cavity-induced modifications of molecular structure in the...

14
Cavity-Induced Modifications of Molecular Structure in the Strong-Coupling Regime Javier Galego, 1 Francisco J. Garcia-Vidal, 1,2 and Johannes Feist 1,* 1 Departamento de Física Teórica de la Materia Condensada and Condensed Matter Physics Center (IFIMAC), Universidad Autónoma de Madrid, E-28049 Madrid, Spain 2 Donostia International Physics Center (DIPC), E-20018 Donostia/San Sebastián, Spain (Received 10 June 2015; revised manuscript received 17 August 2015; published 9 November 2015) In most theoretical descriptions of collective strong coupling of organic molecules to a cavity mode, the molecules are modeled as simple two-level systems. This picture fails to describe the rich structure provided by their internal rovibrational (nuclear) degrees of freedom. We investigate a first-principles model that fully takes into account both electronic and nuclear degrees of freedom, allowing an exploration of the phenomenon of strong coupling from an entirely new perspective. First, we demonstrate the limitations of applicability of the Born-Oppenheimer approximation in strongly coupled molecule-cavity structures. For the case of two molecules, we also show how dark states, which within the two-level picture are effectively decoupled from the cavity, are indeed affected by the formation of collective strong coupling. Finally, we discuss ground-state modifications in the ultrastrong-coupling regime and show that some molecular observables are affected by the collective coupling strength, while others depend only on the single-molecule coupling constant. DOI: 10.1103/PhysRevX.5.041022 Subject Areas: Chemical Physics, Plasmonics, Quantum Physics I. INTRODUCTION Strong coupling in quantum electrodynamics is a well- known phenomenon that occurs when the coherent energy exchange between a light mode and quantum emitters is faster than the decay and decoherence of either constituent [1,2]. The excitations of the system are then hybrid light- matter excitations, so-called polaritons, that combine the properties of both constituents. Exploiting these properties enables new applications, such as polariton condensation under collective strong coupling to excitons (excited electron-hole pairs) in semiconductors [3,4] and organic materials [57]. Organic materials present a particularly favorable case, as the Frenkel excitons in these materials possess large binding energies, large dipole moments, and can reach high densities. This enables Rabi splittings Ω R (the energy splitting between the polaritons) up to more than 1 eV [810], a significant fraction of the uncoupled transition energy. These properties allow for strong cou- pling to many kinds of electromagnetic (EM) modes [11], such as cavity photons [8,9,12], surface plasmon polaritons [1316], surface lattice resonances [17,18], or localized surface plasmons [19,20]. While organic molecules are thus uniquely suited to achieving strong coupling, they are not simple two-level quantum emitters, but rather have a complicated level structure including not only electronic excitations but also rovibrational degrees of freedom (schematically depicted in Fig. 1). It has been experimentally demonstrated that strong coupling can modify this structure, in the sense that material properties and chemical reaction rates change [2123]. However, the models used to describe strong coupling are often focused on macroscopic descriptions [24], and most microscopic models do treat organic molecules as two-level systems (see Ref. [25] for a recent review). When the rovibrational degrees of freedom are taken into account, this is often done using effective decay and dephasing rates [26], with a few works explicitly including a phononic degree of freedom [2730]. All of FIG. 1. Illustration of energy level structure of a bare complex molecule and the hybrid states that result in the strong-coupling regime with a photonic mode of energy ω c , resonant with the molecular excitation. * [email protected] Published by the American Physical Society under the terms of the Creative Commons Attribution 3.0 License. Further distri- bution of this work must maintain attribution to the author(s) and the published articles title, journal citation, and DOI. PHYSICAL REVIEW X 5, 041022 (2015) 2160-3308=15=5(4)=041022(14) 041022-1 Published by the American Physical Society

Upload: others

Post on 23-Dec-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

Cavity-Induced Modifications of Molecular Structure in the Strong-Coupling Regime

Javier Galego,1 Francisco J. Garcia-Vidal,1,2 and Johannes Feist1,*1Departamento de Física Teórica de la Materia Condensada and Condensed Matter Physics Center

(IFIMAC), Universidad Autónoma de Madrid, E-28049 Madrid, Spain2Donostia International Physics Center (DIPC), E-20018 Donostia/San Sebastián, Spain

(Received 10 June 2015; revised manuscript received 17 August 2015; published 9 November 2015)

In most theoretical descriptions of collective strong coupling of organic molecules to a cavity mode,the molecules are modeled as simple two-level systems. This picture fails to describe the rich structureprovided by their internal rovibrational (nuclear) degrees of freedom. We investigate a first-principlesmodel that fully takes into account both electronic and nuclear degrees of freedom, allowing an explorationof the phenomenon of strong coupling from an entirely new perspective. First, we demonstrate thelimitations of applicability of the Born-Oppenheimer approximation in strongly coupled molecule-cavitystructures. For the case of two molecules, we also show how dark states, which within the two-level pictureare effectively decoupled from the cavity, are indeed affected by the formation of collective strongcoupling. Finally, we discuss ground-state modifications in the ultrastrong-coupling regime and show thatsome molecular observables are affected by the collective coupling strength, while others depend only onthe single-molecule coupling constant.

DOI: 10.1103/PhysRevX.5.041022 Subject Areas: Chemical Physics, Plasmonics,Quantum Physics

I. INTRODUCTION

Strong coupling in quantum electrodynamics is a well-known phenomenon that occurs when the coherent energyexchange between a light mode and quantum emitters isfaster than the decay and decoherence of either constituent[1,2]. The excitations of the system are then hybrid light-matter excitations, so-called polaritons, that combine theproperties of both constituents. Exploiting these propertiesenables new applications, such as polariton condensationunder collective strong coupling to excitons (excitedelectron-hole pairs) in semiconductors [3,4] and organicmaterials [5–7]. Organic materials present a particularlyfavorable case, as the Frenkel excitons in these materialspossess large binding energies, large dipole moments, andcan reach high densities. This enables Rabi splittings ΩR(the energy splitting between the polaritons) up to morethan 1 eV [8–10], a significant fraction of the uncoupledtransition energy. These properties allow for strong cou-pling to many kinds of electromagnetic (EM) modes [11],such as cavity photons [8,9,12], surface plasmon polaritons[13–16], surface lattice resonances [17,18], or localizedsurface plasmons [19,20].While organic molecules are thus uniquely suited to

achieving strong coupling, they are not simple two-level

quantum emitters, but rather have a complicated levelstructure including not only electronic excitations but alsorovibrational degrees of freedom (schematically depicted inFig. 1). It has been experimentally demonstrated that strongcoupling can modify this structure, in the sense thatmaterial properties and chemical reaction rates change[21–23]. However, the models used to describe strongcoupling are often focused on macroscopic descriptions[24], and most microscopic models do treat organicmolecules as two-level systems (see Ref. [25] for a recentreview). When the rovibrational degrees of freedom aretaken into account, this is often done using effective decayand dephasing rates [26], with a few works explicitlyincluding a phononic degree of freedom [27–30]. All of

FIG. 1. Illustration of energy level structure of a bare complexmolecule and the hybrid states that result in the strong-couplingregime with a photonic mode of energy ℏωc, resonant with themolecular excitation.

*[email protected]

Published by the American Physical Society under the terms ofthe Creative Commons Attribution 3.0 License. Further distri-bution of this work must maintain attribution to the author(s) andthe published article’s title, journal citation, and DOI.

PHYSICAL REVIEW X 5, 041022 (2015)

2160-3308=15=5(4)=041022(14) 041022-1 Published by the American Physical Society

Page 2: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

these approaches only provide limited insight into theeffects of strong coupling on molecular structure.Furthermore, even the models that include phononic modestreat them as harmonic oscillators with a quadratic poten-tial. This is a good approximation close to the potentialminimum, but breaks down when the nuclei start movingsignificantly, e.g., in a chemical reaction.In the present work, we thus aim for a microscopic

description of strong coupling with organic molecules. Weintroduce a simple first-principles model that fullydescribes nuclear, electronic, and photonic degrees offreedom, but can be solved without approximations. Thisallows us to provide a simple picture for understanding theinduced modification of molecular structure.In Sec. II, after introducing the model, we discuss

under which conditions and in which form the Born-Oppenheimer approximation (BOA) [31,32] is valid inthe strong-coupling regime for a single molecule. The BOAis widely used in molecular and solid-state physics andquantum chemistry and provides a simple picture of nucleimoving on effective potential energy surfaces (PES)generated by the electrons, which underlies most of thecurrent understanding of chemical reactions [32]. However,the BOA depends on the separation of electronic andnuclear energy scales, i.e., the fact that electrons typicallymove much faster than nuclei. It could thus conceivablybreak down when an additional, intermediate time scale isintroduced under strong coupling to an EM mode. Thespeed of energy exchange between field and molecules isdetermined by the Rabi frequency ΩR, and typical exper-imental values of hundreds of meV land squarely betweentypical nuclear (≃100 meV) and electronic (≃2 eV) ener-gies. We show that the BOA indeed breaks down atintermediate Rabi splittings but remains valid when ΩRbecomes large enough. For cases where it breaks down, weshow that the non-BO coupling terms can be obtained to agood approximation without requiring knowledge of theelectronic wave functions.In Sec. III, we focus on the effects of strong coupling

when more than one molecule is involved, using twomolecules as the simplest test case. In experiments, strongcoupling is achieved by collective coupling to a largenumber of molecules, under which the Rabi frequency isenhanced by a factor of

ffiffiffiffiN

p. The number of emitters N is

on the order of ≳109 within cavities [8–10,12], withplasmonic nanoparticle modes allowing us to reduce thisto N ∼ 100 [20]. In this context, it is well known that only asmall fraction of the collective electronic excitations arestrongly coupled [25,33,34], with a large number of “dark”or “uncoupled” modes that show no mixing with the EMmode and no energy shift. We show that even these darkmodes are affected by strong coupling, with the nuclearmotion of separated molecules becoming correlated.In Sec. IV, we focus on the so-called ultrastrong-

coupling regime, where the Rabi frequency reaches a

significant fraction of the electronic transition energy, asachieved in experiments. In this regime, not just excited-state, but also ground-state properties are modified—forexample, the ground state acquires a photonic contribution[24,35]. Accordingly, we discuss whether ground-statechemical properties of organic molecules could be modi-fied by strong coupling. This also allows us to partiallyanswer the open question of what strong coupling meansfor modifications of chemical structure [36], i.e., whether“all”molecules are modified by it, or only a small subset, orwhether we necessarily have to invoke collective modeseven when discussing “single-molecule” effects. We showthat some observables, such as energy shifts, are deter-mined by the collective Rabi frequency, but other observ-ables, such as the shift in ground-state bond length, areinstead determined by the single-molecule couplingstrength ∝ ΩR=

ffiffiffiffiN

p.

For simplicity, we only treat a single EM mode andcompletely neglect dissipation in the following. We useatomic units unless stated otherwise (4πε0 ¼ ℏ ¼ me ¼e ¼ 1, with electron mass me and elementary charge e).

II. SINGLE MOLECULE

In this section, we introduce our model for a singlemolecule coupled to an EM mode. Because of the expo-nential scaling with the degrees of freedom, solving the fulltime-independent Schrödinger equation for an organic mol-ecule without the BOA is an extremely challenging task thateven modern supercomputers can handle for only very smallmolecules. We thus employ a reduced-dimensionality modelthat we can easily solve, both for the bare molecule and aftercoupling to an EM mode.

A. Method

1. Bare molecule

We work within the single-active-electron approxima-tion, in which all but one electron are frozen around thenuclei, and additionally restrict the motion of the activeelectron to one dimension x. Furthermore, we treat only onenuclear degree of freedom, the reaction coordinate R. Thiscould correspond to the movement of a single bond in amolecule, but can equally well represent collective motion,e.g., the breathing mode of a carbon ring. The effectivemolecular Hamiltonian then highly resembles that of a one-dimensional diatomic molecule,

Hm ¼ Tn þ Te þ Venðx; RÞ þ VnnðRÞ; ð1Þwhere Tn ¼ ðP2=2MÞ and Te ¼ ðp2=2Þ are the nuclear andelectronic kinetic energy operators (with P, p the corre-sponding momenta) and M is the nuclear mass. Thepotentials Venðx; RÞ and VnnðRÞ represent the effectiveelectron-nuclei and internuclear interactions, where weassume two nuclei located at x ¼ �R=2. These potentials

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-2

Page 3: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

encode the information about the frozen electrons as well asthe nuclear structure of the molecule and can be adjusted toapproximately represent different molecules.The electron-nucleus interaction Ven contains the inter-

action of the active electron with each nucleus, as well aswith the frozen electrons surrounding it. Assuming anuclear charge of Z, we have 2Z − 1 frozen electronsdistributed across the two nuclei. For large distances, theactive electron should thus feel a Coulomb potential withan effective charge of 1=2 from each nucleus. Conversely,at very small distances, the active electron is not affectedby the cloud of frozen electrons and feels an effectivecharge of Z. Since we are working within one dimension,we use a soft Coulomb potential to take into account thatthe electron avoids the singularity at the nucleus. Wechoose a simple model potential fulfilling these conditions:

VenðrÞ ¼ −12þ ðZ − 1

2Þe−r=r0ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

r2 þ α2p ; ð2Þ

where α is the softening parameter, r0 describes thelocalization of the frozen electrons around the nucleus,and r is the electron-nucleus distance. The total potential isthus Venðx; RÞ ¼ Venðjx − R=2jÞ þ Venðjxþ R=2jÞ.The internuclear potential VnnðRÞ represents the inter-

action between the nuclei and the 2Z − 1 frozen electrons,i.e., the ground-state potential energy surface of themolecular ion. We model this surface by a Morse potential,

VnnðRÞ ¼ Deð1 − eAðR−R0ÞÞ2; ð3Þ

which adds three new parameters: the dissociation energyDe, the equilibrium distance R0, and the width of thepotential well A. By tuning the seven free parameters wehave at our disposal (M, Z, α, r0, De, R0 and A), we canapproximately fit both the electronic and vibrationalstructure and absorption spectrum to those of real organicmolecules.We can now solve the stationary Schrödinger equation

HmΨðx; RÞ ¼ EΨðx; RÞ for the bare-molecule HamiltonianEq. (1) without further approximations by representing Hmon a two-dimensional grid in x and R. For a bare molecule,the results are virtually identical to those obtained withinthe BOA and thus not shown here.We next give a short description of the Born-Oppenheimer

approximation for completeness (see Refs. [31,32] for moredetails). As mentioned above, the basic idea is to exploit theseparation between nuclear and electronic time scales and toassume that the electrons perfectly follow nuclear rearrange-ments without changing state (i.e., adiabatically). This isachieved by separating the Hamiltonian into the nuclearkinetic energy Tn and an electronic Hamiltonian Heðx;RÞ ¼Hmðx; RÞ − Tn that only depends on R parametrically.Diagonalizing He yields a set fϕkg of electronic eigenstatesfor every R, with Heðx;RÞϕkðx;RÞ ¼ EkðRÞϕkðx;RÞ.

Without loss of generality, each total eigenstate Ψi can berepresented by Ψiðx; RÞ ¼ P

kϕkðx;RÞχikðRÞ. Inserting thisexpansion into the Hamiltonian Eq. (1) leads to a set ofcoupled differential equations,

ðTn þ EkÞχikðRÞ þXk0Ckk0n χik0 ¼ EχikðRÞ; ð4Þ

with nuclear motion taking place on potential energysurfaces EkðRÞ that are coupled through correction termsCkk0n ¼ hϕkjTnjϕk0 ix þ hϕkjP=Mjϕk0 ixP, where the sub-

script x indicates that the integration in the bra-kets is onlyover the electronic coordinate. The Born-Oppenheimerapproximation now consists in neglecting the intersurfacecouplings Ckk0

n , which can be shown to be small when theelectronic levels are well separated. This gives a set ofindependent PES EkðRÞ on which the nuclei move, whereeach eigenstate is a product of a single electronic and nuclearwave function, Ψi

kðx; RÞ ¼ ϕkðx;RÞχikðRÞ. The differentnuclear functions on each electronic curve correspond torotational or vibrational excitation. This picture of nuclearmotion on PES is extremely powerful and underlies mostof the current understanding of chemical reactions [32]. Thequestion of its validity in the strong-coupling regime is thusof central importance for the possible modification ofchemical reactions and structure through strong coupling.In the following, we focus on two model molecules,

which approximately reproduce the absorption spectra ofrhodamine 6G (R6G) and anthracene molecules that arecommonly used in experimental realizations of strongcoupling [12,15,17]. Only the first two PES, correspondingto the ground EgðRÞ and first electronically excited EeðRÞstate, play a role in the results discussed in the following.They are shown in Figs. 2(a) and 2(c), together with thenuclear probability densities jχðRÞj2 for the lowest vibra-tional levels on each PES. Importantly, the two modelsdiffer significantly in two relevant quantities: the vibra-tional mode frequency ωvib and the offset ΔR, i.e., thechange in equilibrium distance between the ground andexcited PES. This offset is related to the strength of theelectron-phonon interaction and influences the Stokesshift between emission and absorption [37]. The R6G-likemodel has relatively small vibrational spacing ωvib ≈70 meV and small offset ΔR ≈ 0.018 a:u:, while theanthracenelike model has large vibrational spacing ωvib ≈180 meV and large offset ΔR ≈ 0.092 a:u:. Accordingly,their absorption spectra [Figs. 2(b) and 2(d), obtained fromEq. (6)] are qualitatively different, with anthracene showinga broader absorption peak with well-resolved vibronicsubpeaks.

2. Molecule-photon coupling

We now add a photonic mode and its coupling to themolecule (within the dipole approximation) into the molecu-lar Hamiltonian,

CAVITY-INDUCED MODIFICATIONS OF MOLECULAR … PHYS. REV. X 5, 041022 (2015)

041022-3

Page 4: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

Hmc ¼ Hm þ ωca†aþ gμða† þ aÞ; ð5Þ

where μ is the dipole operator of the molecule (μ ¼ xin our case), a† and a are the creation and annihilationoperators for the bosonic EM field mode, ωc is its frequency,and g is the coupling strength constant, given by the electricfield amplitude (along the x axis) of a single photon. In thefollowing, we always set the photon energy ωc to achieve“zero detuning,” with ωc at the absorption maximum of themolecule. This gives ωc ≈ EeðReÞ − EgðReÞ, where Re isthe equilibrium position at which EgðRÞ has its minimum.To provide some context for the field strengths used in

the following, we note that for a typical microcavity witha mode volume V ≈ λ3c, one obtains g ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

ℏωc=2ε0Vp

≈1.34 × 10−7ω2

c a:u: (for ωc given in eV). However, for aneffective mode volume close to the current record achievedin plasmonic nanoantennas, V ≈ 1.3 × 10−7λ3c [38], thesingle-particle coupling reaches g ≈ 3.72 × 10−4ω2

c a:u:(ωc again in eV). We furthermore note that the ground-to-excited-state dipole transition moments of our modelmolecules are on the order of 1 a:u: ≈ 2.54 D, i.e., almostan order of magnitude smaller than in typical organicmolecules [39].Compared to the bare-molecule case, the Hamiltonian

now includes a new degree of freedom, the photon numbern ∈ f0; 1; 2;…g, with the system eigenstates defined byHmcΨðx; n; RÞ ¼ EΨðx; n; RÞ. As discussed above, the

typical energies associated with strong coupling in organicmolecules are somewhere between the nuclear and elec-tronic energies. A priori, this suggests two options ofperforming the BOA: the additional terms introduced bythe photonic degree of freedom could be grouped eitherwith the “slow” nuclear motion or with the “fast” electronicHamiltonian. However, as the photon couples to theelectron, grouping it with the nuclear terms necessarilyleads to additional off-diagonal terms in Eq. (4), and noindependent PES on which the nuclei move could beobtained. Consequently, the only way to maintain theusefulness of the BOA is to include the photonic degreeof freedom within the electronic Hamiltonian, leading to anew set of “strongly coupled PES.”We first focus on the singly excited subspace, within

which the splitting between polaritons is observed. Here,either the molecule is electronically excited and no photonsare present or the molecule is in its electronic ground stateand the photon mode is singly occupied. At zero coupling(g ¼ 0), this gives two uncoupled PES [EeðRÞ andEgðRÞ þ ωc, dashed curves in Fig. 3] that cross close toRe for our choice ofωc. When the electron-photon couplingis nonzero but small, a narrow avoided crossing developsinstead [solid lines in Fig. 3(a)], while for large couplingstrengths, the energy exchange between photonic andelectronic degrees of freedom is so fast that we observetwo entirely new PES [Fig. 3(b)], which cannot be easilyassociated with either of the uncoupled PES. Instead, theybecome hybrid polaritonic PES that contain a mixture ofelectronic and photonic excitation, the hallmark of thestrong-coupling regime.As discussed above, the BOA is known to be valid when

two PES are sufficiently separated from each other. Thisimplies that the BOA breaks down when g is small and thetwo PES possess a narrow avoided crossing. This in itself isnot a surprising result—when the electron-photon couplingis very small, the system is not even in the strong-couplingregime, and the photon mode is better treated as a small

FIG. 2. Bare-molecule potential energy surfaces of the two firstelectronic states in the BOA for (a) the rhodamine 6G-like modelmolecule and (c) the anthracenelike model molecule. The vibra-tional levels and associated nuclear probability densities arerepresented on top of the PES. (b), (d) Absorption spectrum forthe (b) R6G-like and (d) anthracenelike molecule in arbitraryunits.

FIG. 3. Strongly coupled electronic PES (solid lines) in thesingly excited subspace, for the anthracenelike molecule for(a) g ¼ 0.001 a:u: and (b) g ¼ 0.008 a:u:. The dashed lines showthe corresponding uncoupled states: A molecule in the firstexcited state, EeðRÞ, and a molecule in the ground state with onephoton present, EgðRÞ þ ωc.

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-4

Page 5: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

perturbation. Fortunately, the weak-coupling regime is alsonot interesting from the standpoint of understanding ormodifying molecular structure through strong coupling.The real question thus must be: How strong does theelectron-photon coupling have to be for the BOA to bevalid, and is this condition fulfilled for realistic experi-mental parameters? In order to better quantify the agree-ment between the BOA and the full solution, we next turnto an easily measured physical observable, the absorptionspectrum.

B. Absorption

In order to calculate the absorption spectrum that wouldbe observed under driving by an external field, the details ofthe experimental setup would have to be taken into account.For example, for a microcavity, an input-output formalism[40], in which the cavity mode is driven by externalphotons through the cavity mirrors, would be most appro-priate. On the other hand, if the molecules are placed nextto a metallic nanoparticle, an external field would typicallydrive both the molecules and the localized surface plasmonresonance. For easier comparison between different valuesof g, we calculate the absorption under the assumption thatonly the molecules are directly coupled to the external lightsource. This allows us to focus on the influence of themolecular structure on the absorption spectrum, withoutcontamination from a peak due to the bare EM mode atlow coupling g. We explicitly checked that our conclusionsconcerning the validity of the BOA do not depend onwhether the molecule or the cavity mode is externallydriven. Under these assumptions, the absorption crosssection at frequency ω can be calculated using the opticaltheorem as [41,42]

σðωÞ ¼ 4πω

cIm lim

ε→0

Xk

jhψkjμjψ0ij2ωk − ω0 − ω − iε

; ð6Þ

where the sum runs over all eigenstates jψki of the system(with energies ωk) and jψ0i is the ground state. As we donot include incoherent processes in our calculation, thiswould give δ-like peaks in the absorption cross section. Inthe plots shown in the following, we instead introduce aphenomenological width representing losses and puredephasing by setting ε to a small nonzero value, such thatthe absorption cross section becomes a sum of Lorentzians.For the bare-molecule case without coupling to an EMmode, the absorption spectra of our two model moleculesapproximately agree with those of R6G [Fig. 2(b)] [17] andanthracene [Fig. 2(d)] [12].In Fig. 4, we compare the absorption cross sections

under strong coupling as obtained from a full calculationwithout approximations to those obtained within the BOA,for a range of coupling strengths g to the EM mode. Evenfor relatively small g, the BOA is found to agree almostperfectly with the full results for the R6G-like molecule

with small vibrational spacing [Fig. 4(a)]. However, for theanthracenelike molecule with a high-frequency vibrationalmode and large offset ΔR, the BOA only agrees with thefull result for relatively large values of g, where the Rabisplitting ΩR (as defined by the energy difference betweenthe two “polariton” peaks in the absorption spectrum) isappreciably larger than the vibrational frequency ωvib ≈180 meV [Fig. 4(b)]. As an aside, we note here that forintermediate values of the coupling strength [e.g., forg ¼ 0.002 a:u: in Fig. 4(b)], the EMmode strongly coupleswith the individual vibronic subpeaks, as observed inexperiments using anthracene [5,12].This qualitative observation can be quantified by com-

paring the non-Born-Oppenheimer correction terms Ckk0n in

Eq. (4) with the energy difference between the anticrossingPES at the point of closest approach. In Appendix A, wepresent a model that achieves this without any explicitknowledge of the electronic wave functions. It relies on theobservation that close to the anticrossing, the coupled(polariton) states switch character between the twouncoupled states, while the “intrinsic” R dependence ofthe uncoupled electronic states can be neglected. Thecorrection terms Ckk0

n can then be obtained just from theknowledge of EgðRÞ, EeðRÞ, and μegðRÞ, where μegðRÞ isthe electronic transition dipole between the ground andexcited state. By approximating EgðRÞ and EeðRÞ asharmonic oscillators, the correction terms can be analyti-cally evaluated and are found to be negligible under thecondition that ΔRω2

vib=Ω2R is small compared to the nuclear

momentum of the relevant eigenstates. This demonstratesthat the model molecules present two opposite cases forthe applicability of the BO approximation: our R6G-likemolecule has a relatively small vibrational spacingωvib ≈ 70 meV and small electron-phonon coupling,ΔR ≈ 0.018 a:u:, while our anthracenelike model molecule

FIG. 4. Absorption cross sections of a single molecule, calcu-lated using the full Hamiltonian without approximation (solidgreen lines) and within the BOA (dashed black lines). Resultsare shown for the (a) R6G-like and (b) anthracene-like modelmolecules, for several values of the coupling strength g.

CAVITY-INDUCED MODIFICATIONS OF MOLECULAR … PHYS. REV. X 5, 041022 (2015)

041022-5

Page 6: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

has a large vibrational spacing ωvib ≈ 180 meV and largeelectron-phonon coupling, ΔR ≈ 0.092 a:u:. We note thatin many experiments involving organic molecules, ΩR ≳500 meV [9,10] is significantly larger than typical vibra-tional frequencies ωvib ≲ 200 meV [43]. This shows thatthe intuitive picture of nuclear dynamics unfolding onuncoupled Born-Oppenheimer potential energy surfacescan often be applied to understand the modification ofmolecular chemistry induced by strong coupling.Additionally, even when the BOA breaks down, the modelpresented in Appendix A can be used to obtain the non-BOcoupling terms without requiring knowledge of the elec-tronic wave functions. The only necessary inputs are theuncoupled PES and the associated transition dipolemoments. Even for relatively large molecules, these canbe obtained using the standard methods of quantumchemistry or density functional theory (DFT).

III. TWO MOLECULES

In the previous section, we show that on the single-molecule level, the BOA is valid as long as the Rabisplitting ΩR is large enough. However, current experimentsare performed with a large number of molecules, wherecoherent superpositions of electronic excitations (bright“Dicke states” [44]) couple strongly to the photonicmode(s), while other superpositions give uncoupled ordark modes. It is thus important to consider if and how ourconclusions have to be modified when more than a singlemolecule is involved in strong coupling.For later reference, we give a quick overview of the

theory when using an ensemble of two-level emitterscoupled to a photonic mode, i.e., the many-particleJaynes-Cummings (JC) model [45], also known as theTavis-Cummings model [46]. Its Hamiltonian within therotating wave approximation is

HJC ¼ ωca†aþXi

ωic†i ci þ

Xi

giðac†i þ a†ciÞ; ð7Þ

where ωi is the energy of emitter i with destruction(creation) operator ci (c

†i ) and the gi describe the emitter-

photon couplings. For identical emitters (ωi ¼ ωm, gi ¼ g),the resulting eigenstates in the single-excitation subspaceare given by two polaritons j�i¼ð1= ffiffiffi

2p Þða†j0i�jBiÞ,

symmetric and antisymmetric combinations of the photonicmode with the emitter bright state jBi ¼ ð1= ffiffiffiffi

Np ÞPic

†i j0i.

At zero detuning (ωc ¼ ωm), the polariton energies aregiven by ω� ¼ ωm �ΩR=2, where ΩR ¼ 2g

ffiffiffiffiN

pis the

collective Rabi splitting. The N − 1 superpositions ofemitter states orthogonal to jBi are dark states that arenot coupled to the photonic mode, with energies identical tothe uncoupled emitters, ωDS ¼ ωm. Note that in configu-rations with many photonic modes (e.g., planar cavities),more than one emitter state is coupled to the photonic mode

(typically at low in-plane momentum), but there remainmany uncoupled (dark) modes at higher in-plane momen-tum [25,34]. There is an ongoing discussion in the literatureon whether the dark modes are affected by strong couplingas well, or whether they should be thought of as completelyunmodified emitter states. We show below that when takingthe internal structure of the emitters (molecules) intoaccount, even the dark modes are affected by strongcoupling and the nuclear dynamics of separate moleculesbecome correlated.

A. Method

We now treat the case of two model molecules, whichcan still be solved exactly within our approach, but whichdisplays many of the effects of many-molecule strongcoupling. As in the JC model, we assume that the twomolecules both couple to the same photonic mode, but donot directly interact with each other, giving

H2mmc ¼ ωca†aþ

Xj¼1;2

½HðjÞm þ gμðjÞða† þ aÞ�; ð8Þ

where the superscripts j indicate the molecule on which theoperator acts. Directly diagonalizing this Hamiltonian inthe “raw” basis fx1; R1; x2; R2; ng is already a formidablecomputational task for typical grid sizes. We thus calculatethe full solution by first diagonalizing the single-moleculeHamiltonian, Hm ¼ P

kEkjkihkj, and including only arelevant subset of eigenstates fkg for each molecule inthe total basis fk1; k2; ng. The number of necessaryeigenstates to obtain completely converged results is quitesmall (≈30 per molecule). However, this approach onlyprovides limited insight into the dynamics of the stronglycoupled system, especially regarding nuclear motion.We thus again apply the Born-Oppenheimer approxi-

mation by separating the nuclear kinetic energy terms anddiagonalizing the remaining Hamiltonian parametrically asa function of R1 and R2. Similar to above, instead ofworking in the fx1; x2; ng basis for each combinationðR1; R2Þ, we prediagonalize the single-molecule electronicHamiltonian Heðx;RÞ ¼

PkEkðRÞjkðRÞihkðRÞj, where

(for the cases discussed here) the sum only has to includethe ground and first excited states to achieve convergence,k ∈ fg; eg. If we additionally allow at most one photon inthe system, n ∈ f0; 1g, we obtain an 8 × 8 Hamiltonian foreach combination of nuclear coordinates R1, R2.The electronic Hamiltonian consists of all possible

combinations of electronic states Eg, Ee of the twomolecules with 0 or 1 photons. A further simplificationis achieved by taking into account that the Hamiltonianconserves parity Π ¼ ð−1Þπ1þπ2þn, with πj the parity of thestate of molecule j (gerade or ungerade). For large couplingg, this separation by parity avoids some accidental degen-eracies between uncoupled PES and thus improves theBOA. We now obtain two independent 4 × 4Hamiltonians,

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-6

Page 7: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

HevenðR1; R2Þ ¼

0BBBBB@

Egg0 gdð1Þ gdð2Þ 0

gdð1Þ Eeg1 0 gdð1Þ

gdð2Þ 0 Ege1 gdð2Þ

0 gdð1Þ gdð2Þ Eee0

1CCCCCA; ð9aÞ

HoddðR1; R2Þ ¼

0BBBBB@

Egg1 gdð1Þ gdð2Þ 0

gdð1Þ Eeg0 0 gdð1Þ

gdð2Þ 0 Ege0 gdð2Þ

0 gdð1Þ gdð2Þ Eee1

1CCCCCA; ð9bÞ

where the uncoupled PES are represented by the compactnotation Eabn ¼ EaðR1Þ þ EbðR2Þ þ nωc and the single-molecule dipole transition moment between the ground andfirst excited state is denoted by dðjÞ ¼ hϕgðRjÞjμjϕeðRjÞi.Diagonalizing these Hamiltonians for each ðR1; R2Þ resultsin a set of strongly coupled two-dimensional PES. In Fig. 5,we show the three surfaces in the single-excitation sub-space, corresponding to the three lowest states of Eq. (9b).For zero molecule-photon coupling (g ¼ 0) [Fig. 5(a)],there are now a number of one-dimensional seams wherethe three PES cross. When the molecule-photon coupling isturned on, these crossings again turn into avoided cross-ings, as shown in Figs. 5(b) and 5(c) for two differentcoupling strengths g. Following the conventions used in theJaynes-Cummings model, we label the three coupled PESin order of energy as the “lower polariton (LP) PES,” the“dark-state (DS) PES,” and the “upper polariton (UP) PES.”We first address the applicability of the BOA, which

breaks down when two PES are close in energy, for the caseof two molecules. Within the single-excitation subspace(which determines the linear properties of the system, such

as absorption), there are now a range of (avoided) cross-ings. They occur when (i) all three surfaces approach eachother, Egg1 ≈ Ege0 ≈ Eeg0, (ii) the photonically excited PESis close to only one of the electronically excited PES,Egg1 ≈ Ege0 or Egg1 ≈ Eeg0, or (iii) only the two electroni-cally excited states cross, Ege0 ≈ Eeg0. Case (i) correspondsto the JC model at zero detuning, giving the two polaritonicPES at energy shifts of�ΩR=2, and an additional dark statethat is unshifted from the bare-molecule case. The BOA inthis region is thus valid for similar conditions as in thesingle-molecule case, although the PES separation isreduced by half due to the additional dark-state surface.Case (ii) corresponds exactly to the single-molecule case,with the second molecule acting as a “spectator” thatinduces only additional energy shifts. The BOA shouldthus again be valid for similar conditions as with a singlemolecule, albeit with the coupling reduced by 1=

ffiffiffi2

pfor a

fixed total Rabi splitting. Finally, case (iii) presents thebiggest challenge, as the two electronically excited PES,Eeg0 and Ege0, are not directly coupled, but only splitindirectly through coupling to the photonically excitedsurface Egg1. The splitting between the two surfaces is thussmall for large detuning, ΔE ≈ ðgdÞ2=4ðEgg1 − Eeg0Þ. Thisis clearly observed in Fig. 5(b) along the line R1 ¼ R2,where the dark-state PES almost touches the upper polar-iton PES for small R’s and the lower polariton PES forlarge R’s.

B. Absorption

The discussion above implies that, for almost anycoupling strength, there will be regions in the nuclearconfiguration space R1, R2 where the BOA breaks down.However, not all parts of the PES are visited by the nuclei

FIG. 5. (a) Uncoupled potential energy surfaces of two anthracenelike molecules in the singly excited subspace: Eeg0ðR1; R2Þ(orange), Eeg0ðR1; R2Þ (blue), and Egg1ðR1; R2Þ (green). (b) Coupled PES for g ¼ 0.002 a:u: and (c) g ¼ 0.013 a:u:, corresponding tothe lower polariton (orange), dark state (blue), and upper polariton (green). For clarity, only parts where R1 < R2 are shown (note thatthe system is symmetric under the exchange R1 ↔ R2).

CAVITY-INDUCED MODIFICATIONS OF MOLECULAR … PHYS. REV. X 5, 041022 (2015)

041022-7

Page 8: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

during a given physical process. To explicitly check theBOA in the subspace relevant for polaritonic physics, inFig. 6 we thus again compare the absorption with thatobtained by a full diagonalization of the HamiltonianEq. (8). Compared to the single-molecule case, many moremolecular levels are present in the system, leading to smallchanges in the absorption spectra. In order to properlycompare the results, we take into account the

ffiffiffiffiN

pscaling of

the total Rabi frequency and reduce the coupling strengthsby

ffiffiffi2

pto produce the same total splitting. The BOA is

shown to again become valid for large enough coupling, butthe minimum coupling required is increased compared tothat for a single molecule. In the common case of slownuclear motion, as for our R6G-like model in Fig. 6(a), theBOA is already valid for relatively small Rabi splitting ofΩR ≈ 250 eV. However, in the anthracenelike case ofvery fast vibrational motion, Fig. 6(b), the BOA still doesnot give perfect agreement with the full model forg ¼ 0.0057 a:u: (ΩR ≈ 600 meV), and agreement is onlyreached at roughly twice that value.

C. Nuclear correlation

Having established the validity of the BOA for manyrelevant cases and Rabi splittings comparable to experimen-tal values, we now discuss the implications of collectivestrong coupling for the internal molecular (nuclear) dynam-ics. Note that this question, by design, cannot be addressedwithin the JC model, where emitters are two-level systemswithout any internal degrees of freedom. In contrast, theBOA provides a straightforward approach to this problem.Any two-dimensional PES can be decomposed into a sum ofindependent single-molecule potentials plus a remainder that

describes the coupling between the nuclear motion of themolecules:

EðR1; R2Þ ¼ E1ðR1Þ þ E2ðR2Þ þ E12ðR1; R2Þ: ð10Þ

The nuclear motion of two molecules is independent ifand only if the coupled part E12ðR1; R2Þ is identically zero.In order to quantify this coupling, we expand each of thecoupled PES in the single-excitation subspace around itsminimum ðR0

1; R02Þ, giving

EðR1; R2Þ ≈ E0 þ αδR21 þ αδR2

2 þ βδR1δR2; ð11Þ

with E0 ¼ EðR01; R

02Þ and δRi ¼ Ri − R0

i . Note that due tosymmetry under the exchange R1 ↔ R2, the prefactor α isthe same for δR2

1 and δR22. As can be seen in Fig. 7, both the

polariton and even the dark-state PES show significantcoupling of the nuclear degrees of freedom, with valuesof β=α on the order of a few percent for values ofg≲ 0.01 a:u: giving Rabi splittings of ≲1 eV (see Fig. 6).Interestingly, the coupling is much larger for the lowerpolariton PES than for either the upper polariton or thedark-state PES, and decreases with increasing g for allthree PES. We therefore conclude that even dark states thathave negligible mixing with photonic modes are affectedby strong coupling, in the sense that the nuclear degreesof freedom of separate molecules behave like coupledharmonic oscillators, and their motion becomes correlated.This implies that, e.g., local excitation of nuclear motionwithin one molecule could affect the nuclear motion inanother, spatially separated molecule, even when no photonis ever present in the EM mode of the system.Note that the BOA results predict monotonically increas-

ing correlation for arbitrarily small (but nonzero) values ofg. This again shows that the BOA is not correct in the limitof small coupling g → 0, where the correlation should also

FIG. 6. Absorption cross section of two molecules drivencoherently, calculated using the full Hamiltonian withoutapproximation (solid green lines) and within the BOA (dashedblack lines). Results are shown for the (a) R6G-like and(b) anthracenelike model molecules, for several values of thecoupling strength g. The values of g are scaled by 1=

ffiffiffi2

pwith

respect to the single-molecule case (Fig. 4) in order to obtain thesame total Rabi frequency ΩR.

FIG. 7. Coupling between nuclear motion in different mole-cules for the lower (LP) and upper polariton (UP) and dark-state(DS) PES. Results are shown as the ratio β=α between theprefactors of the off-diagonal R1R2 and diagonal R2

i terms inEq. (11), for the R6G-like model molecule.

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-8

Page 9: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

go to zero as the molecules are completely uncoupled. Wethus start Fig. 7 at g ¼ 0.002 a:u:, for which the BOAalready produces good agreement with the full result in theabsorption cross section (cf. Fig. 6), and note that ourresults indicate that there is a maximum of correlation inthe nuclear motion at intermediate coupling strengths. InAppendix B, we additionally present results for the mutualinformation of the nuclear probability distribution obtainedunder driving of a single molecule, which allow us tocompare the BOA prediction with the results of a fulldiagonalization of the Hamiltonian, and also allow us tophenomenologically include decay processes.All results above are obtained using two molecules. An

important question is thus whether the same correlationcould be observed when using macroscopically largenumbers of molecules. While an exhaustive answer isoutside the scope of the current paper, initial tests usingthree molecules indicate that the observed correlation innuclear motion does survive as the number of moleculesis increased.Note that all of the results discussed here apply within

the singly excited subspace, i.e., the coupled nuclearmotion is observed only when electronic excitation ispresent, not in the ground state. In the next section, wediscuss which modifications of the ground state propertiescould be observed in the ultrastrong-coupling regime.

IV. ULTRASTRONG COUPLING ANDGROUND-STATE MODIFICATIONS

Up to now, we have focused on the molecular propertiesin the singly excited subspace, which are probed in linearresponse measurements such as absorption and transmission,and where the effect of strong coupling is immediatelyapparent. However, when the Rabi frequency, i.e., the energyexchange rate between the molecules and the photonicmode, becomes significant compared to the frequencies ofthese two modes, the so-called ultrastrong-coupling regimeis entered [21–24,35]. In this regime, the rotating waveapproximation for the emitter-light interaction (under whichthe number of excitations is conserved) becomes invalid. Inour approach, the rotating wave approximation is not used,and we can thus naturally explore the ultrastrong-couplingregime. One of its most intriguing predictions is that even theground-state properties of the system should be significantlymodified. For example, the ground state is shifted in energyand acquires a photonic component, such that a number ofvirtual photons are present in the system even without anyexternal excitations. This raises the question of how theinternal degrees of freedom of organic molecules are affectedwhen this regime is entered.The BOA is well suited to explore this regime. In contrast

to the singly excited subspace, where narrow avoidedcrossings can affect its validity, the ground-state PES isenergetically well separated from all other PES. This remainstrue even under ultrastrong coupling, and consequently the

BOA is valid for all coupling strengths. The ground-statepotential energy surface EgðRÞ is coupled to the doublyexcited surface EeðRÞ þ ωc [cf. Eq. (9a)], with the stronglycoupled ground-state PES given to lowest order byESCg ðRÞ ≈EgðRÞ−g2μ2egðRÞ=½EeðRÞþωc−EgðRÞ�þOðg4Þ.

Ground-state properties such as the bond length are deter-mined by the shape of the PES. The largest modification canthus be expected when the R dependence of the ground andexcited PES is as different as possible. This occurs for largeelectron-phonon coupling, i.e., a large value of ΔR, such asin our anthracenelike molecule. For a coupling strength ofg ¼ 0.016 a:u:, corresponding to a Rabi splitting of ΩR ≈1.2 eV in absorption, we obtain a shift in the ground-statebond length of ΔR0 ≈ 0.84mÅ ¼ 84 fm. While small, sucha change in bond length could be detectable using x-rayabsorption fine-structure spectroscopy or x-ray crystallog-raphy, which can obtain few- or even subfemtometerprecision for measuring bond-length shifts [47,48].However, the previous paragraph applies only for a

single molecule under strong coupling. Repeating thecalculation using two molecules and taking into accountthe reduced single-molecule coupling strength (for fixedRabi splitting, ΩR ∝

ffiffiffiffiN

pg) reveals a reduction of the bond-

length change by a factor of 2, ΔR2mol0 ≈ 0.42 mÅ. This is

confirmed by using a similar analytical model as presentedin Appendix A, in which the bare-molecule potentialenergy surfaces are approximated as harmonic oscillators.Because of the simple analytical structure, perturbationtheory can be applied to obtain results for arbitrary numbersof molecules, and the change in ground-state bond lengthis found to be proportional to the squared single-moleculecoupling g2, not to the squared collective Rabi frequencyΩ2

R. We note that, in contrast, the ground-state energy shiftis indeed determined by the collective coupling strength,ΔE0 ∝ Ω2

R. In realistic experimental configurations involv-ing large numbers of molecules, the change in ground-statebond length is thus expected to be minuscule and extremelychallenging to measure. This demonstrates that the influ-ence of strong coupling on any specific observable is notimmediately obvious, and has to be checked case by case.For some properties, the molecules will behave as if theyfeel the full collective coupling ΩR, while for others, theywill show only the change induced by the single-moleculecoupling g. These results are also compatible with theexperimental observation that the vibrational frequenciesin surface-enhanced Raman scattering, which probe theground-state PES, are not strongly modified under strongcoupling [49].We thus check another observable, and ask whether the

ground state will show correlated nuclear motion betweendistant molecules, as observed in the dark-state surface.This can again be quantified using the expansion of thePES in Eq. (11). Doing so reveals a very small couplingparameter β that to lowest order is proportional to g4=ω5

c[close to zero detuning, ωc ∼ EeðRÞ − EgðRÞ]. This

CAVITY-INDUCED MODIFICATIONS OF MOLECULAR … PHYS. REV. X 5, 041022 (2015)

041022-9

Page 10: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

corresponds to an even higher-order correction than thealready small energy or bond-length shifts. Furthermore,like the bond-length shift, it depends on the single-molecule coupling instead of the collective couplingstrength. We can thus conclude that in contrast to theexcited states, the ground-state nuclear motion of themolecules does not become correlated even in the ultra-strong-coupling limit.

V. CONCLUSIONS AND OUTLOOK

We present a simple model that offers a new perspectiveon strong coupling with organic molecules. We show underwhich conditions the molecular properties under strongcoupling can be understood by the modification of thepotential energy surfaces determining nuclear dynamicsunder the Born-Oppenheimer approximation. In particular,we find that in many cases of experimental interest wherethe Rabi splitting is large, the BOA is applicable andprovides an intuitive picture of the strongly coupleddynamics. However, we also show that for molecules withfast vibrational modes and large phonon-exciton couplings,the BOA can break down and a more involved picture isnecessary. We furthermore demonstrate that the non-BOcoupling terms between PES in this case are dominantlydue to the change of character between light and matterexcitations which can be obtained from simple few-levelmodels without requiring knowledge of the electronic wavefunctions.In addition, we show that under collective strong

coupling involving more than one molecule, the nucleardynamics of the molecules in electronic “dark states” thatare only weakly coupled to the photonic mode are none-theless affected by the formation of strong coupling.In particular, we find that the dark state PES describescoupling between the nuclear degrees of freedom of thedifferent molecules.Finally, we investigate the change of the ground-state

properties under ultrastrong coupling, where the Rabisplitting becomes a significant fraction of the transitionenergy. Using our numerical calculations and a simpleanalytical model, we show that while the ground-stateenergy shift is affected by the collective Rabi frequency(which is enhanced by

ffiffiffiffiN

pfor N molecules), other proper-

ties such as the ground-state bond length depend on thesingle-molecule coupling strength and are not significantlyaffected for experimentally realistic parameters.Our results also lay the groundwork for a further in-depth

exploration of the modification of molecular propertiesunder strong coupling. In particular, they provide a simplepicture that can be used to understand the modificationof chemical reactions, e.g., by lowering a potential barrieralong a reaction coordinate. There are also a number ofobvious extensions of the simple model presented here thatwill be explored in the future. These include more realisticmodels of organic molecules using more degrees of

freedom (for example, employing the PES obtained usingquantum chemistry packages) and the inclusion of inco-herent processes such as decay and decoherence. We notethat there has been some recent progress on combiningQED with density functional theory [50,51], which couldprovide complementary information to the model pre-sented here.

ACKNOWLEDGMENTS

This work has been funded by the European ResearchCouncil (ERC-2011-AdG Proposal No. 290981), by theEuropean Union Seventh Framework Programme underGrant Agreement FP7-PEOPLE-2013-CIG-618229, andthe Spanish MINECO under Contracts No. MAT2011-28581-C02-01 and No. MAT2014-53432-C5-5-R.

Note added.—Recently, another paper [52] appearedthat reports the fact some molecular observables dependon the single-molecule coupling g, not the collectivecoupling g

ffiffiffiffiN

p.

APPENDIX A: MODEL FOR NON-BORN-OPPENHEIMER CORRECTIONS

In this appendix, we derive an analytical model for thenon-Born-Oppenheimer corrections Ckk0

n under molecularstrong coupling, for a single molecule. We treat the twoPES in the single-excitation subspace, EgðRÞ þ ωc andEeðRÞ, coupled by the term gμegðRÞ. This leads to a 2 × 2

BO Hamiltonian of the form

HðRÞ ¼�EgðRÞ þ ωc gμegðRÞgμegðRÞ EeðRÞ

�; ðA1Þ

which can be easily diagonalized to obtain polaritoneigenstates jþi¼ cosθjg1iþ sinθje0i and j−i¼sinθjg1i−cosθje0i, where jani is short for jϕaðx;RÞ; ni, and

tan 2θ ¼ 2hðRÞδEðRÞ ; ðA2Þ

where we define δEðRÞ¼EgðRÞþωc−EeðRÞ and hðRÞ ¼gμegðRÞ. Using EavðRÞ ¼ ½EgðRÞ þ ωc þ EeðRÞ�=2, theeigenenergies are given by

E�ðRÞ ¼ EavðRÞ �1

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi4h2ðRÞ þ δEðRÞ2

q: ðA3Þ

We can now evaluate the non-Born-Oppenheimer cou-pling terms Ckk0

n ¼ hkjTnjk0i þ hkjP=Mjk0iP within thismodel, using a series of approximations to obtain simpleanalytical results. First, we linearize δEðRÞ ≈ a0ðR − RcÞaround the point of intersection between the two PES,where EgðRcÞ þ ωc ¼ EeðRcÞ. Second, in the spirit of theFranck-Condon approximation, we assume that the dipole

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-10

Page 11: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

coupling is constant over the range of relevant R values,and set hðRÞ ¼ h0. Following the same idea, we addition-ally assume that the electronic wave functions do notchange significantly with R; i.e., ð∂=∂RÞjϕaðx;RÞi ≈ 0.This implies that the change in the polaritonic eigenfunc-tions j�i close to the avoided crossing at Rc is mostly dueto the switchover between the two uncoupled surfaces, i.e.,the change in θðRÞ, not because of an intrinsic change ofelectronic state with R. With these approximations, thecorrection terms become

h−jPjþi ¼ −ia0h04h20 þ a20ðR − RcÞ2

; ðA4aÞ

h−jP2jþi ¼ 2a30h0ðR − RcÞ½4h20 þ a20ðR − RcÞ2�2

; ðA4bÞ

h�jP2j�i ¼ a20h20

½4h20 þ a20ðR − RcÞ2�2; ðA4cÞ

with the diagonal terms h�jPj�i identically zero. Notethat diagonal terms correspond only to energy shifts anddo not induce additional transitions [32]. The non-Born-Oppenheimer coupling between the polariton surfaces hasa Lorentzian shape around the avoided crossing, and asexpected, only becomes non-negligible close to it. As shownin Fig. 8, the non-Born-Oppenheimer corrections obtainedfrom this simple model agree almost perfectly with thoseobtained from the full numerical calculation for our anthra-cenelike model molecule. The only molecule-specific infor-mation entering the model is the PES of the uncoupledmolecule and the dipole moment between the coupledsurfaces. Specifically, the electronic wave functions are

never used here, and their derivative as a function of thenuclear coordinates is not required. This implies that thismodel could be used to obtain accurate non-BO correctionsthat describe the transitions between potential surfaces evenwhen the full electronic wave functions of a molecule are notavailable (e.g., in DFT calculations). The dynamics of themolecule could thus be fully recovered within a potentialenergy surface picture even when the BOA per se is notapplicable.We now exploit this model to derive a condition for when

the BOA becomes applicable, i.e., when the non-BO termsbecome negligible. We approximate the bare molecularpotential energy surfaces as two harmonic oscillators withthe same vibrational frequency ωvib, but with an offset inenergy and equilibrium position,

EgðRÞ ≈Mω2

vib

2R2; ðA5Þ

EeðRÞ ≈Mω2

vib

2ðR − ΔRÞ2 þ ΔE; ðA6Þ

where without loss of generality, we choose the origin in RandE at the minimum ofEgðRÞ. Note that this model exactlyresults from the common approximation of linear couplingbetween a single electronic excitation and a bosonic vibra-tional mode [53,54]. Within this model, δEðRÞ ¼ EgðRÞ þωc − EeðRÞ ¼ a0ðR − RcÞ is already exactly linear; i.e.,the linearization of the energy difference performed aboveis not an approximation. The constants are given by a0¼Mω2

vibΔR and Rc ¼ ΔR=2þ ðΔE − ωcÞ=a0. The maxi-mum value of jhþjP=Mj−ij, reached at R¼Rc, is givenby ΔRω2

vib=ð4h0Þ. Comparing this with the energy splittingat that point, EþðRcÞ − E−ðRcÞ ¼ 2h0, gives the conditionthat ΔRω2

vib=ð8h20Þ must be small compared to the momen-tum of the respective nuclear wave function (due to theadditional P operating on the nuclear wave function). Theoff-diagonal terms h−jP2=2Mjþi reach a maximum value(again relative to the detuning) of MΔR2ω4

vib=ð25ffiffiffi5

ph30Þ

at R ¼ Rc þ h0=ðMΔRω2vibÞ.

APPENDIX B: CORRELATION OF STRONGLYCOUPLED STATES

As discussed in Sec. III C, the BOA predicts correlationin the nuclear motion under local excitation of a singlemolecule, even in the dark-state PES with only smallcontribution of the photonic mode. In order to verify this,and to quantify the range of applicability of the BOA, herewe calculate the steady-state wave function of the twomolecules under external driving of a single molecule.From first-order perturbation theory, the driven steady stateis given by

jψdr1 ðωÞi ¼

1

H − ω0 − ω − iϵμ1jψ0i; ðB1Þ

FIG. 8. Non-Born-Oppenheimer correction terms couplingthe “lower polariton” and “upper polariton” PES for a singleanthracenelike model molecule for a coupling strength ofg ¼ 0.002 a:u:. Solid colored lines: results from a full numericalcalculation. Dashed black lines: results from the modelEq. (A4). Note that while all results are given in atomic units,the units of the P and P2 terms are not identical, and thus notdirectly comparable.

CAVITY-INDUCED MODIFICATIONS OF MOLECULAR … PHYS. REV. X 5, 041022 (2015)

041022-11

Page 12: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

which we solve using the full Hamiltonian. We again use anonzero ϵ to artificially represent losses in the system (for theresults below, we choose ϵ ¼ 2.5 meV, corresponding to aneffective decay rate of 5 meV). While the eigenstates of theHamiltonian split into quasidegenerate symmetric and anti-symmetric superpositions (which show large correlation) forany nonzero g, the nonzero value of ϵ leads to a smearing ofthe energy resolution, such that the degeneracy is effectivelylifted and the superposition of only a single molecule beingexcited is observed in the steady state for small enough g.We then calculate the joint nuclear probability distributionPðR1; R2Þ of the driven steady-state wave function jψdr

1 ðωÞiand evaluate its mutual information I ¼ ∬PðR1; R2Þlog2½PðR1; R2Þ=PðR1ÞPðR2Þ�dR1dR2 [55]. For the groundstate of two coupled harmonic oscillators, I can beanalytically calculated as

I0 ¼ log2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2 − β=α

p þ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiβ=αþ 2

p2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi4 − β2=α24

p : ðB2Þ

In order to compare with the predictions obtained from theratio β=α for the PES, we choose driving frequenciesω equalto the vibrational ground-state energies of each PES. Thedashed lines in Fig. 9 show that, as could be expected, at zerocoupling (g ¼ 0) there is no correlation under driving of asingle molecule. As g increases, the mutual informationquickly increases and actually becomes significantly largerthan the BO ground-state values for the DS and UP PES.In this region, there is a series of avoided crossings, and theresults are expected to depend strongly on the correctdescription of decay and dephasing, which we treat onlyphenomenologically. For larger g, where the BOA becomesvalid, the mutual information in the driven steady stateagrees very well with the mutual information as predicted

from the coupling β=α in the Taylor expansion of the PES.Interestingly, the agreement between the full calculation andthe BO result for the LP PES is very good even at relativelylow coupling strengths. This is a consequence of the fact thatthe LP ground state is well isolated in energy, while theDS and UP surfaces are not. We believe that this property isalso related to the experimentally observed fast nonradiativedecay of upper polariton states, which can take placeefficiently close to avoided crossings of the PES (wherethe BOA breaks down).

[1] R. J. Thompson, G. Rempe, and H. J. Kimble, Observationof Normal-Mode Splitting for an Atom in an Optical Cavity,Phys. Rev. Lett. 68, 1132 (1992).

[2] C. Weisbuch, M. Nishioka, A. Ishikawa, and Y. Arakawa,Observation of the Coupled Exciton-Photon Mode Splittingin a Semiconductor Quantum Microcavity, Phys. Rev. Lett.69, 3314 (1992).

[3] J. Kasprzak, M. Richard, S. Kundermann, A. Baas, P.Jeambrun, J. M. J. Keeling, F. M. Marchetti, M. H.Szymańska, R. André, J. L. Staehli, V. Savona, P. B.Littlewood, B. Deveaud, and Le Si Dang, Bose-EinsteinCondensation of Exciton Polaritons, Nature (London)443, 409 (2006).

[4] R. Balili, V. Hartwell, D. Snoke, L. Pfeiffer, and K. West,Bose-Einstein Condensation of Microcavity Polaritons in aTrap, Science 316, 1007 (2007).

[5] S. Kéna-Cohen and S. R. Forrest, Room-TemperaturePolariton Lasing in an Organic Single-Crystal Microcavity,Nat. Photonics 4, 371 (2010).

[6] J. D. Plumhof, T. Stöferle, L. Mai, U. Scherf, and R. F.Mahrt, Room-Temperature Bose-Einstein Condensation ofCavity Exciton-Polaritons in a Polymer, Nat. Mater. 13, 247(2014).

[7] K. S. Daskalakis, S. A. Maier, R. Murray, and S. Kéna-Cohen, Nonlinear Interactions in an Organic PolaritonCondensate, Nat. Mater. 13, 271 (2014).

[8] D. G. Lidzey, D. D. C. Bradley, M. S. Skolnick, T. Virgili,S. Walker, and D. M. Whittaker, Strong Exciton-PhotonCoupling in an Organic Semiconductor Microcavity, Nature(London) 395, 53 (1998).

[9] T. Schwartz, J. A. Hutchison, C. Genet, and T.W. Ebbesen,Reversible Switching of Ultrastrong Light-MoleculeCoupling, Phys. Rev. Lett. 106, 196405 (2011).

[10] S. Kéna-Cohen, S. A. Maier, and D. D. C. Bradley,Ultrastrongly Coupled Exciton-Polaritons in Metal-CladOrganic Semiconductor Microcavities, Adv. Opt. Mater. 1,827 (2013).

[11] P. Törmä and W. L. Barnes, Strong Coupling betweenSurface Plasmon Polaritons and Emitters: A Review,Rep. Prog. Phys. 78, 013901 (2015).

[12] S. Kéna-Cohen, M. Davanćo, and S. R. Forrest, StrongExciton-Photon Coupling in an Organic Single CrystalMicrocavity, Phys. Rev. Lett. 101, 116401 (2008).

[13] J. Bellessa, C. Bonnand, J. C. Plenet, and J. Mugnier, StrongCoupling between Surface Plasmons and Excitons in anOrganic Semiconductor, Phys. Rev. Lett. 93, 036404 (2004).

FIG. 9. Solid lines: mutual information in the nuclear prop-ability distribution of the vibrational ground states of the coupledPES as predicted by Eq. (B2) using the values of β=α shown inFig. 7. Dashed lines: mutual information in the nuclear proba-bility distribution of the steady state under driving of a singlemolecule, Eq. (B1).

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-12

Page 13: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

[14] J. Dintinger, S. Klein, F. Bustos, W. L. Barnes, and T.W.Ebbesen, Strong Coupling between Surface Plasmon-Polaritons and Organic Molecules in Subwavelength HoleArrays, Phys. Rev. B 71, 035424 (2005).

[15] T. K. Hakala, J. J. Toppari, A. Kuzyk, M. Pettersson, H.Tikkanen, H. Kunttu, and P. Törmä, Vacuum Rabi Splittingand Strong-Coupling Dynamics for Surface-Plasmon Po-laritons and Rhodamine 6G Molecules, Phys. Rev. Lett.103, 053602 (2009).

[16] P. Vasa, R. Pomraenke, G. Cirmi, E. De Re, W. Wang,S. Schwieger, D. Leipold, E. Runge, G. Cerullo, andC. Lienau, Ultrafast Manipulation of Strong Coupling inMetal-Molecular Aggregate Hybrid Nanostructures, ACSNano 4, 7559 (2010).

[17] S. R. K. Rodriguez, J. Feist, M. A. Verschuuren, F. J. GarcíaVidal, and J. Gómez Rivas, Thermalization and Cooling ofPlasmon-Exciton Polaritons: Towards Quantum Conden-sation, Phys. Rev. Lett. 111, 166802 (2013).

[18] A. I. Väkeväinen, R. J. Moerland, H. T. Rekola, A.-P.Eskelinen, J.-P. Martikainen, D.-H. Kim, and P. Törmä,Plasmonic Surface Lattice Resonances at the StrongCoupling Regime, Nano Lett. 14, 1721 (2014).

[19] A. L. Baudrion, A. Perron, A. Veltri, A. Bouhelier, P. M.Adam, and R. Bachelot, Reversible Strong Coupling inSilver Nanoparticle Arrays Using Photochromic Molecules,Nano Lett. 13, 282 (2013).

[20] G. Zengin, M. Wersäll, S. Nilsson, T. J. Antosiewicz,M. Käll, and T. Shegai, Realizing Strong Light-MatterInteractions between Single-Nanoparticle Plasmons andMolecular Excitons at Ambient Conditions, Phys. Rev. Lett.114, 157401 (2015).

[21] J. A. Hutchison, T. Schwartz, C. Genet, E. Devaux, andT.W. Ebbesen, Modifying Chemical Landscapes by Cou-pling to Vacuum Fields, Angew. Chem. 124, 1624 (2012).

[22] J. A. Hutchison, A. Liscio, T. Schwartz, A. Canaguier-Durand, C. Genet, V. Palermo, P. Samorì, and T.W.Ebbesen, Tuning the Work-Function via Strong Coupling,Adv. Mater. 25, 2481 (2013).

[23] S. Wang, A. Mika, J. A. Hutchison, C. Genet, A. Jouaiti,M.W. Hosseini, and T. W. Ebbesen, Phase Transitionof a Perovskite Strongly Coupled to the Vacuum Field,Nanoscale 6, 7243 (2014).

[24] I. Carusotto and C. Ciuti, Quantum Fluids of Light, Rev.Mod. Phys. 85, 299 (2013).

[25] P. Michetti, L. Mazza, and G. C. La Rocca, StronglyCoupled Organic Microcavities, in Organic Nanophoton-ics, Nano-Optics and Nanophotonics, edited by Yong ShengZhao (Springer, Berlin, 2015) pp. 39–68.

[26] A. González-Tudela, P. A. Huidobro, L. Martín-Moreno, C.Tejedor, and F. J. García-Vidal, Theory of Strong Couplingbetween Quantum Emitters and Propagating Surface Plas-mons, Phys. Rev. Lett. 110, 126801 (2013).

[27] L. Mazza, S. Kéna-Cohen, P. Michetti, and G. C. La Rocca,Microscopic Theory of Polariton Lasing via VibronicallyAssisted Scattering, Phys. Rev. B 88, 075321 (2013).

[28] J. A. Ćwik, S. Reja, P. B. Littlewood, and J. Keeling,Polariton Condensation with Saturable Molecules Dressedby Vibrational Modes, Europhys. Lett. 105, 47009 (2014).

[29] A. Canaguier-Durand, C. Genet, A. Lambrecht, T. W.Ebbesen, and S. Reynaud, Non-Markovian Polariton

Dynamics in Organic Strong Coupling, Eur. Phys. J. D69, 24 (2015).

[30] F. C. Spano, Optical Microcavities Enhance the ExcitonCoherence Length and Eliminate Vibronic Coupling inJ-Aggregates, J. Chem. Phys. 142, 184707 (2015).

[31] M. Born and R. Oppenheimer, Zur Quantentheorie derMolekeln, Ann. Phys. (Berlin) 389, 457 (1927).

[32] J. C. Tully, Perspective on “Zur Quantentheorie derMolekeln,” Theor. Chim. Acta 103, 173 (2000).

[33] R. Houdré, R. P. Stanley, andM. Ilegems, Vacuum-Field RabiSplitting in the Presence of Inhomogeneous Broadening:Resolution of a Homogeneous Linewidth in an Inhomoge-neously Broadened System, Phys. Rev. A 53, 2711 (1996).

[34] V. M. Agranovich, Yu. N. Gartstein, and M. Litinskaya,Hybrid Resonant Organic-Inorganic Nanostructures forOptoelectronic Applications, Chem. Rev. 111, 5179(2011).

[35] S. De Liberato, C. Ciuti, and I. Carusotto,Quantum VacuumRadiation Spectra from a Semiconductor Microcavity with aTime-Modulated Vacuum Rabi Frequency, Phys. Rev. Lett.98, 103602 (2007).

[36] J. George, S. Wang, T. Chervy, A. Canaguier-Durand, G.Schaeffer, J.-M. Lehn, J. A. Hutchison, C. Genet, and T.W.Ebbesen, Ultra-Strong Coupling of Molecular Materials:Spectroscopy and Dynamics, Faraday Discuss. 178, 281(2015).

[37] V. May and O. Kühn, Charge and Energy TransferDynamics in Molecular Systems (Wiley-VCH Verlag GmbHand Co. KGaA, Weinheim, 2011).

[38] M.-K. Kim, H. Sim, S. J. Yoon, S.-H. Gong, C. W. Ahn,Y.-H. Cho, and Y.-H. Lee, Squeezing Photons into a Point-Like Space, Nano Lett. 15, 4102 (2015).

[39] J. Moll, S. Daehne, J. R. Durrant, and D. A. Wiersma,Optical Dynamics of Excitons in J Aggregates of a Carbo-cyanine Dye, J. Chem. Phys. 102, 6362 (1995).

[40] C. Ciuti and I. Carusotto, Input-Output Theory of Cavitiesin the Ultrastrong Coupling Regime: The Case of Time-Independent Cavity Parameters, Phys. Rev. A 74, 033811(2006).

[41] T. N. Rescigno and V. McKoy, Rigorous Method forComputing Photoabsorption Cross Sections from a Basis-Set Expansion, Phys. Rev. A 12, 522 (1975).

[42] K. D. Bonin and V. V. Kresin, Electric-Dipole Polarizabil-ities Of Atoms, Molecules, And Clusters (World Scientific,Singapore, 1997).

[43] T. Shimanouchi, National Bureau of Standards ReportNo. NSRDS-NBS 39, 1972 (http://www.nist.gov/data/nsrds/NSRDS‑NBS‑39.pdf).

[44] R. Dicke, Coherence in Spontaneous Radiation Processes,Phys. Rev. 93, 99 (1954).

[45] E. T. Jaynes and F.W. Cummings, Comparison of Quantumand Semiclassical Radiation Theories with Application tothe Beam Maser, Proc. IEEE 51, 89 (1963).

[46] M. Tavis and F. W. Cummings, The Exact Solution of N TwoLevel Systems Interacting with a Single Mode, QuantizedRadiation Field, Phys. Lett. A 25, 714 (1967).

[47] C. F. J. König, J. A. van Bokhoven, T. J. Schildhauer, andM. Nachtegaal, Quantitative Analysis of Modulated Exci-tation X-Ray Absorption Spectra: Enhanced Precision ofEXAFS Fitting, J. Phys. Chem. C 116, 19857 (2012).

CAVITY-INDUCED MODIFICATIONS OF MOLECULAR … PHYS. REV. X 5, 041022 (2015)

041022-13

Page 14: Cavity-Induced Modifications of Molecular Structure in the …webs.ftmc.uam.es/plasmonanoquanta.group/pdfs/118.pdf · 2017. 11. 16. · Cavity-Induced Modifications of Molecular Structure

[48] M. Kozina, T. Hu, J. S. Wittenberg, E. Szilagyi, M. Trigo,T. A. Miller, C. Uher, A. Damodaran, L. Martin, A. Mehta,J. Corbett, J. Safranek, D. A. Reis, and A. M. Lindenberg,Measurement of Transient Atomic Displacements in ThinFilms with Picosecond and Femtometer Resolution, Struct.Dyn. 1, 034301 (2014).

[49] F. Nagasawa, M. Takase, and K. Murakoshi, RamanEnhancement via Polariton States Produced by StrongCoupling between a Localized Surface Plasmon and DyeExcitons at Metal Nanogaps, J. Phys. Chem. Lett. 5, 14(2014).

[50] I. V. Tokatly, Time-Dependent Density Functional Theoryfor Many-Electron systems Interacting with Cavity Photons,Phys. Rev. Lett. 110, 233001 (2013).

[51] M. Ruggenthaler, J. Flick, C. Pellegrini, H. Appel,I. V. Tokatly, and A. Rubio, Quantum-Electrodynamical

Density-Functional Theory: Bridging Quantum Optics andElectronic-Structure Theory, Phys. Rev. A 90, 012508(2014).

[52] J. A. Ćwik, P. Kirton, S. De Liberato, and J. Keeling,Self-Consistent Molecular Adaptation Induced by StrongCoupling, arXiv:1506.08974.

[53] A. J. Leggett, S. Chakravarty, A. T. Dorsey, Matthew P. A.Fisher, Anupam Garg, and W. Zwerger, Dynamics of theDissipative Two-State System, Rev. Mod. Phys. 59, 1(1987).

[54] R. D. Coalson, D. G. Evans, and A. Nitzan, A Nonequili-brium Golden Rule Formula for Electronic State Popula-tions in Nonadiabatically Coupled Systems, J. Chem. Phys.101, 436 (1994).

[55] R. M. Gray, Entropy and Information Theory, 1st ed.(Springer, New York, 1990).

GALEGO, GARCIA-VIDAL, AND FEIST PHYS. REV. X 5, 041022 (2015)

041022-14