Abstract
Hybridisation of the cavity modes and the excitons to polariton states together with the coupling to the vibrational modes determine the linear optical properties of organic semiconductors in microcavities. In this article we compute the refractive index for such system using the Holstein–Tavis–Cummings model and determine then the linear optical properties using the transfer matrix method. We first extract the parameters for the exciton in our model from fitting to experimentally measured absorption of a 2,7-bis[9,9-di(4-methylphenyl)-fluoren-2-yl]-9,9-di(4-methylphenyl) fluorene (TDAF) molecular thin film. Then we compute the reflectivity of such a thin film in a metal clad microcavity system by including the dispersive microcavity mode to the model. We compute susceptibility of the model systems evolving just a single state vector by using the non-Markovian quantum state diffusion. The computed location and height of the lower and upper polaritons agree with the experiment within the estimated errorbars for small angles
1 Introduction
The localized nature of molecular excitons imparts two crucial attributes: a large binding energy and a dense population of excitons compared to the photonic density of states. The former feature promotes strong light–matter coupling at room temperature, enabling the macroscopic study of Bose–Einstein condensation and superfluidity at elevated temperatures [1], [2], [3], [4]. The latter feature enabled the observation of ultrastrong coupling [5] and single photon nonlinearity [6]. Recently, the concept of polariton chemistry was introduced, which promises to reshape the energy landscape of molecular systems, exerting control over their photochemical and photophysical processes [7], [8], [9], [10], [11]. In optical microcavities filled with molecular absorbers, polariton modes are the eigenstates resulting from the strong coupling between the optical modes and the exciton resonances. They can be observed experimentally as the avoided crossing of the bare exciton and microcavity photon dispersion through reflectivity measurements [12].
An interesting implication of strong coupling at non-zero temperature is that the exciton might also strongly couple to the surrounding molecular vibrations which can be the catalyst for complex relaxation dynamics of polaritons [13]. Currently, there is significant research being conducted to explore the impact of strong collective light–matter interactions on the intersystem crossing (ISC) and reverse intersystem crossing (RISC) rates [14], [15], [16], [17], [18], [19], [20], [21]. Rabi splitting is proportional to the square root of the molecular density that share the microcavity photon
The linear optical properties of organic microcavity polaritons have been investigated in the past using various different methods. The coupling of the exciton to the vibrational modes can be modeled using the Holstein model (independent boson model) [26], [27] which can be exactly solved. When the molecules are placed into a microcavity the polariton dispersion relation can be observed in the collective strong coupling regime. A commonly used approach to extract the Rabi splitting is to fit the dispersion relation of coupled damped harmonic oscillators to the experimentally measured reflectivity spectrum [28], [29]. The coupling to the molecular vibrations can be included with more advanced approaches such as the input–output theory [30], [31], cumulant expansion [32], [33], [34], stochastic sampling of exact wave functions [35], Green’s function methods [36], QM/MM simulations [22], few molecule models [37], and even single molecule models [38] to name a few.
The purpose of this article is to introduce a new method for investigating the polariton dynamics and pave the way for gaining new insights into the so called “large number of molecules N” problem. In this work we take the first steps of applying non-Markovian quantum state diffusion (NMQSD) for computing the linear optical properties of organic microavity polaritons. While a possible drawback of this approach is that it is stochastic and typically requires a large number of trajectories for computing expectation values of observables, for computing linear optical properties using just one trajectory is enough [39], [40], [41]. In contrast to computing the response of polaritons from the response of the cavity modes alone, as in [38], [42] for example, we use NMQSD to compute the linear response of the combined exciton cavity mode system which is probed by classical electromagnetic field. This means that we consider a situation where the light that is transmitted through the cavity mirrors may be directly absorbed by both the cavity modes and the molecules. Excluding the molecular absorption results in high absorption of the upper polariton at higher angles of the incoming light, which is in conflict with experimental data. From the response function (linear susceptibility) we obtain the index of refraction which is used then in solving the Maxwell equations corresponding to the experimental setup.
We present a graphical summary of this article in Figure 1. (a) We use the transfer matrix method (TMM) to compute the reflectivity, transmission, and absorption of the system when it is probed by a classical electromagnetic field. We assume that the response of the system is linear. Then the macroscopic properties of the sample are determined by the index of refraction which is obtained from the linear susceptibility χ(ω). We use NMQSD to compute this in the model systems evolving just a single state vector. (b) The susceptibility of a thin film is modeled by the susceptibility of just one molecule multiplied by the dipole number density. We fit a model to the experimental data and extract the values for the parameters of the organic semiconductor exciton. (c) The microcavity system is modeled by a slab of molecules interacting with a microcavity mode. The susceptibility is computed by evolving now a single pure state of a system of N molecules interacting with K cavity modes using NMQSD.

Measurement and systems considered in this article.(a) We model the linear optical properties of a slab of material in terms of the dielectric function ɛ, which can be computed from the susceptibility χ. The susceptibility is computed from a quantum mechanical model evolving just a single state vector. This leads to a computationally efficient scheme in contrast to approaches where density matrix evolution is needed. In the other panels we describe which kind of system we consider. (b) Susceptibility of a thin film can be modeled as the susceptibility of a single quantum absorber (red circle) multiplied with the number of such absorbers, when spatial disorder can be neglected. We include energetic disorder as indicated by the distribution of exciton energies ɛ s , and a coupling to vibration modes when the system is excited (indicated by the spring). As the molecule absorbs a photon an exciton is formed indicated by the red filled and empty circles. (c) In the case of microcavity polaritons we consider a slice of emitters in the z-direction, whereas the cavity mirrors are located at x = ±L/2 (top and bottom). The cavity mode is represented by the parabolic curve between the mirrors. The incoming light may be absorbed coherently by the molecules and the cavity mode, which is indicated by the superposition of excitons in the molecule and an excitation in the cavity mode.
The structure of this article is the following. In Section 2 we introduce the linear response theory and how the susceptibility and refractive index can be computed. In Section 3 we introduce the NMQSD method. Furthermore, we show how the susceptibility can be computed using the NMQSD approach. Then in Section 4 we fit the model parameters to the single molecule data obtained from experimental measurements and density functional calculations. We will focus on 2,7-bis[9,9-di(4-methylphenyl)-fluoren-2-yl]-9,9-di(4-methylphenyl) fluorene (TDAF) as it is a model system for strong light–matter studies [21], [43], [44]. In Section 5 we construct the model for microcavity polaritons using the same TDAF molecule. We discuss in detail how the susceptibility can be computed in this case when the dispersive microcavity mode is also included. We also present the results of the theoretical calculations. In Section 6 we present our conclusions and outlook. Experimental details are presented in Section 7.
2 Linear response
2.1 Linear optics
In linear materials the polarization field is proportional to the applied electric field
where ɛ 0 is the vacuum permittivity and χ is the susceptibility. We set ℏ = 1 in all subsequent equations. Due to causality polarization can depend only on the fields applied on earlier times and may have non-local spatial dependency [45]. In general, χ is a second rank tensor. In this work we focus only on situations where the polarization is aligned with the applied electric field making χ a scalar. By using the convolution theorem and after dropping the vector notation this relation is
where ω is the angular frequency and k z is the z-component of the wave vector. The dielectric function (or relative permittivity) is
The dielectric function determines the refractive index by
2.2 Dipole density
Polarization corresponds to the dipole density of the medium, which can be computed from a microscopical model. We consider a situation where weak field is used to probe a quantum system. The interaction term between the classical electromagnetic field and the matter is taken to be
where z m is the location of a point dipole and μ m is the dipole operator. The other degrees of freedom are described by a time independent Hamilton operator H. The dynamics generated by H is given by the unitary operator
The linear response can be computed from the dipole correlation function M(t) [39], [40], [47]. The susceptibility of the system can be computed as the Fourier transform of M(t)
where
and |Ψ G ⟩ is the ground state of the Hamiltonian H. We assume that the systems we investigate do not have permanent dipole moment. Under this assumption the dipole correlation function is obtained from perturbation theory and keeping only the positive energy terms and the terms linearly proportional to the applied field [48], [49]. Generalization to anisotropic cases is straightforward. Using Eqs. (2) and (6) gives the polarization density of the system. The macroscopic polarization is obtained by multiplying the polarization density with the sample volume. In this work we neglect any spatial disorder. Finite temperature results can be computed using the ground state initial condition in the NMQSD approach as we show later. Alternative methods for solving Eq. (7) exist such as cumulant expansion techniques [33], [34] or exact stochastic wave function sampling [35].
3 Non-Markovian Quantum State Diffusion
3.1 General theory
The aim of the NMQSD approach is to solve the time evolution of the full Schrödinger equation for the open system and the environment [50], [51], [52]. A typical model consists of an open system with Hamiltonian H S and a coupling operator L. These operators are arbitrary at this point. The environment is assumed to consist of quantum harmonic oscillators with a Hamiltonian
where
The initial state of the bath is the thermal state ρ β and the system and the bath are initially uncorrelated. In the interaction picture with respect to (8) the Schrödinger equation is
The finite temperature NMQSD equation corresponding to the Schrödinger equation (10) is [50]
where
The Hermitian autocorrelation of the process corresponds to the zero temperature bath correlation function (BCF)
We have introduced the spectral density
Finite temperature is incorporated by a “stochastic potential” V(t)
where η t is a zero mean Gaussian stochastic process with correlations [51], [54]
and
The NMQSD approach works in general with an arbitrary bath correlation function. For example, bath correlation functions obtained from surrogate harmonic bath embeddings such as in Ref. [38] could be incorporated into our formalism. The challenging part in using NMQSD is the occurrence of the functional derivative. In recent years a powerful hierarchy of pure states (HOPS) approach has emerged as a general numerical approach to solve the NMQSD equations [51], [55]. We explain next the weak coupling approximation for the functional derivative term we use later in this work. The weak coupling means that we keep terms up to quadratic order in the coupling strength g
λ
. In particular we have
and defining auxiliary states
and the auxiliary states satisfy following equation of motion [55]
We truncate hierarchy to first order, which means that we neglect the last term. We also neglect the term V(t) and
As G
μ
are already proportional to
We stress that these approximations lead to closed NMQSD equations of motion and correspond to a weak system bath coupling approximation [56]. Further refinements are possible in terms of HOPS, but we could explain the experimental data in this work using the weak coupling approximation only.
3.2 Susceptibility using NMQSD
Using the NMQSD we can compute the susceptibility of the system evolving pure states only. In case the coupling operator is Hermitian we need to evolve only one pure state, otherwise we need to average over the thermal noise [41]. In the case that we have multiple transition dipole moments we need to extend the NMQSD to many systems which all couple to their individual environments. This simply means that each subsystem has their own coupling operator L
m
, noise term
where |ψ
0⟩ is the initial state for the NMQSD evolution and |ψ
t
(z* = 0)⟩ is the solution to the NMQSD equation where the driving noise is set to zero, i.e.
If the coupling operator is Hermitian, we can replace the stochastic potential with the finite temperature bath correlation function
Otherwise we need to average over different realizations of the thermal noise [39].
4 Molecular thin film
4.1 Description
The exciton of the TDAF is determined from experimentally measured absorption of a 60-nm-thick film of TDAF on a quartz substrate. The absorption is defined as A = 1 − T − R, where T and R are the fractions of the transmitted and reflected light, respectively. To accomplish the measurement of the reflected and transmitted light from the film without increasing the optical path length, the sample was excited at a 15° angle. We will model this process by computing the refractive index from a microscopic model and then the reflected and transmitted light using TMM [57].
4.2 Model
The dynamics of the molecule is governed by the Holstein model [26]
We denote by K = σ + σ − from now on. The system Hamiltonian and the coupling operator are in this case
where ζ is a disorder parameter. The coupling operator is Hermitian. The NMQSD equation in this case is
where α(t − s) is the thermal BCF. We model the radiative damping by and additional white noise process
where a is parameter additionally controlling the coupling strength. In the limit that the radiative damping is small compared to other parameters of the system the model admits a solution
where
4.3 Susceptibility
The dipole operator for this system is
with the abuse of notation we use the same symbol for the transition dipole operator and the transition dipole moment. Now the initial state given by (19) is the excited state |e⟩ of the molecule. Inserting this and computing the average over the disorder gives the following expression for the dipole correlation function
In the case that g(t) = 0 this corresponds to the Voigt lineshape. The susceptibility is obtained by taking the Laplace transform from the dipole correlation function (27) and the refractive index can be readily computed.
4.4 Thin film absorption
The first singlet excited state is of the system is at ɛ
S
≈ 3.6 eV as can be seen from the experimental trace in Figure 2. The radiative lifetime of the TDAF thin film is reported to be 133 ps (

Absorption of a 60 nm thick thin film of TDAF molecules. The exciton energy is approximately 3.6 eV. Our model (NMQSD) fits well to the measured data, whereas a blind fit with a Voigt lineshape performs weaker in line with reported χ 2 values for each fit. The errorbars in the fit are smaller than the linewidth.
The parameter values found in the fitting process are ɛ
s
= 3.6 eV, σ = 0.14 eV, and ξ = 0.09. We kept the coupling strength parameter at a fixed value a = 1 and γ = 5 × 10−5 eV and included a background refractive index
4.5 Summary
We model the thin film of TDAF molecules as two level systems which each are coupled to their respective molecular vibrations and are probed by a weak classical electromagnetic field. We assume that there is no spatial disorder but the energies of the excitons are distributed according to Gaussian distribution with mean ɛ s and variance σ. The macroscopic polarization is then obtained by multiplying the induced dipole moment with the number of molecules in the sample. The inclusion of the disorder is motivated by the likeness of the experimental absorption lineshape to a Voigt profile and the fact that we found that thermal effects in our model were negligible in this parameter regime. The asymmetry in the absorption lineshape is explained by the coupling to the vibrational bath with spectral density given in Eq. (24). We find the parameter values by fitting the model to the experimentally measured absorption using TMM. The fitted parameters are the excitonic energy (ɛ s ), energy disorder (σ), coupling strength a, and the cut-off ξ. We keep the radiative lifetime of the exciton and the background refractive index constant during the fit. The fit of our model is very good and the obtained parameter values are in agreement with what is obtained from DFT calculations [21]. The quality of the fit justifies the use of approximate Eq. (25), which excludes for example possible direct interactions and cooperative effects between the molecules themselves.
5 Microcavity polariton reflectivity
5.1 System
The system is a 80 nm TDAF film in a cavity formed by two aluminium mirrors with thicknesses 25 nm and 100 nm. The reflectivity of this system can be measured experimentally as a function of the angle and energy of the incoming light. We again compute the refractive index from a microscopic model and then compare the reflectivity calculated using TMM with the experimental data.
5.2 Model
We follow [22], [59] in the construction of the model. The system consists of N molecules in a planar microcavity. The Hamiltonian for the molecules is
Photons in the cavity have the energy
where c is the speed of light in the vacuum and n
r
is the refractive index of the propagation medium. k is the wave vector of the light which is assumed to be in the x–z plane. Mirrors at
Each mode has two orthogonal transverse polarizations u
1 and
where
where u indicates the direction of the electric field of the confined mode, ɛ 0 is the vacuum permittivity, and V the cavity mode volume. We assume that the polarization of the cavity modes u, the dipole moments μ j , and the polarization of any incoming light are all aligned in the y-direction, thus we consider a situation described by Eq. (2). The coupling between the molecules and the cavity modes is given by the Tavis–Cummings interaction
where
The coupling of each molecule to local vibrational modes is the same as in Section 4
where we assume that each molecule couples to its own bath of vibrational modes. The total Hamiltonian for the system is then
where H E is the free Hamiltonian for the vibrational modes
In addition, the cavity modes are damped with a rate that does not depend on k
z
and we denote it by κ. The cavity damping is described in terms of additional zero mean white noise processes
Similarly, the excitons are damped with the rate γ, which is described by white noise processes ξ t,j . The evolution of the system is given by the NMQSD equation which is trivially extended from the ones given earlier in the paper. Namely, we read off the terms of the NMQSD equation from the HTC Hamiltonian (Eq. (35)) and add independent excitonic and cavity dampings described by the white noise terms (Eq. (37)). Differently to the previous case, the thermal effects are significant and we use the non-zero temperature bath correlation function (Eq. (20)) in the subsequent calculations.
5.3 Susceptibility
We write the transition dipole moments for the molecules similarly as in Eq. (26) for each individual molecule. The dipole correlation function of this system can be calculated using (18). In this case the initial state (19)
is used. We take into account the possibility of light being absorbed by the cavity mode by an additional dipole moment
where θ c is the angle inside the cavity and we have used the Snell’s law.
5.4 Reflectivity
We show the computed and measured reflectivity in Figure 3 as a density plot. The quantitative agreement is good at small angles. For larger angles the locations of the polariton modes are in good qualitative agreement. In Figure 4 we show the computed and measured reflectivity as a function of the energy for different angles. There the disagreement at larger angles is more prominently visible. In Figure 4 we also present the estimated errorbars. We can conclude that for angles up to approximately 30° the model and the experiment are within the estimated errorbars. We discuss the errors involved in Section 6. In the TMM calculations the front mirror and the TDAF film is replaced with a material that has the computed refractive index. We use the parameters found in the fitting process in Section 4. However, we set molecular disorder to zero and discuss this point in Section 6. To cover the full range of angles and energies used in the experiment, we use 21 cavity modes. Then the number of molecules is chosen to be high enough where additional increase does not change the reflectivity spectrum significantly. We choose N = 84. The cavity decay rate κ = 0.21 eV is estimated from experimental photoluminescence of the cavity system. We fitted the positions and depths of the reflectivity minima to the experimental data and got the values n r = 2.15 for the refractive index that determines the cavity dispersion relation (30), E(0) = 3.41 eV for the energy of the cavity mode with k z = 0, Ω = 0.91 eV for the Rabi splitting, and μ c = 3.5μ m for the magnitude of the cavity dipole moment.

Experimental and calculated reflectivities for the cavity system. The qualitative agreement of the experiment and the theory is good. The reduced reflectivity of the upper polariton is due to the coupling to the vibrational modes. The agreement between our theory and the experiment is better at smaller angles.

Error estimates for the polariton model. For small angles θ ≤ 30° the location and amplitude of the computed and measured upper and lower polariton peaks are within the estimated error bars. For larger angles the model is in good qualitative agreement with the measured reflectivity and predicts the locations of the polariton peaks.
With this model we can consider the classical field interacting with the system in three different ways. The first is that the light is absorbed by one of the cavity modes and there is no direct molecular absorption (μ m = 0). This makes the upper polariton absorb more and more as the angle increases because it becomes increasingly photonic. The second is that the light can only be absorbed by the molecules and μ c = 0. This results in both the upper and the lower polariton absorbing roughly the same amount. The third option is to include both cavity and molecular absorption which we found to match the experimental results the best. The vibrational coupling reduces the absorption of the upper polariton and this effect is greater at small angles where the polariton is mostly excitonic. Excluding this coupling would lead to pronounced absorption of the upper polariton in comparison to what is observed experimentally. Including quantum mechanical coupling to the vibrational modes and taking in account the possibility for both the molecular and the cavity mode absorption leads to an agreement between the experiment and our theory.
5.5 Summary
We extend the thin film model from Section 4 by including the dispersive cavity mode and spatial locations of the molecules. We have excluded the energy disorder of the excitons from the model. This is done because the effect of the energy disorder led to poor agreement with the model and the data. It may be that the approximations we do in our modeling do not correctly take into account the disorder and the coupling to the cavity modes. On the other hand, neglecting the effects of the exciton broadening are in line with the observations [60], [61], where the coupling to the cavity mode diminishes the exciton broadening effect. There, however, the cavities had higher Q factors than in our case. We estimate the error in our modeling by Monte Carlo sampling. We sample the computed reflectivity with randomly choosing the input parameters. We assume that all of the deviations are independent and distributed normally around the parameter values used in Figure 3. We consider the following deviations: The input energy has a standard deviation σ
E
= 10−6 eV, the input angle σ
θ
= 0.001°, and the thickness of the system σ
d
= 10−9 m. We assume that the computed refractive index has energy dependent standard deviation
6 Conclusions and outlook
We have computed the susceptibility for organic microcavity polaritons from the Holstein–Tavis–Cummings model using non-Markovian quantum state diffusion. This approach is very efficient since we can compute susceptibility from evolving just a single state vector, so that density matrix computations are not needed. We have shown that our model can explain the absorptive properties of TDAF thin films and the reflectivity of the TDAF microcavity polaritons. In the thin film case the model explains the asymmetric shape of the absorption line very well. In the polariton case we do have good quantitative agreement around the polariton peaks at small angles
We have identified several open questions that will be the focus of our forthcoming investigations:
We have seen that energetic disorder leads to a good fit in the thin film case but needs to be removed in the microcavity polariton case. Possible explanations used in the literature are the cavity filtering effect [60], [61] or motional narrowing [62], [63]. The former may not be the right explanation in this case since the cavities used in this work have such poor Q factors. The latter explanation fails as the spectroscopy we use conserves the planar wave vector of the cavity [45].
In this work we have relied on perturbative solutions to the NMQSD equation as they provided reasonable fits to the experiment. It will be interesting to investigate situations where such perturbative approaches fail. In such cases, there may be more complex intramolecular dynamics which may involve also the spin orbit coupling between the singlet and triplet states [23].
Lastly, we point out that state vector based approaches, such as NMQSD, open up a new way to study delocalization degree of the polaritons and polariton transport as it is possible to observe dynamically how the localization due to coupling to vibrational degrees of freedom and delocalization due to cavity coupling compete.
7 Methods
7.1 Fabrication
The samples were fabricated using thermal evaporation at a base pressure below 10−7 Torr (Angstrom Engineering physical vapor deposition system). We used 15 × 15 mm2 quartz substrates that were cleaned by sonication for 10 min in soapy water (3 % Decon 90), acetone, and isopropanol, respectively, and dried with nitrogen. A 100-nm-thick aluminium was deposited on top of the substrate as the bottom mirror, followed by the deposition of 80 nm TDAF as the active layer, 1 nm of LiF, and a 25-nm-thick aluminium layer as a top polariton microcavity mirror.
7.2 Characterization
The TDAF absorption and polariton angle-resolved reflectivity were measured with a spectroscopic ellipsometer (J.A. Woollam VASE) in reflectivity and transmission configuration. To extract the absorption of TDAF film, we measured transmitted and reflected light at a 15° excitation angle, which represents the minimum angle our setup can measure reflectivity and adds only a 2 % increase in the optical path.
Funding source: H2020 European Research Council
Award Identifier / Grant number: 94826
Funding source: Business Finland
Award Identifier / Grant number: 1951/31/2021
Acknowledgments
K.L. would like to thank Walter Strunz, Valentin Link, Kai Müller, and Christian Schäfer for fruitful discussions. The authors would like to thank the anonymous reviewers and Olli Siltanen for useful comments.
-
Research funding: This project has received funding from the European Research Council under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. [948260]) and from Business Finland project Turku-R2B-Bragg WOLED with decision number 1951/31/2021.
-
Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Conflict of interests: Authors state no conflicts of interest.
-
Informed consent: Informed consent was obtained from all individuals included in this study.
-
Ethical approval: The conducted research is not related to either human or animals use.
-
Data availability: The datasets generated during and/or analyzed during the current study are available from the corresponding author upon reasonable request.
References
[1] J. Keeling and S. Kéna-Cohen, “Bose-einstein condensation of exciton-polaritons in organic microcavities,” Annu. Rev. Phys. Chem., vol. 71, no. 1, p. 435, 2020, https://doi.org/10.1146/annurev-physchem-010920-102509.Search in Google Scholar PubMed
[2] J. Tang, et al.., “Room temperature exciton–polariton Bose–Einstein condensation in organic single-crystal microribbon cavities,” Nat. Commun., vol. 12, no. 1, p. 3265, 2021, https://doi.org/10.1038/s41467-021-23524-y.Search in Google Scholar PubMed PubMed Central
[3] A. J. Moilanen, K. S. Daskalakis, J. M. Taskinen, and P. Törmä, “Spatial and temporal coherence in strongly coupled plasmonic bose-einstein condensates,” Phys. Rev. Lett., vol. 127, no. 25, 2021, Art. no. 255301, https://doi.org/10.1103/physrevlett.127.255301.Search in Google Scholar
[4] G. W. Castellanos, M. Ramezani, S. Murai, and J. Gómez Rivas, “Non-equilibrium bose–einstein condensation of exciton-polaritons in silicon metasurfaces,” Adv. Opt. Mater., vol. 11, no. 7, 2023, Art. no. 2202305, https://doi.org/10.1002/adom.202202305.Search in Google Scholar
[5] A. Frisk Kockum, A. Miranowicz, S. De Liberato, S. Savasta, and F. Nori, “Ultrastrong coupling between light and matter,” Nat. Rev. Phys., vol. 1, no. 1, p. 19, 2019, https://doi.org/10.1038/s42254-018-0006-2.Search in Google Scholar
[6] A. V. Zasedatelev, et al.., “Single-photon nonlinearity at room temperature,” Nature, vol. 597, no. 7877, p. 493, 2021, https://doi.org/10.1038/s41586-021-03866-9.Search in Google Scholar PubMed
[7] D. Sanvitto and S. Kéna-Cohen, “The road towards polaritonic devices,” Nat. Mater., vol. 15, no. 10, p. 1061, 2016, https://doi.org/10.1038/nmat4668.Search in Google Scholar PubMed
[8] M. Hertzog, M. Wang, J. Mony, and K. Börjesson, Strong Light-Matter Interactions: A New Direction within Chemistry, London, Royal Society of Chemistry,2019.10.1039/C8CS00193FSearch in Google Scholar PubMed PubMed Central
[9] J. Y. Yuen-Zhou, L. A. Martínez-Martínez, J. B. Pérez-Sánchez, and K. Schwennicke, “Polariton chemistry: controlling organic photophysical processes with strong light-matter coupling,” in Physical Chemistry of Semiconductor Materials and Interfaces IX, August, D. Congreve, C. Nielsen, and A. J. Musser, Eds., Bellingham, Washington, USA,SPIE, 2020, p. 23.10.1117/12.2569171Search in Google Scholar
[10(a)] C. Ciuti and I. Carusotto, “Input-output theory of cavities in the ultrastrong coupling regime: the case of time-independent cavity parameters,” Phys. Rev. A, vol. 74, no. 3, p. 033811, 2006.Search in Google Scholar
(b) F. J. Garcia-Vidal and T. W. Ebbesen, “Manipulating matter by strong coupling to vacuum fields,” Science, vol. 373, no. 6551, p. eabd0336, 2021.10.1126/science.abd0336Search in Google Scholar PubMed
[11] R. Bhuyan, et al.., “The rise and current status of polaritonic photochemistry and photophysics,” Chem. Rev., vol. 123, no. 18, pp.10877–10919, 2023.10.1021/acs.chemrev.2c00895Search in Google Scholar PubMed PubMed Central
[12] P. Törmä and W. L. Barnes, “Strong coupling between surface plasmon polaritons and emitters: a review,” Rep. Prog. Phys., vol. 78, no. 1, 2015, Art. no. 013901, https://doi.org/10.1088/0034-4885/78/1/013901.Search in Google Scholar PubMed
[13] E. Hulkko, S. Pikker, V. Tiainen, R. H. Tichauer, G. Groenhof, and J. J. Toppari, “Effect of molecular Stokes shift on polariton dynamics,” J. Chem. Phys., vol. 154, no. 15, 2021, Art. no. 154303, https://doi.org/10.1063/5.0037896.Search in Google Scholar PubMed
[14] K. Stranius, M. Hertzog, and K. Börjesson, “Selective manipulation of electronically excited states through strong light–matter interactions,” Nat. Commun., vol. 9, no. 1, p. 2273, 2018, https://doi.org/10.1038/s41467-018-04736-1.Search in Google Scholar PubMed PubMed Central
[15] L. A. Martínez-Martínez, M. Du, R. F. Ribeiro, S. Kéna-Cohen, and J. Yuen-Zhou, “Polariton-assisted singlet fission in acene aggregates,” J. Phys. Chem. Lett., vol. 9, no. 8, p. 1951, 2018, https://doi.org/10.1021/acs.jpclett.8b00008.Search in Google Scholar PubMed
[16] A. M. Berghuis, et al.., “Enhanced delayed fluorescence in tetracene crystals by strong light-matter coupling,” Adv. Funct. Mater., vol. 29, no. 36, 2019, Art. no. 1901317, https://doi.org/10.1002/adfm.201901317.Search in Google Scholar
[17] E. Eizner, L. A. Martínez-Martínez, J. Yuen-Zhou, and S. Kéna-Cohen, “Inverting singlet and triplet excited states using strong light-matter coupling,” Sci. Adv., vol. 5, 2019, no. 12, Art. no. eaax4482, https://doi.org/10.1126/sciadv.aax4482.Search in Google Scholar PubMed PubMed Central
[18] D. Polak, et al.., “Manipulating molecules with strong coupling: harvesting triplet excitons in organic exciton microcavities,” Chem. Sci., vol. 11, no. 2, p. 343, 2020, https://doi.org/10.1039/c9sc04950a.Search in Google Scholar PubMed PubMed Central
[19] Y. Yu, S. Mallick, M. Wang, and K. Börjesson, “Barrier-free reverse-intersystem crossing in organic molecules by strong light-matter coupling,” Nat. Commun., vol. 12, no. 1, p. 3255, 2021, https://doi.org/10.1038/s41467-021-23481-6.Search in Google Scholar PubMed PubMed Central
[20] A. Mukherjee, J. Feist, and K. Börjesson, “Quantitative investigation of the rate of intersystem crossing in the strong exciton–photon coupling regime,” J. Am. Chem. Soc., vol. 145, no. 9, p. 5155, 2023, https://doi.org/10.1021/jacs.2c11531.Search in Google Scholar PubMed PubMed Central
[21] A. G. Abdelmagid, et al.., “Identifying the origin of delayed electroluminescence in a polariton organic light-emitting diode,” Nanophotonics, vol. 13, no. 14, pp. 2565–2573, 2024. https://doi.org/10.1515/nanoph-2023-0587.10.1515/nanoph-2023-0587Search in Google Scholar PubMed PubMed Central
[22] R. H. Tichauer, J. Feist, and G. Groenhof, “Multi-scale dynamics simulations of molecular polaritons: the effect of multiple cavity modes on polariton relaxation,” J. Chem. Phys., vol. 154, no. 10, 2021, Art. no. 104112, https://doi.org/10.1063/5.0037868.Search in Google Scholar PubMed
[23] L. A. Martínez-Martínez, E. Eizner, S. Kéna-Cohen, and J. Yuen-Zhou, “Triplet harvesting in the polaritonic regime: a variational polaron approach,” J. Chem. Phys., vol. 151, no. 5, p. 054106, 2019.10.1063/1.5100192Search in Google Scholar
[24] M. Sánchez-Barquilla, F. J. García-Vidal, A. I. Fernández-Domínguez, and J. Feist, “Few-mode field quantization for multiple emitters,” Nanophotonics, vol. 11, no. 19, p. 4363, 2022, https://doi.org/10.1515/nanoph-2021-0795.Search in Google Scholar PubMed PubMed Central
[25] K. Miwa, S. Sakamoto, and A. Ishizaki, “Control and enhancement of single-molecule electroluminescence through strong light–matter coupling,” Nano Lett., vol. 23, no. 8, p. 3231, 2023, https://doi.org/10.1021/acs.nanolett.2c05089.Search in Google Scholar PubMed
[26] T. Holstein, “Studies of polaron motion: Part i. the molecular-crystal model,” Ann. Phys., vol. 8, no. 3, p. 325, 1959, https://doi.org/10.1016/0003-4916(59)90002-8.Search in Google Scholar
[27] B. Krummheuer, V. M. Axt, and T. Kuhn, “Theory of pure dephasing and the resulting absorption line shape in semiconductor quantum dots,” Phys. Rev. B, vol. 65, no. 19, 2002, Art. no. 195313, https://doi.org/10.1103/physrevb.65.195313.Search in Google Scholar
[28] V. M. Agranovich, Excitations in Organic Solids, vol. 142, Oxford, England, OUP Oxford, 2009.10.1093/acprof:oso/9780199234417.001.0001Search in Google Scholar
[29] V. M. Agranovich, M. Litinskaia, and D. G. Lidzey, “Cavity polaritons in microcavities containing disordered organic semiconductors,” Phys. Rev. B, vol. 67, no. 8, 2003, Art. no. 085311, https://doi.org/10.1103/physrevb.67.085311.Search in Google Scholar
[30] C. W. Gardiner and M. J. Collett, “Input and output in damped quantum systems: quantum stochastic differential equations and the master equation,” Phys. Rev. A, vol. 31, no. 6, p. 3761, 1985, https://doi.org/10.1103/physreva.31.3761.Search in Google Scholar PubMed
[31] C. Ciuti and I. Carusotto, “Input-output theory of cavities in the ultrastrong coupling regime: the case of time-independent cavity parameters,” Phys. Rev. A, vol. 74, no. 3, 2006, Art. no. 033811, https://doi.org/10.1103/physreva.74.033811.Search in Google Scholar
[32] N. Makri, “The linear response approximation and its lowest order corrections: an influence functional approach,” J. Phys. Chem. B, vol. 103, no. 15, p. 2823, 1999, https://doi.org/10.1021/jp9847540.Search in Google Scholar
[33] J. Ma and J. Cao, “Förster resonance energy transfer, absorption and emission spectra in multichromophoric systems. I. Full cumulant expansions and system-bath entanglement,” J. Chem. Phys., vol. 142, 2015, no. 9, Art. no. 094106.10.1063/1.4908599Search in Google Scholar PubMed
[34] J. Ma, J. Moix, and J. Cao, “Förster resonance energy transfer, absorption and emission spectra in multichromophoric systems. II. Hybrid cumulant expansion,” J. Chem. Phys., vol. 142, no. 9, 2015, Art. no. 094107.10.1063/1.4908600Search in Google Scholar PubMed
[35] J. M. Moix, J. Ma, and J. Cao, “Förster resonance energy transfer, absorption and emission spectra in multichromophoric systems. III. Exact stochastic path integral evaluation,” J. Chem. Phys., vol. 142, no. 9, 2015, Art. no. 094108.10.1063/1.4908601Search in Google Scholar PubMed
[36] G. Engelhardt and J. Cao, “Polariton localization and dispersion properties of disordered quantum emitters in multimode microcavities,” Phys. Rev. Lett., vol. 130, no. 21, 2023, Art. no. 213602, https://doi.org/10.1103/physrevlett.130.213602.Search in Google Scholar
[37] J. B. Pérez-Sánchez, A. Koner, N. P. Stern, and J. Yuen-Zhou, “Simulating molecular polaritons in the collective regime using few-molecule models,” Proc. Natl. Acad. Sci. U. S. A., vol. 120, no. 15, 2023, Art. no. e2219223120, https://doi.org/10.1073/pnas.2219223120. Search in Google Scholar PubMed PubMed Central
[38] J. Yuen-Zhou and A. Koner, Linear response of Molecular Polaritons, 2023, arXiv:2310.15424 [quant-ph].10.1063/5.0183683Search in Google Scholar PubMed
[39] J. Roden, A. Eisfeld, W. Wolff, and W. T. Strunz, “Influence of complex exciton-phonon coupling on optical absorption and energy transfer of quantum aggregates,” Phys. Rev. Lett., vol. 103, no. 5, 2009, Art. no. 058301, https://doi.org/10.1103/physrevlett.103.058301.Search in Google Scholar
[40] J. Roden, W. T. Strunz, and A. Eisfeld, “Non-Markovian Quantum State Diffusion for absorption spectra of molecular aggregates,” J. Chem. Phys., vol. 134, no. 3, 2011, Art. no. 034902, https://doi.org/10.1063/1.3512979.Search in Google Scholar PubMed
[41] G. Ritschel, D. Suess, S. Möbius, W. T. Strunz, and A. Eisfeld, “Non-Markovian Quantum State Diffusion for temperature-dependent linear spectra of light harvesting aggregates,” J. Chem. Phys., vol. 142, no. 3, 2015, Art. no. 034115.10.1063/1.4905327Search in Google Scholar PubMed
[42] J. A. Ćwik, P. Kirton, S. De Liberato, and J. Keeling, “Excitonic spectral features in strongly coupled organic polaritons,” Phys. Rev. A, vol. 93, no. 3, 2016, Art. no. 033840, https://doi.org/10.1103/physreva.93.033840.Search in Google Scholar
[43] K. S. Daskalakis, A. I. Väkeväinen, J.-P. Martikainen, T. K. Hakala, and P. Törmä, “Ultrafast pulse generation in an organic nanoparticle-array laser,” Nano Lett., vol. 18, no. 4, p. 2658, 2018, https://doi.org/10.1021/acs.nanolett.8b00531.Search in Google Scholar PubMed PubMed Central
[44] E. Palo, et al.., “Developing solution-processed distributed bragg reflectors for microcavity polariton applications,” J. Phys. Chem. C, vol. 127, no. 29, 2023, Art. no. 14255, https://doi.org/10.1021/acs.jpcc.3c01457.Search in Google Scholar PubMed PubMed Central
[45] A. Kavokin, J. Baumberg, G. Malpuech, and F. Laussy, Microcavities, Series on Semiconductor Science and Technology, Oxford, England, OUP Oxford, 2007.10.1093/acprof:oso/9780199228942.001.0001Search in Google Scholar
[46] G. Panzarini, et al.., “Cavity-polariton dispersion and polarization splitting in single and coupled semiconductor microcavities,” Phys. Solid State, vol. 41, no. 8, p. 1223, 1999, https://doi.org/10.1134/1.1130973.Search in Google Scholar
[47] V. May and O. Kühn, Charge and Energy Transfer Dynamics in Molecular Systems, Hoboken, New Jersey, U.S., Wiley, 2011.10.1002/9783527633791Search in Google Scholar
[48] H. Haug and S. Koch, Quantum Theory of the Optical and Electronic Properties of Semiconductors, G - Reference,Information and Interdisciplinary Subjects Series, Singapore, World Scientific, 2004.10.1142/5394Search in Google Scholar
[49] C. Klingshirn, Semiconductor Optics, Springer Study Edition, Berlin, Germany, Springer Berlin Heidelberg, 1997.Search in Google Scholar
[50] L. Diósi, N. Gisin, and W. T. Strunz, “Non-markovian quantum state diffusion,” Phys. Rev. A, vol. 58, no. 3, p. 1699, 1998, https://doi.org/10.1103/physreva.58.1699.Search in Google Scholar
[51] R. Hartmann and W. T. Strunz, “Exact open quantum system dynamics using the hierarchy of pure states (hops),” J. Chem. Theory Comput., vol. 13, no. 12, p. 5834, 2017, https://doi.org/10.1021/acs.jctc.7b00751.Search in Google Scholar PubMed
[52] N. Megier, W. T. Strunz, C. Viviescas, and K. Luoma, “Parametrization and optimization of Gaussian non-markovian unravelings for open quantum dynamics,” Phys. Rev. Lett., vol. 120, no. 15, 2018, Art. no. 150402, https://doi.org/10.1103/physrevlett.120.150402.Search in Google Scholar
[53] V. Link, K. Luoma, and W. T. Strunz, “Non-markovian quantum state diffusion for spin environments,” New J. Phys., vol. 25, no. 9, 2023, Art. no. 093006, https://doi.org/10.1088/1367-2630/aceff3.Search in Google Scholar
[54] P. Goetsch, R. Graham, and F. Haake, “Microscopic foundation of a finite-temperature stochastic Schrödinger equation,” Quantum Semiclassical Opt. J. Eur. Opt. Soc. Part B, vol. 8, no. 1, p. 157, 1996, https://doi.org/10.1088/1355-5111/8/1/012.Search in Google Scholar
[55] D. Suess, A. Eisfeld, and W. T. Strunz, “Hierarchy of stochastic pure states for open quantum system dynamics,” Phys. Rev. Lett., vol. 113, no. 15, 2014, Art. no. 150403, https://doi.org/10.1103/physrevlett.113.150403.Search in Google Scholar
[56] T. Yu, L. Diósi, N. Gisin, and W. T. Strunz, “Non-markovian quantum-state diffusion: perturbation approach,” Phys. Rev. A, vol. 60, no. 1, p. 91, 1999, https://doi.org/10.1103/physreva.60.91.Search in Google Scholar
[57] S. J. Byrnes, Multilayer Optical Calculations, 2020, arXiv:1603.02720 [physics.comp-ph].Search in Google Scholar
[58] S. Toffanin, et al.., “Molecular host-guest energy-transfer system with an ultralow amplified spontaneous emission threshold employing an ambipolar semiconducting host matrix,” J. Phys. Chem. B, vol. 114, no. 1, p. 120, 2010, https://doi.org/10.1021/jp909003n.Search in Google Scholar PubMed
[59] P. Michetti and G. C. La Rocca, “Polariton states in disordered organic microcavities,” Phys. Rev. B, vol. 71, no. 11, 2005, Art. no. 115320, https://doi.org/10.1103/physrevb.71.115320.Search in Google Scholar
[60] M. Wurdack, et al.., “Motional narrowing, ballistic transport, and trapping of room-temperature exciton polaritons in an atomically-thin semiconductor,” Nat. Commun., vol. 12, no. 1, p. 5366, 2021, https://doi.org/10.1038/s41467-021-25656-7.Search in Google Scholar PubMed PubMed Central
[61] R. Pandya, et al.., “Tuning the coherent propagation of organic exciton-polaritons through dark state delocalization,” Adv. Sci., vol. 9, no. 18, 2022, Art. no. 2105569, https://doi.org/10.1002/advs.202105569.Search in Google Scholar PubMed PubMed Central
[62] D. M. Whittaker, et al.., “Motional narrowing in semiconductor microcavities,” Phys. Rev. Lett., vol. 77, no. 23, p. 4792, 1996, https://doi.org/10.1103/physrevlett.77.4792.Search in Google Scholar
[63] A. V. Kavokin, et al.. “Motional narrowing of in homogeneously broadened excitons in a semiconductor microcavity: semiclassical treatment,” Phys. Rev. B, vol. 57, no. 7, p. 3757, 1998, https://doi.org/10.1103/physrevb.57.3757.Search in Google Scholar
© 2024 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Editorial
- Strong Coupling of Organic Molecules 2023 (SCOM23)
- Perspective
- Strong coupling of metamaterials with cavity photons: toward non-Hermitian optics
- Research Articles
- Strong coupling in molecular systems: a simple predictor employing routine optical measurements
- Extracting accurate light–matter couplings from disordered polaritons
- Linear optical properties of organic microcavity polaritons with non-Markovian quantum state diffusion
- Non-Hermitian polariton–photon coupling in a perovskite open microcavity
- Realization of ultrastrong coupling between LSPR and Fabry–Pérot mode via self-assembly of Au-NPs on p-NiO/Au film
- Self-hybridisation between interband transitions and Mie modes in dielectric nanoparticles
- Probing the anharmonicity of vibrational polaritons with double-quantum two-dimensional infrared spectroscopy
- Enhancement of the internal quantum efficiency in strongly coupled P3HT-C60 organic photovoltaic cells using Fabry–Perot cavities with varied cavity confinement
- Active control of polariton-enabled long-range energy transfer
- Coherent transient exciton transport in disordered polaritonic wires
- Identifying the origin of delayed electroluminescence in a polariton organic light-emitting diode
- Extracting kinetic information from short-time trajectories: relaxation and disorder of lossy cavity polaritons
- Exploring the impact of vibrational cavity coupling strength on ultrafast CN + c-C6H12 reaction dynamics
- Resonance theory of vibrational polariton chemistry at the normal incidence
- Investigating the collective nature of cavity-modified chemical kinetics under vibrational strong coupling
- Thermalization rate of polaritons in strongly-coupled molecular systems
- Room temperature polaritonic soft-spin XY Hamiltonian in organic–inorganic halide perovskites
- Electrical polarization switching of perovskite polariton laser
- A mixed perturbative-nonperturbative treatment for strong light-matter interactions
- Few-emitter lasing in single ultra-small nanocavities
- Letters
- Photochemical initiation of polariton-mediated exciton propagation
- Deciphering between enhanced light emission and absorption in multi-mode porphyrin cavity polariton samples
Articles in the same Issue
- Frontmatter
- Editorial
- Strong Coupling of Organic Molecules 2023 (SCOM23)
- Perspective
- Strong coupling of metamaterials with cavity photons: toward non-Hermitian optics
- Research Articles
- Strong coupling in molecular systems: a simple predictor employing routine optical measurements
- Extracting accurate light–matter couplings from disordered polaritons
- Linear optical properties of organic microcavity polaritons with non-Markovian quantum state diffusion
- Non-Hermitian polariton–photon coupling in a perovskite open microcavity
- Realization of ultrastrong coupling between LSPR and Fabry–Pérot mode via self-assembly of Au-NPs on p-NiO/Au film
- Self-hybridisation between interband transitions and Mie modes in dielectric nanoparticles
- Probing the anharmonicity of vibrational polaritons with double-quantum two-dimensional infrared spectroscopy
- Enhancement of the internal quantum efficiency in strongly coupled P3HT-C60 organic photovoltaic cells using Fabry–Perot cavities with varied cavity confinement
- Active control of polariton-enabled long-range energy transfer
- Coherent transient exciton transport in disordered polaritonic wires
- Identifying the origin of delayed electroluminescence in a polariton organic light-emitting diode
- Extracting kinetic information from short-time trajectories: relaxation and disorder of lossy cavity polaritons
- Exploring the impact of vibrational cavity coupling strength on ultrafast CN + c-C6H12 reaction dynamics
- Resonance theory of vibrational polariton chemistry at the normal incidence
- Investigating the collective nature of cavity-modified chemical kinetics under vibrational strong coupling
- Thermalization rate of polaritons in strongly-coupled molecular systems
- Room temperature polaritonic soft-spin XY Hamiltonian in organic–inorganic halide perovskites
- Electrical polarization switching of perovskite polariton laser
- A mixed perturbative-nonperturbative treatment for strong light-matter interactions
- Few-emitter lasing in single ultra-small nanocavities
- Letters
- Photochemical initiation of polariton-mediated exciton propagation
- Deciphering between enhanced light emission and absorption in multi-mode porphyrin cavity polariton samples