Skip to main content
Philosophical transactions. Series A, Mathematical, physical, and engineering sciences logoLink to Philosophical transactions. Series A, Mathematical, physical, and engineering sciences
. 2024 Dec 24;382(2287):20230343. doi: 10.1098/rsta.2023.0343

Multiple interacting photonic modes in strongly coupled organic microcavities

Felipe Herrera 1, William L Barnes 2,
PMCID: PMC11667587  PMID: 39717976

Abstract

Room-temperature cavity quantum electrodynamics with molecular materials in optical cavities offers exciting prospects for controlling electronic, nuclear and photonic degrees of freedom for applications in physics, chemistry and materials science. However, achieving strong coupling with molecular ensembles typically requires high molecular densities and substantial electromagnetic-field confinement. These conditions usually involve a significant degree of molecular disorder and a highly structured photonic density of states. It remains unclear to what extent these additional complexities modify the usual physical picture of strong coupling developed for atoms and inorganic semiconductors. Using a microscopic quantum description of molecular ensembles in realistic multimode optical resonators, we show that the emergence of vacuum Rabi splitting in linear spectroscopy is a necessary but not sufficient metric of coherent admixing between light and matter. In low-finesse multi-mode situations, we find that molecular dipoles can be partially hybridized with photonic dissipation channels associated with off-resonant cavity modes. These vacuum-induced dissipative processes ultimately limit the extent of light-matter coherence that the system can sustain.

This article is part of the theme issue ‘The quantum theory of light’.

Keywords: strong coupling, molecules

1. Introduction

Strong coupling between a large ensemble of molecules and an optical cavity mode is still a rapidly evolving field, despite being more than 25 years old [1], indeed, arguably much older [2]. In strong coupling, hybridization takes place between a molecular resonance and a cavity mode to yield two new polariton modes (states), modes that inherit characteristics of both light and matter. Two broad schemes have attracted the most attention: exciton-polaritons, where an excitonic molecular resonance is coupled to a cavity mode, and vibrational-polaritons, where a molecular vibrational mode is coupled to a cavity mode. Despite considerable progress, an underlying theoretical framework has yet to be established that provides a coherent picture of strong coupling phenomena. Here, we identify and explore one largely ignored ingredient, the multiple photonic modes nature of the cavities typically employed.

One of the main attractions of molecular strong coupling is that the key phenomenon, that of an anti-crossing between a molecular resonance and a photonic mode, can be explained with a very simple model based on two coupled oscillators, one oscillator representing the molecular system (a large number of identical molecules are taken to behave as though they are a single oscillator) the other representing a single photonic (cavity) resonance. This simple picture is a powerful one but can do little to capture a wealth of important features, including dark states, disorder and especially material behaviour such as reactivity. It is for this reason that so much effort has been devoted to developing a wider theoretical framework. Significant progress has been made by building more realistic models of the molecular systems involved, as reviewed recently [35].

Much of the theoretical work on strong coupling in the past few years has been devoted to incorporating the complexities that arise when including more realistic numbers of molecules (typically models have less than 103, whilst in experiments there may be more than 108) [6], and the presence of disorder [7]. Various approaches have been explored, examples include: employing a Holstein–Tavis–Cummings model [811] together with a Markovian [12,13] approach for the dissipative dynamics of organic polaritons; ab-initio studies [14,15]; and multiscale molecular dynamics simulations [16]. Whilst these theoretical approaches strive to include more realistic models for the molecular ensembles involved, little attention appears to have been directed towards including photonic complexities.

It was recognized early on that the dispersion of the photon (cavity) modes was important [17], and recent studies have focused on dispersion in connection with energy transport [18]. Whilst dispersion of a given photon mode is indeed important in several processes of interest such as condensation [19] and lasing [20], the fact that most cavities that are currently employed in strong coupling experiments involving molecules support several discrete cavity modes [21], has, with some exceptions, see for example [2224], been much less explored. A number of studies have looked at this problem in the context of ultrastrong coupling [2527] where the Rabi splitting is comparable or greater than the free-spectral range (FSR). However, our focus here is on strong rather than ultrastrong coupling, and more specifically in the parameter space applicable to recent experiments [28].

An unwritten assumption in the context of polaritons involving organic materials seems to have been that if the spacing between cavity modes (the FSR) is ‘sufficient’, then the presence of many (rather than one) photonic modes can be ignored as having minimal influence on the polariton properties. Indeed, this presumed minimal effect has led to the use of ‘off-resonance’ modes being employed to monitor changes in the constituents within a cavity [29]. Here, we show explicitly that the presence of multiple photonic modes can have a significant effect on the strong coupling process. Recently, the importance of this effect has been recognized in other physical implementations of cavity QED, e.g. artificial atoms in superconducting resonators [30]. The model elucidated here is an expanded version of an outline we recently presented to explain polariton-mediated photoluminescence in low-finesse cavities [28]. Before looking at our multi-mode framework in detail, let us briefly discuss some of the prior work on multimode cavities.

Multi-mode cavities have been employed in a large number of experiments, both for excitonic strong coupling (see for example [23,31,32]) and vibrational strong coupling (see for example [21,3335]), but the implications of there being more than one mode have in general only involved considering the presence of single couplings, i.e. coupling between the molecular mode and each of the photonic modes. For example, when more than one photonic mode is involved then one has to be careful about mode assignment owing to the overlap (in energy) between different polariton bands [21]. However, as we show below, one also needs to consider how the couplings between different photonic modes alter the overall picture. One of our major findings is that such couplings can limit the extent of light-matter mixing and may thus limit the coherence of polaritons, with possible consequences for a variety of phenomena such as photoluminescence [28] and polariton transport [36].

Figure 1 provides a schematic overview of our model. In molecular strong coupling, an excitonic resonance is usually considered to interact with a single cavity mode, depicted in panel (a). Although other cavity modes might be present, they do not spectrally overlap with the cavity mode being considered, they are too far detuned. However, when the finesse is low, adjacent cavity modes may overlap and may thus couple to each other and to the excitonic resonance, see panel (b). In this situation the strong coupling is no longer single mode (SM), the exciton resonance now being ‘spread’ over more than one cavity mode. The work reported here is the result of an investigation to explore the consequences of this ‘spreading’.

Figure 1.

Schematic: Left, panel

Schematic: Left, panel (a). Although several cavity modes are present (C1, C0 and C+1), our excitonic molecular resonance X interacts with a single discrete on-resonance cavity mode, C0. Right, panel (b). Now the finesse is low and there is spectral overlap between adjacent photonic (cavity) modes. The excitonic resonance interacts directly with the various cavity modes, and there is also direct interaction between adjacent photonic modes. The consequence is that the excitonic resonances is now spread out over more than one cavity mode.

2. Multi-mode theory of organic microcavities

Our aim is to build a model for molecular optical cavities that corresponds to an ensemble of N electronic dipole emitters coupled to the full set of resonant optical modes supported by a (planar) cavity structure. In what follows we restrict the discussion to molecular dipoles with a negligible Huang–Rhys factor [37], such that their emission properties are accurately described by considering only the vibration-less ground (S0g) and first excited (S1e) electronic states (see [38,39] for examples). The system is described by a multimode Tavis–Cummings Hamiltonian of the form (1 throughout)

H^=qωqa^qa^q+i=1Nωiσ^i+σ^i+qi=1Ngiqσ^i+a^q+giqσ^ia^q (2.1)

where q={𝐪,q} is generally a composite photonic mode index describing the continuous in-plane component of the wavevector 𝐪 and the discrete transverse component q of confined electromagnetic modes in planar cavities [40]. The local dipole transition operators between ground |gi and excited state |ei are defined as σ^i=|giei| for emission and σ^i+=(σ^i) for absorption. The electronic transition frequency ωi is in general inhomogeneously distributed, although there are examples of organic emitters with negligible inhomogeneous broadening [39]. Bosonic cavity field operators are a^q and the local mode-dependent Rabi couplings are denoted by giq.

Ideal planar cavities of length L have photon dispersion ωq=(c/nd)(q2+q2(m))1/2, where q(m)=mπ/2L, with m1 an integer, which determines the discrete set of allowed cavity mode energies at normal incidence (q=0); c is the speed of light and nd the (real) dielectric constant of the intracavity medium. The mode dispersion with respect to the in-plane wavevector 𝐪 determines the propagation properties of the normal modes of the coupled system, which for strong light-matter coupling correspond to exciton-polaritons [4143], and the FSR between adjacent modes Δqωq+1ωq is controlled at normal incidence (q=0) by the cavity length as Δ=cπ/2ndL. Throughout this work, we neglect dispersion and only study system properties at q=0.

The explicit multimode structure of equation (2.1) generalizes early approaches that simplify the mode structure of the microcavity to having a single dispersionless mode a^ [44], or a dispersive mode a^q in a cavity with infinite FSR. Such simplifications are often introduced as a necessity in favour of capturing relevant aspects of complexity of the internal degrees of freedom of the molecular dipole emitters, such as high-frequency vibrations [812,45], electron tunnelling [46] or the role of static disorder in establishing quantum transport regimes [18,47,48]. Single-mode or single-branch theories are intrinsically limited with respect to their ability to describe the influence of multiple transverse photonic modes in realistic organic microcavities with finite FSR, Δ, as the value of Δ is not much larger than the frequency separation between lower and upper polaritons, i.e. the Rabi splitting ΩR. For some systems Δ is even smaller than ΩR [49].

3. Microcavities as open quantum systems

We model the organic microcavity microscopically as an open quantum system described by a Lindblad quantum master equation for the light-matter reduced density matrix ρ^, given by [50]

ddtρ^=i[H^,ρ^]+qκq2(2a^qρ^a^qa^qa^qρ^ρ^a^qa^q)+iγi2(2σ^iρ^σ^i+σ^i+σ^iρ^ρ^σ^i+σ^i), (3.1)

where κq is the bare radiative decay rate of the q-th cavity mode and γi is the bare spontaneous decay rate of the i-th electronic dipole excitation, which includes radiative and non-radiative contributions. For simplicity, we ignore cross-terms that could dissipatively couple different cavity modes or different molecules, under the assumption that direct diagonal relaxation channels are much faster.

The Lindblad master equation is the basis for deriving effective non-unitary propagators that unravel the state evolution as an ensemble of wavefunction trajectories [51]. For the light-matter state ansatz,

|Ψ(t)1ξ(t)|ψ(0)(t)+ξ(t)|ψ(1)(t), (3.2)

with ξ1, as is relevant for weakly excited microcavities, we can ignore stochastic quantum jumps coming from the recycling terms of the Lindblad equation [12] and account for dissipation as an exponential decay of the excited state population. This is equivalent to rewriting equation (3.1) as

ddtρ^=i[H^eff,ρ^]+L1[ρ^], (3.3)

with an effective non-Hermitian Hamiltonian

H^eff=H^i2qκqa^qa^qi2iγiσ^i+σ^i. (3.4)

As mentioned above, we ignore the recycling terms L1[ρ^]qκqa^qρ^a^q+iγiσ^iρ^σ^i+. This simplified approach to the open system dynamics effectively reduces the problem to one of solving a non-Hermitian Schrodinger equation with the effective Hamiltonian H^eff. The weak excitation assumption that justifies the neglect of recycling terms in equation (3.1) holds in strong coupling as long as the ground state is predominantly populated and has the separable form |G>|0>. For non-separable ground states that can arise in ultrastrong coupling, weak excitation conditions can still be defined, but the quantum master equation needs to be reformulated [52].

Since the ground state |ψ(0) has no electronic or photonic excitations, the dynamics of polaritons is fully determined by the excited electron–photon wavefunction

|ψ(1)=ici(0)|ei|{0q}+qcq(1)|g1g2,,gN|1q, (3.5)

where |{0q} is the multi-mode cavity vacuum, |ei describes a single excitation in the i-th molecule with all other dipoles in the ground state; |1q describes a single photon in the q-th transverse mode, all other modes being empty.

4. Strong coupling in high-finesse cavities

Consider N molecular dipoles coupled near resonantly with a q=0 mode of frequency ω0 and decay rate κ0. Higher- and lower-order cavity modes are detuned from the central frequency by Δq=ωqω0, with Δq>0 for higher-order and Δq<0 for lower-order modes. They also have bandwidths that in general differ from q=0 by Δκq=κqκ0. In the high-finesse limit, |Δq|Ω0, with Ω0=Ng0 being the single-mode Rabi coupling strength for a homogeneous molecular ensemble, dipole excitations cannot exchange energy effectively with higher- and lower-order cavity modes (note, the SM Rabi splitting, ΩRSM, is related to the SM Rabi coupling, Ω0, through ΩRSM = 2Ω0). Consequently, light-matter hybridization leading to polariton formation only occurs in the vicinity of the near-resonant q=0 mode, see panel (a) of figure 1. However, far-detuned modes do have an effect via second-order (two photons) processes, something we look at next.

Far-detuned higher- and lower-order modes evolve on a timescale of 1/Δq, which is much faster than the Rabi oscillation period (τR1/Ω0) between the near-resonant mode and molecular excitations. These fast-oscillating mode variables thus adiabatically adjust to the dynamics of the near-resonant manifold, which affects the process of polariton formation around q=0. This can be understood as the emergence of processes similar to a Lamb shift in which molecules absorb and re-emit virtual photons from higher- and lower-order modes. These two-photon processes result in a change to the energetics; a single-molecule frequency shift of the form

Γj =q0|gjq|2ΔqΔq2+(Δκq/2)2, (4.1)

and an effective inter-molecular coupling with interaction energy given by

Jij=q0giq*gjqΔq/2Δq2+(Δκq/2)2. (4.2)

The sign of the contributions per mode in these expressions is different for lower-order modes (blue shift and repulsive interaction) and higher-order modes (red shift and attractive interactions). In general, Jij is a complex-valued quantity depending on the relative phase of the Rabi frequency at the location of the two dipoles.

Cavity-induced frequency shifts and inter-dipole interactions induced by far-detuned modes are well known from atomic physics [53] and have been used for quantum state preparation in high-quality resonators [54]. Organic microcavities are qualitatively different from atomic cavities in that their quality factors are much lower, typically Q1–10, and changes in bandwidth κq with mode order can be large. This is particularly true for modes close to the region where absorption of metal mirrors cannot be neglected [55]. Therefore, in addition to the changes in frequency and interaction energy discussed above, the dispersive interaction of molecular dipoles with far-detuned lossy modes also changes the single-molecule dipole decay rates by

Γj=q0|gjq|2Δκq/2Δq2+(Δκq/2)2, (4.3)

and establishes the two-body loss rate

Jij=q0giq*gjqΔκq/2Δq2+(Δκq/2)2. (4.4)

These are again signed quantities summed over all available modes, the contribution of which depends on the relative bandwidths Δκq. If all relevant modes have bandwidths κq equal to the resonant (q=0) mode, then Δκq0 and no second-order corrections to the decay rates are expected. The bandwidth mismatch of sub-wavelength cavities thus introduces a phenomenology that is not present in other cavity QED systems, we discuss one key aspect next.

The dispersive relaxation channels discussed above involving material degrees of freedom could, in principle, alter the ability to establish strong coupling with the central q=0 mode. To assess this, consider a simplified homogeneous scenario in which all dipoles are identical (Dicke regime) and the polariton wavefunction in equation (3.5) reduces to

|ψ(1)=β|X|0+α|g1g2,,gN|1, (4.5)

where |X=j|ej/N is the fully symmetric excitonic state and the Fock states {|0,|1} refer to the central q=0 mode. The dynamics of the state vector 𝐱=[α,β]T can be written as 𝐱˙=iM𝐱, where

M=(0Ω0Ω0δNiΔΓN) (4.6)

is the dynamical matrix for which the complex eigenvalues λ=E+iΓ/2 give the polariton energies, E, and bandwidths, Γ. As before, Ω0=Ng0 is the Rabi coupling strength. The effective detuning δN and bandwidth mismatch ΔΓN can be written as

δN=δ0ΓNJ, (4.7)
ΔΓN=Δγ/2+Γ+NJ, (4.8)

where δ0=ω0ωe and Δγ=κ0γ are the bare detuning and bandwidth mismatch between the q=0 mode and the dipole resonance, respectively. The one-body and two-body energy shifts Γ and NJ contribute to the detuning of the q=0 mode from the molecular resonance; the one- and two-body decay rates Γ and NJ contribute to the bandwidth mismatch. The microscopic derivation of equation (4.6) starting from equation (3.4) is given in the appendix.

The real part of the eigenvalues of M give the lower and upper polariton frequencies, ELP and EUP, respectively. The Rabi splitting ΩREUPELP can thus be written as

ΩR=Re(NJδ0i(NJΔγ/2))2+4Ω02, (4.9)

where Γ and Γ are neglected in the thermodynamic limit1. Equation (4.9) gives the usual strong coupling result ΩR=2Ω0 for infinite finesse, Δq and finite N, since NJ1/Δq2 and NJ1/Δq.

To estimate the magnitude of NJ for typical high-finesse cavities (|Δq|>ΩR), consider a model three-mode cavity with a central q=0 mode at ω0, a lower-order mode (q=1) detuned from q=0 by Δ1=Δ and a higher-order mode (q=+1) detuned by Δ+1=Δ, with Δ>0. The mode-dependent decay rates are {κ,κ0,κ+}, respectively. We assume linear scaling of the bandwidth with mode order, i.e. κq=κ0+qζ with positive ζ for increasing bandwidth and negative otherwise. We also allow the Rabi coupling strength to depend on mode order as Ω±1Ng±1, for collective coupling of dipoles to q=±1 cavity modes. From equation (4.4), the two-body rate can be written as

NJ=(Ω12Ω+12)ζ/2Δ2+ζ2/4ζfΩ02Δ2+ζ2/4, (4.10)

where in the second equality we use Ω±1=Ω0(1±f) with |f|<1. This contribution to the polariton decay can either increase or decrease the bandwidth of LP and UP resonances around the q=0 mode, depending on the sign of ζf.

Figure 2 shows the magnitude of NJ from equation (4.10) as a function of the variation in Rabi coupling per mode δΩ=fΩ0 and the variation in bandwidth δκ=ζ, estimated for a system with Ω0=0.35 eV and FSR Δ=1.0 eV. We consider a two-mode cavity (panel a), where q=0 is the lowest-order mode (Ω1=0) and modes beyond q=1 are ignored, as well as the three-mode scenario (panel b). In general, the magnitude of NJ is smaller than kBT (approx. 26 meV) for multi-mode microcavities with relatively weak mode-dependence of the Rabi coupling and photon bandwidth (δΩδκ102 eV [28]), but the analysis above is general and larger polariton bandwidth modifications could be expected for other high-finesse photonic structures with greater cavity bandwidths and coupling variations with mode order.

Figure 2.

Modified polariton decay in high-finesse cavities

Modified polariton decay in high-finesse cavities: Two-body contribution to the polariton decay rate NJ, as a function of the change in Rabi coupling δΩ and change in bandwidth δκ relative to a reference q=0 strongly coupled resonant mode for two cases; (a) two-mode cavity where q=0 is the lowest-order mode; (b) three-mode cavity with one lower and one higher-order mode. We use Ω0=0.35 eV and Δ=1.0 eV.

In summary, the presence of far-off-resonance cavity modes can introduce adiabatic corrections to the Rabi splitting ΩR established in strong coupling. Such corrections originate from coherent and incoherent two-photon processes in which dipoles scatter virtual photons from far-detuned higher- and lower-order modes, primarily leading to changes in the dipole bandwidth (figure 2). Since adiabatic corrections to ΩR scale as (Ω0/Δ)2, it can be difficult to measure their contribution in typical high-finesse Fabry–Perot microcavities (FΔ/κ010, Ω0/Δ0.1).

5. Strong coupling in low-finesse cavities

The adiabatic elimination procedure described in §5 is strictly valid for ΩR/Δ1 and breaks down if the Rabi splitting is no longer much smaller than the FSR, even when the bare cavity finesse is nominally high (F10).

For microcavities with lower finesse, there is no significant separation of scales between κ, γ, Ω0 and Δ, although typically κγ<Ω0<Δ in strong coupling [28]. For the light-matter system discussed above, with N identical dipoles at ωe resonant with a reference q=0 cavity mode at ω0, having Rabi coupling strength Ω0, the frequency separations Δ±1 of adjacent higher-order and lower-order modes (q±1) are comparable with the bare Rabi couplings Ω±1 to those modes and thus the direct coupling of dipoles to neighbouring modes needs to be considered.

For the homogeneous dipole ensemble, the simplest extension of the light-matter state vector includes the next-order modes q=±1, i.e. 𝐱=[α1,α0,α+1,β]T, where αq is the photon amplitude in the q-th mode. The dynamical matrix for this (3+1) system can be written as

M=(d100e1000e000d+1e+1e1e0e+1p), (5.1)

where we use the simplified notation dq=ΔqiΔκq/2, eq=Ωq and p=δ+iΔγ/2. From the eigenvalues of M, λ=E+iΓ/2, the coupled energies E and decay rates Γ are obtained. In general, λ is a root of the polynomial

Φ(λ)pλ+e02λe12d1λe+12d+1λ, (5.2)

which for Ω±1=0 gives the quadratic eigenvalue equation

λ(δiΔγ/2λ)Ω02=0, (5.3)

which is often used to derive conditions for strong coupling in the single-mode picture. In particular, for resonant bandwidth-matched light-matter interaction with the q=0 mode (δ=0, γ=κ0), equation (5.3) gives the LP and UP frequencies E±=±Ω0 (ΩR=2Ω0) with Γ±=κ for the decay rates.2

Direct coupling of dipoles to the q=±1 modes modifies the LP and UP energies and bandwidths. In the appendix, we derive a general expression for the lowest-order corrections to the energies and bandwidths of the single-mode polariton problem, owing to the presence of neighbouring cavity modes. These corrections scale nonlinearly with the Rabi couplings Ω±1 and mode detunings Δ±1, and have a strong dependence on the change in bandwidth Δκq between different cavity modes. For a resonant, bandwidth-matched interaction with the q=0 mode, and ignoring possible changes of the Rabi coupling for different cavity modes (Ωq=Ω), the modified Rabi splitting and polariton bandwidths around the q=0 mode can be approximated as

ΩR=2Ω[1Ω2(Ω2+Δ2)(Ω2+Δ2)2+Δ2ζ2], (5.4)

and

Γ±=κ0±2ΩΩ2Δζ(Ω2+Δ2)2+Δ2ζ2, (5.5)

where again Δ±1=±Δ and κq=κ0+qζ are assumed. These expressions reduce to the single-mode case, when ΩΔ and |ζ|Δ, which is the high-finesse regime discussed in the previous section. Although the correction to the splitting also scales as (Ω/Δ)2 for ΔΩ as in the high-finesse problem, now the presence of nearby modes in finite-finesse cavities directly modifies the polariton energies E± around the dipole resonance via level repulsion. In contrast, adiabatic corrections introduce an overall dipole shift via two-photon processes which can in principle be compensated for by tuning the cavity frequency.

Similar to the adiabatic corrections in equation (4.10), the changes of the polariton decay rates Γ± predicted for low-finesse cavities also scale linearly with the difference in bandwidth Δκq of nearby modes relative to the near-resonant q=0 mode. However, whilst adiabatic bandwidth corrections vanish for systems with mode-independent Rabi couplings, equation (5.5) suggests that the mode-order dependence of the bare bandwidths is more important than variations in the field profile (Rabi coupling) of different cavity modes to establish the bandwidths of the LP and UP resonances.

Figure 3a shows the Rabi splitting ΩR (in eV) predicted by equation (5.4) as function of the bare Rabi coupling Ω0 with the central mode (Ωq=Ω) and the FSR Δ, for a system with decreasing mode bandwidth with increasing mode order (ζ=0.1 eV). Even for relatively large values of Δ (approx. 1 eV), the direct interaction of dipoles with q=±1 modes significantly reduces the multi-mode Rabi splitting from the usual single-mode value (ΩRSM=2Ω0). In practical terms, the requirements for establishing a light-matter interaction strength that gives a desired polariton splitting become more demanding as the finesse decreases.

Figure 3.

Modified Rabi splitting and bandwidth in low-finesse cavities

Modified Rabi splitting and bandwidth in low-finesse cavities: (a) Rabi splitting ΩR as function of the bare Rabi coupling Ω0 and FSR Δ for a three-mode cavity (q={0,±1}) with the central mode on exact resonance with the dipole transition; (b) Decay rates Γ of the Lower Polariton (LP, blue lines) and Upper Polariton (UP, green lines) in units of κ, as functions of Δ for different values of the bare Rabi coupling Ω0, for a linear variation of the mode bandwidths κq=κ0+qζ. We use δ=0, γ=κ=0.1eV and ζ=0.1 eV. We use Ω0=Ω±1=0.35 eV, κ=γ=0.15eV, ζ=0.1eV.

Figure 3b shows the complementary effect on the LP and UP bandwidths Γ± given by equation (5.5), as functions of Δ. The bare mode bandwidth is κ=0.15 eV and a small linear decrease of the bandwidth with mode order is assumed (ζ=50 meV). The polariton level (UP) closer in frequency to the narrower mode (q=+1), becomes narrower as Δ decreases, and the level (LP) closer to the broader mode (q=1) broadens. Even for relatively large mode separations (Δ=1.2 eV, F8), the LP and UP bandwidths are asymmetric and can differ significantly from the single-mode prediction Γ±SM=(κ+γ)/2. The deviation from Γ±SM grows with increasing coupling strength Ω0.

In addition to the changes in Rabi splitting and polariton bandwidths introduced by the direct interaction of dipoles with neighbouring modes, the microscopic composition of the LP and UP polariton wavefunctions can be very different in low-finesse cavities relative to a single-mode picture. Understanding this can improve the ability to control the emission properties of organic microcavities [11,12].

In an ideal single-mode cavity under strong coupling conditions ΩR>(κ0+γ)/2, the LP and UP states can be accurately described by equation (4.5) with |β|=|α|=1/2, i.e. equal exciton and photon content. For the three-mode problem centered around q=0, molecular dipoles can also admix with the lower- and higher-order modes q=±1. Therefore, a more realistic description of LP and UP states would be

|ψ=β|X|00+|g1g2,,gN(α1|11+α0|10+α+1|1+1), (5.6)

where |nq denotes a Fock state of the q-th cavity mode; |β|2 is the exciton fraction and |αq|2 the fraction of the q-th cavity mode in the polariton state. Equation (5.6) highlights the fact that dipoles exchange energy and coherence with a single-photon wavepacket,3 not an individual Fock state. Since |ψ is normalized, the photon fraction associated with the near-resonant mode |α0|2=1|β|2|α1|2|α+1|2 is always smaller than the single-mode limit (Δ,α±10).

Figure 4a shows the material (exciton) content of the LP state as a function of the bare coupling Ω0, obtained from the eigenvectors of the three-mode matrix M in equation (5.1), parametrized with realistic frequencies and bandwidths from [55] (ωe=2.15 eV, γ=0.37 eV, ω1=1.45 eV, ω0=2.14 eV, ω+1=2.76 eV, κ1=38 meV, κ0=90 meV, κ+1=90 meV); Ω0=Ω±1 is a free parameter. The q=1 mode is significantly narrower than q=0, but no significant variation is seen for q=1. The standard single-mode picture of strong coupling suggests that for Rabi splittings ΩR>(κ0+γ)/20.24 eV (Ω0>0.12 eV) the exciton fraction should be |β|20.5. In contrast, the exciton fraction of the LP state in figure 4a decreases with increasing Rabi coupling. We also show results for a hypothetical scenario where the lower- and higher-order mode frequencies are varied as ω±1=ω±1ε (reducing Δ), with all other parameters kept constant.4 We find that even for moderate values of cavity finesse Δ/κ 4–5, the exciton fraction of the LP can be lower than 30% even when the Rabi coupling Ω0 exceeds the bare bandwidths κ0 and γ (Ω0>0.3 eV, ΩR>0.5 eV).

Figure 4.

Modified polariton wavefunctions in low-finesse cavities

Modified polariton wavefunctions in low-finesse cavities: (a) Exciton fraction of the LP state as a function of the bare Rabi coupling for a three-mode cavity with realistic parameters (Δ0.65 eV [55]). The SM limit for q=0 resonant with the dipoles is marked. Results for other values of Δ are also shown, keeping other parameters the same. (b) Photon fraction per q-mode of the LP state as a function bare Rabi coupling for different values of Δ, using the same parameters in (a).

Figure 4b shows a complementary view of the photon content per mode of the LP state, for the same system parameters shown in figure 4a. As the bare Rabi coupling increases, the contribution of the q=0 mode decreases below the single-mode limit and the q=1 contribution increases significantly. The changes in the q=0 and q=1 components with Ω0 are stronger with decreasing inter-mode separation Δ. The higher-order q=+1 state component is less sensitive to Ω0 and Δ, because it is further detuned from the LP. Taken together the results shown in figure 4 indicate that in situations of low finesse a significant reduction in photoluminescence associated with the LP is to be expected, as recently found [28].

6. Conclusion and outlook

It has been understood for many years that taking account of some of the real-world complexities, especially disorder, is important in building up a full conceptual model of molecular strong coupling. In this contribution, we have shown that it is also important to take into account off-resonance photonic modes supported by a cavity if one is to properly understand molecular strong coupling. In particular, we have shown that in low-finesse situations the extent of light-matter mixing (hybridization) is altered. When adjacent photonic modes are spectrally overlapped then the molecular content (here we considered excitons) is spread over several photonic modes, resulting in a lower-matter content in any given polariton mode. This conclusion is supported by a similar finding in the context of circuit QED [30]. We have also shown that the extent of the Rabi-splitting can be curtailed by the effect of extra photonic modes.

Looking ahead it will be important to explore these issues further, especially in conjunction with experiment. We have already made a start in this direction [28], where we monitored luminescence from a range of planar samples, since luminescence probes the extent of light-matter mixing more directly than, for example, reflectivity. It would be useful to build on this start with a more systematic study. For example, one could envisage a series of experiments on Fabry–Perot planar cavities that employ metal mirrors, where the metal-mirror thickness is altered to control the cavity mode-width, and hence the finesse.

Regarding development of the model that we have outlined here, future extensions should include frequency disorder of dipole transitions, which in most cases is the dominant contribution to material absorption lineshape [57]. Disorder leads to the formation of semi-localized states in the spectral region where the uncoupled dipoles also absorb [58,59], adding complexity to the analysis of spectroscopic signals not present in the homogeneous dipole (Dicke) models we discussed here. We anticipate that disorder will introduce quantitative changes to the dependence with finesse of the polariton splitting and polariton bandwidths relative to the homogeneous model predictions here, but the qualitative physics should remain. We already find evidence of the reduction of the exciton content of exciton-polariton states owing to multimode light-matter interaction in a realistically disordered system [28]. A more general theoretical framework should be able to treat the simultaneous coupling of molecular dipoles to multiple discrete transverse modes and continuous in-plane momenta in a microcavity. Such a theory would be complex but may enable studies of controlled excitation transport along a polariton branch by possibly driving off-resonant coupled photonic branches with external fields.

Appendix

Adiabatic elimination of far-detuned cavity modes

In this section, we derive equation (4.6), presented in the main text.

The non-degenerate ground state has no excitations in the electronic and cavity modes and is given by |ψ(0)=|g1g2,,gN|0K,,01,00,0+1,,0K. We consider M=2K+1 discrete transverse modes. The single-excitation electron-photon wavefunction can be written as

|ψ(1)=ici(0)|ei|{0q}+qcq(1)|g1g2,,gN|1q, (A 1)

where |{0q} is the multi-mode cavity vacuum, |ei describes a single excitation in the i-th molecule with all the other dipoles in the ground state, and |1q describes a single photon in the q-th mode, with all the other modes in the vacuum state. There are no other restrictions on the material and photonic coefficients ci(0) and cq(1) except for normalization i|ci(0)|2+q|cq(1)|2=1.

We derive evolution equations for the material and photonic wavepacket amplitudes ci(0) and cq(1) from the single-excitation ansatz (A 1), directly from the non-Hermitian Schrödinger equation (d/dt)|ψ1=iH^eff|ψ(1), with H^eff given by equation (3.4). Given the global phase freedom of the wavefunction, we rewrite the state amplitudes as cj(0)=c~j(0)exp[iω~0t] and cq(1)=c~q(1)exp[iω~0t], where ω~0=ω0iκ0/2 is the complex frequency of the free q=0 mode. In this new q=0 frame, we obtain the following system of (M+N) coupled equations for the amplitudes:

ddtc~j(0)=(iδj+Δγj/2)c~j(0)iqgjqc~q(1), (A 2)
ddtc~q(1)=(iΔq+Δκq/2)c~q(1)ijgjq*c~j(0), (A 3)

where δj=ω0ωj is the detuning of the reference (q=0) mode relative to the i-th dipole frequency, ωq=ω0+Δq is the frequency of the q-th higher (Δq>0) or lower-order mode (Δq<0) relative to q=0; Δγj=κ0γj is the decay mismatch between dipoles and the reference mode, and Δκq=κqκ0 is the decay mismatch relative to q=0. Different equations of motion could be derived by adopting other rotating frame choices, but the resulting spectrum and state populations would not change. The (M+N) coupled equation model is the most general model. Frequently (M+1) coupled equations or (2M) models are used [22,23,32], but these are versions with a restricted range of validity [24], the appropriate choice depending on the detailed structure of the cavities considered. Indeed, interim choices may also be appropriate [24], highlighting the need for a more extensive investigation in the future.

We derive a reduced set of coupled equations of motion for the N excited state amplitudes c~j(0) and the single-photon amplitude of the reference mode c~0(1), under the assumption that only the reference mode q=0 interacts resonantly with the dipolar ensemble, but higher and lower-order modes (q0) are sufficiently detuned from the dipole frequencies ωj to prevent significant exchange of energy and coherence between light and matter. In this case, the off-resonant modes simply follow adiabatically the dipole polarization to the lowest order in gjq. The stationary off-resonant mode amplitude is obtained from equation (A 3) to give

c~q0(1)=ii=1Ngiq*c~i(0)iΔq+Δκq/2. (A 4)

Separating equations (A 2) and (A 3) into resonant (q=0) and non-resonant mode contributions, and inserting the adiabatic solution into equation (A 4) for the lower and higher-order modes, we obtain a reduced set of (1+N) equations of motion of the form

ddtc~j(0)=i(δjΓj )c~j(0)+(Δγj/2Γj)c~j(0)igj0c~0(1)ij(Jij+iJij )c~i(0), (A 5)
ddtc~0(1)=iigi0c~i(0), (A 6)

where we introduce the effective one-body and two-body decay rates,

Γj=q0|gjq|2Δκq/2Δq2+(Δκq/2)2, (A 7)
Jij=q0giq*gjqΔκq/2Δq2+(Δκq/2)2, (A 8)

and effective frequency shifts given by

Γj=q0|gjq|2ΔqΔq2+(Δκq/2)2, (A 9)
Jij=q0giq*gjqΔq/2Δq2+(Δκq/2)2. (A 10)

In these expressions, the sum-over-modes exclude the reference q=0 cavity resonance. The reduced equations of motion equations (A 5) and (A 6) can accurately describe strong coupling between the N oscillators and the near-resonant q=0 mode, and improves over previous treatments in the literature by including the effect of far-detuned higher- and lower-order modes self-consistently. The quasi-static solution for non-resonant modes in equation (A 4) is accurate for modes that do not significantly admix with dipole transitions, which requires |Δq|>j|gjq|2 for each q0.

In the idealized Dicke regime where the N dipoles have equal transition frequencies (ωi=ωe), Rabi couplings (giq=gq) and relaxation rates (γi=γ), the reduced dynamical equations (A 5) and (A 6) further simplify by writing c~i=β/N and c~0=α0. The evolution equation for the collective dipole coherence X~=ic~i(0)=Nβ, can be obtained by summing equation (A 5) over all dipoles, approximating N1N in the two-body terms, to obtain two coupled equations for α0 and β that can be written in matrix form as

(α˙0β˙)=(0iNg0iNg0iδNΔΓN)(α0β), (A 11)

with

δN=δ0ΓNJ, (A 12)
ΔΓN=Δγ/2+Γ+NJ. (A 13)

This is equation (4.6), as presented in the main text; δ0=ω0ωe, Δγ=κ0γ, and g0 refer to properties of the near-resonant q=0 mode.

Approximate Rabi splitting and polariton bandwidths in low-finesse three-mode cavities

In this section, we derive equations (5.4) and (5.5), presented in the main text.

We consider a three-mode cavity model with a central q=0 mode at ω0 tuned near resonance with a homogeneous ensemble of N dipole transitions at ωe. The dipoles also couple to the lower-order mode q=1 and a higher-order mode q=+1, with frequencies ω±1=ω0+Δ±1, respectively. The evolution equations (A 2) and (A 3) reduce in this case to a (3+1) system for the collective coherence X~=jc~j0=Nβ, (δj=δ0, Δγj=Δγ, gjq=gq) and photon amplitudes c~q(1)αq, with q={1,0,1}. The resulting system of equations can be written in matrix form as 𝐱˙=i𝐌𝐱, with 𝐱=[α+1,α0,α1,β]T and 𝐌 written in arrowhead form as

𝐌=(Δ1iΔκ1/200Ω1000Ω000Δ+1iΔκ+1/2Ω+1Ω1Ω0Ω+1δ+iΔγ/2)(d100e1000e000d+1e+1e1e0e+1p), (A 14)

where Ωq=Ngq (with gq=gq*), δ=ω0ωe, Δγ=κ0γ, Δκq=κqκ0 and Δq=ωqω0. We introduced a simplified notation in the second equality. The arrowhead interaction matrix is not the only possibility, other alternatives have also been explored, see [24] and references therein. For our purposes, the arrowhead approach provides a convenient starting point to explore the physics we wish to investigate. In the cases we study here the molecules often occupy a limited part of the cavity mode, so that orthogonality does not strictly hold, as a consequence there will be cross-coupling of cavity modes induced by excitons. Future work will need to explore other choices of the interaction matrix 𝐌 and examine the ensuing consequences.

The complex eigenvalues λ=E+iΓ/2 of 𝐌 are roots of the characteristic polynomial [60]

Φ(λ)pλ+e02λe12d1λe+12d+1λ. (A 15)

We are interested in corrections to the polariton splitting ΩRRe[λUP]Re[λLP] and bandwidths ΓLP=2Im[λLP], ΓUP=2Im[λUP], from their values in a single-mode scenario where q=0 is resonant with the dipoles (δ0). Corrections are due to the presence of the adjacent q=±1 modes. We thus solve for x in Φ(λ±+x)=0, where λ± denotes the pair of coupled single-mode solutions to equation (A 15) obtained for Ω±1=0, giving λ±(pλ±)+e02=0. For Δγ=δ=0 (p=0), the single-mode LP and UP solutions are λ±=±Ω0, giving the bare splitting ΩR=2Ω0 and bandwidths ΓLP=ΓUP=κ.

Linearizing the polynomial Φ(λ±+x)=0 around x=0 gives the general solution

x±=λ±[e12d+1+e+12d1(e12+e+12)λ±]P(λ±), (A 16)

with

P(λ±)=d+1(e02+e12)+d1(e02+e+12)d1d+1p2(e02+e12+e+12d+1pd1pd1d+1)λ±3(d1+d+1+p)λ±2+4λ±3. (A 17)

The solution λLP=λ+x gives the LP frequency and decay rate and λUP=λ++x+ gives the UP properties. The denominator P(λ±) can be reduced by setting p=0, which holds for a resonant single-mode scenario (δ0=0) without decay mismatch (Δγ=0), giving

x±=λ±×d+1e12+d1e+12λ±(e12+e+12)d+1(e02+e12)+d1(e02+e+12)2(e02+e+12+e12d1d+1)λ±3(d1+d+1)λ±2+4λ±3. (A 18)

For the special case where there are no mode-order variations of the Rabi coupling and cavity mode bandwidth, i.e. Ωq=Ω and κq=κ for all q, and assuming modes that are equally spaced (Δ±=±Δ), we have d+1e12+d1e+12=+ΔΩ2ΔΩ2=0 in equation (A 18) and corrections to the LP and UP eigenvalues become purely real:

x±=λ±Ω23Ω2+Δ22λ±2=Ω(Ω2Ω2+Δ2), (A 19)

with Δ>Ω and λ±=±Ω. Therefore, without mode-order variations of the cavity parameters, level pushing from the higher-order and lower-order modes gives the reduced Rabi splitting

ΩR=2Ω+x+x=2Ω(Δ2Δ2+Ω2), (A 20)

but there are no changes to the polariton bandwidths relative to the single-mode scenario.

The simplest ideal scenario with corrections to the LP and UP bandwidths is one where the Rabi couplings do not depend on mode order (Ωp=Ω), but cavity modes can differ in bandwidth (Δκq0). From equation (A 18), neglecting quadratic terms in Δκ±1 relative to Δ2+Ω2, we obtain in this case,

Re[x±]ΩΩ2(Ω2+Δ2)2(Ω2+Δ2)2+(Δ(Δκ1Δκ+1)Ω(Δκ1+Δκ+1)/2)2, (A 21)

and,

2Im[x±]=ΩΩ(Δκ1+Δκ+1)(Ω2+Δ2)/2±Ω2[Δ(Δκ1Δκ+1)Ω(Δκ1+Δκ+1)/2](Ω2+Δ2)2+(Δ(Δκ1Δκ+1)Ω(Δκ1+Δκ+1)/2)2. (A 22)

From these expressions, we may obtain the lower and upper polariton energies E±=±Ω+Re[x±] and decay rates Γ±=κ+2Im[x±]. For linear changes in mode bandwidth of the form Δκq=qζ, with ζ positive or negative, the LP and UP energies can then be written as

E±=±Ω[1Ω2(Ω2+Δ2)(Ω2+Δ2)2+Δ2ζ2], (A 23)

from where equation (5.4) in the main text is obtained. From equation (A 22), the polariton bandwidths for linear bandwidth changes with mode order are thus

Γ±=κ±2ΩΩ2Δζ(Ω2+Δ2)2+Δ2ζ2, (A 24)

which is equation (5.5) in the main text.

1

The thermodynamic limit being N, with Ω0=Ng0, NJ and NJ finite.

2

In the frame where M in equation (5.1) is defined, coupled bandwidths are given by Γ=κ0+2Im[λ], where λ is an eigenvalue of M, see the appendix.

3

The continuous-variable version of the single-photon wavepacket in equation (5.6) arises naturally when describing dipole emission in macroscopic QED [56].

4

In real microcavity structures, the mode bandwidths and Rabi couplings also vary with mode frequency.

Contributor Information

Felipe Herrera, Email: felipe.herrera.u@usach.cl.

William L. Barnes, Email: w.l.barnes@exeter.ac.uk.

Data accessibility

This will be via the University of Exeter Repository (ORE).

Declaration of AI use

We have not used AI-assisted technologies in creating this article.

Authors’ contributions

F.H.: conceptualization, data curation, formal analysis, funding acquisition, investigation, project administration, resources, software, validation, visualization, writing—original draft, writing—review and editing; W.L.B.: conceptualization, funding acquisition, investigation, methodology, project administration, validation, visualization, writing—original draft, writing—review and editing.

Both authors gave final approval for publication and agreed to be held accountable for the work performed therein.

Conflict of interest declaration

We declare we have no competing interests.

Funding

WLB acknowledges the support of European Research Council through the photmat project (ERC-2016-AdG-742222 :www.photmat.eu). FH acknowledges the support of the Royal Society through the award of a Royal Society Wolfson Visiting Fellowship, and through grants Fondecyt Regular 1221420, Millennium Science Initiative Program ICN17_012, and the Air Force Office of Scientific Research under award number FA9550-22-1-0245. For the purpose of open access, the authors have applied a Creative Commons Attribution (CC BY) licence to any Author AcceptedManuscript version arising. WLB would like to record his good fortune to have met Rodney Loudon several times. His unique combination of friendship, professionalism and curiosity, always came with a twinkle in his eye; humour was never far from the surface. Research data associated with this article can be found at: https://doi.org/10.24378/exe.5446.

References

  • 1. Lidzey DG, Bradley DDC, Virgili T, Armitage A, Skolnick MS, Walker S. 1999. Room temperature polariton emission from strongly coupled organic semiconductor microcavities. Phys. Rev. Lett. 82, 3316–3319. ( 10.1103/PhysRevLett.82.3316) [DOI] [Google Scholar]
  • 2. Pockrand I, Brillante A, Möbius D. 1982. Exciton–surface plasmon coupling: an experimental investigation. J. Chem. Phys. 77, 6289–6295. ( 10.1063/1.443834) [DOI] [Google Scholar]
  • 3. Herrera F, Owrutsky J. 2020. Molecular polaritons for controlling chemistry with quantum optics. J. Chem. Phys. 152, 100902. ( 10.1063/1.5136320) [DOI] [PubMed] [Google Scholar]
  • 4. Sánchez-Barquilla M, Fernández-Domínguez AI, Feist J, García-Vidal FJ. 2022. A theoretical perspective on molecular polaritonics. ACS Photonics 9, 1830–1841. ( 10.1021/acsphotonics.2c00048) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Mandal A, Taylor MAD, Weight BM, Koessler ER, Li X, Huo P. 2023. Theoretical advances in polariton chemistry and molecular cavity quantum electrodynamics. Chem. Rev. 123, 9786–9879. ( 10.1021/acs.chemrev.2c00855) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Bhuyan R, Lednev M, Feist J, Börjesson K. 2024. The effect of the relative size of the exciton reservoir on polariton photophysics. Adv. Opt. Mater. 12, 2301383. ( 10.1002/adom.202301383) [DOI] [Google Scholar]
  • 7. Khazanov T, Gunasekaran S, George A, Lomlu R, Mukherjee S, Musser AJ. 2023. Embrace the darkness: an experimental perspective on organic exciton–polaritons. Chem. Phys. Rev. 4, 041305. ( 10.1063/5.0168948) [DOI] [Google Scholar]
  • 8. Spano FC. 2015. Optical microcavities enhance the exciton coherence length and eliminate vibronic coupling in J-aggregates. J. Chem. Phys. 142, 184707. ( 10.1063/1.4919348) [DOI] [PubMed] [Google Scholar]
  • 9. Ćwik JA, Reja S, Littlewood PB, Keeling Jonathan. 2014. Polariton condensation with saturable molecules dressed by vibrational modes. EPL 105, 47009. ( 10.1209/0295-5075/105/47009) [DOI] [Google Scholar]
  • 10. Herrera F, Spano FC. 2016. Cavity-controlled chemistry in molecular ensembles. Phys. Rev. Lett. 116, 238301. ( 10.1103/PhysRevLett.116.238301) [DOI] [PubMed] [Google Scholar]
  • 11. Herrera F, Spano FC. 2017. Dark vibronic polaritons and the spectroscopy of organic microcavities. Phys. Rev. Lett. 118, 223601. ( 10.1103/PhysRevLett.118.223601) [DOI] [PubMed] [Google Scholar]
  • 12. Herrera F, Spano FC. 2017. Absorption and photoluminescence in organic cavity QED. Phys Rev A 95, 053867. ( 10.1103/PhysRevA.95.053867) [DOI] [Google Scholar]
  • 13. del Pino J, Schröder FAYN, Chin AW, Feist J, Garcia-Vidal FJ. 2018. Tensor network simulation of non-markovian dynamics in organic polaritons. Phys. Rev. Lett. 121, 227401. ( 10.1103/PhysRevLett.121.227401) [DOI] [PubMed] [Google Scholar]
  • 14. Ruggenthaler M, Flick J, Pellegrini C, Appel H, Tokatly IV, Rubio A. 2014. Quantum-electrodynamical density-functional theory: bridging quantum optics and electronic-structure theory. Phys. Rev. A 90, 012508. ( 10.1103/PhysRevA.90.012508) [DOI] [Google Scholar]
  • 15. Flick J, Ruggenthaler M, Appel H, Rubio A. 2017. Atoms and molecules in cavities, from weak to strong coupling in quantum-electrodynamics (QED) chemistry. Proc. Natl Acad. Sci. USA 114, 3026–3034. ( 10.1073/pnas.1615509114) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Luk HL, Feist J, Toppari JJ, Groenhof G. 2017. Multiscale molecular dynamics simulations of polaritonic chemistry. J. Chem. Theory Comput. 13, 4324–4335. ( 10.1021/acs.jctc.7b00388) [DOI] [PubMed] [Google Scholar]
  • 17. Agranovich VM, Litinskaia M, Lidzey DG. 2003. Cavity polaritons in microcavities containing disordered organic semiconductors. Phys. Rev. B 67, 085311. ( 10.1103/PhysRevB.67.085311) [DOI] [Google Scholar]
  • 18. Ribeiro RF. 2022. Multimode polariton effects on molecular energy transport and spectral fluctuations. Commun. Chem. 5, 48. ( 10.1038/s42004-022-00660-0) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Keeling J, Kéna-Cohen S. 2020. Bose-einstein condensation of exciton-polaritons in organic microcavities. Annu. Rev. Phys. Chem. 71, 435–459. ( 10.1146/annurev-physchem-010920-102509) [DOI] [PubMed] [Google Scholar]
  • 20. Arnardottir KB, Moilanen AJ, Strashko A, Törmä P, Keeling J. 2020. Multimode organic polariton lasing. Phys. Rev. Lett. 125, 233603. ( 10.1103/PhysRevLett.125.233603) [DOI] [PubMed] [Google Scholar]
  • 21. George J, Chervy T, Shalabney A, Devaux E, Hiura H, Genet C, Ebbesen TW. 2016. Multiple rabi splittings under ultrastrong vibrational coupling. Phys. Rev. Lett. 117, 153601. ( 10.1103/PhysRevLett.117.153601) [DOI] [PubMed] [Google Scholar]
  • 22. Balasubrahmaniyam M, Genet C, Schwartz T. 2021. Coupling and decoupling of polaritonic states in multimode cavities. Phys. Rev. B 103, L241407. ( 10.1103/PhysRevB.103.L241407) [DOI] [Google Scholar]
  • 23. Godsi M, Golombek A, Balasubrahmaniyam M, Schwartz T. 2023. Exploring the nature of high-order cavity polaritons under the coupling-decoupling transition. J. Chem. Phys. 159, 134307. ( 10.1063/5.0167945) [DOI] [PubMed] [Google Scholar]
  • 24. Mandal A, Xu D, Mahajan A, Lee J, Delor M, Reichman DR. 2023. Microscopic theory of multimode polariton dispersion in multilayered materials. Nano Lett. 23, 4082–4089. ( 10.1021/acs.nanolett.3c01017) [DOI] [PubMed] [Google Scholar]
  • 25. Załużny M, Zietkowski W. 2009. Intersubband cavity polaritons: the role of higher photonic modes. Phys. Rev. B 80, 245301. ( 10.1103/PhysRevB.80.245301) [DOI] [Google Scholar]
  • 26. Ao Z, Ashhab S, Yoshihara F, Fuse T, Kakuyanagi K, Saito S, Aoki T, Semba K. 2023. Extremely large lamb shift in a deep-strongly coupled circuit QED system with a multimode resonator. Sci. Rep. 13, 11340. ( 10.1038/s41598-023-36547-w) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Mornhinweg J, Diebel L, Halbhuber M, Riepl J, Cortese E, De Liberato S, Bougeard D, Huber R, Lange C. 2024. Sculpting ultrastrong light-matter coupling through spatial matter structuring. Nanophotonics 13, 1909–1915. ( 10.1515/nanoph-2023-0604) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Menghrajani KS, Vasista AB, Tan WJ, Thomas PA, Herrera F, Barnes WL. 2024. Molecular strong coupling and cavity finesse. J. Phys. Chem. Lett. 15, 7449–7457. ( 10.1021/acs.jpclett.4c00782) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Thomas PA, Barnes WL. 2024. Strong coupling-induced frequency shifts of highly detuned photonic modes in multimode cavities. J. Chem. Phys. 160, 204303. ( 10.1063/5.0208379) [DOI] [PubMed] [Google Scholar]
  • 30. Sundaresan NM, Liu Y, Sadri D, Szőcs LJ, Underwood DL, Malekakhlagh M, Türeci HE, Houck AA. 2015. Beyond strong coupling in a multimode cavity. Phys. Rev. X 5, 021035. ( 10.1103/PhysRevX.5.021035) [DOI] [Google Scholar]
  • 31. Coles DM, Lidzey DG. 2014. A ladder of polariton branches formed by coupling an organic semiconductor exciton to a series of closely spaced cavity-photon modes. Appl. Phys. Lett. 104, 191108. ( 10.1063/1.4876604) [DOI] [Google Scholar]
  • 32. Georgiou K, McGhee KE, Jayaprakash R, Lidzey DG. 2021. Observation of photon-mode decoupling in a strongly coupled multimode microcavity. J. Chem. Phys. 154, 124309. ( 10.1063/5.0038086) [DOI] [PubMed] [Google Scholar]
  • 33. Shalabney A, George J, Hutchison J, Pupillo G, Genet C, Ebbesen TW. 2015. Coherent coupling of molecular resonators with a microcavity mode. Nat. Commun. 6, 5981. ( 10.1038/ncomms6981) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Simpkins BS, Fears KP, Dressick WJ, Spann BT, Dunkelberger AD, Owrutsky JC. 2015. Spanning strong to weak normal mode coupling between vibrational and fabry–pérot cavity modes through tuning of vibrational absorption strength. ACS Photonics 2, 1460–1467. ( 10.1021/acsphotonics.5b00324) [DOI] [Google Scholar]
  • 35. Hertzog M, Börjesson K. 2020. The effect of coupling mode in the vibrational strong coupling regime. Chem. Photochem. 4, 612–617. ( 10.1002/cptc.202000047) [DOI] [Google Scholar]
  • 36. Balasubrahmaniyam M, Simkhovich A, Golombek A, Sandik G, Ankonina G, Schwartz T. 2023. From enhanced diffusion to ultrafast ballistic motion of hybrid light-matter excitations. Nat. Mater. 22, 338–344. ( 10.1038/s41563-022-01463-3) [DOI] [PubMed] [Google Scholar]
  • 37. Hestand NJ, Spano FC. 2018. Expanded theory of h- and j-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer. Chem. Rev. 118, 7069–7163. ( 10.1021/acs.chemrev.7b00581) [DOI] [PubMed] [Google Scholar]
  • 38. Hobson PA, Barnes WL, Lidzey DG, Gehring GA, Whittaker DM, Skolnick MS, Walker S. 2002. Strong exciton–photon coupling in a low-Q all-metal mirror microcavity. Appl. Phys. Lett. 81, 3519–3521. ( 10.1063/1.1517714) [DOI] [Google Scholar]
  • 39. Wang D, Kelkar H, Martin-Cano D, Rattenbacher D, Shkarin A, Utikal T, Götzinger S, Sandoghdar V. 2019. Turning a molecule into a coherent two-level quantum system. Nat. Phys. 15, 483–489. ( 10.1038/s41567-019-0436-5) [DOI] [Google Scholar]
  • 40. Skolnick MS, Fisher TA, Whittaker DM. 1998. Strong coupling phenomena in quantum microcavity structures. Semicond. Sci. Technol. 13, 645–669. ( 10.1088/0268-1242/13/7/003) [DOI] [Google Scholar]
  • 41. Lidzey DG, Bradley DDC, Skolnick MS, Virgili T, Walker S, Whittaker DM. 1998. Strong exciton–photon coupling in an organic semiconductor microcavity. Nature 395, 53–55. ( 10.1038/25692) [DOI] [Google Scholar]
  • 42. Barnes WL. 1998. Fluorescence near interfaces: the role of photonic mode density. J. Mod. Opt. 45, 661–699. ( 10.1080/09500349808230614) [DOI] [Google Scholar]
  • 43. Kéna-Cohen S, Davanço M, Forrest SR. 2008. Strong exciton-photon coupling in an organic single crystal microcavity. Phys. Rev. Lett. 101, 116401. ( 10.1103/PhysRevLett.101.116401) [DOI] [PubMed] [Google Scholar]
  • 44. Herrera F, Spano FC. 2018. Theory of nanoscale organic cavities: the essential role of vibration-photon dressed states. ACS Photonics 5, 65–79. ( 10.1021/acsphotonics.7b00728) [DOI] [Google Scholar]
  • 45. Spano FC. 2020. Exciton-phonon polaritons in organic microcavities: testing a simple ansatz for treating a large number of chromophores. J. Chem. Phys. 152, 204113. ( 10.1063/5.0002164) [DOI] [PubMed] [Google Scholar]
  • 46. Hagenmüller D, Schachenmayer J, Schütz S, Genes C, Pupillo G. 2017. Cavity-enhanced transport of charge. Phys. Rev. Lett. 119, 223601. ( 10.1103/PhysRevLett.119.223601) [DOI] [PubMed] [Google Scholar]
  • 47. Engelhardt G, Cao J. 2023. Polariton localization and dispersion properties of disordered quantum emitters in multimode microcavities. Phys. Rev. Lett. 130, 213602. ( 10.1103/PhysRevLett.130.213602) [DOI] [PubMed] [Google Scholar]
  • 48. Chávez NC, Mattiotti F, Méndez-Bermúdez JA, Borgonovi F, Celardo GL. 2021. Disorder-enhanced and disorder-independent transport with long-range hopping: application to molecular chains in optical cavities. Phys. Rev. Lett. 126, 153201. ( 10.1103/PhysRevLett.126.153201) [DOI] [PubMed] [Google Scholar]
  • 49. Kéna‐Cohen S, Maier SA, Bradley DDC. 2013. Ultrastrongly coupled exciton–polaritons in metal‐clad organic semiconductor microcavities. Adv. Opt. Mater. 1, 827–833. ( 10.1002/adom.201300256) [DOI] [Google Scholar]
  • 50. Breuer HP, Petruccione F. 2002. The theory of open quantum systems. Oxford, UK: Oxford University Press. ( 10.1007/3-540-44874-8_4) [DOI] [Google Scholar]
  • 51. Mølmer K, Castin Y, Dalibard J. 1993. Monte carlo wave-function method in quantum optics. J. Opt. Soc. Am. B 10, 524. ( 10.1364/JOSAB.10.000524) [DOI] [Google Scholar]
  • 52. Lednev M, García-Vidal FJ, Feist J. 2024. Lindblad master equation capable of describing hybrid quantum systems in the ultrastrong coupling regime. Phys. Rev. Lett. 132, 106902. ( 10.1103/PhysRevLett.132.106902) [DOI] [PubMed] [Google Scholar]
  • 53. Raimond JM, Brune M, Haroche S. 2001. Manipulating quantum entanglement with atoms and photons in a cavity. Rev. Mod. Phys. 73, 565–582. ( 10.1103/RevModPhys.73.565) [DOI] [Google Scholar]
  • 54. Leroux ID, Schleier-Smith MH, Vuletić V. 2010. Implementation of cavity squeezing of a collective atomic spin. Phys. Rev. Lett. 104, 073602. ( 10.1103/PhysRevLett.104.073602) [DOI] [PubMed] [Google Scholar]
  • 55. Thomas PA, Tan WJ, Kravets VG, Grigorenko AN, Barnes WL. 2024. Non-polaritonic effects in cavity-modified photochemistry. Adv. Mater. Weinheim 36, e2309393. ( 10.1002/adma.202309393) [DOI] [PubMed] [Google Scholar]
  • 56. Feist J, Fernández-Domínguez AI, García-Vidal FJ. 2021. Macroscopic qed for quantum nanophotonics: emitter-centered modes as a minimal basis for multiemitter problems. Nanophotonics 10, 477–489. ( 10.1515/nanoph-2020-0451) [DOI] [Google Scholar]
  • 57. Spano FC. 2010. The spectral signatures of frenkel polarons in h- and j-aggregates. Acc. Chem. Res. 43, 429–439. ( 10.1021/ar900233v) [DOI] [PubMed] [Google Scholar]
  • 58. Dubail J, Botzung T, Schachenmayer J, Pupillo G, Hagenmüller D. 2022. Large random arrowhead matrices: multifractality, semilocalization, and protected transport in disordered quantum spins coupled to a cavity. Phys. Rev. A 105, 023714. ( 10.1103/PhysRevA.105.023714) [DOI] [Google Scholar]
  • 59. Wellnitz D, Pupillo G, Schachenmayer J. 2022. Disorder enhanced vibrational entanglement and dynamics in polaritonic chemistry. Commun. Phys. 5, 120. ( 10.1038/s42005-022-00892-5) [DOI] [Google Scholar]
  • 60. O’Leary DP, Stewart GW. 1990. Computing the eigenvalues and eigenvectors of symmetric arrowhead matrices. J. Comput. Phys. 90, 497–505. ( 10.1016/0021-9991(90)90177-3) [DOI] [Google Scholar]

Associated Data

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

Data Availability Statement

This will be via the University of Exeter Repository (ORE).


Articles from Philosophical transactions. Series A, Mathematical, physical, and engineering sciences are provided here courtesy of The Royal Society

RESOURCES