visualizing the contributions of virtual states to two...

10
Visualizing the Contributions of Virtual States to Two-Photon Absorption Cross Sections by Natural Transition Orbitals of Response Transition Density Matrices Kaushik D. Nanda and Anna I. Krylov* Department of Chemistry, University of Southern California, Los Angeles, California 90089-0482, United States * S Supporting Information ABSTRACT: Observables such as two-photon absorption cross sections cannot be computed from the wave functions of initial and nal states alone because of their nonlinear nature. Rather, they depend on the entire manifold of the excited states, which follows from the familiar sum-over-states expressions of second- and higher-order properties. Consequently, the interpretation of the computed nonlinear optical properties in terms of molecular orbitals is not straightforward and usually relies on approximate few-states models. Here, we show that the two-photon absorption (2PA) transitions can be visualized using response one-particle transition density matrices, which are dened as transition density matrices between the zero-order and rst-order perturbed states. We also extend the concept of natural transition orbitals to 2PA transitions. We illustrate the utility of this new tool, which provides a rigorous black box alternative to traditional qualitative few-states analysis, by considering 2PA transitions in ethylene, trans-stilbene, and para-nitroaniline. N onlinear optical properties enable novel technologies and powerful spectroscopic techniques. In particular, two- photon absorption (2PA) spectroscopy is exploited in photo- dynamic cancer therapy, optogenetics, 3D optical storage, high- resolution bioimaging, optical limiting, etc. 15 Owing to dierent selection rules, 69 2PA spectroscopy provides rich information about molecular electronic structure, which is complementary to the information provided by one-photon absorption (1PA) spectroscopy. Theoretical modeling allows one to interpret the observed 1PA and 2PA spectra and to connect the spectral features with the underlying molecular structure and wave functions. 10 The ability to understand spectroscopic properties in terms of molecular orbital theory is critically important for rationalizing the trends and developing guidelines for the design of new chromophores with desired properties. 1013 Reduced transition density matrices, 14,15 such as transition 1PDM (one-particle density matrix, γ pq ), enable concise description of the electronic transition between the two states, Ψ i and Ψ f γ = ⟨Ψ| ̂ ̂|Ψpq pq f i f i (1) where p̂ and q̂ are the creation and annihilation operators corresponding to ϕ p and ϕ q molecular orbitals. In the case of pseudoindependent electrons, Ψ i is just a single Slater determinant (Φ 0 ), Ψ f is a linear combination of singly excited determinants, C j b Φ j b , and γ pq is equal to the wave function amplitudes, γ pq = δ pb δ qj C j b . Thus, the individual elements of γ correspond to the weights of the electronic transitions between the respective molecular orbitals. Equation 1 allows one to extend this simple molecular orbital picture of electronic transitions to general correlated wave functions: 15,16 γ fi gives a map of a change in one-electron density between Ψ f and Ψ i , which can be represented as a linear combination of one-electron excitations: γ |Ψ⟩= ̂ ̂|Ψ⟩ + pq higher excitations f pq pq f i i (2) One can also interpret the transition density as the excitons wave function 1517 γ ϕ ϕ Ψ = rr r r ( , ) ( ) ( ) pq pq f i exc h p p h (3) where r p and r h denote particle (electron) and hole coordinates. The plot of γ pq in spatial representation provides a picture of the exciton, i.e., the change in electronic distribution between the initial and nal states. 15,16,18,19 One can also compute various quantities characterizing the exciton, e.g., average separation between the hole and the electron = ⟨Ψ | d rr r r rr ( , )( ) ( , ) r r exc exc h p h p 2 exc h p , h p (4) This quantity is related to the exciton size. By using orbital transformations, one can represent γ in the most compact way, such that the electronic transition between the two states is described by the least number of molecular orbital pairs. The transformation that achieves this goal is given by the singular value decomposition (SVD) of γ, and the resulting pairs of Received: June 5, 2017 Accepted: June 26, 2017 Published: June 26, 2017 Letter pubs.acs.org/JPCL © XXXX American Chemical Society 3256 DOI: 10.1021/acs.jpclett.7b01422 J. Phys. Chem. Lett. 2017, 8, 32563265

Upload: others

Post on 04-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

Visualizing the Contributions of Virtual States to Two-PhotonAbsorption Cross Sections by Natural Transition Orbitals of ResponseTransition Density MatricesKaushik D. Nanda and Anna I. Krylov*

Department of Chemistry, University of Southern California, Los Angeles, California 90089-0482, United States

*S Supporting Information

ABSTRACT: Observables such as two-photon absorption cross sections cannot becomputed from the wave functions of initial and final states alone because of their nonlinearnature. Rather, they depend on the entire manifold of the excited states, which follows fromthe familiar sum-over-states expressions of second- and higher-order properties.Consequently, the interpretation of the computed nonlinear optical properties in terms ofmolecular orbitals is not straightforward and usually relies on approximate few-states models.Here, we show that the two-photon absorption (2PA) transitions can be visualized usingresponse one-particle transition density matrices, which are defined as transition densitymatrices between the zero-order and first-order perturbed states. We also extend the conceptof natural transition orbitals to 2PA transitions. We illustrate the utility of this new tool,which provides a rigorous black box alternative to traditional qualitative few-states analysis, byconsidering 2PA transitions in ethylene, trans-stilbene, and para-nitroaniline.

Nonlinear optical properties enable novel technologies andpowerful spectroscopic techniques. In particular, two-

photon absorption (2PA) spectroscopy is exploited in photo-dynamic cancer therapy, optogenetics, 3D optical storage, high-resolution bioimaging, optical limiting, etc.1−5 Owing to differentselection rules,6−9 2PA spectroscopy provides rich informationabout molecular electronic structure, which is complementary tothe information provided by one-photon absorption (1PA)spectroscopy. Theoretical modeling allows one to interpret theobserved 1PA and 2PA spectra and to connect the spectralfeatures with the underlying molecular structure and wavefunctions.10 The ability to understand spectroscopic propertiesin terms of molecular orbital theory is critically important forrationalizing the trends and developing guidelines for the designof new chromophores with desired properties.10−13

Reduced transition density matrices,14,15 such as transition1PDM (one-particle density matrix, γpq), enable concisedescription of the electronic transition between the two states,Ψi and Ψf

γ = ⟨Ψ | |Ψ⟩← †p qpqf i

f i (1)

where p† and q are the creation and annihilation operatorscorresponding to ϕp and ϕq molecular orbitals. In the case ofpseudoindependent electrons, Ψi is just a single Slaterdeterminant (Φ0), Ψf is a linear combination of singly exciteddeterminants, ∑ Cj

b Φjb, and γpq is equal to the wave function

amplitudes, γpq = δpbδqjCjb. Thus, the individual elements of γ

correspond to the weights of the electronic transitions betweenthe respective molecular orbitals. Equation 1 allows one toextend this simple molecular orbital picture of electronictransitions to general correlated wave functions:15,16 γf←i gives

a map of a change in one-electron density between Ψf and Ψi,which can be represented as a linear combination of one-electronexcitations:

∑ γ|Ψ ⟩ = |Ψ⟩ +← †p q higher excitationsfpq

pqf i

i(2)

One can also interpret the transition density as the exciton’s wavefunction15−17

∑ γ ϕ ϕΨ = ←r r r r( , ) ( ) ( )pq

pqf i

exc h p p h(3)

where rp and rh denote particle (electron) and hole coordinates.The plot of γpq in spatial representation provides a picture of theexciton, i.e., the change in electronic distribution between theinitial and final states.15,16,18,19 One can also compute variousquantities characterizing the exciton, e.g., average separationbetween the hole and the electron

= ⟨Ψ | − |Ψ ⟩d r r r r r r( , ) ( ) ( , ) r rexc exc h p h p2

exc h p ,h p (4)

This quantity is related to the exciton size. By using orbitaltransformations, one can represent γ in the most compact way,such that the electronic transition between the two states isdescribed by the least number of molecular orbital pairs. Thetransformation that achieves this goal is given by the singularvalue decomposition (SVD) of γ, and the resulting pairs of

Received: June 5, 2017Accepted: June 26, 2017Published: June 26, 2017

Letter

pubs.acs.org/JPCL

© XXXX American Chemical Society 3256 DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

Page 2: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

molecular orbitals are called natural transition orbitals(NTOs)15,16,20

γ = Σ← V Uf i T (5)

where the diagonal matrix Σ contains the singular values σK andmatrices V and U contain the molecular orbital coefficients forthe particle and hole states

ψ ϕ

ψ ϕ

=

=

r V r

r U r

( ) ( )

( ) ( )

Kq

qK q

Kq

qK q

p

h

(6)

where indexKmarks theNTO pair corresponding to the singularvalue σK. The NTO representation of the electronic transitionsremoves the arbitrariness associated with a specific choice ofmolecular orbitals and allows one to describe the exciton directlyand in the most compact form.15,16,20 Orbital choice affects thewave function amplitudes (and, consequently, the resultingtransition 1PDMs), such that a wave function with a singlenonzero amplitude in one basis can be transformed into a wavefunction with multiple amplitudes with similar weights in anotherbasis. The SVD procedure produces an essential description ofthe transition. Usually, only a small number of singular values aresubstantial; so by using the NTOs, one can represent thetransition density ρf←i(r) and the exciton wave function by a(very) small number of terms

∑ ∑ρ γ ϕ ϕ σ ψ ψ= =← ←r r r r r( ) ( ) ( ) ( ) ( )f i

pqpqf i

p qK

K K Kp h

(7)

∑ σ ψ ψΨ =r r r r( , ) ( ) ( )K

K K Kexc h pp

ph

h(8)

NTO analysis allows one to distinguish between the states thatcan be described by a single NTO pair (such as a pure valence π→ π* excitation) or the states that have mixed character (e.g., amixed π → π* and π → Ry(3s) excitation). Quantitatively, thenumber of distinct configurations contributing to a particulartransition is given by the participation ratio, defined as

σσ σ

=∑∑

= Ω∑

PR( )K K

K K K KNTO

2 2

4

2

4(9)

whereΩ = ∥γ∥. For example, forΩ = 1 (which is the case for thetransition 1PDM for a transition between a reference and a CIStarget state), PRNTO = 1 for a CIS target state with exactly oneconfiguration, 2 for a CIS state represented by the twoconfigurations with equal weights, and so on. PRNTO is closelyrelated to the number of entangled states, an alternative metric17

also defined on the basis of the transition 1PDMs. Bothquantities provide a measure of collectivity of electronicexcitation and quantum entanglement.As per eq 2, the norm of transition 1PDM (Ω) quantifies the

degree of one-electron character of a transition; it equals 1 forpurely one-electron transitions and is smaller than 1 fortransitions involving changes of state of more than one electron(e.g., Ω = 0 for purely two-electron transitions). This quantitycan be used to estimate an upper bound of certain transitionproperties.21,22 NTO analysis also enables comparisons of thecharacters of excited states computed by different electronicstructure methods, including the comparison between the wavefunction and density functional theories.

Importantly, γ contains all the information about electronictransition needed to compute one-electron transition properties.For example, the transition dipole moment is

∑μ γ μ α⟨Ψ | |Ψ⟩ = =α α← x y z, ,f ipq

pqf i

pq(10)

where μα is the matrix of the dipole moment operator, μpqα =

⟨ϕp|μα|ϕq⟩. Since γ is directly related to the experimental

observable (e.g., oscillator strength), its analysis providesimportant insight regarding the nature of electronic transition.For example, by visualizing γ or NTOs,14,15,18−20,23−26 one canidentify local and charge-transfer excitations, distinguish betweenRydberg and valence excited states, and connect the strength ofthe transition with the shape of the exciton (e.g., increaseddelocalization of the exciton in conjugated molecules leads to anincrease in oscillator strength, as can be readily seen from eq 10).The NTO representation allows one to express the observablesin terms of matrix elements between just a few pairs of orbitals

∑μ σ ψ μ ψ⟨Ψ | |Ψ⟩ = ⟨ | | ⟩α αf i

KK K K

p h

(11)

thus representing matrix elements between many-electron statesas matrix elements between the orbitals.Transition 1PDMs and the NTOs can be computed for

electronic states probed by 2PA spectroscopy, as was done, forexample, in refs 26 and 27. Whereas these quantities correctlyrepresent the differences between Ψi and Ψf, they are no longerrelated to the underlying observables, the 2PA transitionmoments, which are given by sum-over-all-states expres-sions:28−31

ω ω

μ μω

μ μω

= −⟨Ψ | |Ψ ⟩⟨Ψ | |Ψ⟩

Ω −+

⟨Ψ | |Ψ ⟩⟨Ψ | |Ψ⟩Ω −

⎛⎝⎜⎜

⎞⎠⎟⎟

M ( , )xyf i

x y

n

fy

n nx

i

ni x

fx

n ny

i

ni y

1 2

1 2

(12)

ω ω

μ μω

μ μω

− −

= −⟨Ψ| |Ψ ⟩⟨Ψ | |Ψ ⟩

Ω −+

⟨Ψ| |Ψ ⟩⟨Ψ | |Ψ ⟩Ω −

⎛⎝⎜⎜

⎞⎠⎟⎟

M ( , )xyi f

x y

n

ix

n ny

f

ni x

iy

n nx

f

ni y

1 2

1 2

(13)

Here, Ωni = En − Ei and ω1 and ω2 are the frequencies of the twoabsorbed photons such that ω1 + ω2 =Ωf i. Note that the index xin ω1x indicates the polarization of this photon and ω1x = ω1 andω2y = ω2.Equations 12 and 13 reflect important physics of the 2PA cross

sections, as well as other nonlinear optical properties, i.e., theproperties of the transition can no longer be understood byconsidering only the initial and final states. Rather, all excitedstates contribute to the observable. In order to gain insight intothese phenomena, few-states models (in which the sum-over-states expressions are approximated by a few leading terms) areoften exploited.11−13,19,26,27,31 For example, ref 19 described the2PA transitions in donor-bridge-acceptor chromophores byplotting transition 1PDMs corresponding to the leading terms insum-over-states expressions and rationalized the respective 2PAcross sections in terms of the spatial representation of thetransition 1PDMs. Such analysis, however, relies on identifyingthe intermediate states that provide the leading contributionsand is rather tedious and approximate. It is well-known that theconvergence of the sum-over-states expressions is slow and that

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3257

Page 3: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

estimates of 2PA cross sections obtained from few-states modelsdiffer significantly from the values computed using completeexpressions.26,27 In this Letter, we introduce a new approach foranalyzing 2PA transitions by extending the NTO concept to theperturbed transition density matrices. This analysis allows us todescribe the character of 2PA transitions without resorting tofew-states models. Rather, the analysis produces a molecularorbital picture of the virtual states that provide the leadingcontributions to the 2PA transition moments.We begin by recasting the 2PA transition moments as

ω

μ μω

μ μω

= −⟨Ψ | |Ψ ⟩⟨Ψ | |Ψ⟩

Ω −+

⟨Ψ | |Ψ ⟩⟨Ψ | |Ψ⟩Ω +

⎛⎝⎜⎜

⎞⎠⎟⎟

M ( )xyf i

x

n

fy

n nx

i

ni x

fx

n ny

i

nf x

1

1 1

(14)

ω

μ μω

μ μω

= −⟨Ψ| |Ψ ⟩⟨Ψ | |Ψ ⟩

Ω ++

⟨Ψ| |Ψ ⟩⟨Ψ | |Ψ ⟩Ω −

⎛⎝⎜⎜

⎞⎠⎟⎟

M ( )xyi f

y

n

ix

n ny

f

nf y

iy

n nx

f

ni y

2

2 2

(15)

where we use the resonant condition ω1x + ω2y = Ωf i to expressthe 2PA transition moments as a function of just one photonfrequency. To derive practical expressions that circumventexplicit calculations of all excited states, the above expressions arerewritten in the following form

ω μ μ= − ⟨Ψ | | ⟩ + ⟨ | |Ψ⟩ω ω←M X X( ) ( )xyf i

x fy

i fy

i1x x1 1

(16)

ω μ μ− = − ⟨Ψ| | ⟩ + ⟨ | |Ψ ⟩ω ω← − −M X X( ) ( )xyi f

y ix

f ix

f2y y2 2

(17)

where the response vectors, X and X, are computed by solvingauxiliary response equations.28−30,32 The resulting approach isformally and numerically equivalent to the sum-over-statesexpressions. Within the formalism outlined in refs 30 and 32, theresponse vectors are

ω μ − + = ⟨Ψ| |Φ ⟩ωX H E( )k k x kx

I1x1 (18)

ω μ− − = ⟨Φ | |Ψ⟩ωH E X( )k x k Ix

k1x1 (19)

where ΦI’s denote Slater determinants from the target-statesmanifold. The response vectors are related to the first-orderresponse33 of Ψk to the perturbation with frequency ωx:

∑ μω

| ⟩ ≡ |Ψ ⟩ = |Ψ ⟩⟨Ψ | |Ψ⟩Ω −

ωXk kn

nn

xk

nk x

(1)

1

x1

(20)

∑ μω

⟨ | ≡ ⟨Ψ | =⟨Ψ | |Ψ⟩Ω +

⟨Ψ |ωXk kn

nx

k

nk xn

(1)

1

x1

(21)

(The first-order response vector for state k is not orthogonal tothe zero-order state k unless one uses the fluctuation operator,μ − ⟨Ψk|μ|Ψk⟩, instead of the dipole operator in perturbationtheory.)Within the equation-of-motion coupled-cluster singles and

doubles (EOM-CCSD) framework,34−36 we rewrite eqs 16−19by replacing the Hamiltonian H with the similarity-transformedHamiltonian (H = e−THeT), the dipole operator μwith similarity-transformed dipole operator (μ = e−TμeT), and ⟨Ψk| and |Ψk⟩withthe left and right EOM amplitudes (⟨Φ0L

k| and |RkΦ0⟩,respectively).30

The inspection of eqs 16 and 17 reveals that the 2PA transitionmoments can be computed as the sum of the two transitiondipole moments: one between Ψi and the first-order perturbedfinal state (Ψf

(1)) and another between Ψf and the first-orderperturbed initial state (Ψi

(1)). Equations 16 and 17 can then becast in the following form

∑ω γ μ= ω←

+←M ( ) [ ]xy

f ix

pq

f ipq pq

y1 x1

(22)

∑ω γ μ− = ω←

−←M ( ) [ ]xy

i fy

pq

i fpq pq

x2 y2

(23)

where we introduced new 1PDM-like quantities, the responsetransition 1PDMs, γ±ω (denoted ωDM below)

γ γ γ= ∥ ∥ + ∥ ∥ωω

ωω

ω+← +

+← +

+ ←X X[ ] [ ] [ ]f i

pq if X

pq fX i

pqx

x

x

i x

x

f

1

1

1

1

1 (24)

γ γ γ= ∥ ∥ + ∥ ∥ωω

ωω

ω−← −

−← −

− ←X X[ ] [ ] [ ]i f

pq fi X

pq iX f

pqy

y

y

f y

y

i

2

2

2

2

2 (25)

where ∥Xi,f∥ denotes the norms of the response vectors and

γ

γ

γ

γ

= −⟨Ψ | |∥ ∥

= −⟨

∥ ∥| |Ψ⟩

= −⟨Ψ| |∥ ∥

= −⟨

∥ ∥| |Ψ ⟩

ω

ω

ω

ω

ω

ω

ω

ω

ω

ω

ω

ω

+← †

+

+

+ ←

+

+†

−← †

− ←

−†

p qXX

X

Xp q

p qX

X

XX

p q

[ ]

[ ]

[ ]

[ ]

f Xpq f

i

i

X ipq

f

fi

i Xpq i

f

f

X fpq

i

if

x

ix

x

x

f

x

x

y

f

y

y

y

i

y

y

1

1

1

1

1

1

2

2

2

2

2

2(26)

Here we use the normalized response states so that the analogywith the interstate 1PDMs is preserved. Thus, ωDM for aparticular f ← i transition is a weighted sum of two terms: thetransition density matrix between the initial state and thenormalized perturbed final state and the transition density matrixbetween the normalized perturbed initial state and the final state.The weights of these partialωDMs (e.g., γ+ω1x

f←Xi and γ+ω1x

Xf←i) are givenby the norms of the first-order perturbed states (responsevectors). The magnitudes of these norms quantify the magnitudeof the response, e.g., the first-order response function becomeslarge when the perturbing frequency is resonant with the zero-order states. The norms of partial ωDM quantify one-electroncharacter of the “transition” between the respective zero-orderand (normalized) first-order perturbed state. As in the case ofone-electron interstate properties, one-electron character is arequisite quantity determining the magnitude of the observ-ables.21,22 ωDMs and the norms of response vectors arefrequency-dependent.As in the case of 1PA transition properties, all molecular orbital

information related to the 2PA transition is encoded in theseωDMs, e.g., the orbitals that give rise to the response wavefunctions due to electric field of frequency ω1 or ω2, the orbitalsinvolved in the transitions between the first- and zero-order wavefunctions, the dependence of 2PA response on frequency ω1 orω2 (and indirect dependence on ω2 or ω1 due to the resonancecondition for 2PA), etc. Equations 24 and 25 allow one torationalize the 2PA transition moment as a first-order perturbed(by field ω1 along x) transition density between the initial andfinal states interacting with the dipole field of the second photon(with ω2 polarized along y), with the two photons satisfying the

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3258

Page 4: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

resonance condition, ω1 + ω2 =Ωf i. Because of the tensor natureof the 2PA process, there are three ωDMs (for x, y, zpolarizations) for each f ← i and i ← f transition. Although theanalysis and visualization of theseωDMs can be performed in thebasis of canonical molecular orbitals, here we perform SVD ofωDMs to obtain NTOs for 2PA transitions.Before proceeding to the numerical examples, let us discuss

eqs 24 and 25 in more detail. In Hermitian theory, eq 25 wouldbe a Hermitian conjugate of eq 24 for degenerate photons (ω1 =ω2) with the same polarization. In the EOM-CC framework, thisis not the case, but as will be shown below, the differencesbetween the NTOs computed for γ+ω1x

f←i and γ−ω1x

i←f are very small,which is similar to other examples of left-right and right-leftinterstate properties.37,38 Thus, we focus our analysis on eq 24.The two partial ωDMs contributing to eq 24 can be analyzedseparately. The SVD decomposition of γ+ω

f←Xi yields NTOsconnecting the final state and the perturbed initial state (Xi),whereas γ+ω

Xf←i yields NTOs connecting the initial state and theperturbed final state (Xf). It is instructive to consider two limitingcases, when either ∥Xi∥ or ∥Xf∥ is zero. In such cases, the 2PAresponse is dominated by either the response of the final state orthe response of the initial state; the NTOs of the full ωDM arethe same as those of the respective partialωDM, and the singularvalues of ωDM are equal to the respective singular values of thepartial ωDMs scaled by the norm of the corresponding responsevector: ∥Xi

+ω1x∥σf←Xi or ∥Xf+ω1x∥σXf←i. The participation ratio of

the full ωDM then equals the participation ratio of the partialωDMs for which the response-vector norm is not zero. In ageneral case, the magnitudes of ∥Xi∥ and ∥Xf∥ might becomparable. Then the NTOs of the full ωDMs are notnecessarily related to the NTOs of the partial ωDMs, especiallywhen both partial ωDMs contribute to the NTOs of the fullωDM (this case is illustrated below by the π→ π* 2PA transitionin para-nitroaniline). The situation is different when the normsof the response vectors are nonzero but only one partial ωDMcontributes to a given dominant NTO pair of the full ωDM (asillustrated by the 2PA transitions for ethylene and trans-stilbene).If the singular vectors of the full transition ωDMs are

approximately the same as the singular vectors of the partialtransition ωDMs, then the singular values of ωDM are related tothe singular values of the partial ωDMs by

σ σ σ σ≈ ≡ ∥ ∥ + ∥ ∥ω ω← ← + ← + ←X Xf i f ii

f Xf

X iappr

x i x f1 1(27)

σ σ σ σ≈ ≡ ∥ ∥ + ∥ ∥ω ω← ← − ← − ←X Xi f i ff

i Xi

X fappr

y f y i2 2(28)

Thus, one can use the difference between the exact singularvalues of ωDMs and the approximate singular values (σappr)

computed using eqs 27 and 28 as a simple numerical diagnostic ofthe similarity between the NTOs of the full and partial ωDMs.Alternatively, one can compute the overlap between the NTOsthemselves.We implemented the calculation of the NTOs for ωDMs in a

development version of Q-Chem39,40 using the existing infra-structure for wave function analysis.15,41 Below, we illustrate theutility of the NTO analysis ofωDMs computed within the EOM-EE-CCSD framework by considering the 1Ag→ 2Ag transition inethylene and the 1Ag → 5Ag transition in trans-stilbene. Thepeaks corresponding to these transitions dominate the 2PAspectra.26,27 We also analyze the π → π* charge-transferexcitation (1A1 → 2A1) in para-nitroaniline. In contrast to thetwo transitions in ethylene and stilbene, the large 2PA crosssection of this transition arises due to the difference in the dipolemoments of the ground and excited states and not due to a strongcoupling with a third state. We use the daug-cc-pVTZ basis forethylene, the daug(-d)-cc-pVDZ basis for trans-stilbene, and theaug-cc-pVDZ basis for pNA. We freeze the core electrons fortrans-stilbene and pNA. Ethylene (D2h point group) and trans-stilbene (C2h point group) are in the xy-plane with the x-axisbeing the long axis, and pNA (C2v point group) is in the xz-planewith the dipole moment along the z-direction. The Cartesiancoordinates of these systems are taken from refs 26 and 42; theyare also given in the Supporting Information. We use Gabedit43

to visualize NTOs.The 1Ag → 2Ag 2PA transition in ethylene with degenerate

photons26 has largeMzz andmoderateMxxmoments, as shown inTable 1. The relative magnitude of Mzz and Mxx changes whendifferent photon frequencies are used. With the photonfrequencies chosen here (see Table 1), the Mxx momentsincrease considerably and become larger than Myy and Mzzmoments. Below, we analyze the orbital picture of the Mzzmoments for the degenerate 2PA case andMxxmoments for bothdegenerate and nondegenerate 2PA.We begin by analyzing the dominant contributions to theMzz

andMxx 2PA moments in ethylene by using a three-states model,as was done in ref 26. The NTOs for regular 1PDM of the 1Agand 2Ag pair (Figure S1 in the Supporting Information) showthat this transition has Rydberg π → Ry(pz) character. Byanalyzing the symmetries and transition dipole momentsbetween these two states and other nearby states, one canidentify the 1B1u state as the state providing large contributionsto Mzz via the 1Ag → 1B1u → 2Ag channel. The dominant NTOpairs of the individual 1Ag → 1B1u and 1B1u → 2Ag transitionshave π → Ry(s) and Ry(s) → Ry(pz) characters, respectively.From this three-states picture, one can infer that the dominantcontribution toMzz arises from the π→ Ry(s)→ Ry(pz) orbitals.

Table 1. Photon Frequencies (in eV), 2PA Transition Moments (in au), and Microscopic Cross Sections (in au) for Selected 2PATransitions in Ethylene, trans-Stilbene, and para-Nitroaniline

system transition ω1 ω2 Mxx Myy Mzz δ2PA

ethylene 1Ag → 2Ag 4.20 4.20 19.6 11.4 46.1 7272Ag → 1Ag 19.0 11.3 44.7

ethylene 1Ag → 2Ag 7.81 0.60 −599.1 −28.5 94.4 649812Ag → 1Ag −571.7 −27.9 91.7

ethylene 1Ag → 2Ag 0.60 7.81 −599.1 −28.5 94.4 649812Ag → 1Ag −571.7 −27.9 91.7

trans-stilbene 1Ag → 5Ag 3.15 3.15 475.9 −3.4 10.1 415915Ag → 1Ag 432.6 −3.8 9.2

para-nitroaniline 1A1 → 2A1 2.31 2.31 7.2 1.0 −139.0 36222A1 → 1A1 1.8 −0.9 −133.1

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3259

Page 5: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

In a similar fashion, the three-states model analysis ofMxx revealslarge contribution from the terms involving the 1B3u state, owingto the strong dipole coupling of 1B3u with both the initial and thefinal states. The dominant NTO pairs of the individual one-photon transitions, 1Ag → 1B3u and 1B3u → 2Ag, have π → π*and π*→ Ry(pz) characters, respectively. Thus, one can ascribethe π → π* → Ry(pz) orbital character to the Mxx moment.The dominant NTOs of ωDMs for the 1Ag → 2Ag 2PA

transition are shown in Figure 1 (the NTOs for 2Ag → 1Ag areshown in the Supporting Information). The quantitative analysisis given in Tables 2 and 3. The z-component of the 1Ag → 2Ag

and 2Ag → 1Ag ωDMs gives rise to two NTO pairs: pair (a)represents a transition between π and Ry(s) orbitals, and pair (b)represents a transition between Ry(s) and Ry(pz) orbitals. Thedominant NTO pairs for the x-component of the 1Ag→ 2Ag and2Ag→ 1AgωDMs represent two transitions: pair (a) represents a

transition between π and π* orbitals, and pair (b) represents atransition between π* and Ry(pz) orbitals. Thus, the orbitaldescription of the 2PA transition inferred from three-statesmodels is obtained directly from the NTO analysis of thedifferent components of the transition ωDMs, without makingany assumptions on the nature of the electronic statesdominating the sum-over-states. The singular values for theseNTO pairs are given in Tables 2 and 3. To understand the originof these NTO pairs, we deconstruct the ωDM into the partialtransition ωDMs, according to eqs 24 and 25, and perform NTOanalysis of these partial transitionωDMs. For both the i→ f and f→ i transitions, the dominant NTO pair of the partial transitionωDMs corresponds to one (or the other) of the NTOpairs of thefull transition ωDM. Further, the singular values of the NTOpairs (a) and (b) of the full transition ωDMs can be reproducedaccurately with the norms of the response vectors and the

Table 2. NTO Analysis of the z-Components of ωDMs (Ethylene, ω1 = ω2 = 33911 cm−1)

transition NTO σ ∥Xi±ωz∥ σpartial

f←Xi ∥Xf±ωz∥ σpartial

Xf←i σappra PRNTO

full

i → f pair

1Ag → 2Ag 8.5 41.9 1.04(a) 40.44 ≈0 0.97 40.62(b) 5.45 0.64 ≈0 5.47

2Ag → 1Ag 41.8 8.3 1.04(a) 38.97 0.94 ≈0 39.12(b) 5.28 ≈0 0.65 5.38

aSee eqs 27 and 28 for the definition of σappr.

Table 3. NTO Analysis of the x-Components of ωDMs (Ethylene)

transition NTO σ ∥Xi±ωx∥ σpartial

f←Xi ∥Xf±ωx∥ σpartial

Xf←i σappra PRNTO

full

i → f pair

ω1 = ω2 = 33911 cm−1

1Ag → 2Ag 12.2 23.1 1.20(a) 22.10 ≈0 0.96 22.27(b) 7.00 0.59 ≈0 7.24

2Ag → 1Ag 23.1 11.6 1.20(a) 21.32 0.93 ≈0 21.46(b) 6.64 ≈0 0.60 6.94

ω1 = 63000 cm−1, ω2 = 4821 cm−1

1Ag → 2Ag 373.0 13.3 1.01(a) 14.70 ≈0 0.95 12.67(b) 250.16 0.67 ≈0 250.36

2Ag → 1Ag 448.0 6.8 1.09(a) 409.29 0.91 ≈0 409.29(b) 2.28 ≈0 0.55 3.72

ω1 = 4821 cm−1, ω2 = 63000 cm−1

1Ag → 2Ag 7.0 441.9 1.09(a) 423.66 ≈0 0.96 423.66(b) 2.71 0.54 ≈0 3.78

2Ag → 1Ag 13.3 349.9 1.01(a) 13.19 0.92 ≈0 12.22(b) 236.30 ≈0 0.68 236.50

aSee eqs 27 and 28 for the definition of σappr.

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3260

Page 6: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

singular values of the partial transition ωDMs, according to eqs27 and 28. Thus, the dominant NTO pairs for the components ofthe full transition ωDMs correspond to the NTOs of the partialtransition ωDMs.

Figure 1. Dominant NTOs (isovalue = 0.003) for the Mzz (left) andMxx (right) 2PA transition moments for the 1Ag → 2Ag ωDMs(ethylene, ω1 = ω2 = 33911 cm−1).

The singular values of the two dominant NTO pairs in thedegenerate 2PA case are different due to the different magnitudesof the dominant amplitudes of the response vectors X and X.When the photon frequencies are different, the denominator,Ωni− ω or Ωnf + ω, of one of the sum-over-states terms becomessmaller in the right-hand side term of eqs 12 and 13. Theamplitudes of the response vectors that correspond to thetransition between the orbitals involved in the transitionsbetween the states (i and n or f and n) would be enhanced bythe amplified norm of the response vector. The singular values ofthe NTO pairs arising from this response vector (or due to thetransition between the perturbed state and the zeroth-orderstate) would be dominant. For example, in the case ofnondegenerate 2PA in ethylene with ω1 = 63000 cm−1 and ω2= 4821 cm−1, the sum-over-states is dominated by the terminvolving the B3u state due to smaller Ωni − ω1 and Ωnf + ω2relative toΩnf +ω1 andΩni−ω2, which is reflected in large normsof the X-vectors in eqs 24 and 25. These large norms result inlarge singular values for the NTOs of the full ωDMscorresponding to the f ← Xi and i ← Xf transitions for the f ←i and i← f 2PA transitions, respectively. Similarly, in the case ofnondegenerate 2PA in ethylene with ω1 = 4821 cm−1 and ω2 =63000 cm−1, the sum-over-states is dominated by the terminvolving the B3u state due to smaller Ωnf + ω1 and Ωni − ω2relative to Ωni − ω1 and Ωnf + ω2, which leads to large norms ofthe X-vectors in eqs 24 and 25. These large norms result in largesingular values for the NTOs of the full ωDMs corresponding tothe Xf ← i and Xi ← f transitions for the f ← i and i ← f 2PAtransitions, respectively. This suggests that the orbital characterof the 2PA transition produced by NTO analysis does notdepend on the choice of ω1 and ω2 when both γ+ω1

f←i and γ−ω2

i←f areanalyzed together (only the directions of the NTOs arereversed). We note that PRNTO forMxx drops for nondegenerate2PA cases relative to the degenerate 2PA case due to theenhancement of one of the two components of the sum-over-states term involving the 1B3u state in the nondegenerate case.Moreover, the PRNTO values for the f ← i and i ← f 2PAtransitions are swapped when ω1 and ω2 are swapped. Thechanges in the 2PA regime can be quantified by analyzing the size

of the exciton;17 see eq 4. The values for dexc for the 2PA andrelevant 1PA transitions are given in Table S2 in the SupportingInformation. As one can see, dexc values for the partial and fullωDMs depend strongly on the values of two frequencies. Inparticular, we observe that asω1 approaches the frequency of the1Ag → 1B3u transition, dexc for the ωDMs approach dexc for therespective 1PA transitions, 1B3u → 2Ag and 1B3u → 1Ag. Thisreflects the change in the 2PA regime; as one approaches theresonant condition with the 1B3u state, the 2PA transitionbecomes increasingly more like the two 1PA transitions via thisstate. Thus, one can use the comparison between the exciton sizefor the 2PA transitions and the exciton size of the 1PA transitions(between the initial/final states and the intermediate state) as asimple indication of the validity of the three-states model.The EOM-CCSD 2PA spectra with degenerate photons for

trans-stilbene have been reported recently in ref 26. Thedominant 2PA peak corresponds to the 1Ag → 5Ag transition.Unlike the 1Ag→ 2Ag 2PA transition with degenerate photons inethylene, this transition is characterized by large Mxx transitionmoments (the molecule is in the xy-plane in our calculations), asshown in Table 1. The three-states model analysis in ref 26reveals that the terms involving the 1Bu state provide leadingcontributions toMxx, primarily due to the large transition dipolemoments between this state and the two Ag states.

Figure 2. Dominant NTOs (isovalue = 0.003) for the 1Ag → 1Bu and1Bu → 5Ag transitions of trans-stilbene. The weights of each NTO aregiven σK

2 .

The dominant NTO pair(s) of 1PDMs of the 1Ag → 1Bu and1Bu → 5Ag transitions in trans-stilbene are shown in Figure 2.Whereas the 1Ag → 1Bu 1PA transition is dominated by oneπ(au) → π*(bg) NTO pair, the 1Bu → 5Ag 1PA transition ischaracterized by two NTO pairs that correspond to the π*(bg)→ Ry(pz) and π(bg) → π(au) transitions. Whereas the three-states model analysis of the 2PA transition reveals a dominant1Ag → 1Bu → 5Ag channel, it is not clear which orbitals areimportant for the 2PA transition. Because two NTO pairscontribute to the 1Bu → 5Ag 1PA transition, it is not obviouswhether both pairs are also important in the 1Ag → 5Ag 2PAtransition. Deeming only the π*(bg)→ Ry(pz) NTO pair with aslightly larger singular value in the 1PA transition to be importantfor the 2PA transition based on the three-states model analysiscould be misleading as the sum-over-states terms not included inthe few-states analysis could increase the overall contribution ofthe π(bg)→ π(au) NTOpair relative to the contribution from theπ*(bg) → Ry(pz) NTO pair. In addition, within such a three-states model, it is difficult to quantify contributions from different1PA NTO pairs to the 2PA transition moment.The NTO analysis of the x-component of the ωDMs for the

1Ag → 5Ag 2PA transition with degenerate photons in trans-stilbene reveals the orbitals involved in the 2PA transition in a

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3261

Page 7: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

straightforward manner, as shown in Figure 3 and Table 4.Similarly to the NTO analysis of the x-component of the ωDMsfor the 1Ag → 2Ag 2PA transition in ethylene, here the twodominant NTO pairs represent either a π→ π*(bg) (NTO pair(a)) or π*(bg)→ Ry(pz) (NTO pair (b)) transition. In terms ofthe dominant contributions to the 2PAmoment, these two NTOpairs result in the dominant π → π* (bg) → Ry(pz) channel. Inaddition, there is a third non-negligible contribution from NTOpair (c), which originates from the π(bg) → π(au) NTO pair inthe 1Bu → 5Ag transition. NTO pairs (c) and (a) represent the2PA channel π(bg) → π(au) → π*(bg). The larger norms of theresponse vectors compared to other systems studied in thisLetter are consistent with the larger 2PA cross sections for thistransition. The origin of the dominant NTO pairs in the 1Ag →5Ag and 5Ag → 1Ag transition ωDMs can be traced back to thepartial transitionωDMs. As shown in Table 4, the singular valuesof these two dominant NTO pairs of the full ωDMs can beapproximated using the norms of the response vectors and thesingular values of the dominant NTO pair(s) of each of the twocomponent partial transition ωDMs, according to eqs 27 and 28.We also note a larger PRNTO (1.6 in trans-stilbene versus 1.0−1.2in ethylene), which reflects a more complex character of the 1Ag→ 5Ag transition in trans-stilbene relative to the 1Ag → 2Agtransition in ethylene.

Figure 3. Dominant NTOs (isovalue = 0.003) for the x-componentsof the ωDMs for the 1Ag → 5Ag transition (trans-stilbene, degeneratephotons).

The π → π* charge-transfer 2PA transition between 1A1 and2A1 states in pNA with degenerate photons is characterized bylargeMzz transition moments (z is along the dipole moment), asshown in Table 1. These large 2PA transition moments are aconsequence of the large difference in the dipole moments of theinitial ground and final excited states. Within the two-statesmodel, Mzz transition moments can be approximated as

ωμ μ

ωμ μ

ω

μ μω

μ μω

≈ −⟨Ψ | |Ψ⟩⟨Ψ| |Ψ⟩

−+

⟨Ψ | |Ψ⟩⟨Ψ| |Ψ⟩Ω +

−⟨Ψ | |Ψ ⟩⟨Ψ | |Ψ⟩

Ω −+

⟨Ψ | |Ψ ⟩⟨Ψ | |Ψ⟩+

←⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜⎜

⎞⎠⎟⎟

M ( )zzf i f

zi i

zi f

zi i

zi

if

fz

f fz

i

fi

fz

f fz

i

(29)

ωμ μ

ωμ μ

ω

μ μω

μ μω

− ≈ −⟨Ψ| |Ψ⟩⟨Ψ| |Ψ ⟩

Ω ++

⟨Ψ| |Ψ⟩⟨Ψ| |Ψ ⟩−

−⟨Ψ| |Ψ ⟩⟨Ψ | |Ψ ⟩

++

⟨Ψ| |Ψ ⟩⟨Ψ | |Ψ ⟩Ω −

←⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜⎜

⎞⎠⎟⎟

M ( )zzi f i

zi i

zf

if

iz

i iz

f

iz

f fz

f iz

f fz

f

fi

(30)

Rearranging the above expressions, Mzz transition moments canbe expressed in terms of the difference in the dipole moments ofthe initial and final states as follows

ωμω

μ μ≈ −⟨Ψ | |Ψ⟩

⟨Ψ | |Ψ ⟩ − ⟨Ψ| |Ψ⟩←M ( ) 2 ( )zzf i f

zi

fz

f iz

i (31)

ωμω

μ μ− ≈ −⟨Ψ| |Ψ ⟩

⟨Ψ | |Ψ ⟩ − ⟨Ψ| |Ψ⟩←M ( ) 2 ( )zzi f i

zf

fz

f iz

i

(32)

Mzzf←i andMzz

i←f values computed using the above two-states modelexpressions equal 146.3 and 134.5 au, respectively, which aresimilar to the EOM-EE-CCSD values given in Table 1. Theseexpressions suggest that this 2PA transition can be characterizedby the same single NTO pair, π→ π*, as the corresponding 1A1→ 2A1 1PA transition (Figure S2 in the SupportingInformation). These expressions also suggest that the singularvalues of the NTO pair for the 2PA transition can beapproximated in terms of the singular values for this NTO pairfor the 1PA transition and the change in the dipole moment uponexcitation as follows:

σσ

ωμ μ≈ ⟨Ψ | |Ψ ⟩ − ⟨Ψ| |Ψ⟩←

←2 ( )f i

f i

fz

f iz

i2PA1PA

(33)

Table 4. NTO Analysis for x-Components of the ωDMs of trans-Stilbene (ω1 = ω2 = 25400 cm−1)

transition NTO σ ∥Xi±ωx∥ σpartial

f←Xi ∥Xf±ωx∥ σpartial

Xf←i σappra PRNTO

full

i → f pair

1Ag → 5Ag 62.8 85.4 1.62(a) 73.99 ≈0 0.93 79.65(b) 33.08 0.50 ≈0 31.59(c) 19.40 0.32 ≈0 20.25

5Ag → 1Ag 85.6 57.6 1.58(a) 70.00 0.87 ≈0 74.75(b) 30.65 ≈0 0.51 29.30(c) 17.04 ≈0 0.31 17.86

aSee eqs 27 and 28 for the definition of σappr.

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3262

Page 8: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

σσ

ωμ μ≈ ⟨Ψ | |Ψ ⟩ − ⟨Ψ| |Ψ⟩←

←2 ( )i f

i f

fz

f iz

i2PA1PA

(34)

This qualitative picture is confirmed by the NTO analysis ofthe z-components of 2PA transition ωDM, which produces asingle NTO pair involving the π and π* orbitals, as shown inFigure 4. These NTOs are practically the same as the regularNTOs (shown in Figure S2 in the Supporting Information). Thesingular values for this NTO pair in the 2PA and 1PA transitionsare 58.4 and 0.85, and the dipole moments of the initial and finalstates are 2.7 and 5.7 au, respectively. Using eq 34, a singularvalue for the f→ i transition, and the dipole moments of the twostates, we calculate the singular value for the dominant 2PANTOpair as 61.2, which is similar to the actual singular value of 58.4.

Figure 4. Dominant NTOs (isovalue = 0.005) for the z-componentsof the ωPDM for the 1A1 → 2A1 2PA transition (pNA, degeneratephotons).

The PRNTO for the 2PA i→ f and f→ i transitions are given inTable 5. These values are close to 1. In contrast to ethylene andstilbene examples, the PRNTOs of the partial transitionωDMs arelarge (>20) as these partial transition DMs have significant i→ ior f → f character. This large difference between the PRNTOs ofthe full and partial ωDMs indicates that this 2PA transition hascharge-transfer character that is driven by the change in thedipole moments of the initial and final states (and not by thedipole coupling to a third state). The dominant NTO pair for thefull 2PA transition ωDMs is not the dominant NTO pair inpartial transition ωDMs (γf←Xi and γXi←f) as the Xi and Xi

response states have dominant Ψf-like character. This can beexplained using the components of the sum-over-states terms ineqs 29 and 30. The first and third right-hand-side (RHS) termscontribute to the partial transition ωDMs involving the X-vectors, while the second and fourth RHS terms contribute to thepartial transition ωDMs involving the X-vectors. The third term(with sum-over-states index n = f) in eq 29 is larger than the firstterm (with sum-over-states index n = i), and hence, the NTOsfrom the f → f transition dominate the NTOs from the i → ftransition. Similarly, the fourth term (with sum-over-states indexn = f) in eq 30 is larger than the second term (with sum-over-states index n = i), and hence, again the NTOs from the f → ftransition dominate the NTOs from the i → f transition. As aresult, NTOs corresponding to π → π* transitions aresuppressed in γf←Xi and γXi←f. In fact, the π → π* NTO pair isdifficult to identify for γf←Xi and γXi←f through visualization (evenfor γi←Xf and γXf←i, the dominant NTO pair looks slightly differentfrom the dominant NTO pair for the corresponding full ωDMfor 2PA transition), which invalidates the use of eqs 27 and 28.

Table 5. NTO Analysis for z-Components of the ωDMs forthe 2PA Transition of para-Nitroaniline (ω1 = ω2 = 18626cm−1)

transition i → f σ PRNTOfull PRNTO

f←Xi PRNTOXf←i

1A1 → 2A1 41.03 1.20 26.14 20.612A1 → 1A1 58.41 1.12 20.30 25.90

In summary, we presented the extension of the theory oftransition 1PDMs and their analyses in terms of NTOs to the2PA transitions. This new tool can accurately elucidate theorbital picture of 2PA transitions in a compact way and withoutguessing which electronic states provide dominant contributionsto the 2PA transition moments. It also allows one to distinguishbetween two types of bright 2PA transitions: those that aredriven by the strong dipole coupling with a particular virtual stateversus those that have charge-transfer character and are driven bythe large difference in the dipole moment in the initial and finalstates. The NTOs of ωDMs also aid the visualization of 2PAtransitions. Similar to the standard NTO analysis, we find that2PA transitions can often be described using just a few pairs ofhole and particle NTOs of ωDM. In addition, this analysiscaptures the changes in orbital characters of 2PA transitions fordifferent photon frequencies, which cannot be captured in astraightforward manner using the traditional few-states analysis.This analysis tool can help to elucidate the role of molecularstructure and intermolecular interactions on the nonlinearspectra. Accurate electronic structure methods for calculating2PA cross sections augmented by this analysis tool forinterpreting the calculated results not only reveal the orbitalcharacter of the important peaks in experimental 2PA spectra butcan also aid the design of new molecules with desirable opticalproperties.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.jpclett.7b01422.

NTOs for 1PDMs of ethylene and para-nitroaniline andNTOs for the f→ iωDMs for all transitions considered inthis Letter (PDF)

■ AUTHOR INFORMATIONORCIDKaushik D. Nanda: 0000-0002-3447-6678NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSA.I.K. acknowledges support by the U.S. National ScienceFoundation (No. CHE-1566428). We thank Dr. Marc deWergifosse for his feedback about the manuscript.

■ REFERENCES(1) Zipfel, W. R.; Williams, R. M.; Webb, W. W. Nonlinear Magic:Multiphoton Microscopy in the Biosciences. Nat. Biotechnol. 2003, 21,1369−1376.(2) Kawata, S.; Kawata, Y. Three-Dimensional Optical Data StorageUsing Photochromic Materials. Chem. Rev. 2000, 100, 1777−1788.(3) Konig, K. Multiphoton Microscopy in Life Sciences. J. Microsc.2000, 200, 83−104.

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3263

Page 9: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

(4) Helmchen, F.; Denk, W. Deep Tissue Two-Photon Microscopy.Nat. Methods 2005, 2, 932−940.(5) Kachynski, A. V.; Pliss, A.; Kuzmin, A. N.; Ohulchanskyy, T. Y.;Baev, A.; Qu, J.; Prasad, P. N. Photodynamic Therapy by In SituNonlinear Photon Conversion. Nat. Photonics 2014, 8, 455−461.(6) McClain, W. M. Excited State Symmetry Assignment ThroughPolarized Two-Photon Absorption Studies of Fluids. J. Chem. Phys.1971, 55, 2789−2796.(7) McClain, W. M. Two-Photon Molecular Spectroscopy. Acc. Chem.Res. 1974, 7, 129−135.(8) Swofford, R. L.; Albrecht, A. C. Nonlinear Spectroscopy. Annu. Rev.Phys. Chem. 1978, 29, 421−440.(9) Elles, C. G.; Rivera, C.; Zhang, X.; Pieniazek, P. A.; Bradforth, S. E.Electronic Structure of Liquid Water from Polarization-DependentTwo-Photon Absorption Spectroscopy. J. Chem. Phys. 2009, 130,084501.(10) Terenziani, F.; Katan, C.; Badaeva, E.; Tretiak, S.; Blanchard-Desce, M. Enhanced Two-Photon Absorption of Organic Chromo-phore: Theoretical and Experimental Assessments.Adv.Mater. 2008, 20,4641−4678.(11) Kogej, T.; Beljonne, D.; Meyers, F.; Perry, J. W.; Marder, S. R.;Bredas, J. L. Mechanisms for Enhancement of Two-Photon Absorptionin Donor-Acceptor Conjugated Chromophores. Chem. Phys. Lett. 1998,298, 1−6.(12) Marder, S. R.; Torruellas, W. E.; Blanchard-Desce, M.; Ricci, V.;Stegeman, G. I.; Gilmour, S.; Bredas, J.-L.; Li, J.; Bublitz, G. U.; Boxer, S.G. Large Molecular Third-Order Optical Nonlinearities in PolarizedCarotenoids. Science 1997, 276, 1233−1236.(13) Albota, M.; Beljonne, D.; Bredas, J.-L.; Ehrlich, J. E.; Fu, J.-Y.;Heikal, A. A.; Hess, S. E.; Kogej, T.; Levin, M. D.; Marder, S. R.;McCord-Maughon, D.; Perry, J. W.; Rockel, H.; Rumi, M.;Subramaniam, G.; Webb, W. W.; Wu, X.-L.; Xu, C. Design of OrganicMolecules with Large Two-Photon Absorption Cross Sections. Science1998, 281, 1653−1656.(14) Luzanov, A. V.; Zhikol, O. A. In Practical Aspects of ComputationalChemistry I: An Overview of the Last Two Decades and Current Trends;Leszczynski, J., Shukla, M. K., Eds.; Springer, 2012; p 415.(15) Plasser, F.; Wormit, M.; Dreuw, A. New Tools for the SystematicAnalysis and Visualization of Electronic Excitations. I. Formalism. J.Chem. Phys. 2014, 141, 024106.(16) Bappler, S. A.; Plasser, F.; Wormit, M.; Dreuw, A. Exciton Analysisof Many-Body Wave Functions: Bridging the Gap Between theQuasiparticle and Molecular Orbital Pictures. Phys. Rev. A: At., Mol.,Opt. Phys. 2014, 90, 052521.(17) Plasser, F. Entanglement Entropy of Electronic Excitations. J.Chem. Phys. 2016, 144, 194107.(18) Tretiak, S.; Mukamel, S. Density Matrix Analysis and Simulationof Electronic Excitations in Conjugated and Aggregated Molecules.Chem. Rev. 2002, 102, 3171.(19) Sun, M.; Chen, J.; Xu, H. Visualizations of Transition Dipoles,Charge Transfer, and Electron-Hole Coherence on Electronic StateTransitions Between Excited States for Two-Photon Absorption. J.Chem. Phys. 2008, 128, 064106.(20) 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.(21) Feng, X.; Luzanov, A. V.; Krylov, A. I. Fission of Entangled Spins:An Electronic Structure Perspective. J. Phys. Chem. Lett. 2013, 4, 3845−3852.(22) Matsika, S.; Feng, X.; Luzanov, A. V.; Krylov, A. I. What We canLearn from the Norms of One-Particle Density Matrices, and What WeCan’t: Some Results for Interstate Properties in Model Singlet FissionSystems. J. Phys. Chem. A 2014, 118, 11943−11955.(23) Reisler, H.; Krylov, A. I. Interacting Rydberg and Valence States inRadicals and Molecules: Experimental and Theoretical Studies. Int. Rev.Phys. Chem. 2009, 28, 267−308.

(24) Dreuw, A.; Head-Gordon, M. Single-Reference Ab InitioMethods for the Calculation of Excited States of Large Molecules.Chem. Rev. 2005, 105, 4009−4037.(25) Luzanov, A. V.; Casanova, D.; Feng, X.; Krylov, A. I. QuantifyingCharge Resonance and Multiexciton Character in Coupled Chromo-phores by Charge and Spin Cumulant Analysis. J. Chem. Phys. 2015, 142,224104.(26) de Wergifosse, M.; Elles, C. G.; Krylov, A. I. Two-PhotonAbsorption Spectroscopy of Stilbene and Phenanthrene: Excited-StateAnalysis and Comparison with Ethylene and Toluene. J. Chem. Phys.2017, 146, 174102.(27) de Wergifosse, M.; Houk, A. L.; Krylov, A. I.; Elles, C. G. Two-Photon Absorption Spectroscopy of Trans-stilbene, Cis-stilbene, andPhenanthrene: Theory and Experiment. J. Chem. Phys. 2017, 146,144305.(28) Olsen, J.; Jørgensen, P. Linear and Non-Linear ResponseFunctions for an Exact State and for MCSCF State. J. Chem. Phys.1985, 82, 3235.(29) Paterson, M. J.; Christiansen, O.; Pawlowski, F.; Jørgensen, P.;Hattig, C.; Helgaker, T.; Salek, P. Benchmarking Two-PhotonAbsorption with CC3 Quadratic Response Theory, and Comparisonwith Density-Functional Response Theory. J. Chem. Phys. 2006, 124,054322.(30) Nanda, K. D.; Krylov, A. I. Two-Photon Absorption CrossSections within Equation-Of-Motion Coupled-Cluster FormalismUsing Resolution-of-the-Identity and Cholesky Decomposition Repre-sentations: Theory, Implementation, and Benchmarks. J. Chem. Phys.2015, 142, 064118.(31) Nanda, K. D.; Krylov, A. I. Effect of the Diradical Character onStatic Polarizabilities and Two-Photon Absorption Cross-Sections: ACloser Look with Spin-Flip Equation-Of-Motion Coupled-ClusterSingles and Doubles Method. J. Chem. Phys. 2017, 146, 224103.(32) Knippenberg, S.; Rehn, D. R.; Wormit, M.; Starcke, J. H.;Rusakova, I. L.; Trofimov, A. B.; Dreuw, A. Calculations of NonlinearResponse Properties Using the Intermediate State Representation andthe Algebraic-Diagrammatic Construction Polarization PropagatorApproach: Two-Photon Absorption Spectra. J. Chem. Phys. 2012, 136,064107.(33) Kallay, M.; Gauss, J. Calculation of Frequency-DependentPolarizabilities Using General Coupled-Cluster Models. J. Mol. Struct.:THEOCHEM 2006, 768, 71−77.(34) Krylov, A. I. Equation-Of-Motion Coupled-Cluster Methods forOpen-Shell and Electronically Excited Species: The Hitchhiker’s Guideto Fock Space. Annu. Rev. Phys. Chem. 2008, 59, 433−462.(35) Sneskov, K.; Christiansen, O. Excited State Coupled ClusterMethods. WIREs Comput. Mol. Sci. 2012, 2, 566.(36) Bartlett, R. J. Coupled-Cluster Theory and its Equation-Of-Motion Extensions. WIREs Comput. Mol. Sci. 2012, 2, 126−138.(37) Stanton, J. F.; Bartlett, R. J. The Equation of Motion Coupled-Cluster Method. A Systematic Biorthogonal Approach to MolecularExcitation Energies, Transition Probabilities, and Excited StateProperties. J. Chem. Phys. 1993, 98, 7029−7039.(38) Ichino, T.; Gauss, J.; Stanton, J. F. Quasidiabatic States Describedby Coupled-Cluster Theory. J. Chem. Phys. 2009, 130, 174105.(39) Shao, Y.; Gan, Z.; Epifanovsky, E.; Gilbert, A. T. B.; Wormit, M.;Kussmann, J.; Lange, A. W.; Behn, A.; Deng, J.; Feng, X.; et al. AdvancesinMolecular QuantumChemistry Contained in the Q-Chem 4 ProgramPackage. Mol. Phys. 2015, 113, 184−215.(40) Krylov, A. I.; Gill, P. M. W. Q-Chem: An Engine for Innovation.WIREs Comput. Mol. Sci. 2013, 3, 317−326.(41) Plasser, F.; Bappler, S. A.; Wormit, M.; Dreuw, A. New Tools forthe Systematic Analysis and Visualization of Electronic Excitations. II.Applications. J. Chem. Phys. 2014, 141, 024107.(42) Nanda, K. D.; Krylov, A. I. Static Polarizabilities for Excited Stateswithin the Spin-Conserving and Spin-Flipping Equation-Of-MotionCoupled-Cluster Singles and Doubles Formalism: Theory, Implemen-tation, and Benchmarks. J. Chem. Phys. 2016, 145, 204116.

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3264

Page 10: Visualizing the Contributions of Virtual States to Two ...iopenshell.usc.edu/pubs/pdf/jpcl-8-3256.pdf · Rather, they depend on the entire manifold of the excited states, which follows

(43) Allouche, A. Gabedit - A Graphical User Interface forComputational Chemistry Softwares. J. Comput. Chem. 2011, 32,174−182.

The Journal of Physical Chemistry Letters Letter

DOI: 10.1021/acs.jpclett.7b01422J. Phys. Chem. Lett. 2017, 8, 3256−3265

3265