Abstract
We present a theory that explains the resonance effect of the vibrational strong coupling (VSC) modified reaction rate constant at the normal incidence of a Fabry–Pérot (FP) cavity. This analytic theory is based on a mechanistic hypothesis that cavity modes promote the transition from the ground state to the vibrational excited state of the reactant, which is the rate-limiting step of the reaction. This mechanism for a single molecule coupled to a single-mode cavity has been confirmed by numerically exact simulations in our recent work in [J. Chem. Phys. 159, 084104 (2023)]. Using Fermi’s golden rule (FGR), we formulate this rate constant for many molecules coupled to many cavity modes inside a FP microcavity. The theory provides a possible explanation for the resonance condition of the observed VSC effect and a plausible explanation of why only at the normal incident angle there is the resonance effect, whereas, for an oblique incidence, there is no apparent VSC effect for the rate constant even though both cases generate Rabi splitting and forming polariton states. On the other hand, the current theory cannot explain the collective effect when a large number of molecules are collectively coupled to the cavity, and future work is required to build a complete microscopic theory to explain all observed phenomena in VSC.
1 Introduction
Recent experiments [1], [2], [3], [4], [5], [6] have demonstrated that chemical reaction rate constants can be suppressed [1], [2], [3], [4], [7], [8], [9] or enhanced [5], [6], [10] by resonantly coupling molecular vibrations to quantized radiation modes inside a Fabry–Pérot (FP) microcavity [11], [12], [13]. This effect has the potential to selectively slow down competing reactions [3] or speed up a target reaction, thus achieving mode selectivity and offering a paradigm shift in chemistry. Despite extensive theoretical efforts [8], [14], [15], [16], [17], [18], [19], [20], [21], [22], [23], [24], [25], [26], [27], [28], [29], [30], [31], [32], [33], [34], [35], [36], [37], [38], [39], [40], [41], the fundamental mechanism and theoretical understanding of the cavity-modified ground-state chemical kinetics remain elusive [14], [42], [43], [44]. To the best of our knowledge, there is no unified theory that can explain all of the observed phenomena in the vibrational strong coupling (VSC) experiments [14], including (1) The resonance effect, which happens when the cavity frequency matches the bond vibrational frequency, ω
c = ω
0, but also only happens when the in-plane photon momentum is k
‖ = 0 (the normal incidence), (2) The collective effect [1], [4], [5] which is the increase in the magnitude of VSC modification when increasing the number of molecules N (or concentration
We aim to develop a microscopic theory to explain these observed VSC effects, especially focusing on understanding the resonance effect under normal incidence. Experimentally, only the resonance at normal incidence (k ‖ = 0) gives rise to VSC effects on the rate constant, while a red-detuned cavity that has a light–matter resonance at k ‖ > 0 (oblique incidence) does not give any VSC effect. This observation strongly suggests that forming Rabi splitting is not a sufficient condition for achieving the VSC-modified rate effect. Despite recent theoretical progress [18], [45], [46], the resonant condition under normal incidence remains an unresolved question.
In this work, we generalized our recently developed analytic Fermi’s golden rule (FGR) rate theory of VSC in Ref. [47] by incorporating many molecules and many cavity modes for both 1D FP cavities (with only 1D in the in-plane direction) and 2D FP cavities (with 2D in the in-plane direction) cases. In particular, we evaluated the photonic mode density of states (DOS) inside a 1D FP cavity and found that it gives rise to a van-Hove-type singularity at k ‖ = 0; for a 2D FP cavity, it is found that due to the cavity modes with k ‖ > 0 propagating outside a given cavity mode extent area, the modified photon mode DOS still remains dominant around the bottom of the dispersion band where k ‖ = 0, which are the keys to account for the normal incidence condition of the VSC-modified chemical reaction rate constant. The current theory provides a possible explanation of the resonance condition for the observed VSC effect and provides a microscopic understanding of why only at the normal incident angle there is a resonance effect.
2 Model system
Let us consider N identical molecules coupled to many radiation modes inside a FP cavity,
where

Top: Schematic illustration of the normal incidence condition for VSC-modified reactions. Bottom: (a) Schematic illustration of the dispersion relations of the cavity (red dashed line), the vibrational energy (gray dashed line), and hybrid polariton states (solid lines). (b) Schematic plot of reaction rate modification as a function of the cavity frequency ω c.
The photonic wavevector k (also referred to as the field propagation direction) has two components, one perpendicular to the cavity mirror k ⊥, and the other coplanar with the cavity k ‖. The FP cavity has the following dispersion relation,
where c is the speed of light in vacuum, n c is the refractive index of the cavity, c/n c is the speed of the light inside the cavity, and θ is commonly referred to as the incident angle (where tanθ = k ‖/k ⊥), which is the angle of the photonic mode wavevector k relative to the norm direction of the mirrors (see the top panel of Figure 1 for a schematic illustration). In most of the VSC experiments, n c ≈ 1.5 for the solution used inside the microcavity. Because n c ≈ 1, it will not influence the order of the magnitude of our discussion. For simplicity, we explicitly drop n c throughout this paper. Later, whenever we write c in an expression we should replace it with c/n c, in principle. When k ‖ = 0 (or θ = 0), the photon frequency is
The cavity frequency ω
k
in Eq. (1) is associated with the wavevector k, according to Eq. (2). Furthermore,
where ϵ
0 is the effective permittivity inside the cavity and
Figure 1(a) presents a schematic illustration of the cavity dispersion relation in Eq. (2) (red dashed line). The molecular excitation dispersion (black dashed line) is insensitive to the incident angle and is a straight line, with energy ℏω 0 (see Eq. (7)). These two dispersions hybridize due to the light–matter interactions, generating polariton dispersions (the upper and lower branches with solid curves) with the color coding indicating the character of the states, with purely photonic (red), purely vibrational (black), and hybridized (yellow to orange). Figure 1(b) presents a schematic illustration of the typical cavity detuning dependence of the rate constant modifications, with the highest intensity of the modification arising at the frequency when ω c = ω 0 (resonance condition at the normal incidence).
In this paper, we consider a reaction using a thermal barrier crossing model. Figure 2(a) presents the first few vibrational states of the double well model, where |ν
L⟩ denotes the vibrational ground state of the reactant (left well),

Potential energy surface for the reaction model. The red arrows represent the thermal activation process from the vibrational ground state, |ν
L⟩, to the vibrationally excited state,
Consider a simplified reaction mechanism outside the cavity as
Considering many molecules, we focus on the single excitation subspace. This includes the ground state |G⟩ and N singly excited states |ν j ⟩ (where j ∈ [1, N] labels the molecules), defined as
The vibrational transition dipole matrix element is
which is identical for all molecules j. When measuring the absorption spectra of the molecule, the optical response shows a peak at the quantum vibrational frequency
In the singly excited manifold, the light–matter coupling term,
where
The formation of Rabi splitting/polariton states comes from a collective phenomenon, resulting in the well-known dependence of
3 Theoretical results
3.1 Analytic rate constant expression
To provide a microscopic mechanism of VSC-modified reactions, we hypothesize that the cavity modes enhance the transition from ground states to a vibrationally excited state manifold of the reactant, leading to an enhancement of the steady-state population of both the delocalized states on the reactant side and the excited states manifold on the product side (right well)
among which k
1 ≪ k
2, k
3. Note that in the current work, we only consider the single excitation subspace (where one particular vibration is excited). In real experiments, many molecules could be simultaneously excited [13], with a number
When the molecular system is originally in the Kramers low friction regime (before the Kramers turnover [53], [54], or the so-called energy diffusion-limited regime), the cavity enhancement of the rate constant k
1 will occur [30], [31], [32], [34], [35], [39]. This has been extensively discussed in recent theoretical work [39], [47]. If we explicitly assume that k
1 ≪ k
2, k
3, then
where k 0 is the chemical reaction rate constant outside the cavity, and k VSC accounts for the pure cavity-induced effect. As this is a thermally activated reaction, there already exist some excited-state populations and transitions outside the cavity, which k 0 accounts for. Note that Eq. (10) assumes that the pure cavity effect k VSC can be added with k 0, which is a fundamental assumption in the current theory.
To quantitatively express k
VSC, we analyze the overall effect of the cavity and the photon loss environment on molecular systems by performing a normal mode transformation [55], [56], [57] to the Hamiltonian in Eq. (1) and obtaining an effective Hamiltonian, where now the cavity modes
where
where τ c is the cavity lifetime. Detailed derivation is provided in Supplementary Material, Section IV.
The rate constant change k
VSC in Eq. (10) originates from a purely cavity-induced effect, which promotes the transition from |G⟩ to the singly excited states manifold {|ν
j
⟩}. Note that this transition is mediated by the cavity operators
where D denotes the dimension of the in-plane direction in a FP cavity. The collective Jaynes-Cummings-type [50] coupling strength
and the 1/N factor accounts for the normalized rate constant per molecule. Furthermore,
is the Bose–Einstein distribution function, where β = 1/(k
B
T) with k
B as the Boltzmann constant and T as the temperature. For the typical parameters in VSC experiments, ω
0 ≈ 1200 cm−1 and room temperature 1/β = k
B
T ≈ 200 cm−1, such that βℏω
0 ≫ 1. Finally,
and
Under the continuous k
‖ limit, one can replace the sum in Eq. (13) with an integral as
where
Note that when all molecules are aligned with the cavity field polarization direction, such that cos φ
j
= 1,
For 2D FP cavities, similarly, one has (cf. Eq. (13))
where
whereas the standard 2D-DOS is defined as
Note that
We further define the accumulated spectral function
and
3.2 The resonance effect at the normal incidence
Next, we work to provide an analytic expression of
For the one-dimensional FP cavity [46], if we ignore the influence of cavity loss (
where Θ(ω − ω c) is the Heaviside step function. Details of the derivations are provided in the Supplementary Material, Section VI. The DOS, g 1D(ω), in Eq. (24) has a singularity at ω = ω c, which is known as (the first type of) the van-Hove-type singularity [58]. Such a concentrated peak in g 1D(ω) at ω = ω c has been numerically observed in Figure 1 of Ref. [46]. We will turn to the case of including the effects of photon propagation in the in-plane direction later in this section.
By using Eq. (24), we have the spectral function
where ω
m
→ ∞ is the cutoff frequency. The integral in Eq. (25) gives a finite value despite the singularity in g
1D(ω), because only the contribution from ω = ω
c survives. At the same time,
We have also numerically evaluated Eq. (25) and compared it with Eq. (26) for the VSC rates, presented in Figure S2 of the Supplementary Material, which shows a nearly identical behavior. The above theoretical results also suggest that for a 1D cavity, the commonly used single mode approximation [22], [27], [39] is indeed valid, because only the mode of frequency ω
c survives. Using the expression of
Figure 3 presents the cavity dispersion relation of ω
k
(θ) (see Eq. (2)) in panels (a) and (d), the 1D DOS g
1D(ω) (see Eq. (24)) in panels (b) and (e), and the 1D accumulated spectral function

Dispersion relation, DOS, and the accumulated spectral function
This analysis also provides a possible explanation for the resonance effect at normal incidence (k
‖ = 0) for a 1D FP cavity. In Eq. (26), it is clear that the peak of this function is located at ω
c = ω
0 for k
‖ = 0. Thus, the VSC-modified rate constant occurs only when ω
c = ω
0. This is because there is a van-Hove-type singularity [58] in the 1D DOS, g
1D(ω), which manifests itself as the
However, directly extending this simple consideration for the DOS cannot explain the normal incidence condition for a 2D FP cavity (even when only considering the TE polarization direction). This is because the 2D DOS g 2D(ω) does not have any singularity. Specifically, the DOS for the photonic modes inside a 2D FP cavity is expressed as
where Section VI of the Supplementary Material contain more details of this derivation.
For the 2D cavity case, one needs to consider beyond the simple DOS argument. Note that the photon loss associated with the lifetime τ
c only considers the loss in the k
⊥ direction. What we have not explicitly considered before was the photon traveling outside a mode area along the k
‖ direction. Let
which is propotional to
An estimation for
Note that the term
where Γ01 is the rate for the |0
k
⟩ → |1
k
⟩ photonic Fock states transition due to thermal excitation, and Γ10 = 1/τ
c is the cavity loss rate along the k
⊥ direction (associated with |1
k
⟩ → |0
k
⟩), which was assumed to be identical for all k modes. Note that all of the above-mentioned excitation and decay processes are related to the thermally activated radiation (thermal photon), and not related to the pumping with an external radiation field. To account for the additional effect of photon propagating outside a given area associated with a specific mode ω
k
, we modify the detailed balance relation (in Eq. (29)) by replacing the original
where τ
‖ (defined in Eq. (28)) is k
‖-dependent. This can also be viewed as putting a
Using
where we used
This
Figure 4 presents the cavity dispersion relation of ω
k
(θ) (see Eq. (2)) in panels (a) and (d), the weighting factor

Same as Figure 3, but with a 2D FP cavity. (a) Same as Figure 3(a). (b) The weighting factor
Figure 4(c) presents the behavior of the accumulated spectral function
With the above analysis, we have theoretically justified why the VSC-modified chemical kinetics only occurs at the normal incidence when ω
c = ω
0 for a 2D FP cavity, which agrees with experimental observations [1], [11], [12], [13]. This is because even though there is no singularity in g
2D(ω), the photons propagating outside the mode area along the k
‖ direction force the 2D cavity spectra function
3.3 No apparent collective effect
For our discussion on collectivity, we begin by considering the FGR expression in Eq. (13). For simplicity, we just focus on the 1D cavity case, since for 2D cavity there is no apparent collective effect either. If all the molecules’ dipoles are perfectly aligned with the cavity field polarization direction, then cosφ
j
= 1 for all molecules, j, and
where we have explicitly approximated
where
However, for the current theory in Eq. (34), the overall rate constant would not explicitly depend on N (Eq. (33)), meaning that only for the small N and strong coupling between molecules and the cavity mode there will be an appreciation amount of the cavity-modified effect. This is in contrast to the experimental observation of the collective effect and should be viewed as a major limitation of current theory. This limitation could be related to the fact that we have only considered the case of single excitation subspace in our theory, whereas in the experiments, a total of
When considering the disorder of the orientation between the dipole and the cavity field polarization direction, the FGR rate in Eq. (33) becomes
upon statistical averaging of dipole orientations. For fully isotropically distributed dipoles, ⟨ cos2 φ⟩ = 1/3.
3.4 Resonance behavior of k VSC
We want to demonstrate the numerical behavior of the current theory predicted by Eqs. (25) and (31). Because the current theory lacks the collective effect, we take the N = 1 limit and scale up the coupling strength between a single molecule and the cavity modes, as most previous work does [22], [23], [39]. This leads to the expression of (cf. Eq. (33))
under the single mode limit (or under the 1D cavity case, see Eqs. (25) and (26)). When further considering the presence of homogeneous or inhomogeneous broadening of the molecular system, the FGR expression will be a convolution between the original FGR expression, which does not considering the broadening for the ω 0 (for example, Eq. (33)), and a broadening function (assumed to be a Gaussian), expressed as follows [47]
where
where σ is the variance of the Gaussian.
As expected, the k VSC expression in Eq. (33) should contain several characteristic physical constants, including the speed of light c in ω c (see Eq. (3)) as it is related to light–matter interaction, Planck’s constant ℏ in g c (see Eq. (8)) as it should be a quantum theory, and Boltzmann’s constant k B in n(ω 0) as it is a thermally activated theory. We adopt a model system used in Ref. [39] to demonstrate the basic trend of k VSC predicted by the current theory. The schematic of the model is provided in Figure 2, whereas the details are provided in Supplementary Material, Section II.
To obtain the numerically exact rate constant for the same model, we use hierarchical equations of motion (HEOM) to simulate the population dynamics and obtain the VSC-modified rate constant, with the details provided in Section VII of the Supplementary Material. The HEOM simulation requires a linear system-bath coupling Hamiltonian. To this end, we follow the previous work [22], [39] and assume that the dipole operator is linear,
We use a similar range of η c as used in Ref. [39].
The forward rate constant from the HEOM simulation is obtained by evaluating [39], [47]
where
We report the numerical value of k/k 0 as a function of the cavity frequency ω c. For the rate constant predicted by FGR, we only report the value of k/k 0 = 1 + k VSC/k 0 (see Eq. (10)), where k VSC is evaluated using Eq. (36), and the variance defined in Eq. (37b) is estimated as σ = 30.74 cm−1 for the model parameters we used. See Supplementary Material, Section VII for details. And we directly use the numerical result of k 0 obtained from the HEOM simulation.
Figure 5 presents the numerical simulations of the rate constant from HEOM as well as the FGR results. Figure 5(a) presents k(t) for the resonant case when ω c = ω 0, at various light–matter coupling strengths η c. One can see the plateau value of k(t) increases as η c increases. Figure 5(b) presents the case where ω c < ω 0 where ω c = 1000 cm−1, and there is no apparent η c dependence of k(t), indicating that the coupling to the cavity has no effect. Figure 5(c) presents the value of k/k 0 from Eq. (36) (scaled by 0.4) as a function of ω c, depicted by the thick solid lines. A range of light–matter coupling strength η c is explored. The FGR expression shows the sharp resonance behavior of the VSC-modified rate profile at ω c = ω 0 = 1190 cm−1. A similar sharp resonance has been observed in VSC experiments [1], [5], [6] and quantum dynamics simulations [39]. Further, we provide the rate constant calculated from the numerically exact HEOM simulations (see Section VII of the Supplementary Material), depicted by the open circles with a thin guiding line. Although the analytic FGR expression overestimates the rate constant by about two times, the overall agreement between the FGR expression and the HEOM numerical results is remarkable, across the range of ω c and η c we explored.

Numerically exact simulation and the analytic FGR results of the rate constant. (a) The flux-side correlation functions computed by HEOM at resonance (with ω c = ω 0 = 1190 cm−1). (b) The flux-side correlation functions are calculated by HEOM but off-resonance (with ω c = 1000 cm−1). (c) The profile of the resonant VSC rate constant k/k 0 as a function of ω c with different light–matter coupling strengths, η c, obtained by FGR expressed in Eq. (36) (solid lines) and HEOM simulations (open circles with guiding thin lines), respectively. The cavity lifetime is set to be τ c = 200 fs.
Next, we explicitly consider going beyond the single-mode limit. For the 1D FP cavity,
where
Figure 6 presents the FGR rates under different η
c values. Figure 6(a) is the same as Figure 5(a), which corresponds to the single-mode case (or the many-mode case inside a 1D FP cavity). Figure 6(b) presents the estimated value of k/k
0 using k
VSC expression in Eq. (40), corresponding to the case of many modes inside a 2D FP cavity. Here, we choose

FGR rate profiles of k/k
0 as a function of ω
c. (a) FGR rate profiles for the single mode case (or the many modes case inside a 1D FP cavity) calculated using Eq. (36) (same as the solid lines in Figure 5(c)). (b) FGR rate profiles
4 Conclusions
We present a theory to explain the current VSC experiments, focusing on the origin of the resonance condition at normal incidence. The theory provides a possible explanation to the resonance condition for the observed VSC effect and of why the resonance effect occurs only at the normal incident angle. In particular, we find that the cavity-modified rate constant k
VSC can be expressed as the coupling strength multiplied by the accumulated spectral function
Under the normal incidence condition, k
VSC will peak at ω
c = ω
0. For the 1D cavity case,
On the other hand, the current theory cannot explain the observed collective effect, and only when a few molecules are strongly coupled to the cavity can the current theory predict the cavity modifications to the rate constant. This is the limitation of the current theory, and future work is needed to fully address these issues. However, the current work provides significant progress toward building the ultimate theory for understanding VSC effects. Future work will focus on developing a microscopic theory that can explain the collective effect.
Supplementary Material
See Supplementary Material for additional information on detailed derivations of the Hamiltonian; details of the molecular system; analysis of the Rabi splitting; the effective Hamiltonian and effective spectral density derived by applying harmonic analysis to classical equations of motion; derivation of the VSC-modified rate constant expression in Eq. (13) of the main text; DOS analysis for the 1D and 2D FP cavity; details of the quantum dynamics simulation results; effects of the
Funding source: National Science Foundation
Award Identifier / Grant number: CHE-2244683
Award Identifier / Grant number: DGE-1939268
Funding source: Research Corporation for Science Advancement
Award Identifier / Grant number: Cottrell Scholar Award
Acknowledgments
W.Y. appreciates the support of his Esther M. Conwell Graduate Fellowship at the University of Rochester. M.A.D.T. appreciates the support from the National Science Foundation Graduate Research Fellowship Program under Grant No. DGE-1939268. P.H. appreciates the support of the Cottrell Scholar Award (a program by the Research Corporation for Science Advancement). We appreciate valuable discussions with Eric Koessler, Arkajit Mandal, Raphael Ribeiro, Tao Li, Jino George, Blake Simpkins, and Igor Vurgaftman.
-
Research funding: This work was supported by the National Science Foundation Award under Grant No. CHE-2244683, National Science Foundation Graduate Research Fellowship Program (DGE-1939268), Research Corporation for Science Advancement: Cottrell Scholar Award.
-
Author contributions: W.Y., M.A.D.T., and P.H. designed the project; W.Y. and M.A.D.T. performed the research; W.Y., M.A.D.T., and P.H. wrote the paper.
-
Conflict of interest: Authors state no conflicts of interest.
-
Data availability: The data that support the findings of this work are available from the corresponding author under reasonable request.
References
[1] A. Thomas, et al.., “Ground-state chemical reactivity under vibrational coupling to the vacuum electromagnetic field,” Angew. Chem., Int. Ed., vol. 55, no. 38, pp. 11462–11466, 2016. https://doi.org/10.1002/anie.201605504.Search in Google Scholar PubMed PubMed Central
[2] R. M. A. Vergauwe, et al.., “Modification of enzyme activity by vibrational strong coupling of water,” Angew. Chem., Int. Ed., vol. 58, no. 43, pp. 15324–15328, 2019. https://doi.org/10.1002/anie.201908876.Search in Google Scholar PubMed PubMed Central
[3] A. Thomas, et al.., “Tilting a ground-state reactivity landscape by vibrational strong coupling,” Science, vol. 363, no. 6427, pp. 615–619, 2019. https://doi.org/10.1126/science.aau7742.Search in Google Scholar PubMed
[4] A. Thomas, et al.., “Ground state chemistry under vibrational strong coupling: dependence of thermodynamic parameters on the rabi splitting energy,” Nanophotonics, vol. 9, no. 2, pp. 249–255, 2020. https://doi.org/10.1515/nanoph-2019-0340.Search in Google Scholar
[5] J. Lather, P. Bhatt, A. Thomas, T. W. Ebbesen, and J. George, “Cavity catalysis by cooperative vibrational strong coupling of reactant and solvent molecules,” Angew. Chem., Int. Ed., vol. 58, no. 31, pp. 10635–10638, 2019. https://doi.org/10.1002/anie.201905407.Search in Google Scholar PubMed PubMed Central
[6] J. Lather, A. N. K. Thabassum, J. Singh, and J. George, “Cavity catalysis: modifying linear free-energy relationship under cooperative vibrational strong coupling,” Chem. Sci., vol. 13, no. 1, pp. 195–202, 2022. https://doi.org/10.1039/d1sc04707h.Search in Google Scholar PubMed PubMed Central
[7] K. Hirai, R. Takeda, J. A. Hutchison, and H. Uji-i, “Modulation of prins cyclization by vibrational strong coupling,” Angew. Chem., Int. Ed., vol. 59, no. 13, pp. 5332–5335, 2020. https://doi.org/10.1002/anie.201915632.Search in Google Scholar PubMed
[8] W. Ahn, J. F. Triana, F. Recabal, F. Herrera, and B. S. Simpkins, “Modification of ground state chemical reactivity via light-matter coherence in infrared cavities,” Science, vol. 380, no. 6650, pp. 1165–1168, 2023. https://doi.org/10.1126/science.ade7147.Search in Google Scholar PubMed
[9] K. Gu, Q. Si, N. Li, F. Gao, L. Wang, and F. Zhang, “Regulation of recombinase polymerase amplification by vibrational strong coupling of water,” ACS Photonics, vol. 10, no. 5, pp. 1633–1637, 2023. https://doi.org/10.1021/acsphotonics.3c00243.Search in Google Scholar
[10] J. Lather and J. George, “Improving enzyme catalytic efficiency by co-operative vibrational strong coupling of water,” J. Phys. Chem. Lett., vol. 12, no. 1, pp. 379–384, 2021. https://doi.org/10.1021/acs.jpclett.0c03003.Search in Google Scholar PubMed
[11] K. Hirai, J. A. Hutchison, and H. Uji-i, “Recent progress in vibropolaritonic chemistry,” ChemPlusChem, vol. 85, no. 9, pp. 1981–1988, 2020. https://doi.org/10.1002/cplu.202000411.Search in Google Scholar PubMed
[12] K. Nagarajan, A. Thomas, and T. W. Ebbesen, “Chemistry under vibrational strong coupling,” J. Am. Chem. Soc., vol. 143, no. 41, pp. 16877–16889, 2021. https://doi.org/10.1021/jacs.1c07420.Search in Google Scholar PubMed
[13] B. S. Simpkins, A. D. Dunkelberger, and I. Vurgaftman, “Control, modulation, and analytical descriptions of vibrational strong coupling,” Chem. Rev., vol. 123, no. 8, pp. 5020–5048, 2023. https://doi.org/10.1021/acs.chemrev.2c00774.Search in Google Scholar PubMed
[14] J. A. Campos-Gonzalez-Angulo, Y. R. Poh, M. Du, and J. Yuen-Zhou, “Swinging between shine and shadow: theoretical advances on thermally activated vibropolaritonic chemistry,” J. Chem. Phys., vol. 158, no. 23, p. 230901, 2023. https://doi.org/10.1063/5.0143253.Search in Google Scholar PubMed
[15] J. Galego, C. Climent, F. J. Garcia-Vidal, and J. Feist, “Cavity casimir-polder forces and their effects in ground-state chemical reactivity,” Phys. Rev. X, vol. 9, no. 2, p. 021057, 2019. https://doi.org/10.1103/physrevx.9.021057.Search in Google Scholar
[16] J. A. Campos-Gonzalez-Angulo, R. F. Ribeiro, and J. Yuen-Zhou, “Resonant catalysis of thermally activated chemical reactions with vibrational polaritons,” Nat. Commun., vol. 10, no. 1, p. 4685, 2019. https://doi.org/10.1038/s41467-019-12636-1.Search in Google Scholar PubMed PubMed Central
[17] A. Semenov and A. Nitzan, “Electron transfer in confined electromagnetic fields,” J. Chem. Phys., vol. 150, no. 17, p. 174122, 2019. https://doi.org/10.1063/1.5095940.Search in Google Scholar PubMed
[18] I. Vurgaftman, B. S. Simpkins, A. D. Dunkelberger, and J. C. Owrutsky, “Negligible effect of vibrational polaritons on chemical reaction rates via the density of states pathway,” J. Phys. Chem. Lett., vol. 11, no. 9, pp. 3557–3562, 2020. https://doi.org/10.1021/acs.jpclett.0c00841.Search in Google Scholar PubMed
[19] T. E. Li, A. Nitzan, and J. E. Subotnik, “On the origin of ground-state vacuum-field catalysis: equilibrium consideration,” J. Chem. Phys., vol. 152, no. 23, p. 234107, 2020. https://doi.org/10.1063/5.0006472.Search in Google Scholar PubMed
[20] V. P. Zhdanov, “Vacuum field in a cavity, light-mediated vibrational coupling, and chemical reactivity,” Chem. Phys., vol. 535, no. 1, p. 110767, 2020. https://doi.org/10.1016/j.chemphys.2020.110767.Search in Google Scholar
[21] J. A. Campos-Gonzalez-Angulo and J. Yuen-Zhou, “Polaritonic normal modes in transition state theory,” J. Chem. Phys., vol. 152, no. 16, p. 161101, 2020. https://doi.org/10.1063/5.0007547.Search in Google Scholar PubMed
[22] X. Li, A. Mandal, and P. Huo, “Cavity frequency-dependent theory for vibrational polariton chemistry,” Nat. Commun., vol. 12, no. 1, p. 1315, 2021. https://doi.org/10.1038/s41467-021-21610-9.Search in Google Scholar PubMed PubMed Central
[23] C. Schäfer, J. Flick, E. Ronca, P. Narang, and A. Rubio, “Shining light on the microscopic resonant mechanism responsible for cavity-mediated chemical reactivity,” Nat. Commun., vol. 13, no. 1, p. 7817, 2022. https://doi.org/10.1038/s41467-022-35363-6.Search in Google Scholar PubMed PubMed Central
[24] X. Li, A. Mandal, and P. Huo, “Theory of mode-selective chemistry through polaritonic vibrational strong coupling,” J. Phys. Chem. Lett., vol. 12, no. 29, pp. 6974–6982, 2021. https://doi.org/10.1021/acs.jpclett.1c01847.Search in Google Scholar PubMed
[25] T. E. Li, A. Nitzan, and J. E. Subotnik, “Collective vibrational strong coupling effects on molecular vibrational relaxation and energy transfer: numerical insights via cavity molecular dynamics simulations,” Angew. Chem., Int. Ed., vol. 60, no. 28, pp. 15533–15540, 2021. https://doi.org/10.1002/anie.202103920.Search in Google Scholar PubMed
[26] T. E. Li, A. Nitzan, and J. E. Subotnik, “Polariton relaxation under vibrational strong coupling: comparing cavity molecular dynamics simulations against Fermi's golden rule rate,” J. Chem. Phys., vol. 156, no. 13, p. 134106, 2022. https://doi.org/10.1063/5.0079784.Search in Google Scholar PubMed
[27] A. Mandal, X. Li, and P. Huo, “Theory of vibrational polariton chemistry in the collective coupling regime,” J. Chem. Phys., vol. 156, no. 1, p. 014101, 2022. https://doi.org/10.1063/5.0074106.Search in Google Scholar PubMed
[28] M. Du and J. Yuen-Zhou, “Catalysis by dark states in vibropolaritonic chemistry,” Phys. Rev. Lett., vol. 128, no. 9, p. 096001, 2022. https://doi.org/10.1103/physrevlett.128.096001.Search in Google Scholar
[29] J. P. Philbin, Y. Wang, P. Narang, and W. Dou, “Chemical reactions in imperfect cavities: enhancement, suppression, and resonance,” J. Phys. Chem. C, vol. 126, no. 35, pp. 14908–14913, 2022. https://doi.org/10.1021/acs.jpcc.2c04741.Search in Google Scholar
[30] D. S. Wang, T. Neuman, S. F. Yelin, and J. Flick, “Cavity-modified unimolecular dissociation reactions via intramolecular vibrational energy redistribution,” J. Phys. Chem. Lett., vol. 13, no. 15, pp. 3317–3324, 2022. https://doi.org/10.1021/acs.jpclett.2c00558.Search in Google Scholar PubMed PubMed Central
[31] D. S. Wang, J. Flick, and S. F. Yelin, “Chemical reactivity under collective vibrational strong coupling,” J. Chem. Phys., vol. 157, no. 22, p. 224304, 2022. https://doi.org/10.1063/5.0124551.Search in Google Scholar PubMed
[32] J. Sun and O. Vendrell, “Suppression and enhancement of thermal chemical rates in a cavity,” J. Phys. Chem. Lett., vol. 13, no. 20, pp. 4441–4446, 2022. https://doi.org/10.1021/acs.jpclett.2c00974.Search in Google Scholar PubMed
[33] E. W. Fischer, J. Anders, and P. Saalfrank, “Cavity-altered thermal isomerization rates and dynamical resonant localization in vibro-polaritonic chemistry,” J. Chem. Phys., vol. 156, no. 15, p. 154305, 2022. https://doi.org/10.1063/5.0076434.Search in Google Scholar PubMed
[34] L. P. Lindoy, A. Mandal, and D. R. Reichman, “Resonant cavity modification of ground-state chemical kinetics,” J. Phys. Chem. Lett., vol. 13, no. 28, pp. 6580–6586, 2022. https://doi.org/10.1021/acs.jpclett.2c01521.Search in Google Scholar PubMed
[35] S. Mondal, D. S. Wang, and S. Keshavamurthy, “Dissociation dynamics of a diatomic molecule in an optical cavity,” J. Chem. Phys., vol. 157, no. 24, p. 244109, 2022. https://doi.org/10.1063/5.0124085.Search in Google Scholar PubMed
[36] J. Cao, “Generalized resonance energy transfer theory: applications to vibrational energy flow in optical cavities,” J. Phys. Chem. Lett., vol. 13, no. 47, pp. 10943–10951, 2022. https://doi.org/10.1021/acs.jpclett.2c02707.Search in Google Scholar PubMed
[37] K. S. U. Kansanen and T. T. Heikkilä, “Cavity-induced bifurcation in classical rate theory,” 2023, arXiv, 10.48550/arXiv.2202.12182 (accessed 2023–05–12).Search in Google Scholar
[38] M. Du, Y. R. Poh, and J. Yuen-Zhou, “Vibropolaritonic reaction rates in the collective strong coupling regime: Pollak–Grabert–Hänggi theory,” J. Phys. Chem. C, vol. 127, no. 11, pp. 5230–5237, 2023. https://doi.org/10.1021/acs.jpcc.3c00122.Search in Google Scholar
[39] L. P. Lindoy, A. Mandal, and D. R. Reichman, “Quantum dynamical effects of vibrational strong coupling in chemical reactivity,” Nat. Commun., vol. 14, no. 1, p. 2733, 2023. https://doi.org/10.1038/s41467-023-38368-x.Search in Google Scholar PubMed PubMed Central
[40] M. C. Anderson, E. J. Woods, T. P. Fay, D. J. Wales, and D. T. Limmer, “On the mechanism of polaritonic rate suppression from quantum transition paths,” J. Phys. Chem. Lett., vol. 14, no. 30, pp. 6888–6894, 2023. https://doi.org/10.1021/acs.jpclett.3c01188.Search in Google Scholar PubMed
[41] M. R. Fiechter, J. E. Runeson, J. E. Lawrence, and J. O. Richardson, “How quantum is the resonance behavior in vibrational polariton chemistry?” J. Phys. Chem. Lett., vol. 14, no. 36, pp. 8261–8267, 2023. https://doi.org/10.1021/acs.jpclett.3c01154.Search in Google Scholar PubMed PubMed Central
[42] D. S. Wang and S. F. Yelin, “A roadmap toward the theory of vibrational polariton chemistry,” ACS Photonics, vol. 8, no. 10, pp. 2818–2826, 2021. https://doi.org/10.1021/acsphotonics.1c01028.Search in Google Scholar
[43] D. Sidler, M. Ruggenthaler, C. Schäfer, E. Ronca, and A. Rubio, “A perspective on ab initio modeling of polaritonic chemistry: the role of non-equilibrium effects and quantum collectivity,” J. Chem. Phys., vol. 156, no. 23, p. 230901, 2022. https://doi.org/10.1063/5.0094956.Search in Google Scholar PubMed
[44] A. Mandal, M. A. Taylor, B. M. Weight, E. R. Koessler, X. Li, and P. Huo, “Theoretical advances in polariton chemistry and molecular cavity quantum electrodynamics,” Chem. Rev., vol. 123, no. 16, pp. 9786–9879, 2023. https://doi.org/10.1021/acs.chemrev.2c00855.Search in Google Scholar PubMed PubMed Central
[45] I. Vurgaftman, B. S. Simpkins, A. D. Dunkelberger, and J. C. Owrutsky, “Comparative analysis of polaritons in bulk, dielectric slabs, and planar cavities with implications for cavity-modified reactivity,” J. Chem. Phys., vol. 156, no. 3, p. 034110, 2022. https://doi.org/10.1063/5.0078148.Search in Google Scholar PubMed
[46] R. F. Ribeiro, “Multimode polariton effects on molecular energy transport and spectral fluctuations,” Commun. Chem., vol. 5, no. 1, p. 48, 2022, https://doi.org/10.1038/s42004-022-00660-0.Search in Google Scholar PubMed PubMed Central
[47] W. Ying and P. Huo, “Resonance theory and quantum dynamics simulations of vibrational polariton chemistry,” J. Chem. Phys., vol. 159, no. 8, p. 084104, 2023. https://doi.org/10.1063/5.0159791.Search in Google Scholar PubMed
[48] J. J. Hopfield, “Theory of the contribution of excitons to the complex dielectric constant of crystals,” Phys. Rev., vol. 112, no. 5, pp. 1555–1567, 1958. https://doi.org/10.1103/physrev.112.1555.Search in Google Scholar
[49] M. Tavis and F. Cummings, “Exact solution for an n-molecule-radiation-field Hamiltonian,” Phys. Rev., vol. 170, no. 2, pp. 379–384, 1968. https://doi.org/10.1103/physrev.170.379.Search in Google Scholar
[50] E. Jaynes and F. Cummings, “Comparison of quantum and semiclassical radiation theories with application to the beam maser,” Proc. IEEE, vol. 18, no. 1, pp. 89–109, 1963. https://doi.org/10.1109/proc.1963.1664.Search in Google Scholar
[51] K. Kim, et al.., “Sensitive control of broad-area semiconductor lasers by cavity shape,” APL Photonics, vol. 7, no. 5, p. 056106, 2022. https://doi.org/10.1063/5.0087048.Search in Google Scholar
[52] J. del Pino, J. Feist, and F. J. Garcia-Vidal, “Quantum theory of collective strong coupling of molecular vibrations with a microcavity mode,” New J. Phys., vol. 17, no. 5, p. 053040, 2015. https://doi.org/10.1088/1367-2630/17/5/053040.Search in Google Scholar
[53] P. Hänggi, P. Talkner, and M. Borkovec, “Reaction-rate theory: fifty years after kramers,” Rev. Mod. Phys., vol. 62, no. 2, pp. 251–341, 1990. https://doi.org/10.1103/revmodphys.62.251.Search in Google Scholar
[54] E. Pollak, H. Grabert, and P. Hänggi, “Theory of activated rate processes for arbitrary frequency dependent friction: solution of the turnover problem,” J. Chem. Phys., vol. 91, no. 7, pp. 4073–4087, 1989. https://doi.org/10.1063/1.456837.Search in Google Scholar
[55] A. J. Leggett, “Quantum tunneling in the presence of an arbitrary linear dissipation mechanism,” Phys. Rev. B, vol. 30, no. 3, pp. 1208–1218, 1984. https://doi.org/10.1103/physrevb.30.1208.Search in Google Scholar
[56] A. Garg, J. N. Onuchic, and V. Ambegaokar, “Effect of friction on electron transfer in biomolecules,” J. Chem. Phys., vol. 83, no. 9, pp. 4491–4503, 1985. https://doi.org/10.1063/1.449017.Search in Google Scholar
[57] M. Thoss, H. Wang, and W. H. Miller, “Self-consistent hybrid approach for complex systems: application to the spin-boson model with debye spectral density,” J. Chem. Phys., vol. 115, no. 7, pp. 2991–3005, 2001. https://doi.org/10.1063/1.1385562.Search in Google Scholar
[58] L. V. Hove, “The occurrence of singularities in the elastic frequency distribution of a crystal,” Phys. Rev., vol. 89, no. 6, pp. 1189–1193, 1953. https://doi.org/10.1103/physrev.89.1189.Search in Google Scholar
[59] A. Shalabney, J. George, J. Hutchison, G. Pupillo, C. Genet, and T. W. Ebbesen, “Coherent coupling of molecular resonators with a microcavity mode,” Nat. Commun., vol. 6, no. 1, p. 5981, 2015. https://doi.org/10.1038/ncomms6981.Search in Google Scholar PubMed PubMed Central
[60] B. Xiang, et al.., “Intermolecular vibrational energy transfer enabled by microcavity strong light-matter coupling,” Science, vol. 368, no. 6491, pp. 665–667, 2020. https://doi.org/10.1126/science.aba3544.Search in Google Scholar PubMed
[61] M. A. C. Saller, Y. Lai, and E. Geva, “Cavity-modified Fermi’s golden rule rate constants: beyond the single mode approximation,” J. Chem. Phys., vol. 159, no. 15, p. 151105, 2023. https://doi.org/10.1063/5.0172265.Search in Google Scholar PubMed
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2023-0685).
© 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