ ′ Ψ ′ | ⟩·

+ − + − − + −

† †

⎛⎝⎜

⎞⎠⎟

G N T T N(12; 1 2 ) ( i/ ) [ (1) (2)] [ (2 ) (1 )]

2 212 2 2

12 2 2

2V 2

1 2 2 1 2 1

(22)

θτ τ

ττ τ

ττ τ

′ ′ = − ℏ ⟨ | Ψ ′ Ψ ′ Ψ Ψ | ⟩·

− − − + − − + −

† †

⎛⎝⎜

⎞⎠⎟

G N T T N(12; 1 2 ) ( i/ ) [ (2 ) (1 )] [ (1) (2)]

2 212 2 2

12 2 2

2VI 2

1 2 2 1 2 1

(23)

It is clear that from the combination of the creation andannihilation operators that G2

I−IV correspond to the electron−hole portion of the two-body Green’s function (i.e., one electronand one hole are being propagated in the N-body system). G2

V

and G2VI correspond to the electron−electron and hole−hole

portion of the two-body Green’s function, respectively, which donot involve the excitation energies of the N-body system (butrather the (N ± 2)-body system). Thus, to locate the excitedstates of the N-body system, one needs to consider only G2

I−IV.To construct the eigenvalue equation for BSE, one performs a

Fourier transform on both sides of eq 14. Here we take G2I (12;

1′2′) as an example to compute the Fourier transform results.First we define the right- and left-hand electron−hole amplitudesas

Figure 1. Schematic graph of the time ordering for the two-bodyGreen’s function. Note that the time order for the electrons and/orholes within the same curly brace is random. Each class of the Green’sfunction therefore has four different time combinations, making a totalof 24 terms in the two-body Green’s function. The time increases fromright to left.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

C

Page 4: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

χ −

≡ ⟨ | Ψ Ψ ⟩ − + ℏ†

t t

N T N E E t t

x x( , ; )

[ (1) (2)] exp[i( )( )/2 ]

S

S S

eh1 2 1 2

0 1 2(24)

χ −

≡ ⟨ | Ψ Ψ ⟩ − − + ℏ†

t t

N T N E E t t

x x( , ; )

[ (1) (2)] exp[ i( )( )/2 ]

S

S S

eh1 2 1 2

0 1 2(25)

where |NS⟩ denotes the Sth excited state of the N-body system.42

Then G2I (12; 1′2′) can be rewritten as

∑ τ χ τ

χ τ θ τ τ τ

′ ′ = − ℏ − ℏ

× · − | | − | |

′ ⎜ ⎟⎛⎝

⎞⎠

G E E x x

x x

(12; 1 2 ) ( i/ ) exp[i( ) / ] ( , ; )

( , ; )12

12

SS S

S

2I 2

0eh

1 1 1

eh2 2 2 1 2 (26)

Second, upon setting t2′ = t2 + δ(δ→ 0+) (which gives τ2 =−δand τ = t1− t2) and Fourier transforming the variable t2, the resultgiven by G2

I (12; 1′2′) can be evaluated as

∫ ∫

θ τ χ τ χ δ

χ τ χ δ

χ τ χ δω η

′ ′ = − ℏ − − | | × −

= −ℏ

× −

= −ℏ

· −

ℏ − − +

ω ω

τω

ω ω τ

−∞

+∞−

−∞

+∞− − − ℏ

′ ′

−∞

− | |− − − ℏ

′ ′

− − ℏ − + | | ℏ ′ ′

⎜ ⎟

⎜ ⎟⎛⎝

⎞⎠

⎛⎝

⎞⎠

t G t t t

t

E E

x x x x

x x x x

x x x x

d e (12; 1 2 ) d e ( i/ )12

e ( , ; ) ( , ; ) (27)

id e e ( , ; ) ( , ; ) (28)

ie

( , ; ) ( , ; )

( ) i(29)

t t

S

E E t tS S

tt

S

E E t tS S

t E E

S

S S

S

2i

2I

2i 2

1 2 1i( )( )/ eh

1 1 1eh

2 2

2 (1/2)

2i i( )( )/ eh

1 1 1eh

2 2

i[ ( )( /2 )]eh

1 1 1eh

2 2

0

S

S

S

2 2 0 1 2

1 12 0 1 2

1 0 1

An imaginary infinitesimal has been included in thedenominator of eq 29 to ensure the convergence of the integral.Performing the same transformation for G2

I−IV, one finds that the

electron−hole correlation function (eq 3) in the frequencydomain Leh(x1, x2; x1′, x2′|t1, t1′; ω) contains only thecontributions from G2

I (12; 1′2′) and G2II(12; 1′2′).43 Thus

ω

χ τ χ δω η

χ δ χ τω η

| = − · ′ ′ + ′ ′

=ℏ

· −

ℏ − − +

− · −

ℏ + − −

ω

τ ω τ

ω τ

′ ′ ′−∞

+∞−

− − | | ℏ − − | | ′ ′

− + | | ′ ′

⎡⎣⎢⎢

⎤⎦⎥⎥

L t t t G G

E E

E E

x x x x

x x x x

x x x x

( , ; , , ; ) d e [ (12; 1 2 ) (12; 1 2 )] (30)

ie e

( , ; ) ( , ; )

( ) i

e( , ; ) ( , ; )

( ) i

(31)

t

E E t

S

S S

S

t

S

S S

S

eh1 2 1 2 1 1 2

i2I

2II

i( ) /2 i ( ( /2))eh

1 1 1eh

2 2

0

i ( ( /2))eh

2 2eh

1 1 1

0

S

2

0 1 1 1

1 1

Note that the first term in eq 3, the product of one-bodyGreen’s functions, does not contribute because its transformvanishes when ω is nonzero. Thus, eq 31 gives the Fouriertransform for the left-hand side of eq 14. Performing the sameFourier transform on the right-hand side of eq 14 gives56 (with

= + ′t t t4 2

4 4 and τ4 = t4 − t4′)

χ τ χ δω η

χ δ χ τω η

ℏ∬∬ ′ ′ ′ ′ ′ ′

· −

ℏ − − +

− · −

ℏ + − −

τ

ω τ

ω τ

− − | | ℏ

− − | | ′ ′

− + | | ′ ′

⎡⎣⎢⎢

⎤⎦⎥⎥

G G K

E E

E E

x x x x

x x x x

id3 d3 d4 d4 (1, 3) (3 , 1 ) (34 ; 3 4)e

e( , ; ) ( , ; )

( ) i

e( , ; ) ( , ; )

( ) i

E E

t

S

S S

S

t

S

S S

S

1 1i( ) /2

i ( ( /2))eh

4 4 4eh

2 2

0

i ( ( /2))eh

2 2eh

4 4 4

0

S 0 4

4 4

4 4

(32)

Evaluating the residue of eqs 31 and 32 at a particular pole ℏω

= (ES − E0) ≡ ΩS, one has

χ τ

χ τ

= ∬∬ ′ ′ ′ ′

× ′ ′

′− Ω ℏ

′− Ω ℏ

G G

K

x x

x x

( , ; )e

d3 d3 d4 d4 (1, 3) (3 , 1 )

(34 ; 3 4) ( , ; )e

St

St

eh1 1 1

i /

1 1

eh4 4 4

i /

S

S

1

4 (33)

We now perform a series of transformations upon eq 33 in

order to get the final working equations:1. Plug the expression of K (eq 15) into eq 33, which gives

∬∬

χ τ δ δδ δ χ τ

χ δ

χ τ

= ℏ ∬∬ ′ ′ ′ ′ − ′ ′+ ′ ′ ′

= − ℏ ′ ′ ′ −

+ ℏ ′ ′ ′ ′

′− Ω ℏ +

+′

− Ω ℏ

′ ′− Ω ℏ

+′

− Ω ℏ

G G vW

G G v

G G W

x xx x

x x

x x

( , ; )e i d3 d3 d4 d4 (1, 3) (3 , 1 )[ (3, 3 ) (4 , 4 ) (3, 4)(3, 4) (3 , 4 ) (3 , 3 )] ( , ; )e

(34)

i d3 d3 (1, 3) (3,1 ) (3, 3 ) ( , ; )e

i d3 d3 (1, 3) (3 , 1 ) (3 , 3 ) ( , ; )e

(35)

St

St

St

St

eh1 1 1

i /1 1

eh4 4 4

i /

1 1eh

3 3i /

1 1eh

3 3 3i /

S

S

S

S

1

4

3

3

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

D

Page 5: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

Here we have used the fact that v(3, 4) is local in time, i.e., v(3, 4)= v(r3 − r4)δ(t3 − t4).2. Plug the quasiparticle expression for the single-particle

Green’s function in the time domain

ϕ ϕ θ

ϕ ϕ θ

′ − ′ = −ℏ

* ′ − ′

− * ′ ′ −

ε

ε

− − ′ ℏ

− − ′ ℏ

⎡⎣⎢⎢

⎤⎦⎥⎥

G t t t t

t t

x x x x

x x

( , , )i

( ) ( ) ( )e

( ) ( ) ( )e

aa a

t t

ii i

t t

1i ( )/

i ( )/

a

i

QP

QP

(36)

into eq 35, where εsQP stands for the quasiparticle energy of

orbital s. Here we assume that ϕi and ϕa (and x and x′) have the

same spin, i.e., σ = σ′. After plugging in, we multiply both sides of

eq 33 by ϕa*(x1)ϕi(x1′) and integrate over ∫ ∫ dx1 dx1′. Setting t1′

= t1 + δ(δ → 0+), we find

∬ ∬

∬ ∫

ϕ ϕ χ δϕ ϕ

ε εχ δ

ϕ ϕε ε

τ τ θ τ θ τ

χ τ

* − =*

Ω − +− −

+*

Ω − +× + −

×

τ ε τ ε

′ ′ ′− Ω ℏ

′ ′ ′ ′− Ω ℏ

′′

−∞

+∞

′+ + Ω ℏ − Ω ℏ

′− Ω ℏ

v

W

x x x x x x x xx x

r r x x

x xx x

r r

x x

d d ( ) ( ) ( , ; )e d d( ) ( )

( ) ( , ; )e

d d( ) ( )

d ( , ; ) ( ( )e ( )e )

( , ; )e

i a St i a

S a iS

t

i a

S a i

St

1 1 1 1eh

1 1i /

3 33 3

QP QP 3 3eh

3 3i /

3 33 3

QP QP 3 3 3 3 3i ( ( /2))/

3i ( ( /2))/

eh3 3 3

i /

S S

i S a S

S

1 1

3QP

3QP

1 (37)

To derive eq 37, we have performed a few integrations over time.On the right-hand side, in the first term, the only relevant termsrequire t1 > t3 = t3′; in the second term, the only relevant terms aret1 > t3, t3′. To tackle the second term, we change integrationvariables from t3, t3′ to τ3, t3, where

τ τ= + = −′t t t t12

123 3 3 3 3 3

For the case τ3 > 0, the relevant terms satisfy τ < −t t3 112 3; for

the case τ3 < 0, the relevant terms satisfy τ < +t t3 112 3.

3. Multiply both sides of eq 37 byΩS− εaQP + εi

QP and drop the

common factor − Ω ℏe ti /S 1 . We find

∬ ∬∬ ∫

ε ε ϕ ϕ χ δ ϕ ϕ χ δ

ϕ ϕ τ τ θ τ θ τ χ τ

Ω − + * − = * − −

+ * × + −τ ε τ ε

′ ′ ′ ′ ′ ′ ′

′ ′−∞

+∞

′+ + Ω ℏ − Ω ℏ

v

W

x x x x x x x x x x r r x x

x x x x r r x x

( ) d d ( ) ( ) ( , ; ) d d ( ) ( ) ( ) ( , ; )

d d ( ) ( ) d ( , ; ) ( ( )e ( )e ) ( , ; )

S a i i a S i a S

i a S

QP QP1 1 1 1

eh1 1 3 3 3 3 3 3

eh3 3

3 3 3 3 3 3 3 3 3i ( ( /2))/

3i ( ( /2))/ eh

3 3 3i S a S3QP

3QP

(38)

4. Make the ansatz that χSeh(x, x′; τ) can be expanded into an

incomplete MO basis set that contains only the occupied−virtualand virtual−occupied parts with simple time dependence. Thisansatz corresponds to Assumption #3 of the BSE approach

χ τ ϕ ϕ ϕ ϕ θ τ θ τ

ϕ ϕ ϕ ϕ θ τ θ τ

′ = − ⟨ | | ⟩ * ′ + ⟨ | | ⟩ * ′ × + −

≡ − * ′ + * ′ × + −

τ τε τε

τ τε τε

Ω | | ℏ † † − ℏ − ℏ

Ω | | ℏ − ℏ − ℏ

N a a N N a a N

X Y

x x x x x x

x x x x

( , ; ) e [( ( ) ( ) ( ) ( )) ( ( )e ( )e )] (39)

e [( ( ) ( ) ( ) ( )) ( ( )e ( )e )] (40)

Sjb

j b S b j b j S j b

jbjbS

b j jbS

j b

eh i /2 i / i /

i /2 i / i /

S b j

S b j

QP QP

QP QP

where XjbS and Yjb

S are defined as XjbS ≡ ⟨N|aj

†ab|NS⟩ andYjbS ≡ ⟨N|a b†aj|NS⟩.5. By definition, ϕi and ϕa have the same spin in

eq 38, i.e., x1 and x1′ have the same spin on the left-hand side ofeq 38 (inside of the integral). Now, we plug eq 40 intoeq 38 and Fourier transform the screened potential

∫τ ω ω=π

ω−∞

∞ −W W( ) d e ( )t12

i . Next, we perform the integral

over τ3 and add an infinitesimal imaginary amplitude e−iη|τ3|with η> 0 to force the integration over τ3 to converge. Repeat the exactsame analysis for χS

eh(x, x′; τ). The generalized eigenvalueproblem for BSE finally reads

− −

= Ω

⎛⎝⎜⎜

⎞⎠⎟⎟⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜⎜

⎞⎠⎟⎟A B

B A

X

Y

X

Y

S S

S S

S

S S

S

S(41)

for electron−hole excitation amplitudes XS and YS and excitationenergy ΩS for BSE excited state S. AS and BS have the followingmatrix elements

δ δ ε ε = − +A K( )iajbS

ij ab a i iajbSQP QP

(42)

=B KiajbS

iabjS

(43)

where

= + ΩK K K ( )iabjS

iabjx

iabjd

S (44)

∬ ϕ ϕ ϕ ϕ= ′ ′ ′ ′

= |

K v

ia jb

r r r r r r r rd d ( ) ( ) ( ) ( ) ( , ) (45)

( ) (46)

iajbx

i a j b

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

E

Page 6: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

ϕ ϕ ϕ ϕ

πω ω

ω ε ε η

ω ε ε η

Ω = − ′ ′ ′

× ′ ×Ω − ℏ − − +

+Ω + ℏ − − +

ωη−⎡⎣⎢⎢

⎤⎦⎥⎥

K

W

r r r r r r

r r

( ) d d ( ) ( ) ( ) ( )

i2

d e ( , , )1

( ) i

1( ) i

iajbd

S a b i j

S b i

S a j

iQP QP

QP QP(47)

In eqs 45 and 46, ϕi, ϕa, ϕj, and ϕb are spin orbitals that we haveassumed to be real valued. Furthermore, in eq 45, we require onlythat σi = σa and σj = σb such that same spin and opposite spincases are allowed; in eq 47, we require complete spin alignment,σi = σa = σj = σb. The factor e

−iωη is included in eq 47 to remind usonly that, if we seek to evaluate the integral over a complex, thatcontour must run below the real axis.2.A.3. Quasiparticle Energies from the GW Approximation.

To evaluate eq 47 in practice, we require aGW calculation on topof a Hartree calculation to extract the quasiparticle energy.40 Toavoid a self-consistent calculation, the standard approach is toapply a linear expansion for the real part of the correlation part ofthe self-energy Σc(ω)

ε εε ε ε

ωΣ ≈ Σ +

−ℏ

∂ Σ∂

Re ( ) Re ( )Re ( )c

sGW c

ssGW

sc

sHH H

(48)

which leads to

ε ε ε≈ + Σ + ΣZ Re ( )sGW

s sc

s sxH H

(49)

where εsH denotes the Hartree orbital energy and εs

GW denotes theGW orbital energy. The quasiparticle renormalization factor Zs isgiven by

εω

≡ −∂ Σ

−⎛⎝⎜

⎞⎠⎟Z 1

Re ( )s

csH 1

(50)

In ref 40, for a finite Hubbard model, we found that the Z-factor values for orbitals other than the HOMO and LUMOwereusually far from unity with correspondingly large errorscompared to exact attachment−detachment energies. As a result,in this work, we calculate Zs Re Σc(εs

H) for only the HOMO andLUMO. After adding in the exchange part of the self-energy Σs

x,we shift all of the virtual (occupied) orbitals by Zlumo Re Σc(εlumo

H )(Zhomo Re Σc(εhomo

H )), i.e.

ε ε ε≈ + Σ + ΣZ Re ( )aGW

ac

axH

lumo lumoH

(51)

ε ε ε≈ + Σ + ΣZ Re ( )iGW

ic

ixH

homo homoH

(52)

2.A.4. Matrix Elements for the Electron−Hole InteractingKernel. After the approximate quasiparticle energy εp

GW isobtained, the next step is to build the electron−hole interactionkernel K. The matrix elements of Kx(35, 46) (in the MO basis)are simply given by eq 46. In order to obtain the matrix elementsof Kd, we recall the expression for the screened Coulombinteraction W40,44,45

∫ω χ ω

ω

′ = ′ + ″ ‴ ″ ″ ‴

‴ ′

≡ ′ + ′

W v v v

v W

r r r r r r r r r r

r r

r r r r

( , ; ) ( , ) d d ( , ) ( , , )

( , )

(53)

( , ) ( , ; ) (54)p

where Wp(r, r′; ω) is defined as

∫ω χ ω′ ≡ ″ ‴ ″ ″ ‴ ‴ ′W v vr r r r r r r r r r( , ; ) d d ( , ) ( , , ) ( , )p

(55)

Here χ(r″, r‴, ω) is the time-dependent Hartree (TDH)polarizability in the frequency domain and is computed via thefollowing sum-over-states expression46 in a MO basis

χ ω ω

ω ω

= +−

=Ω + ℏ

+Ω − ℏ

⎜ ⎟⎡⎣⎢⎛⎝

⎞⎠

⎛⎝⎜

⎞⎠⎟⎤⎦⎥

⎡⎣⎢⎢

⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜⎜

⎞⎠⎟⎟

⎤⎦⎥⎥

A BB A

II

X

YX Y

Y

XY X

( )0

0(56)

1( )

1( ) (57)

I I

I

II I

I

I

II I

1

TDH TDH

where XI and YI correspond to the TDH excitation amplitudes of

excited state I with excitation energy ΩITDH.

In terms of MOs, if the perturbing potential is independent of

spin and of the form ujb(t)(ab†aj + aj

†ab), one finds that the

response function is

∑χ ωω η ω η

=Ω + ℏ −

+Ω − ℏ −

⎡⎣⎢⎢

⎤⎦⎥⎥

M M( )

i iiajbI

iajbI

I

iajbI

ITDH TDH

(58)

where

= + + +M X X X Y Y X Y YiajbI

iaI

jbI

iaI

jbI

iaI

jbI

iaI

jbI

(59)

In real space

∑ ∑χ ω ϕ ϕ ϕ ϕ

ω η ω η

′ = ′ ′

×Ω + ℏ −

+Ω − ℏ −

⎡⎣⎢⎢

⎤⎦⎥⎥

M M

r r r r r r( , , ) ( ) ( ) ( ) ( )

i i

I iajbi a j b

iajbI

I

iajbI

ITDH TDH

(60)

For this paper, we assume that the ground state of the system is

a closed-shell singlet so that all up-spin and down-spin orbitals

have the same spatial components. Thus, in eq 60, one can sum

over the spin components and produce a spatial MI matrix

∑ =σ σ σ σ= =

M MiajbI

iajbI

,i a j b (61)

We can then rewrite eq 60 as follows

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

F

Page 7: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

∑ ∑χ ω ϕ ϕ ϕ ϕ

ω η ω η

′ = ′ ′ ′

×

Ω + ℏ −+

Ω − ℏ −

⎡⎣⎢⎢

⎤⎦⎥⎥

M M

r r r r r r( , , ) ( ) ( ) ( ) ( )

i i

I iajbi a j b

iajbI

I

iajbI

ITDH TDH

(62)

where∑′ denotes a sum over only spatial orbitals. Note that onlysinglet states contribute to the sum-over-states in eq 62; all tripletcontributions cancel exactly.Finally, using eq 55, the matrix elements for Wp in real space

are

∑ ∑

ω ϕ ϕ ϕ ϕ

ω η ω η

′ = ′ ′ ′

× ′ | |

Ω + ℏ −+

Ω − ℏ −

⎡⎣⎢⎢

⎤⎦⎥⎥

W

ia kc jb ldM M

r r r r r r( , , ) ( ) ( ) ( ) ( )

( )( )i i

p

I iajbi a j b

kcld

kcldI

I

kcldI

ITDH TDH

(63)

The matrix elements in a spatial MO basis are

∑ ∑ω η

ω η

= ′ | |

Ω + ℏ −

+

Ω − ℏ −

⎡⎣⎢⎢⎤⎦⎥⎥

W ia kc jb ldM

M

( )( )i

i

iajbp

I kcld

kcldI

I

kcldI

I

TDH

TDH(64)

and the matrix elements in an AO basis are

∑= ′μνλσ μ λ ν σW C C C C Wp

iajba b i j iajb

p

(65)

Equation 47 is now ready to be evaluated with the help of eqs54, 64, and 65. Obviously, if one neglects the polarizable part ofthe screened Coulomb interactionWp in eqs 54, one will recoverTDHF exactly from BSE (eqs 41−43).2.A.5. Frequency-Independent Electron−Hole Interaction

Kernel. In many cases (e.g., in most semiconductor crystals), theexcitations in state |ΨS⟩ are mainly composed of electron−holepair configurations |Φi

a⟩ whose transition energies (εaQP − εi

QP)are close to the resulting excitation energy ΩS, which means thatΩS − (εa

QP− εiQP) is much smaller thanΩI

TDH.35 In such cases, eq47 can be approximated by

ϕ ϕ ϕ ϕ

ω

ω

= − ′ ′ ′

× ′ =

= − ′ =μνλσ

μ λ ν σ μνλσ

K

W

C C C C W

r r r r r r

r r

d d ( ) ( ) ( ) ( )

( , , 0)

(66)

( 0) (67)

iajbd

i a j bp

a b ip

j

where

∑ ∑ ∑ω

η η

= = ′ ′ | |

×

Ω −+

Ω −

μνλσ μ λ ν σ

⎡⎣⎢⎢

⎤⎦⎥⎥

W C C C C ia kc jb ld

M M

( 0) ( )( )

i i

p

iajba b i j

I kcld

kcldI

I

kcldI

ITDH TDH

(68)

2.A.6. Summary. The derivation of BSE (eqs 41−43)reviewed above is summarized in Figure 2. BSE can be solvedeither directly (without further approximation, i.e., the randomphase approximation (RPA)) or within the Tamm−Dancoffapproximation (TDA), which sets the matrix B = 0. In this work,we will test both versions. We also will study the self-consistent

(using eq 47 for Kd) and the non-self-consistent (using eq 66 forKd) versions of BSE. This completes our (tedious butcomprehensive) derivation of the BSE approach.

2.B. Configuration Interaction Approaches. Our goal inthis paper is to compare BSE with CI approaches. Forcompleteness, we now review the relevant quantum chemistryapproaches.

2.B.1. CIS/TDHF. The simplest CI method, CIS, includes onlythe single-excitation manifold. The matrix elements of the CISHamiltonian are

δ δ ε ε= ⟨Φ | |Φ ⟩ = − + ⟨ || ⟩A H aj ib( )iajb ia

jb

ij ab a iCIS

(69)

where ⟨aj||ib⟩≡ ⟨aj|ib⟩− ⟨aj|bi⟩ is the two-electron integral. TheCIS wave function can be written as

∑|Ψ ⟩ = |Φ ⟩tIia

iIa

iaCIS

(70)

where tiIa is the excitation coefficients for CIS excited state I. With

low computational cost, CIS is applicable to very large systems.However, in general, CIS only gives qualitatively correct results.The same qualitative features apply equally to TDHF, which is

a slight generalization of CIS. According to TDHF, which is aresponse theory based on top of a HF (not Hartree) groundstate, the excitation energies are computed by diagonalizing

− −

⎛⎝⎜⎜

⎞⎠⎟⎟A B

B A

TDHF TDHF

TDHF TDHF (71)

where AiajbTDHF = Aiajb

CIS and BiajbTDHF = ⟨ab||ij⟩. Like CIS, TDHF is

usually not accurate quantitatively.47

2.B.2. CIS(D). CIS(D) serves as a nondegenerate perturbativesecond-order correction to CIS that approximately introduceseffects of electron correlation and double-excitations for theexcited states in a noniterative scheme.22 In CIS(D), MP2 theory

Figure 2. Flowchart for how to calculate BSE energies in practice.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

G

Page 8: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

is employed in order to include electron correlation in theground state. The original unperturbed Hamiltonian H0 ischosen as the Fock operator. Then, with perturbation V, the totalHamiltonian becomes

λ = + H H V0 (72)

and performing second-order time-independent perturbationtheory, one recovers the MP2 correction to the ground-stateenergy

ω

ε ε ε ε

ε ε ε ε

= ⟨Φ | |Φ ⟩

= −|⟨Φ | |Φ ⟩|+ − −

= −|⟨ || ⟩|

+ − −

VT

V

ij ab

(73)

14

(74)

14

(75)

ijab

ijab

a b i j

ijab a b i j

0MP2

0 2 0

02

2

where T2 is the double-excitation operator that yields the double-excitation manifold |Φij

ab⟩with the effective excitation coefficients

ε ε ε ε

ε ε ε ε

=

≡ −⟨Φ | |Φ ⟩+ − −

= −⟨ || ⟩

+ − −

† †T d a a a a

dV

ij ab

14

(76)

(77)

(78)

ijabijab

a b i j

ijab

ijab

ijab

a b i j

ijab a b i j

2

0

Note that the single-excitation manifold does not appear in theMP2 energy because of Brillouin’s theorem.Now, CIS(D) is consistent with MP2 theory performed on

CIS excited states and includes the corrections to the CIS

excitation energy from three sources: single- and double-excitations with respect to the CIS wave function and again theMP2 correction to the ground-state energy. Therefore, we maywrite

ω ω= ⟨Ψ | |Ψ ⟩ + ⟨Ψ | |Ψ ⟩ −VT VTI I I I ICIS(D) CIS

1CIS CIS

2CIS

0MP2

(79)

Overall, the single-excitation operator acting on the CIS statein the first term gives an overall double-excitation

∑ε ε ε ε

≡ −|Φ ⟩⟨Φ |

+ − − − ΩT V

14 ijab

ijab

ijab

a b i j I1 CIS

(80)

Plugging eq 80 into eq 79, one finds that the first term in eq 79finally yields

ε ε ε ε⟨Ψ | |Ψ ⟩ = −

+ − − − Ω

≡ ⟨Ψ | |Φ ⟩

= ⟨Φ | |Φ ⟩

= ⟨ || ⟩ − ⟨ || ⟩

+ ⟨ || ⟩ − ⟨ || ⟩

VTs

s V

t V

ab cj t ab ci t

ka ij t kb ij t

14

( )(81)

(82)

(83)

[ ]

[ ]

(84)

I Iijab

ijab

a b i j I

ijab

I ijab

kckIc

kc

ijab

ciIc

jIc

kkIb

kIa

CIS1

CIS2

CIS

CIS

The second term in eq 79 gives a triple-excitation with respectto the ground state. Note that in CIS(D), the double-excitationoperator that acts on CIS excited states T2 remains the same asthe operator acting on the ground state in MP2 theory, i.e., thecorresponding excitation coefficients dij

ab are not changed.Therefore, the second term in eq 79 finally yields

∑ ∑

∑ ∑ω

⟨Ψ | |Ψ ⟩ = ⟨Φ | |Φ ⟩

= ⟨Φ | |Φ ⟩ + ⟨ || ⟩ + +

= + ⟨ || ⟩ + +

VT t t VT

VT t kl cd t d t d t d

t kl cd t d t d t d

(85)

2 ( 2 ) (86)

12

( 2 ) (87)

I Iklcd

kIc

lId

kc

ld

iaiIa

klcdiIc

klda

kIa

ildc

kIc

ilad

iaiIa

klcdiIc

klda

kIa

ildc

kIc

ilad

CIS2

CIS2

0 2 0

0MP2

Note that the MP2 correction energy cancels within CIS(D)and the final CIS(D) correction for CIS state I can berepresented as

∑ ∑

ωε ε ε ε

= −+ − − − Ω

+ ⟨ || ⟩ + +

s

t kl cd t d t d t d

14

( )

12

( 2 )

Iijab

ijab

a b i j I

iaiIa

klcdiIc

klda

kIa

ildc

kIc

ilad

CIS(D)2

CIS

(88)

CIS(D) is a second-order nondegenerate perturbationmethod on top of CIS and requires no extra diagonalizationbeyond the CIS Hamiltonian.

2.B.3. CIS(Dn). To generalize CIS(D) to the near-degeneratecase, there is a well-known family of quasi-degenerate second-order perturbation theories known as CIS(Dn).

28 WithinCIS(Dn), one diagonalizes a perturbed Hamiltonian thatcontains only single- and double-excitations to second order

=+†

⎝⎜⎜

⎠⎟⎟H

H H H

H H( )

SS SS SD

SD DD

CIS(D )(0) (2) (1)

(1) (0)n

(89)

where

= ⟨Φ | |Φ ⟩H H( )SS iajb ia

jb(0)

0 (90)

= ⟨Φ | |Φ ⟩H VT( )SS iajb ia

jb(2)

2 (91)

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

H

Page 9: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

= ⟨Φ | |Φ ⟩H V( )SD ia jbkc ia

jkbc(1)

, (92)

= ⟨Φ | |Φ ⟩H F( )DD iajb kcld ijab

klcd(0)

, (93)

Note that the ground-state response is taken into consid-eration through the HSS

(2) sub-block by including the double-excitation operator T2 from ground-state MP2 theory. Note alsothat a zeroth-order approximation is applied to the HDD

(0) sub-block so that only the diagonal elements appear in this sub-block.Thus, the HDD

(0) block can be formally inverted, leading to a muchsmaller, energy-dependent Hamiltonian matrix (and a corre-sponding self-consistent energy-dependent eigenvalue equation)

ω ω= + − − − †H H H H H( ) ( ) ( )SS SS SD DD SDCIS(D ) (0) (2) (1) (0) 1 (1)n

(94)

Note that (HDD(0) − ω)−1 can be written as an expansion

ω− = − Δ

= + Δ + Δ + ···

− − −

H H

H

1

1

( ) ( ) ( ) (95)

( ) ( ) (96)

DD DD

DD

(0) 1 (0) 1 1

(0) 1 2

where Δis defined as

ωΔ ≡ −H( )DD(0) 1

(97)

The CIS(Dn) hierarchy at level n is defined as the method thatresults from truncating the expansion after the Δn term. As aresult, the eigenvalue problem for CIS(D0) simply becomes

+ − |Ψ ⟩

= Ω |Ψ ⟩

− †H H H H H( ( ) ( ) )SS SS SD DD SD I

I I

(0) (2) (1) (0) 1 (1) CIS(D )

CIS(D ) CIS(D )

0

0 0 (98)

where we define the CIS(D0) Hamiltonian CIS(D )0 as

≡ + − − †H H H H H( ) ( )SS SS SD DD SDCIS(D ) (0) (2) (1) (0) 1 (1)0 (99)

∑ ∑

δ δ ε ε

δ δ

= − + ⟨ || ⟩

+ ⟨ || ⟩ + +

− ⟨ || ⟩ + ⟨ || ⟩

ja bi

jk bc d d d

ja bc b jk ib b

( )12

( 2 )

12

[ ]

iajb ij ab a i

kcij jk

caab ik

cbikac

cdlijlbcd

dkljklabd

CIS(D )0

(100)

where dikcb is defined in eqs 77 and 78 and bijk

abc is defined as

δ δ δ δ

ε ε ε ε≡

⟨ || ⟩ − ⟨ || ⟩ + ⟨ || ⟩ − ⟨ || ⟩+ − −

bab ci ab cj kb ij ka ij

ijkabc jk ik ac bc

a b i j

(101)

The effective CIS(D1) Hamiltonian CIS(D )1 is defined as

≡ − −CIS(D ) 1/2 CIS(D ) 1/21 0 (102)

where

≡ + − †H H H1 ( ) ( )SD DD SD(1) (0) 2 (1)

(103)

ε ε ε ε= +

∑ ⟨ || ⟩ + ∑ ⟨ || ⟩+ − −

ja bc b jk ib b1

2( )iajbcdl ijl

bcddkl jkl

abd

a b i j (104)

Finally, the full CIS(Dn) Hamiltonian given by eq 94 (withouttruncating (HDD

(0) − ω)−1) corresponds to the CIS(D∞)Hamiltonian. Solving the resulting energy-dependent eigenvalue

equation of the CIS(D∞) Hamiltonian (eq 105) self-consistentlygives the CIS(D∞) excitation energies

ω ω|Ψ ⟩ = |Ψ ⟩∞ ∞ ∞( ) I ICIS(D ) CIS(D ) CIS(D )

(105)

2.B.4. Relation between CIS(Dn) and CC2 and ADC(2).Although in this paper we will focus exclusively on the CIS(Dn)suite of excited states, it is important to recall that the CIS(D∞)Hamiltonian is nearly identical to several other excited-stateHamiltonians that are common in the literature.24−27,48 Amongthem, the CC2 model, which is also know as the approximatecoupled-cluster singles and doubles model, iteratively determinesthe singles and doubles substitutions as the poles of a true linearresponse function.24,25,48

Let us now remind the reader briefly of a few relevant detailspertaining to CC2. In standard second-order coupled-clustertheory (CCSD) for the electronic ground state,49 one performs asimilarity transformation to the unperturbed Hamiltonian suchthat

= − + + H He eT T T T( ) ( )1CC

2CC

1CC

2CC

(106)

where T1CC and T2

CC are first- and second-order coupled-clusteroperators

∑ = †T t a aia

ia a i1CC CC

(107)

∑ = † †T t a a a aijab

ijab a b i j2CC CC

(108)

The corresponding coefficients tiaCC and tijab

CC are determined viathe CCSD amplitude equations

∑ ⟨Φ | |Φ ⟩ =t H 0ia

ia iaCC

HF(109)

∑ ⟨Φ | |Φ ⟩ =t H 0ijab

ijab ijabCC

HF(110)

In CC2, the similarity transformation is approximated by

≈ + + Te e (1 )T T T( )2CC

1CC

2CC

1CC

(111)

and then one determines a ground-state CC2 energy and wavefunction in the same spirit as a CCSD calculation. Now, todetermine excited-state energies, with CC2, a standard equation-of-motion calculation is performed on top of the CC2 groundstate.50

With this background in mind, we remind the reader that, asshown by Hattig,48 the only essential difference between CC2and CIS(D∞) is that the double-excitation amplitudes are solvedvia ground-state MP2 theory in CIS(D∞), while they are solvedvia approximated coupled-cluster theory as the ground-statecluster amplitudes in CC2

ω ω= + − − − †H H H H H( ) ( ) ( )SS SS SD DD SDCC2 (0) (2) (1) (0) 1 (1)

(112)

= ⟨Φ | |Φ ⟩H VT( )SS iajb ia

jb(2)

2CC

(113)

Similarly, the model of algebraic diagrammatic constructionthrough second order (ADC(2)) is effectively the symmetric orthe Hermitian part of the CIS(D∞) model. More precisely, notethat HSS

(2) in eq 94 gives rise to a nonsymmetric form for theCIS(D∞) Hamiltonian. The ADC(2) Hamiltonian is just thesymmetrized form of the CIS(D∞) Hamiltonian

26,27,48

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

I

Page 10: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

ω ω ω= +†∞ ∞( )12

(( ( )) ( ))ADC(2) CIS(D ) CIS(D )

(114)

2.C. Model System. The Hubbard model offers one of thesimplest ways to get insight into how the interactions betweenelectrons can give rise to insulating, doped, and conductingeffects in a solid. In this work, we take a finite 1DHubbard model(without periodic boundary conditions) as our testing system,which is described by the Hamiltonian

∑ ∑

τ≡ + + +

+

μνμ ν μ ν μ

μμ μ μ μ

μμ μ μ μ

⟨ ⟩

† † † †

† †

a a a a V a a a a

U a a a a

( ) ( )

(115)

where μ denotes the site of the system; τ is the hopping integralbetween neighboring sites, Vμ is the on-site energy (for site μ),and U is the repulsion energy between two electrons of oppositespin occupying the same site μ. (Here we keep the repulsionenergy the same for all sites.) A bar indicates spin down. Notethat for all of the calculations below, we calculate only theexcitation energy of the first and second excited singlet states andwe set η = 0 in eq 64. Note also that, for this model problem,Wp(r, r′) (eq 55) is replaced by Wμμνν

p (eq 65).

3. RESULTSWe apply the theory above to an eight-site 1D Hubbard model,and we assume half-filling with eight electrons. The hoppingintegral between neighboring sites τ is set to be 1.0 for all

systems, while the electron repulsion energy U is varied. Allquantities will be calculated in au henceforward. The exactexcitation energies are obtained by diagonalizing the Hamil-tonian with direct diagonalization.

3.A. Metallic System. We first study metallic systems. Inmetallic systems, every site has the same on-site energy (Vμ = 0for all sites μ).51 The electron repulsion energyU is varied from 0to 0.7. Results are shown in Figure 3. One can see that for thissystem none of these approximated methods gives perfectresults. For both excited states, all of the CI methodsunderestimate the energy, and TDHF gives the largest negativeerrors here. Note that CIS generates larger negative errorscompared with correlated CI methods. We believe that thisfailure is really a failure of the HF calculation, whichoverestimates the ground-state energy: the difference betweenthe exact ground-state energy and the answer given by HF is aslarge as 0.08. In general, CIS is known to overestimate theexcitation energy.47,52 Pertubatively including the double-excitations gives slightly better results but still with an error aslarge as −0.02. Although the BSE-TDA method generallyoverestimates the energy (except for U = 0.7), this method givesrelatively smaller errors compared to CI methods. The BSE-RPAmethod generates themost accurate results for smallU values butunderestimates the energy whenU gets larger (e.g., whenU = 0.6and 0.7 for S1 and U = 0.7 for S2).

3.B. Doped Impurity Systems. By lowering the energy of asingle site in themetallic system, one canmodel an impurity. Thisdopes an otherwise metallic system51 and can influence theoverall photoactivitiy or electrical properties of the total system.

Figure 3. Errors for the excitation energy of the first (left, S1) and second (right, S2) excited singlet states for the metallic system (Vμ = 0 for all sites μ, τ =1.0). All errors are relative to the exact diagonalization. None of these methods gives perfect results. All of the CI methods underestimate the energywhile the BSE-TDAmethod generally overestimates the energy (except forU = 0.7). BSE-RPA is accurate for smallUs but underestimates the energy forlarge Us.

Figure 4. Errors for the excitation energy of the first (left, S1) and second (right, S2) excited singlet states for the doped system (with V5 = −0.5). Allerrors are relative to the exact diagonalization. Results in this doped system are very similar to those in the metallic system.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

J

Page 11: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

In Figure 4, we calculate results for electron repulsion energies Ubetween 0 and 0.7. Here the on-site energy Vμ is set to be−0.5. Itcan be seen in Figure 4 that the results are roughly identical to themetallic case: All of the CI methods underestimate the energywhile BSE-TDA overestimates the energy (except for U = 0.7)for both excited states. The BSE-TDA method again yields asmaller error compared to CI approaches. BSE-RPA is accuratefor small U values but gives large errors when U gets larger. Thesame conclusion holds for the case Vμ = −1.0 (where the on-siteenergy is comparable with the hopping integral).3.C. Molecular (Semiconducting/Insulating) Systems.

To simulate a “molecular” Hubbard model, we construct analternating Hamiltonian, whereby Veven is significantly lowered

relative toVodd = 0. In doing so, we expect that orbitals with lowerenergies will be doubly occupied and well separated from virtualorbitals. Thus, by creating an energy gap, we should be simulatingclosed-shell insulators and/or semiconductors (depending onthe gap size).For “molecular” systems with a small energy gap (Veven =

−0.5), the electron repulsion energy U is varied from 0 to 0.8. Asshown in Figure 5, the error given by the BSE-TDA method isfurther reduced, while the error of BSE-RPA increases. For the CIapproaches, we find a very unusual result: the CIS energyoutperforms the correlated methods. Clearly there is anonintuitive cancellation of error in these calculations.

Figure 5. Errors for the excitation energy of the first (left, S1) and second (right, S2) excited singlet states for the “molecular” system (Veven =−0.5,Vodd =0, τ = 1.0). All errors are relative to the exact diagonalization. BSE-TDAmaintains the same tendency and provides the smallest error whenU reaches itsmaximum. BSE-RPA underestimates the energy too much for large U. Higher-level CI approaches give worse results compared to CIS.

Figure 6. Errors for the excitation energy of the first (left, S1) and second (right, S2) excited singlet states for the “molecular” system (Veven =−1.0,Vodd =0, τ = 1.0). All errors are relative to the exact diagonalization. CIS overestimates the energy for S1 but gives the most accurate results for S2. BSE-TDA isonly better than CIS(D) and BSE-RPA, which generates the largest error in this case.

Figure 7. Errors for the excitation energy of the first (left, S1) and second (right, S2) excited singlet states for the “molecular” system (Veven =−2.0,Vodd =0, τ = 1.0). All errors are relative to the exact diagonalization. Both BSE methods underestimate the energy, while CIS and TDHF overestimate theenergy. CIS(D∞) gives the best results for S1, and yet CIS(D0) generates the best results for S2.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

K

Page 12: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

For “molecular” systems with an energy gap equal to thehopping integral (Veven = −1.0), Figure 6 demonstrates that ourresults are very different than all previous cases. Though CIS(D)still generates a large negative error, the CIS(Dn) approachesbehave much better now. Interestingly, CIS(D0) gives the bestresults in this case (and not CIS(D∞)). The performance of theBSE-TDA method is no longer as good, though it is slightlybetter than CIS(D). The BSE-RPA is the worst and generates thelargest error. Finally, CIS performs poorly for S1 but quite well forS2; again, there must be a nonintuitive cancellation of errors here.Finally, we consider the final, extreme “molecular” case

whereby the on-site energy difference is now twice as large as thehopping integral (Veven = −2.0). As Figure 7 shows, for S1, CISand TDHF generate quite large errors (as large as 0.05) as Uincreases from 0 to 1.2, overestimating the energy. CIS(D) andCIS(Dn) give smaller errors compared with the BSE methods asthe latter underestimate the energy. CIS(D∞) gives the mostaccurate results for S1; however, surprisingly, CIS(D0) providesthe most accurate results for S2. For S2, CIS and TDHF alsooverestimate the energy, but the errors are smaller comparedwith S1. BSE methods are not as reliable as CI methods in thiscase.Overall, clearly BSE methods perform best for metallic and

semiconducting systems, and CI methods perform best forinsulating systems. That being said, though, there is no obviouspattern for identifying the exact performance of each specificalgorithm.

4. DISCUSSION

The results above (in section 3) have compared CI methodsversus BSE methods for calculating the excitation energy of thefirst and second excited singlet states. For all of the cases studied,we did not find a universally optimal approach that can generaterelatively small errors for every system, though it is clear that theBSE-TDA method behaves better for metallic and semi-conducting systems and gradually loses its advantage when thesystem becomes more “molecular”. Here we analyze severalaspects that affect the performance of BSE: (i) the reliability ofthe GW approximation, (ii) the influence of the screenedcorrelation strength, and (iii) the effect of the non-self-consistentapproximation (i.e., ignoring the frequency dependence of Wp)for BSE.4.A.GWApproximation: Strengths andWeaknesses. In

ref 40, we demonstrated that the GWmethod holds comparativeadvantages versus traditional quantum chemistry approaches forcalculating the ionization potentials and electron affinities acrossa large range of Hubbard-like Hamiltonians. However, the sameconclusion did not hold for all orbitals (unpublished); in fact, wefound that GW can generate accurate orbital energies only forHOMO and LUMO for these model systems. To demonstratethis state of affairs, we consider the metallic system as an exampleand plot the Z-factor for the metallic system in Figure 8a. Onecan see in Figure 8a that only the HOMO and LUMO Z-factorsare consistently very close to 1, indicating that the GWapproximation works well for these two orbitals. For otherorbitals, however, the Z-factor in some cases is much smaller than1, which implies that a non-self-consistent G0W0 calculationcannot be reliable. To further demonstrate this, we plot thefrequency dependence of the self-energy for HOMO andHOMO−3 at U = 0.7 in Figure 8b. According to ref 40, weexpress the dynamic part of the self-energy for HOMO andHOMO−3 as

∑ ∑ ∑

ωω ε η

ω ε η

Σ =| |

ℏ + Ω − −

+| |

ℏ − Ω − +

⎛⎝⎜⎜

⎞⎠⎟⎟

ip jb ip kc

ap jb ap kcM

( )( )( )

i

( )( )i

ppc

jbkc I i I i

a I ajbkcI

TDH H

TDH H(116)

where p = HOMO or HOMO−3. (Note that ΩITDH in eq 116

corresponds to positive TDH eigenvalues.) It can been seen fromFigure 8b thatΣhomo

c (ω) (blue line) is very flat and smooth atω =

εhomo, giving rise to a very small derivativeε

ω∂ Σ

∂Re ( )c

homoH

. According

to eq 50, the Z-factor for HOMO is therefore very close to 1. Incontrast, Σhomo−3

c (ω) is very sharp at ω = εhomo−3, leading to avery small Z-factor (∼0.2 from Figure 8a), and indicating thefailure of the GW approximation for HOMO−3 at U = 0.7. Thesame conclusion holds for doped and molecular systems.In summary, the GW approximation works well when the

quantities −ΩITDH + εi

H and ΩITDH + εa

H are far away from theHartree orbital energies, which is true only for the HOMO andLUMO in our tested systems. (See eq 116.) Thus, as a

Figure 8. (a) Z-factors (eq 50) for the metallic systems and (b) thedynamic correlation part of the self-energyΣc (eq 116) as a function ofωfor HOMO and HOMO−3 at U = 0.7. In (a), only the HOMO andLUMO Z-factors are consistently very close to 1 for any repulsionenergy. In (b), Σhomo

c (ω) is very flat and smooth at ω = εhomo, whileΣhomo−3c (ω) is very sharp at ω = εhomo−3, indicating that GW (or, really,

G0W0) will be meaningful for the HOMO but not for HOMO−3 whenU = 0.7.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

L

Page 13: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

compromise, we shifted the orbital energy by the same amount(eqs 51 and 52) when performing all BSE calculations. The BSEresults might be improved in this paper if we can find moreaccurate quasiparticle energies.4.B. Screened Correlation Effects. The strong perform-

ance of BSE-TDA for the metallic and the semiconductingsystems in Figures 3−7 is consistent with the fact that BSE is

popular in the solid-state community where screening isessential. To visualize screening, in Figure 9, we plot the dynamicscreening interactionWμμνν

p for all systems, setting the repulsionenergy to U = 0.7. As shown in Figure 9, the metallic system anddoped systems (left panel) show more screening effects than themolecular systems (right panel), especially when the HOMO/LUMO gap is large. The effective screening for the extreme

Figure 9. Dynamic screening effects for various 1D Hubbard chains. Here we plotWμμννp (ω = 0) (eq 68). U = 0.7 for all systems. Note that the more

insulating the system is, the less screening it shows. The screening of the extreme molecular system (bottom-right) is only one-third in strengthcompared to the metallic and doped systems.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

M

Page 14: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

molecular system (bottom-right) is only one-third in strengthcompared to that of the metallic and doped systems. Thisdecrease in screening must degrade the performance of the BSE-TDA method. Future work must investigate whether thispreliminary conclusions holds up for ab initio calculations ormodel problems with long-range Coulombic electron−electronrepulsion energies (i.e., beyond the local interaction U of theHubbard model).4.C. Self-consistent and nonself-consistent BSE. Finally,

a few words are in order about self-consistency. In this work, wehave performed a non-self-consistent calculation for BSE byignoring the frequency dependence ofWp, i.e., expressing Kd viaeq 66. Alternatively, one can take the frequency dependence ofWp into account and apply the full expression of Kd (eq 47) tosolve the BSE equation self-consistently. To compare the results,we have in fact performed several self-consistent calculations, andwe find that for both TDA and RPA the self-consistent versionand the non-self-consistent version of BSE generate almostexactly the same results. Thus, any BSE errors reported here arenot from a lack of self-consistency.

5. CONCLUSIONS AND FUTURE DIRECTIONS

In this paper, we have reviewed BSE theory as well as a few widelyused CI theories and we have systematically compared theirperformances for the calculation of the first two excitationenergies for several finite 1D Hubbard chains. While BSE-RPAalways underestimates the excitation energy for systems withlarge repulsion energy and behaves poorly compared with CImethods, BSE-TDA slightly outperforms the CI methods formetallic and semiconducting systems. For insulating molecularsystems, BSE-TDA is not as accurate as CIS(D) andCIS(Dn). Toexplain these differences, the most obvious reason would be thechoice of initial orbitals and orbital energies, Hartree or Hartree−Fock. Note that in this study our starting point is a direct Hartreecalculation without any DFT functionals, and we have shown inref 40 that these Hartree orbitals can be unreliable without anycorrection. Thus, there is always the question of whether BSEwould be improved if more accurate quasiparticle energies can beobtained in some other fashion; recall that our GW calculationsyield meaningful orbital corrections only for the HOMO andLUMO. We intend to address this question shortly.In the future, several other questions must also be addressed.

First, in order to make contact with the solid-state and materialsliterature, it will be essential to work with larger model systems.In such a case, in order to push finite Hubbard models beyondeight sites and recover exact excitation energies, one will need toimplement a more sophisticated eigensolver, e.g., a densitymatrix renormalization solver.53 Alternatively, it will also becrucial to implement periodic boundary conditions.Second, as we have already demonstrated, the performance of

BSE (relative to quantum chemistry CI methods) is inextricablytied to the strength of the screening tensor (i.e., how well oneelectron’s charge is screened by the other electrons). For such aquantity, one may question whether the Hubbard model is thecorrect starting point for our analysis. How will our resultschange when we apply BSE and CI methods to truly ab initiocalculations or model problems with long-range Coulombicforces, where the screening tensor may look very different fromthe present article?Both of these outstanding questions will be addressed in a

future publication.

■ APPENDIX AHere we give a brief derivation of eq 4a starting from eq 3. Toderive eqs 4a and 4b, we first apply an external time-local one-electron potential u (x2, x2′; t). In the formulation of secondquantization, assuming a Schrodinger representation, thispotential takes the form

∫ ∫ = ′ Ψ ′ Ψ ′†H t u tx x x x x x( ) d d ( ) ( , ; ) ( )S

ext2 2 2 2 2 2 2 2 (A1)

Vice versa, if we define a quantity u(x2, t2; x2′, t2′) = u (x2, x2′;t2)δ(t2 − t2′), we may write

∫ ∫ ∫ = ′ ′ Ψ ′ ′ Ψ ′†H t t u t tx x x x x x( ) d d d ( ) ( , ; , ) ( )S

ext2 2 2 2 2 2 2 2 2 2

(A2)

For shorthand, as usual, we will now denote u(2, 2′)≡ u(x2, t2; x2′,t2′).Now, the total Hamiltonian is Htotal = H + Hext, where H is the

standard time-independent electronic Hamiltonian that includesall Coulomb interactions. In the interaction picture, the externalapplied Hamiltonian becomes

∫ ∫ ∫ =

= ′ ′ Ψ ′ Ψ ′ ′

H t H

t t u tx x x x

( ) e e (A3)

d d d ( , ) (2, 2 ) ( , ) (A4)

IHt

SHtext

2i ext i

2 2 2 2 2 2 2

2 2

At this point, if we invoke the adiabatic Gell-Mann and Lowtheorem,55 we find that the one- and two-particle Green’sfunctions can be written as

′ = −ℏ

⟨ | Ψ Ψ ′ | ⟩⟨ | | ⟩

GN T S N

N T S N(1, 1 )

i [ (1) (1 )][ ]1

0 0

0 0 (A5)

′ ′ = −ℏ

⟨ | Ψ Ψ Ψ ′ Ψ ′ | ⟩⟨ | | ⟩

† †

⎜ ⎟⎛⎝

⎞⎠G

N T S NN T S N

(12; 1 2 )i [ (1) (2) (2 ) (1 )]

[ ]2

20 0

0 0

(A6)

Here we have replaced |N⟩ (the true many-body ground state) by|N0⟩ (the ground state without any applied perturbation). Thefunction S is defined by

= −ℏ

= −ℏ

Ψ ′ Ψ

−∞

−∞

∞ †

⎡⎣⎢

⎤⎦⎥

⎡⎣⎢

⎤⎦⎥

S t H t

u

expi

d ( ) (A7)

expi

d2 (2) (2, 2 ) (2) (A8)

I2ext

2

In all expressions above, T denotes time ordering.Finally, it is straightforward to derive eq 4a in the paper. From

eqs A5 and A6, it is clear that the functional differential of G1(1,1′) is

δδ

δ

′ = −ℏ

⟨ | Ψ Ψ ′ | ⟩⟨ | | ⟩

− ′⟨ | | ⟩⟨ | | ⟩

GN T S N

N T S N

GN T S NN T S N

(1, 1 )i [ (1) (1 )]

[ ]

(1, 1 )[ ][ ]

10 0

0 0

10 0

0 0 (A9)

Furthermore, to evaluate δS , we differentiate eq A8 and find

∫δ δ δ = −ℏ

′ Ψ ′ ′ Ψ ′†

S T ui

d2 d2 (2) (2, 2 ) (2, 2 ) (2 )

(A10)

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

N

Page 15: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

and therefore

∫δ δ′ = ′ ′ − ′ ′

+ ′ ′

+

+

G u G

G G

(1, 1 ) d2 d2 (2, 2 )[ (12; 1 2 )

(1, 1 ) (2, 2 )]

1 2

1 1 (A11)

Note that we have added a positive sign to the t2′ time index toensure the correct ordering of field operators. The functionalderivative of G1(1, 1′) with respect to u(2, 2′) is thus

δδ

′′

= − ′ ′ + ′ ′+ +Gu

G G G(1, 1 )

(2, 2 )(12; 1 2 ) (1, 1 ) (2, 2 )1

2 1 1

(A12)

which completes the derivation of eq 4b. Finally, to derive eq 4a,we simply insert the definition u(2, 2′)≡ u(x2, t2; x2′, t2′) = u (x2, x2′;t2)δ(t2 − t2′), integrate over t2′, and take the derivative

δδ

′ ′G

u tx x(1, 1 )

( , ; ).1

2 2 2

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected]. Phone: (215)746-7078.ORCIDQi Ou: 0000-0002-6400-7522Present Address†Q.O.: Department of Mechanical and Aerospace Engineering,Princeton University, Princeton, NJ 08544-5263.FundingThis work was supported by NSF CAREER Grant CHE-1150851. J.E.S. gratefully acknowledges support from theStanford PULSE Institute and a John Simon GuggenheimMemorial fellowship.NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThe authors thank Leeor Kronik for very useful conversations.

■ REFERENCES(1) Foresman, J. B.; Head-Gordon, M.; Pople, J. A.; Frisch, M. J.Toward a Systematic Molecular Orbital Theory for Excited States. J.Phys. Chem. 1992, 96, 135−149.(2) Stratmann, R. E.; Scuseria, G. E.; Frisch, M. J. An efficientimplementation of time-dependent density-functional theory for thecalculation of excitation energies of large molecules. J. Chem. Phys. 1998,109, 8218−8224.(3) Onida, G.; Reining, L.; Rubio, A. Electronic excitations: density-functional versus many-body Green’s-function approaches. Rev. Mod.Phys. 2002, 74, 601−659.(4) Adler, D.; Feinleib, J. Electrical and Optical Properties of Narrow-Band Materials. Phys. Rev. B 1970, 2, 3112−3134.(5) Scholes, G. D. Long-Range Resonance Energy Transfer inMolecular Systems. Annu. Rev. Phys. Chem. 2003, 54, 57−87.(6) Autschbach, J.; Ziegler, T.; van Gisbergen, S. J. A.; Baerends, E. J.Chiroptical properties from time-dependent density functional theory. I.Circular dichroism spectra of organic molecules. J. Chem. Phys. 2002,116, 6930−6940.(7) Chang, E.; Bussi, G.; Ruini, A.; Molinari, E. Excitons in CarbonNanotubes: An Ab Initio Symmetry-Based Approach. Phys. Rev. Lett.2004, 92, 196401.(8) Spataru, C. D.; Ismail-Beigi, S.; Capaz, R. B.; Louie, S. G. Theoryand Ab Initio Calculation of Radiative Lifetime of Excitons inSemiconducting Carbon Nanotubes. Phys. Rev. Lett. 2005, 95, 247402.

(9) Lopez-Acevedo, O.; Tsunoyama, H.; Tsukuda, T.; Hakkinen, H.;Aikens, C. M. Chirality and Electronic Structure of the Thiolate-Protected Au38 Nanocluster. J. Am. Chem. Soc. 2010, 132, 8210−8218.(10) Gierschner, J.; Cornil, J.; Egelhaaf, H.-J. Optical Bandgaps of π-Conjugated Organic Materials at the Polymer Limit: Experiment andTheory. Adv. Mater. 2007, 19, 173−1912.(11) Elward, J. M.; Irudayanathan, F. J.; Nangia, S.; Chakraborty, A.Optical Signature of Formation of Protein Corona in the FireflyLuciferase-CdSe Quantum Dot Complex. J. Chem. Theory Comput.2014, 10, 5224−5228.(12) Berkelbach, T. C.; Hybertsen, M. S.; Reichman, D. R. Bright anddark singlet excitons via linear and two-photon spectroscopy inmonolayer transition-metal dichalcogenides. Phys. Rev. B: Condens.Matter Mater. Phys. 2015, 92, 085413.(13) Kronik, L.; Neaton, J. B. Excited-State Properties of MolecularSolids from First Principles. Annu. Rev. Phys. Chem. 2016, 67, 587−616.(14) Yarkony, D. R. Nonadiabatic effects in the vicinity of multiplesurface crossings. Evaluation of derivative couplings with respect torotational and internal degrees of freedom. Application to the chargetransfer reaction H++NO → H+NO+. J. Chem. Phys. 1989, 90, 1657−1665.(15) Yarkony, D. R. On the consequences of nonremovable derivativecouplings 0.1. The geometric phase and quasidiabatic states: Anumerical study. J. Chem. Phys. 1996, 105, 10456−10461.(16) Matsika, S.; Yarkony, D. R. Spin-orbit coupling and conicalintersections in molecules with an odd number of electrons. III. Aperturbative determination of the electronic energies, derivativecouplings and a rigorous diabatic representation near a conicalintersection. J. Chem. Phys. 2002, 116, 2825−2835.(17) Dallos, M.; Lischka, H.; Shepard, R.; Yarkony, D. R.; Szalay, P. G.Analytic evaluation of nonadiabatic coupling terms at the MR-CI level.II. Minima on the crossing seam: Formaldehyde and the photo-dimerization of ethylene. J. Chem. Phys. 2004, 120, 7330−7339.(18) Marques, M. A.; Maitra, N. T.; Nogueira, F. M.; Gross, E.; Rubio,A. Fundamentals of Time-Dependent Density Functional Theory; SpringerScience & Business Media: Berlin, Germany, 2012.(19) Ullrich, C. A. Time-Dependent Density-Functional Theory: Conceptsand Applications; Oxford University Press: New York, 2012.(20) Bene, J. E. D.; Ditchfield, R.; Pople, J. A. Self-ConsistentMolecular Orbital Methods. X. Molecular Orbital Studies of ExcitedStates with Minimal and Extended Basis Sets. J. Chem. Phys. 1971, 55,2236−2241.(21) McLachlan, A. D.; Ball, M. A. Time-Dependent Hartree-FockTheory for Molecules. Rev. Mod. Phys. 1964, 36, 844−855.(22) Head-Gordon, M.; Rico, R. J.; Oumi, M.; Lee, T. J. A doublescorrection to electronic excited states from configuration interaction inthe space of single substitutions. Chem. Phys. Lett. 1994, 219, 21−29.(23) Head-Gordon, M.; Grana, A. M.; Maurice, D.; White, C. A.Analysis of Electronic Transitions as the Difference of ElectronAttachment and Detachment Densities. J. Phys. Chem. 1995, 99,14261−14270.(24) Christiansen, O.; Koch, H.; Jørgensen, P. The second-orderapproximate coupled cluster singles and doubles model CC2. Chem.Phys. Lett. 1995, 243, 409−418.(25) Christiansen, O.; Koch, H.; Jørgensen, P.; Helgaker, T. Integraldirect calculation of CC2 excitation energies: singlet excited states ofbenzene. Chem. Phys. Lett. 1996, 263, 530−539.(26) Schirmer, J. Beyond the random-phase approximation: A newapproximation scheme for the polarization propagator. Phys. Rev. A: At.,Mol., Opt. Phys. 1982, 26, 2395−2416.(27) Trofimov, A. B.; Schirmer, J. An efficient polarization propagatorapproach to valence electron excitation spectra. J. Phys. B: At., Mol. Opt.Phys. 1995, 28, 2299−2324.(28) Head-Gordon, M.; Oumi, M.; Maurice, D. Quasidegeneratesecond-order perturbation corrections to single excitation configurationinter action. Mol. Phys. 1999, 96, 593−602.(29) Heyd, J.; Scuseria, G. E. Efficient hybrid density functionalcalculations in solids: Assessment of the HeydScuseriaErnzerhof

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

O

Page 16: Comparison between the Bethe Salpeter ... - Subotnik Groupsubotnikgroup.chem.upenn.edu/publications/qi_bse_2018.pdffrom the Coulomb potential in eq 13, is usually called the exchangeterm,whilethesecondtermKd,whichresultsfromthe

screened Coulomb hybrid functional. J. Chem. Phys. 2004, 121, 1187−1192.(30) Aufray, B.; Kara, A.; Vizzini, S.; Oughaddou, H.; Leandri, C.; Ealet,B.; Le Lay, G. Graphene-like silicon nanoribbons on Ag(110): Apossible formation of silicene. Appl. Phys. Lett. 2010, 96, 183102.(31) Mosconi, E.; Amat, A.; Nazeeruddin, M. K.; Gratzel, M.; DeAngelis, F. First-Principles Modeling of Mixed Halide OrganometalPerovskites for Photovoltaic Applications. J. Phys. Chem. C 2013, 117,13902−13913.(32) Fuchs, F.; Bechstedt, F. Indium-oxide polymorphs from firstprinciples: Quasiparticle electronic states. Phys. Rev. B: Condens. MatterMater. Phys. 2008, 77, 155107.(33) Aryasetiawan, F.; Gunnarsson, O. The GW method. Rep. Prog.Phys. 1998, 61, 237.(34) Hedin, L. On Correlation Effects in Electron Spectroscopies andthe GW Approximation. J. Phys.: Condens. Matter 1999, 11, R489.(35) Rohlfing, M.; Louie, S. G. Electron-hole excitations and opticalspectra from first principles. Phys. Rev. B: Condens. Matter Mater. Phys.2000, 62, 4927−4943.(36) Rabani, E.; Baer, R.; Neuhauser, D. Time-dependent stochasticBethe-Salpeter approach. Phys. Rev. B: Condens. Matter Mater. Phys.2015, 91, 235302.(37) Neuhauser, D.; Gao, Y.; Arntsen, C.; Karshenas, C.; Rabani, E.;Baer, R. Breaking the Theoretical Scaling Limit for PredictingQuasiparticle Energies: The Stochastic GW Approach. Phys. Rev. Lett.2014, 113, 076402.(38) Hedin, L. New Method for Calculating the One-Particle Green’sFunction with Application to the Electron-Gas Problem. Phys. Rev.1965, 139, A796−A823.(39) Salpeter, E. E.; Bethe, H. A. A Relativistic Equation for Bound-State Problems. Phys. Rev. 1951, 84, 1232−1242.(40) Ou, Q.; Subotnik, J. E. Comparison between GW and Wave-Function-Based Approaches: Calculating the Ionization Potential andElectron Affinity for 1D Hubbard Chains. J. Phys. Chem. A 2016, 120,4514−4525.(41) Strinati, G. Application of the Green’s Function Method to theStudy of the Optical Properties of Semiconductors. Riv. Nuovo Cim.1988, 11, 1.(42) Note that if we express the time-dependent creationand annih i l a t ion operator as ψ ψ = −tx x( , ) e ( )eHt Hti i and

ψ ψ = † ℏ † − ℏtx x( , ) e ( )eHt Hti / i / , we can rewrite χSeh and χ Seh as

χ ψ ψ

χ ψ ψ

− = ⟨ | | ⟩+ − ℏ

− = ⟨ | | ⟩− + − ℏ

− − ℏ †

− − ℏ †

t t N NE E t t

t t N NE E t t

x x x x

x x x x

( , ; ) ( )e ( )exp[i( )( )/2 ]

( , ; ) ( )e ( )exp[ i( )( )/2 ]

SH t t

S

S

S SH t t

S

eh1 2 1 2 1

i ( )/2

0 1 2

eh1 2 1 2 1

i ( )/2

0 1 2

1 2

1 2

for t1 > t2 and

χ ψ ψ

χ ψ ψ

− = −⟨ | | ⟩+ − ℏ

− = −⟨ | | ⟩− + − ℏ

† − − ℏ

† − − ℏ

t t N NE E t t

t t N NE E t t

x x x x

x x x x

( , ; ) ( )e ( )exp[i( )( )/2 ]

( , ; ) ( )e ( )exp[ i( )( )/2 ]

SH t t

S

S

S SH t t

S

eh1 2 1 2 2

i ( )/1

0 2 1

eh1 2 1 2 2

i ( )/1

0 2 1

2 1

2 1

for t1 < t2. This representation proves that χSeh and χ Seh are functions of

only t1 − t2.(43) For GIII and GIV, we must have t2 > t2′ or t2 < t2′, and therefore, thelimit of t2′ = t2 + δ is not meaningful.(44) Inkson, J. C.Many-Body Theory of Solids: An Introduction; PlenumPress: New York, 1984.(45) Bruneval, F.; Gonze, X. AccurateGW self-energies in a plane-wavebasis using only a few empty states: Towards large systems. Phys. Rev. B:Condens. Matter Mater. Phys. 2008, 78, 085125.(46) Furche, F. On the density matrix based approach to time-dependent density functional response theory. J. Chem. Phys. 2001, 114,5982−5992.

(47) Dreuw, A.; Head-Gordon, M. Single reference ab initio methodsfor the calculation of excited states of large molecules. Chem. Rev. 2005,105, 4009.(48) Hattig, C. Structure Optimizations for Excited States withCorrelated Second-Order Methods: CC2 and ADC(2). Adv. QuantumChem. 2005, 50, 37−60.(49) Purvis, G. D., III; Bartlett, R. J. A full coupled-cluster singles anddoubles model: The inclusion of disconnected triples. J. Chem. Phys.1982, 76, 1910−1918.(50) One should note that the terms introduced via the transformationon H0 in CC2 contribute to only the third and higher orders to theexcitation energies of single-excitation dominated transitions. To obtainthe excitation energies correct through the second order, thistransformation can safely be discarded.(51) In truth, it can be shown that an infinite half-filled Hubbard chainis always insulating.54 Thus, formally, any “metallic” features of thisHubbard chain exist only because of the chain’s finite size. That beingsaid, the main conclusions presented here continue to hold if weconsider a chain with less than half filling.(52) Subotnik, J. E. Communication: Configuration interaction singleshas a large systematic bias against charge-transfer states. J. Chem. Phys.2011, 135, 071104.(53) Chan, G. K.-L.; Sharma, S. The Density Matrix RenormalizationGroup in Quantum Chemistry. Annu. Rev. Phys. Chem. 2011, 62, 465−481.(54) Lieb, E. H.; Wu, F. Y. Absence of Mott Transition in an ExactSolution of the Short-Range, One-BandModel in One Dimension. Phys.Rev. Lett. 1968, 20, 1445−1448.(55) Fetter, A. L.; Walecka, J. D. Quantum Theory of Many-articleSystems; Dover Publications, Inc.: New York, 2003.(56) Note that we are not including here the Fourier transform of thefirst term on the right-hand side of eq 14. This term has been omittedbecause there is no residue at ℏω = Es − E0.

Journal of Chemical Theory and Computation Article

DOI: 10.1021/acs.jctc.7b00246J. Chem. Theory Comput. XXXX, XXX, XXX−XXX

P