This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Catalysis by Dark States in Vibropolaritonic Chemistry

Matthew Du, Joel Yuen-Zhou [email protected] Department of Chemistry and Biochemistry, University of California San Diego, La Jolla, California 92093, United States
Abstract

Collective strong coupling between a disordered ensemble of NN localized molecular vibrations and a resonant optical cavity mode gives rise to 2 polariton and N12N-1\gg 2 dark modes. Thus, experimental changes in thermally-activated reaction kinetics due to polariton formation appear entropically unlikely and remain a puzzle. Here we show that the overlooked dark modes, while parked at the same energy as bare molecular vibrations, are robustly delocalized across \sim2-3 molecules, yielding enhanced channels of vibrational cooling, concomitantly catalyzing a chemical reaction. As an illustration, we theoretically show a \approx50% increase in an electron transfer rate due to enhanced product stabilization. The reported effects can arise when the homogeneous linewidths of the dark modes are smaller than their energy spacings.

The past decade has seen much interest in the control of chemical phenomena via the strong coupling of matter to confined electromagnetic modes [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18]. An exciting prospect in this direction is vibropolaritonic chemistry, that is, the use of collective vibrational strong coupling (VSC) [19, 20, 21] to modify thermally-activated chemical reactivity without external pumping (e.g., laser excitation) [22]. While collective VSC involves a large number of molecules per photon mode, it has been observed to substantially alter the kinetics of organic substitution [23, 24], cycloaddition [25], hydrolysis [26], enzyme catalysis [27, 28], and crystallization [29], among other electronic ground-state chemical processes.

However, such modified reactivity under VSC is still not well understood. Studies [30, 31, 32, 33] show that the observed kinetics cannot be explained with transition state theory (TST) [34], the most commonly used framework to predict and interpret reaction rates. Breakdowns of TST, including recrossing the activation barrier [35], deviation from thermal equilibrium [32, 36], and quantum/nonadiabatic phenomena [37, 38, 32, 36, 39], have also been considered. The aforementioned works regard all NN vibrations coupled to the cavity mode to be identical. Under this assumption, VSC forms two polaritons and N1N-1 optically dark vibrational modes, where the latter remain unchanged from the cavity-free system. It follows that VSC-induced changes to thermally-activated reactivity must arise from the polaritons. In fact, one study from our group highlighted the molecular parameter space where polaritons dominate the kinetics [37] with respect to dark modes; however, this hypothesis has been questioned as entropically unlikely [40].

Disorder, despite its ubiquity in molecular systems, has often been ignored when modeling molecules under strong light-matter coupling. Only recently has it been shown that the strong coupling of disordered chromophores to an optical cavity mode can produce dark states which are delocalized on multiple molecules [41, 42] (hereafter, referred as semilocalized). This semilocalization is predicted to improve or even enable coherent energy transport [41, 43]. Other findings hint at adding sample impurities to help strong coupling modify local molecular properties [44].

In this Letter, we demonstrate that the VSC of a disordered molecular ensemble (Fig. 1) can significantly modify the kinetics of a thermally-activated chemical reaction. The altered reactivity is attributed to the semilocalized dark modes. The semilocalization affects the reaction rate by changing the efficiency with which a reactive mode dissipates energy.

Refer to caption
Figure 1: Schematic of the model setup. Inside an optical cavity, a cavity mode collectively interacts with an energetically disordered ensemble of molecular vibrations, each belonging to a separate molecule. When a molecule reacts, its cavity-coupled vibrational mode concomitantly experiences a displacement in equilibrium geometry. Depicted here are molecules that undergo intramolecular electron transfer, the reaction studied in this work.

Consider a disordered ensemble of NN molecular vibrations, respectively corresponding to NN independent molecules, inside an optical cavity (Fig. 1). The system is described by the Hamiltonian H=ωca0a0+i=1Nωiaiai+i=1Ngi(aia0+h.c.)H=\hbar\omega_{c}a_{0}^{\dagger}a_{0}+\hbar\sum_{i=1}^{N}\omega_{i}a_{i}^{\dagger}a_{i}+\hbar\sum_{i=1}^{N}g_{i}(a_{i}^{\dagger}a_{0}+\text{h.c.}). Vibrational mode ii is represented by annihilation operator aia_{i} and has frequency ωi=ω¯v+δωi\omega_{i}=\overline{\omega}_{v}+\delta\omega_{i}, where ω¯v\overline{\omega}_{v} is the mean vibrational frequency and δωi\delta\omega_{i} is the frequency offset of mode ii. Reflecting inhomogeneous broadening (static diagonal disorder), δωi\delta\omega_{i} is a normally distributed random variable with mean zero and standard deviation σv\sigma_{v}. The cavity mode is represented by annihilation operator a0a_{0}, has frequency ωc\omega_{c}, and couples to vibration ii with strength gig_{i}. For simplicity, we hereafter take gi=gg_{i}=g for all ii.

Using HH, we investigate the physicochemical properties of a disordered molecular system under VSC. Unless otherwise noted, calculations assume that the cavity is resonant with the average vibration, ωc=ω¯v\omega_{c}=\overline{\omega}_{v}, and couples to the vibrations with collective strength gN=8σvg\sqrt{N}=8\sigma_{v} (for all NN). Numerical values reported below are obtained by averaging over 5000 disorder realizations, i.e., sets {ωi}i=1N\{\omega_{i}\}_{i=1}^{N}. In plots versus the HH eigenfrequencies, each data point is an average over the HH eigenmodes—from all disorder realizations—whose frequency lies in the bin (Δω(l12),Δω(l+12)]\left(\Delta\omega(l-\frac{1}{2}),\Delta\omega(l+\frac{1}{2})\right] for Δω=101σv\Delta\omega=10^{-1}\sigma_{v} and ll\in\mathbb{Z}.

We first study the eigenmodes of HH. Formally, mode q=1,,N+1q=1,\dots,N+1 is represented by operator αq=i=0Ncqiai\alpha_{q}=\sum_{i=0}^{N}c_{qi}a_{i} and has frequency ωq\omega_{q}. Figs. 2(a) and 2(b) show the probability distribution and photon fraction (|cq0|2|c_{q0}|^{2}), respectively, of the eigenmodes with respect to eigenfrequency. The majority of modes form a broad distribution in frequency around ω¯v\overline{\omega}_{v} and are optically dark. A minority of modes are polaritons, which have frequency ω¯v±8σv\overline{\omega}_{v}\pm 8\sigma_{v}, photon fraction 0.5\approx 0.5, and a lineshape minimally affected by inhomogeneous broadening [45]. As NN rises, the eigenmodes become increasingly composed of dark modes, whose probability distribution approaches that of the bare vibrational modes [Fig. 2(a), gray dashed line].

Refer to caption
Figure 2: (a) Probability distribution and (b) photon fraction of the eigenmodes of HH. Values are plotted versus eigenfrequency ωq\omega_{q} for various NN. In (a), the gray dashed line is the probability density function [P(ωq)=(σv2π)1exp((ωqω¯v)2/(2σv2))P(\omega_{q})=(\sigma_{v}\sqrt{2\pi})^{-1}\exp\left(-(\omega_{q}-\overline{\omega}_{v})^{2}/(2\sigma_{v}^{2})\right), displayed in units of 1/Δω1/\Delta\omega] of the bare vibrational modes.

Next, we examine the delocalization of the dark modes. For the purpose of studying chemical reactions, it is useful to compute the molecular participation ratio (PR) [46, 42]. This measure, defined as

molecular PR=1/i=1N|cqii=1N|cqi|2|4\text{molecular PR}=1/\sum_{i=1}^{N}\left|\frac{c_{qi}}{\sqrt{\sum_{i=1}^{N}|c_{qi}|^{2}}}\right|^{4} (1)

and analogous to the usual PR [47], estimates the number of molecules over which eigenmode qq is delocalized. According to Fig. 3(a), the average dark mode has molecular PR \sim2-3, and this semilocalization persists as NN increases. These phenomena were first noted independently by Scholes [42] and Schachenmayer and co-workers [41]. For additional insight, we plot the squared overlap [|cqi|2|c_{qi}|^{2}, Fig. 3(b)] and frequency difference [|ωqi||\omega_{qi}|, where ωqi=ωqωi\omega_{qi}=\omega_{q}-\omega_{i}; Fig. 3(c)] between each dark mode and each bare vibrational mode. The typical dark mode has sizable overlap with the bare modes that are nearest to it in frequency. The frequency difference between the dark mode and any of its major constituents is O(σv/N)O(\sigma_{v}/N), i.e., negligible for large NN.

Refer to caption
Figure 3: (a) Average molecular PR of the dark modes as a function of NN. (b) Squared overlap |cqi|2|c_{qi}|^{2} (log10 scale) and (c) frequency difference |ωqi||\omega_{qi}| between each eigenmode qq and each bare vibrational mode ii. Each group of modes is ordered from low to high frequency, i.e., dark (polariton) modes have index q=2,,Nq=2,\dots,N (q=1,N+1q=1,N+1). The quantities in (b)-(c) are plotted for various NN.

We now explore how VSC influences the kinetics of a thermally-activated chemical reaction. Consider a reactive molecule under collective VSC (Fig. 1). The molecule undergoes nonadiabatic intramolecular electron transfer, and a cavity mode interacts collectively with a reactive vibrational mode and N1N-1 nonreactive vibrational modes. The model here considered is general enough that it should also be applicable to the case of “solvent-assisted VSC” [48, 49, 50, 51].

To model the reaction of one molecule in the ensemble, we employ the Hamiltonian Hrxn=H+X=R,P|XX|{EX+ωr[λX(ar+ar)+λX2]}+V+Hs(l)H_{\text{rxn}}=H+\sum_{X=R,P}|X\rangle\langle X|\{E_{X}+\hbar\omega_{r}[\lambda_{X}(a_{r}+a_{r}^{\dagger})+\lambda_{X}^{2}]\}+V+H_{s}^{(l)}. Note that vibrational and cavity modes are still described by HH. In writing HrxnH_{\text{rxn}}, we have changed the numerical index of the bare vibrational mode (i=1i=1) involved in the reaction to the letter rr (i.e., ω1ωr\omega_{1}\rightarrow\omega_{r}, a1ara_{1}\rightarrow a_{r}); hereafter, we refer to this mode as vrv_{r}. The electronic subspace consists of reactant |R|R\rangle and product |P|P\rangle states. Electronic state |X|X\rangle has energy EXE_{X} and couples to vrv_{r} with dimensionless strength λX\lambda_{X}. As a result of the vibronic coupling, vrv_{r} experiences a displacement in its equilibrium position upon electron transfer. The interaction between |R|R\rangle and |P|P\rangle is represented by V=JRP(|PR|+h.c.)V=J_{RP}(|P\rangle\langle R|+\text{h.c.}), where JRPJ_{RP} is the interaction strength. Through Hs(l)H_{s}^{(l)} [52], Hamiltonian HrxnH_{\text{rxn}} also accounts for low-frequency vibrational modes of the solvent that help mediate electron transfer. There is no direct coupling, though, of the cavity mode to the |R|P|R\rangle\rightleftharpoons|P\rangle electronic transitions, which we assume are dipole-forbidden.

Since we are considering a nonadiabatic reaction, we treat VV perturbatively and calculate rates of reactive transitions between the zeroth-order electronic-vibrational-cavity eigenstates of HrxnH_{\text{rxn}}. These states take the form |X,χ=|X|χ~(X)|X,\chi\rangle=|X\rangle\otimes|\tilde{\chi}_{(X)}\rangle. Belonging to the subspace of vibrational and cavity modes, |χ~(X)=(q=1N+1Dq(λXq))|χ|\tilde{\chi}_{(X)}\rangle=\left(\prod_{q=1}^{N+1}D_{q}^{\dagger}(\lambda_{Xq})\right)|\chi\rangle is a displaced Fock state with mq(χ)m_{q}^{(\chi)} excitations in HH eigenmode qq. The undisplaced Fock state |χ|\chi\rangle is an eigenstate of HH, and Dq(λ)=exp(λαqλαq)D_{q}(\lambda)=\exp(\lambda\alpha_{q}^{\dagger}-\lambda^{*}\alpha_{q}) is a displacement operator. Mode qq has equilibrium (dimensionless) position λXq=λXcqr(ωr/ωq)\lambda_{Xq}=\lambda_{X}c_{qr}(\omega_{r}/\omega_{q}) when the system is in electronic state |X|X\rangle. Returning to the electronic-vibrational-cavity state |X,χ|X,\chi\rangle, we can write its energy as E(X,χ)=EX+q=1N+1mq(χ)ωq+ΔXE_{(X,\chi)}=E_{X}+\sum_{q=1}^{N+1}m_{q}^{(\chi)}\hbar\omega_{q}+\Delta_{X}, where ΔX=λX2ωrq=1N+1|λXq|2ωq\Delta_{X}=\lambda_{X}^{2}\hbar\omega_{r}-\hbar\sum_{q=1}^{N+1}|\lambda_{Xq}|^{2}\omega_{q} is the difference in reorganization energy—namely, that due to vibronic coupling between vrv_{r} and |X|X\rangle—with and without VSC.

Following extensions [37, 38, 36] of Marcus-Levich-Jortner theory [53, 54, 55] to electron transfer under VSC, the rate of the reactive transition from |R,χ|R,\chi\rangle to |P,χ|P,\chi^{\prime}\rangle can be expressed as

k(R,χ)(P,χ)\displaystyle k_{(R,\chi)\rightarrow(P,\chi^{\prime})} =𝒜Fχ,χexp(βEaχ,χ),\displaystyle=\mathcal{A}F_{\chi,\chi^{\prime}}\exp\left(-\beta E_{a}^{\chi,\chi^{\prime}}\right), (2)

where 𝒜=πβ/λs|JRP|2/\mathcal{A}=\sqrt{\pi\beta/\lambda_{s}}|J_{RP}|^{2}/\hbar, λs\lambda_{s} is the reorganization energy associated with low-frequency solvent modes [52], and β\beta is the inverse temperature For this transition, the activation energy is Eaχ,χ=(E(P,χ)E(R,χ)+λs)2/(4λs)E_{a}^{\chi,\chi^{\prime}}=(E_{(P,\chi^{\prime})}-E_{(R,\chi)}+\lambda_{s})^{2}/(4\lambda_{s}). Through the Franck-Condon (FC) factor Fχ,χ=|χ~(R)|χ~(P)|2F_{\chi,\chi^{\prime}}=|\langle\tilde{\chi}_{(R)}|\tilde{\chi}^{\prime}_{(P)}\rangle|^{2}, the transition rate depends on the overlap between initial and final vibrational-cavity states |χ~(R)|\tilde{\chi}_{(R)}\rangle and |χ~(P)|\tilde{\chi}^{\prime}_{(P)}\rangle, respectively. It can be shown that the rate k(P,χ)(R,χ)k_{(P,\chi^{\prime})\rightarrow(R,\chi)} corresponding to the backward transition, |P,χ|R,χ|P,\chi^{\prime}\rangle\rightarrow|R,\chi\rangle, is related to Eq. (2) by detailed balance [56].

We specifically study a reaction where, in the absence of light-matter coupling, reactive transitions occur on the same timescale as internal thermalization (i.e., thermalization of states having the same electronic component). In such cases, internal thermal equilibrium is not maintained throughout the reaction, and the reaction rate (i.e., net rate of reactant depletion) may not be approximated by a thermal average of reactant-to-product transition rates. Instead, the reaction rate can also depend on, e.g., backward reactive transitions (from product to reactant) or vibrational relaxation.

With this in mind, we numerically simulate the bare (g=0g=0) and VSC reactions using a kinetic model [52, 36], which includes forward and backward reactive transitions [Eq. (2)], vibrational and cavity decay, and energy exchange among dark and polariton states [57, 58, 59]. The third set of processes results from vibrational dephasing interactions (i.e., homogeneous broadening) of the molecular system [57, 52]. The reaction parameters [52]—in particular ω¯v=2000\overline{\omega}_{v}=2000 cm-1, σv=10\sigma_{v}=10 cm-1, N32N\leq 32, gN=8σvg\sqrt{N}=8\sigma_{v} (for all NN), EPER=0.6ω¯vE_{P}-E_{R}=-0.6\overline{\omega}_{v}, and the chosen temperature values—are such that the population dynamics proceeds almost completely through states |X,χ|X,\chi\rangle with zero or one excitation in the vibrational-cavity modes. To reduce computational cost, the kinetic model includes only these states, which are denoted by χ=0\chi=0 and χ=1q\chi=1_{q}, respectively, where qq is an eigenmode of the vibrational-cavity subspace. The kinetic master equation [52] is numerically solved with the initial population being a thermal distribution of reactant states (|R,χ|R,\chi\rangle). Then, the apparent reaction rate is obtained by fitting the reactant population as a function of time [52]. Note that, for large enough NN (we estimate N>72N>72 for the chosen parameters), the energy spacing between dark modes becomes smaller than their decay linewidths; under this condition, our kinetic model is not valid [52]. Interestingly, such invalidation suggests that, within our model, VSC-modified chemistry might not occur in the weak light-matter coupling regime, where the polaritons and dark modes cannot be spectrally resolved.

Even though we run the full numerical simulations as explained above, we now introduce approximate models that shed conceptual intuition on the calculated kinetics. First, the bare reaction can be essentially captured by Fig. 4(a), described as follows. Starting from its vibrational ground state, the reactant converts to product mainly by a 010\rightarrow 1 vibronic transition, which excites the reactive mode and has rate kfk(R,0)(P,1r)k_{f}\equiv k_{(R,0)\rightarrow(P,1_{r})}, where

kf\displaystyle k_{f} =𝒜F0,1rexp(βEa),\displaystyle=\mathcal{A}F_{0,1_{r}}\exp\left(-\beta E_{a}\right), (3)

EaEa0,1rE_{a}\equiv E_{a}^{0,1_{r}}. The vibrationally hot product either reverts to the reactant at rate kbk(P,1r)(R,0)kfk_{b}\equiv k_{(P,1_{r})\rightarrow(R,0)}\gg k_{f}, where

kb\displaystyle k_{b} =kfexp[β(EP+ωrER)],\displaystyle=k_{f}\exp\left[\beta(E_{P}+\hbar\omega_{r}-E_{R})\right], (4)

or decays to its vibrational ground state at rate γkb\gamma\approx k_{b}. Once the product reaches its vibrational ground state, it effectively stops reacting due to the high reverse activation energy. This kinetic scheme leads to a bare reaction rate [52]

kbare(analytical)=kf(γγ+kb).k_{\text{bare}}^{(\text{analytical})}=k_{f}\left(\frac{\gamma}{\gamma+k_{b}}\right). (5)

Second, under VSC, the primary reaction pathway of the bare case is split into multiple pathways, each involving the (de)excitation of a dark or polariton eigenmode qq. For the VSC reaction channels, the forward and backward rates take the form

kf(q)\displaystyle k_{f}^{(q)} =𝒜F0,1qexp(βEa(q)),\displaystyle=\mathcal{A}F_{0,1_{q}}\exp\left(-\beta E_{a}^{(q)}\right), (6)
kb(q)\displaystyle k_{b}^{(q)} =kf(q)exp{β[EP+ωq+ΔP(ER+ΔR)]},\displaystyle=k_{f}^{(q)}\exp\left\{\beta\left[E_{P}+\hbar\omega_{q}+\Delta_{P}-(E_{R}+\Delta_{R})\right]\right\}, (7)

respectively, where kf(q)k(R,0)(P,1q)k_{f}^{(q)}\equiv k_{(R,0)\rightarrow(P,1_{q})}, kb(q)k(P,1q)(R,0)k_{b}^{(q)}\equiv k_{(P,1_{q})\rightarrow(R,0)}, and Ea(q)Ea0,1qE_{a}^{(q)}\equiv E_{a}^{0,1_{q}}. Now, consider the following argument, which holds strictly for large NN. As NN increases, the average bare mode becomes localized on dark modes that have essentially the same frequency as it [Figs. 3(b)-3(c)], and its overlap with the polariton modes vanishes [i.e., cqr1N0c_{qr}\propto\frac{1}{\sqrt{N}}\to 0 for q=1,N+1q=1,N+1; see Fig. 3(b)]. These observations suggest that cqr0c_{qr}\not\approx 0 only for modes qq that are dark and have frequency ωqωr\omega_{q}\approx\omega_{r}. It is then straightforward to show that Ea(q)|cqr0Ea\left.E_{a}^{(q)}\right|_{c_{qr}\not\approx 0}\approx E_{a}, F0,1q|cqr|2F0,1rF_{0,1_{q}}\approx|c_{qr}|^{2}F_{0,1_{r}}, and

kf/b(q)\displaystyle k_{f/b}^{(q)} |cqr|2kf/b.\displaystyle\approx|c_{qr}|^{2}k_{f/b}. (8)

Thus, VSC leads to reaction channels that have lower rates of reactive transitions, due to changes not in activation energies but in FC factors, which are smaller as a result of the semilocalization of dark modes. From Eq. (8), it is evident that the total forward rate is approximately that of the bare reaction (q=1N+1kf(q)kf\sum_{q=1}^{N+1}k_{f}^{(q)}\approx k_{f}, since q=1N+1|cqr|2=1\sum_{q=1}^{N+1}|c_{qr}|^{2}=1). However, once a forward reactive transition happens—and a dark mode is excited—the product either returns to the reactant at a reduced rate (kb(q)<kbk_{b}^{(q)}<k_{b}) or, due to the almost fully vibrational nature of the dark modes, vibrationally decays to its stable form (|P,0|P,0\rangle) at essentially the same bare rate (γ\gamma). In other words, VSC suppresses reverse reactive transitions by promoting the cooling of the reactive mode upon product formation. In analogy to Eq. (5), we determine an effective rate for the VSC reaction [52]:

kVSC(analytical)\displaystyle k_{\text{VSC}}^{(\text{analytical})} =kfγγ+|cqr|2kbdark modes q,\displaystyle=k_{f}\left\langle\frac{\gamma}{\gamma+|c_{qr}|^{2}k_{b}}\right\rangle_{\text{dark modes }q}, (9)

where dark modes q\langle\cdot\rangle_{\text{dark modes }q} is a weighted average over all dark modes qq, each with weight |cqr|2|c_{qr}|^{2}. Since |cqr|2<1|c_{qr}|^{2}<1 for all qq, then kVSC(analytical)>kbare(analytical)k_{\text{VSC}}^{(\text{analytical)}}>k_{\text{bare}}^{(\text{analytical})}. We emphasize that the major contributions to the average in Eq. (9) come from dark modes which are closest in frequency to the bare reactive mode (see above). Further enhancement of the VSC reaction, beyond that given by kVSC(analytical)k_{\text{VSC}}^{(\text{analytical})}, occurs via dissipative scattering from these dark modes to those with cqr0c_{qr}\approx 0. Said differently, the product is protected from reversion to reactant when dark modes with relatively more reactive character lose their energy to those with relatively less. Importantly, this scattering requires dark modes to be delocalized. The VSC reaction kinetics, as described above, is summarized in Fig. 4(b).

Fig. 4(c) shows the ratio of VSC reaction rate to bare reaction rate, as determined from numerical kinetic simulations. As we have shown analytically, VSC significantly accelerates the reaction compared to the bare case. For 8N328\leq N\leq 32 (and gNg\sqrt{N} held constant), the rate enhancement is roughly 50%. Notably, the effect of cavity decay on the reaction is minor and diminishes with NN [Fig. 4(c)]. This behavior supports that the reaction proceeds mainly through the dark modes. The present scenario is quite generic and contrasts with our previous model where extreme geometric parameters are needed for polaritons to dominate the VSC kinetics [37].

Refer to caption
Figure 4: (a) Schematic of the bare reaction kinetics. Parabolas represent potential energy surfaces with respect to the effective low-frequency coordinate xsx_{s}. After the main reactive transition (|R,0|P,1r|R,0\rangle\rightarrow|P,1_{r}\rangle) with rate kfk_{f} (the transition |R,0|P,0|R,0\rangle\rightarrow|P,0\rangle is not shown), the product either reverts to the reactant (|P,1r|R,0|P,1_{r}\rangle\rightarrow|R,0\rangle) at rate kbk_{b} or vibrationally decays to its stable form (|P,1r|P,0|P,1_{r}\rangle\rightarrow|P,0\rangle) at rate γkb\gamma\approx k_{b}. (b) Schematic of the reaction kinetics under VSC. The half-yellow half-gray parabolas qualitatively represent potential energy surfaces of product states with one excitation in a dark mode (|P,1q|P,1_{q}\rangle). The reaction proceeds via multiple reaction channels, each involving a reactive transition (|R,0|P,1q|R,0\rangle\rightarrow|P,1_{q}\rangle) with rate kf(q)<kfk_{f}^{(q)}<k_{f} to the product with one excitation in a dark mode. The total forward rate is approximately the bare rate kfk_{f}. In contrast, the vibrationally hot product formed from each reaction channel either returns to the reactant (|P,1q|R,0|P,1_{q}\rangle\rightarrow|R,0\rangle) at rate kb(q)<kbk_{b}^{(q)}<k_{b} or cools (|P,1q|P,0|P,1_{q}\rangle\rightarrow|P,0\rangle) at the bare molecular rate γ\gamma. There is also scattering (at effective rate ζ\zeta) from dark modes with cqr0c_{qr}\not\approx 0 to those with cqr0c_{qr}\approx 0. Overall, VSC accelerates product thermalization, suppressing backward reactive transitions, and thus enhancing the net reaction rate. (c) kVSC/kbarek_{\text{VSC}}/k_{\text{bare}} as a function of NN for fixed gNg\sqrt{N} and various cavity decay rates κ\kappa. (d) Activation enthalpy ΔH\Delta H^{\ddagger} versus activation entropy ΔS\Delta S^{\ddagger} for reactions with N=32N=32: bare (gray circle), VSC (purple circle, κ=1\kappa=1 ps-1), and bare with vibrational decay rate γ\gamma made 100 times faster (purple diamond). The black dashed line is a fit to the points shown. In (c) and (d), the individual rates (kVSCk_{\text{VSC}}, kbarek_{\text{bare}}) and thermodynamic parameters (ΔH\Delta H^{\ddagger}, ΔS\Delta S^{\ddagger}) are averages over 5000 disorder realizations.

We next look at the dependence on cavity detuning, δ=ωcω¯v\delta=\omega_{c}-\overline{\omega}_{v}, of the reaction rate and reactive-mode delocalization. The lattermost quantity is defined as 1/q=1N+1|cqr|41/\sum_{q=1}^{N+1}|c_{qr}|^{4} (the PR of vrv_{r} when the mode is expressed in the eigenbasis of HH). We find, for various light-matter coupling strengths, that the reactive-mode delocalization is maximum close to resonance and eventually decreases with detuning [Fig. 5(a)]. The rate enhancement due to VSC mostly follows the same trend [Fig. 5(b)]. Deviation from this trend at large negative detunings and collective light-matter couplings is attributed to polariton contributions to the rate which, as discussed in [37], decrease as N increases [52]. The observed correlation between reactivity under VSC and delocalization of the reactive mode corroborates that the reaction is sped up by dark-mode semilocalization. This mechanism is robust to moderate increases in inhomogeneous broadening [52].

Refer to caption
Figure 5: (a) kVSC/kbarek_{\text{VSC}}/k_{\text{bare}} and (b) vrv_{r} delocalization, as a function of cavity detuning δ\delta, for various collective light-matter coupling strengths gNg\sqrt{N} and fixed N=32N=32. In (a), kVSCk_{\text{VSC}} and kbarek_{\text{bare}} are averages over 5000 disorder realizations.

For additional mechanistic insight into VSC catalysis and following the procedures in [23, 48, 25], we plot in Fig. 4(d) the activation enthalpy (ΔH\Delta H^{\ddagger}) versus activation entropy (ΔS\Delta S^{\ddagger}) for multiple cases of VSC and bare reactions. The thermodynamic parameters of activation are computed by calculating the apparent reaction rate for additional temperatures and fitting the obtained values to the Eyring-Polanyi equation [52]. This fit indicates that changes in effective parameters ΔH\Delta H^{\ddagger} and ΔS\Delta S^{\ddagger} can result from dynamical effects such as accelerated vibrational decay, rather than from potential energy changes.

In conclusion, we show that, by forming semilocalized dark modes, the VSC of a disordered molecular ensemble can modify the kinetics of a thermally-activated chemical reaction. For a reactive molecule under collective VSC, we find that the electron transfer rate is significantly increased. The spreading of reactive character across dark modes, as well as the dissipative scattering among these modes, allows the reactive mode to thermalize more efficiently once the product is formed, suppressing dynamical effects, such as reversion to reactant. Although experimental characterization of dark states remains a challenge, the phenomena proposed here might be verified using nonlinear infrared spectroscopy to measure populations [58, 60] and spatially resolved energy transport measurements to detect delocalization [41]. The main mechanisms operating in our model do not seem to be limited to nonadiabatic reactions and might have generalizations in adiabatic reactions; these will be explored in future work. Given that these mechanisms only rely on collective VSC, they should also be operative in the cavity-free polaritonic architectures [61], although experiments along this front have so far not been reported. More broadly, our work highlights that the previously overlooked dark states are the entropically likely channels through which collective light-matter interaction can control chemistry.

Acknowledgements.
We are grateful to Jorge Campos-Gonzalez-Angulo, Arghadip Koner, Luis Martínez-Martínez, Kai Schwennicke, Stephan van den Wildenberg, and Garret Wiesehan for useful discussions. This work employed computational resources of the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation Grant No. ACI-1548562, under allocation No. TG-ASC150024. We also thank Marty Kandes, Nicole Wolter, and Mahidhar Tatineni for assistance in using these resources. Acknowledgment is made to the donors of The American Chemical Society Petroleum Research Fund for partial support of this research through the ACS PRF 60968-ND6 award. M.D. is also supported by a UCSD Roger Tsien Fellowship.

References