Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Polariton assisted incoherent to coherent excitation energy transfer between colloidal nanocrystal quantum dots

Kaiyue Peng kaiyue_peng@berkeley.edu Department of Chemistry, University of California, Berkeley, California 94720, United States    Eran Rabani eran.rabani@berkeley.edu Department of Chemistry, University of California, Berkeley, California 94720, United States Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States The Sackler Center for Computational Molecular and Materials Science, Tel Aviv University, Tel Aviv 69978, Israel
(September 30, 2024)
Abstract

We explore the dynamics of energy transfer between two nanocrystal quantum dots placed within an optical microcavity. By adjusting the coupling strength between the cavity photon mode and the quantum dots, we have the capacity to fine-tune the effective coupling between the donor and acceptor. Introducing a non-adiabatic parameter, γ𝛾\gammaitalic_γ, governed by the coupling to the cavity mode, we demonstrate the system’s capability to shift from the overdamped Förster regime (γ1much-less-than𝛾1\gamma\ll 1italic_γ ≪ 1) to an underdamped coherent regime (γ1much-greater-than𝛾1\gamma\gg 1italic_γ ≫ 1). In the latter regime, characterized by swift energy transfer rates, the dynamics are influenced by decoherence times. To illustrate this, we study the exciton energy transfer dynamics between two closely positioned CdSe/CdS core/shell quantum dots with sizes and separations relevant to experimental conditions. Employing an atomistic approach, we calculate the excitonic level arrangement, exciton–phonon interactions, and transition dipole moments of the quantum dots within the microcavity. These parameters are then utilized to define a model Hamiltonian. Subsequently, we apply a generalized non-Markovian quantum Redfield equation to delineate the dynamics within the polaritonic framework.

Quantum dots, Polariton, Micro–cavity, Coherent energy transfer
preprint: AIP/123-QED

I Introduction

In the realm of nanophotonics and quantum optics, understanding and controlling the transfer of energy between quantum emitters is paramount for advancing technologies ranging from solar cells to quantum information processing.[1, 2] Colloidal nanocrystal (NC) quantum dots (QDs), with unique electronic and optical properties,[3, 4, 5] have emerged as versatile platforms for studying energy transfer processes at the nanoscale.[6, 1, 7, 8] Among these processes, Fluorescence Resonance Energy Transfer (FRET) has garnered considerable attention for its role in mediating efficient energy transfer between quantum dots[1, 9, 8] and other emitters,[10, 11] and has been extensively studied and exploited in various fields, including molecular biology,[12, 13] where it serves as a fundamental tool for probing molecular interactions and dynamics.[11, 14] FRET, an incoherent energy transfer process,[15, 16] relies on the weak coupling between a donor and an acceptor fluorophore, with a transfer rate given by Fermi’s golden rule (FGR):[17]

kFGR=2π𝒟𝒜ρ𝒟|J𝒟𝒜|2δ(E𝒟E𝒜).subscript𝑘FGR2𝜋Planck-constant-over-2-pisubscript𝒟𝒜subscript𝜌𝒟superscriptsubscript𝐽𝒟𝒜2𝛿subscript𝐸𝒟subscript𝐸𝒜k_{\text{FGR}}=\frac{2\pi}{\hbar}\sum_{{\cal D}{\cal A}}\rho_{{\cal D}}\left|J% _{{\cal DA}}\right|^{2}\delta\left(E_{{\cal D}}-E_{{\cal A}}\right).italic_k start_POSTSUBSCRIPT FGR end_POSTSUBSCRIPT = divide start_ARG 2 italic_π end_ARG start_ARG roman_ℏ end_ARG ∑ start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT | italic_J start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ ( italic_E start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT ) . (1)

Here, ρ𝒟subscript𝜌𝒟\rho_{{\cal D}}italic_ρ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT is the population distribution of excited states on the donor, J𝒟𝒜subscript𝐽𝒟𝒜J_{{\cal DA}}italic_J start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT is the electromagnetic coupling between transition densities connecting the ground state and excited state for the donor and the acceptor (approximated by the dipole-dipole term, see Eq. (3) below), and E𝒟subscript𝐸𝒟E_{{\cal D}}italic_E start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT and E𝒜subscript𝐸𝒜E_{{\cal A}}italic_E start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT are the donor and acceptor excited state energies, respectively. When the emission spectrum of the donor overlaps with the absorption spectrum of the acceptor and they are within a certain distance, typically within a few nanometers, energy transfer can occur with high efficiency, but on relatively long timescales (a few tens of nanoseconds).[10, 9, 18] FRET processes are inherently limited by the steep distance dependence (R3superscript𝑅3R^{-3}italic_R start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT) of the dipole-dipole coupling between the donor and acceptor QDs, resulting in small couplings, overdamped energy transfer dynamics, and long transfer rates, necessitating the exploration of alternative mechanisms to accelerate and control the energy transfer mechanism.[19, 20, 21]

One strategy for regulating the exciton energy transfer process involves utilizing a donor–bridge–acceptor configuration, wherein the transition mechanism and rates can be adjusted by manipulating the states of the bridge.[22, 23] Recent advancements have led to the emergence of polariton–mediated energy transfer as a promising alternative to FRET.[24, 20, 25, 26] Polaritons, arising from the strong coupling between photons and excitons,[27] present distinct advantages for facilitating energy transfer, including heightened coupling strength, tunable spectral overlap, and minimal dependence on distance, making efficient remote energy transfer possible with more flexible material design.[28, 26, 29, 21, 30] As efforts intensify to exploit polariton-mediated energy transfer, ongoing research endeavors are dedicated to elucidating the control parameters dictating the efficiency and dynamics of these processes, such as energy decay rate, dephasing rate, and Rabi splitting/light coupling strength, as well as their scaling with number of molecules involved.[31, 32, 21, 33]

In this study, we examine the dynamics of energy transfer between two quantum dots (QDs) situated in an optical microcavity, as depicted in Fig. 1. By adjusting the coupling strength between the cavity photon mode and the QDs, we can manipulate the effective coupling between the donor and acceptor while maintaining the system reorganization energy almost constant. We demonstrate that this ability to control the effective coupling between the donor and acceptor, without altering other aspects of the system, offers a means to transition the system from the overdamped Förster regime to an underdamped coherent regime, where the transfer rate is influenced by decoherence time.

To illustrate this transition to coherent energy transfer dynamics, we examine two closely positioned CdSe/CdS core/shell QDs of sizes and separations relevant to experimental conditions. We utilize an atomistic theory to compute the excitonic level structure,[34, 35] exciton–phonon couplings,[35, 36] and transition dipole moments of the two QDs within a microcavity setting. To capture the intricate interplay among photons, phonons, and excitons, we employ a generalized non-Markovian quantum Redfield equation, which provides a framework for accurately describing both populations and coherences under weak exciton–phonon coupling conditions, suitable for CdSe and CdS materials. We employ a non-adiabatic parameter, γ𝛾\gammaitalic_γ, influenced by the coupling to the cavity mode (among other factors) and demonstrate that in the nonadiabatic regime (γ1much-less-than𝛾1\gamma\ll 1italic_γ ≪ 1), our approach aligns with Förster theory without the necessity of specifying the spectral line width, as it is self-determined by the Redfield approach. As γ𝛾\gammaitalic_γ increases beyond 1111 towards the adiabatic limit, the dynamics exhibit coherent oscillations at the Rabi frequency, with energy transfer rates dictated by the decoherence time.

Refer to caption
Figure 1: An illustration of two core/shell quantum dots in an optical microcavity.

II Model and Dynamics

II.1 Donor-acceptor and cavity Hamiltonians

We write the total Hamiltonian, H=H𝒟𝒜+Hcav𝐻subscript𝐻𝒟𝒜subscript𝐻cavH=H_{{\rm\mathcal{DA}}}+H_{{\rm cav}}italic_H = italic_H start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT, as a sum of the donor–acceptor vibronic Hamiltonian, H𝒟𝒜subscript𝐻𝒟𝒜H_{{\rm\mathcal{DA}}}italic_H start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT, and the contributions due to the cavity mode, Hcavsubscript𝐻cavH_{{\rm cav}}italic_H start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT. The donor–acceptor vibronic Hamiltonian of the two QDs is derived from the model Hamiltonian developed by Jasrasaria and Rabani.[36, 35] This model describes hot exciton relaxation in a single QD, where ultrafast relaxation dynamics are driven by multi–phonon channels.[37] Its generalization for a donor–acceptor system is given by

H𝒟𝒜subscript𝐻𝒟𝒜\displaystyle H_{{\rm\mathcal{DA}}}italic_H start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT =n𝒟,𝒜En|ψnψn|absentsubscript𝑛𝒟𝒜subscript𝐸𝑛ketsubscript𝜓𝑛brasubscript𝜓𝑛\displaystyle=\sum_{n\in\mathcal{D},\mathcal{A}}E_{n}\left|\psi_{n}\right% \rangle\left\langle\psi_{n}\right|= ∑ start_POSTSUBSCRIPT italic_n ∈ caligraphic_D , caligraphic_A end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT |
+n𝒟,m𝒜Jnm(|ψnψm|+h.c.)\displaystyle+\sum_{n\in\mathcal{D},m\in\mathcal{A}}J_{nm}\left(\left|\psi_{n}% \right\rangle\left\langle\psi_{m}\right|+h.c.\right)+ ∑ start_POSTSUBSCRIPT italic_n ∈ caligraphic_D , italic_m ∈ caligraphic_A end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT | + italic_h . italic_c . )
+αωαbαbα+αn,m𝒟,𝒜Vnmα|ψnψm|qα.subscript𝛼Planck-constant-over-2-pisubscript𝜔𝛼superscriptsubscript𝑏𝛼subscript𝑏𝛼subscript𝛼subscriptformulae-sequence𝑛𝑚𝒟𝒜superscriptsubscript𝑉𝑛𝑚𝛼ketsubscript𝜓𝑛brasubscript𝜓𝑚subscript𝑞𝛼\displaystyle+\sum_{\alpha}\hbar\omega_{\alpha}b_{\alpha}^{\dagger}b_{\alpha}+% \sum_{\alpha}\sum_{n,m\in\mathcal{D},\mathcal{A}}V_{nm}^{\alpha}\left|\psi_{n}% \right\rangle\left\langle\psi_{m}\right|q_{\alpha}.+ ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT roman_ℏ italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_n , italic_m ∈ caligraphic_D , caligraphic_A end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT | italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT . (2)

The first and second terms on the right hand side describe the manifold of donor and acceptor excitonic states and their couplings, respectively, and the remaining terms describe the lattice vibrations and the exciton–phonon couplings, approximated to lowest order in the phonons displacement, qαsubscript𝑞𝛼q_{\alpha}italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT. In the above, Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the energy of a singly excited state |ψnketsubscript𝜓𝑛\left|\psi_{n}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩, with n𝒟/𝒜𝑛𝒟𝒜n\in\mathcal{D}/\mathcal{A}italic_n ∈ caligraphic_D / caligraphic_A and 𝒟={𝒟1,𝒟2,}𝒟subscript𝒟1subscript𝒟2\mathcal{D}=\left\{{\cal D}_{1},{\cal D}_{2},\cdots\right\}caligraphic_D = { caligraphic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , caligraphic_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , ⋯ }, 𝒜={𝒜1,𝒜2,}𝒜subscript𝒜1subscript𝒜2\mathcal{A}=\left\{{\cal A}_{1},{\cal A}_{2},\cdots\right\}caligraphic_A = { caligraphic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , caligraphic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , ⋯ }. We use the notation |ψ𝒟1ketsubscript𝜓subscript𝒟1\left|\psi_{{\cal D}_{1}}\right\rangle| italic_ψ start_POSTSUBSCRIPT caligraphic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩ to describe the donor in excitonic state 𝒟1subscript𝒟1{\cal D}_{1}caligraphic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT while the acceptor is in its ground state and vise versa for |ψ𝒜1ketsubscript𝜓subscript𝒜1\left|\psi_{{\cal A}_{1}}\right\rangle| italic_ψ start_POSTSUBSCRIPT caligraphic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ⟩. The excitonic energies and wave functions (Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and |ψnketsubscript𝜓𝑛\left|\psi_{n}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩) were obtained by ignoring the coupling between the donor and acceptor, using the semi–empirical pseudopotential methods combined with the Bethe–Salpeter equation (BSE).[36] The quasiparticle states near the top of the valence band and the bottom of the conduction band were generated for each QD using the filter diagonalization technique[38, 39] with a real–space grid basis of Ng2106subscript𝑁𝑔2superscript106N_{g}\approx 2\cdot 10^{6}italic_N start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ≈ 2 ⋅ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT and a grid spacing of 0.75a0absent0.75subscript𝑎0\approx 0.75a_{0}≈ 0.75 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. A total of 80808080 electron (unoccupied) and 140140140140 hole (occupied) states were generated to construct the Bethe–Salpeter Hamiltonian,[40] within the static screening approximation,[36] providing converged excitonic energies and transition dipole moments for the lowest 30303030 excitonic states that participate in the dynamics, for each QD.

The coupling between the donor in state |ψ𝒟ketsubscript𝜓𝒟\left|\psi_{\mathcal{D}}\right\rangle| italic_ψ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT ⟩ and the acceptor in state |ψ𝒜ketsubscript𝜓𝒜\left|\psi_{\mathcal{A}}\right\rangle| italic_ψ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT ⟩, J𝒟𝒜subscript𝐽𝒟𝒜J_{\mathcal{DA}}italic_J start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT, was approximated to lowest order in the corresponding transition charge densities, with the dipole–dipole leading term given by:[17]

J𝒟𝒜=f𝒟f𝒜|𝝁𝒟ex||𝝁𝒜ex|ϵm|R|3κ(θ𝒟𝒜,θ𝒟,θ𝒜)subscript𝐽𝒟𝒜subscript𝑓𝒟subscript𝑓𝒜superscriptsubscript𝝁𝒟exsuperscriptsubscript𝝁𝒜exsubscriptitalic-ϵmsuperscript𝑅3𝜅subscript𝜃𝒟𝒜subscript𝜃𝒟subscript𝜃𝒜J_{\mathcal{DA}}=f_{{\cal D}}f_{{\cal A}}\frac{\left|\boldsymbol{\mu}_{{\cal D% }}^{{\rm ex}}\right|\left|\boldsymbol{\mu}_{{\cal A}}^{{\rm ex}}\right|}{% \epsilon_{\text{m}}\left|R\right|^{3}}\kappa\left(\theta_{{\cal DA}},\theta_{{% \cal D}},\theta_{{\cal A}}\right)italic_J start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT divide start_ARG | bold_italic_μ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT | | bold_italic_μ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT | end_ARG start_ARG italic_ϵ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT | italic_R | start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG italic_κ ( italic_θ start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT ) (3)

Here, 𝝁𝒟/𝒜ex=ψ𝒟/𝒜|𝝁^|ψgsuperscriptsubscript𝝁𝒟/𝒜exquantum-operator-productsubscript𝜓𝒟/𝒜^𝝁subscript𝜓𝑔\boldsymbol{\mu}_{\mathcal{D\text{/}A}}^{{\rm ex}}=\left\langle\psi_{\mathcal{% D\text{/}A}}\right|\hat{\boldsymbol{\mu}}\left|\psi_{g}\right\ranglebold_italic_μ start_POSTSUBSCRIPT caligraphic_D / caligraphic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT = ⟨ italic_ψ start_POSTSUBSCRIPT caligraphic_D / caligraphic_A end_POSTSUBSCRIPT | over^ start_ARG bold_italic_μ end_ARG | italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ⟩ is the transition dipole moment obtained from the pseudopotential model [36] and |ψgketsubscript𝜓𝑔\left|\psi_{g}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ⟩ is the ground state. The local field factor in the above is given by f𝒟/𝒜=32+ϵ𝒟/𝒜/ϵm=37subscript𝑓𝒟𝒜32subscriptitalic-ϵ𝒟/𝒜subscriptitalic-ϵm37f_{\mathcal{D}/\mathcal{A}}=\frac{3}{2+\epsilon_{\text{$\mathcal{D}$/$\mathcal% {A}$}}/\epsilon_{\text{m}}}=\frac{3}{7}italic_f start_POSTSUBSCRIPT caligraphic_D / caligraphic_A end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 2 + italic_ϵ start_POSTSUBSCRIPT caligraphic_D / caligraphic_A end_POSTSUBSCRIPT / italic_ϵ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT end_ARG = divide start_ARG 3 end_ARG start_ARG 7 end_ARG for [41] ϵ𝒟=ϵ𝒜=5subscriptitalic-ϵ𝒟subscriptitalic-ϵ𝒜5\epsilon_{{\cal D}}=\epsilon_{{\cal A}}=5italic_ϵ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT = italic_ϵ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT = 5 and ϵ=m1\epsilon{}_{\text{m}}=1italic_ϵ start_FLOATSUBSCRIPT m end_FLOATSUBSCRIPT = 1, and |κ(θ𝒟𝒜,θ𝒟,θ𝒜)|2𝜅subscript𝜃𝒟𝒜subscript𝜃𝒟subscript𝜃𝒜2\left|\kappa\left(\theta_{{\cal DA}},\theta_{{\cal D}},\theta_{{\cal A}}\right% )\right|\leq 2| italic_κ ( italic_θ start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT ) | ≤ 2 is an angle dependent term which can be approximated by averaging over the angles, with the well-known results that κ2=23delimited-⟨⟩superscript𝜅223\sqrt{\left\langle\kappa^{2}\right\rangle}=\sqrt{\frac{2}{3}}square-root start_ARG ⟨ italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG = square-root start_ARG divide start_ARG 2 end_ARG start_ARG 3 end_ARG end_ARG.[42] Alternatively, one can compute the angles θ𝒟𝒜subscript𝜃𝒟𝒜\theta_{{\cal DA}}italic_θ start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT, θ𝒟subscript𝜃𝒟\theta_{{\cal D}}italic_θ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT, and θ𝒜subscript𝜃𝒜\theta_{{\cal A}}italic_θ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT, which depend on the relative orientation of the transition dipole moments.[43]

The donor and acceptor excitonic states also couple to the lattice vibrations with frequencies ωαsubscript𝜔𝛼\omega_{\alpha}italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and modes qα=2ωα(bα+bα)subscript𝑞𝛼Planck-constant-over-2-pi2subscript𝜔𝛼superscriptsubscript𝑏𝛼subscript𝑏𝛼q_{\alpha}=\sqrt{\frac{\hbar}{2\omega_{\alpha}}}\left(b_{\alpha}^{\text{$% \dagger$}}+b_{\alpha}\right)italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG roman_ℏ end_ARG start_ARG 2 italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG end_ARG ( italic_b start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT + italic_b start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ), obtained by diagonalizing the dynamical matrix of the QD donor and acceptor using the Stilling–Weber force field parameterized for Cd, Se, S elements.[44] The exciton–phonon couplings were expanded to lowest order in the displacements, qαsubscript𝑞𝛼q_{\alpha}italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and the strength of coupling between exciton |ψnketsubscript𝜓𝑛\left|\psi_{n}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ and |ψmketsubscript𝜓𝑚\left|\psi_{m}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ⟩ to mode α𝛼\alphaitalic_α, Vnmαsuperscriptsubscript𝑉𝑛𝑚𝛼V_{nm}^{\alpha}italic_V start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT, was obtained from the pseudopotential Hamiltonian.[35] We considered both diagonal Vn=mαsuperscriptsubscript𝑉𝑛𝑚𝛼V_{n=m}^{\alpha}italic_V start_POSTSUBSCRIPT italic_n = italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT and off-diagonal Vnmαsuperscriptsubscript𝑉𝑛𝑚𝛼V_{n\neq m}^{\alpha}italic_V start_POSTSUBSCRIPT italic_n ≠ italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT coupling terms and assumed that the excitons couple to the lattice modes on either the donor or the acceptor, depending on whether they localize to the donor or the acceptor. The coupling of excitons to the lattice vibrations is key in accounting for the reorganization and the relaxation of the system to thermal equilibrium.

The optical cavity was modeled by a single–mode Pauli–Fierz Hamiltonian within the dipole approximation and rotation wave approximation:[45, 46]

Hcavsubscript𝐻cav\displaystyle H_{\text{cav}}italic_H start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT =ωcaa+n𝒟,𝒜gn(a|ψgψn|+h.c.)\displaystyle=\hbar\omega_{c}a^{\dagger}a+\sum_{n\in\mathcal{D},\mathcal{A}}% \hbar g_{n}\left(a^{\dagger}\left|\psi_{g}\right\rangle\left\langle\psi_{n}% \right|+h.c.\right)= roman_ℏ italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a + ∑ start_POSTSUBSCRIPT italic_n ∈ caligraphic_D , caligraphic_A end_POSTSUBSCRIPT roman_ℏ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | + italic_h . italic_c . )
+ωcn𝒟,𝒜gn2|ψgψg|Planck-constant-over-2-pisubscript𝜔𝑐subscript𝑛𝒟𝒜superscriptsubscript𝑔𝑛2ketsubscript𝜓𝑔brasubscript𝜓𝑔\displaystyle+\frac{\hbar}{\omega_{c}}\sum_{n\in\mathcal{D},\mathcal{A}}g_{n}^% {2}\left|\psi_{g}\right\rangle\left\langle\psi_{g}\right|+ divide start_ARG roman_ℏ end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_n ∈ caligraphic_D , caligraphic_A end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT |
+ωcn,m𝒟,𝒜gngm(|ψnψm|+h.c.),\displaystyle+\frac{\hbar}{\omega_{c}}\sum_{n,m\in\mathcal{D},\mathcal{A}}g_{n% }g_{m}\left(\left|\psi_{n}\right\rangle\left\langle\psi_{m}\right|+h.c.\right),+ divide start_ARG roman_ℏ end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_n , italic_m ∈ caligraphic_D , caligraphic_A end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT | + italic_h . italic_c . ) , (4)

where a𝑎aitalic_a (asuperscript𝑎a^{\dagger}italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT) is the annihilation (creation) operator of a photon with frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and gnPlanck-constant-over-2-pisubscript𝑔𝑛\hbar g_{n}roman_ℏ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is coupling strength between the cavity mode and an excitonic state |ψnketsubscript𝜓𝑛\left|\psi_{n}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ (n𝒟/𝒜𝑛𝒟𝒜n\in\mathcal{D}/\mathcal{A}italic_n ∈ caligraphic_D / caligraphic_A). The cavity–exciton coupling strength is given by gn=ωc/(2ϵm𝒱)𝝁nex𝜺subscript𝑔𝑛subscript𝜔𝑐2Planck-constant-over-2-pisubscriptitalic-ϵm𝒱superscriptsubscript𝝁𝑛ex𝜺g_{n}=\sqrt{\omega_{c}/\left(2\hbar\epsilon_{\text{m}}\mathcal{V}\right)}% \boldsymbol{\mu}_{n}^{{\rm ex}}\cdot\boldsymbol{\varepsilon}italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = square-root start_ARG italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / ( 2 roman_ℏ italic_ϵ start_POSTSUBSCRIPT m end_POSTSUBSCRIPT caligraphic_V ) end_ARG bold_italic_μ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT ⋅ bold_italic_ε, where as before, 𝝁nexsuperscriptsubscript𝝁𝑛ex\boldsymbol{\mu}_{n}^{{\rm ex}}bold_italic_μ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT is the transition dipole moment from the ground state |ψgketsubscript𝜓𝑔\left|\psi_{g}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ⟩ to exciton |ψnketsubscript𝜓𝑛\left|\psi_{n}\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ (obtained from the BSE calculation) and 𝜺𝜺\boldsymbol{\varepsilon}bold_italic_ε is a unit vector of the polarization direction of the electromagnetic field. ϵmsubscriptitalic-ϵm\epsilon_{{\rm m}}italic_ϵ start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is the effective permittivity inside the cavity (medium) and 𝒱𝒱\mathcal{V}caligraphic_V is the effective cavity quantization volume. The last two terms in the above equation correspond to the quadratic dipole self-energy, which has shown to be important in the ultra-strong coupling limit.[47] The dipole self-energy term arises from the Power-Zienau-Woolley transformation that explicitly mixes the electronic and photonic degrees of freedom.[48] For the results reported in this work, the effect of the dipole-self-energy is negligible since ωcgnmuch-greater-thansubscript𝜔𝑐subscript𝑔𝑛\omega_{c}\gg g_{n}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≫ italic_g start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT.

II.2 Polaritonic representation and polaron transformation

The total Hamiltonian can also be expressed using the polaritonic basis, {|φn=cng|ψg,1+mcnmex|ψm,0}ketsubscript𝜑𝑛subscript𝑐𝑛𝑔ketsubscript𝜓𝑔1subscript𝑚superscriptsubscript𝑐𝑛𝑚exketsubscript𝜓𝑚0\left\{\left|\varphi_{n}\right\rangle=c_{ng}\left|\psi_{g},1\right\rangle+\sum% _{m}c_{nm}^{\text{ex}}\left|\psi_{m},0\right\rangle\right\}{ | italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ = italic_c start_POSTSUBSCRIPT italic_n italic_g end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , 1 ⟩ + ∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ex end_POSTSUPERSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , 0 ⟩ }, where |ψg,1ketsubscript𝜓𝑔1\left|\psi_{g},1\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , 1 ⟩ and |ψm,0ketsubscript𝜓𝑚0\left|\psi_{m},0\right\rangle| italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , 0 ⟩ are the ground state with one photon or an excitonic state with zero photons, respectively. This representation simplifies the description of the dynamics within the Redfield approach described below, and is not limited by the weak donor-acceptor coupling strength (provided that the exciton-phonon couplings are small), as is the case for FRET between CdSe/CdS core-shell QDs. The coefficients cngsubscript𝑐𝑛𝑔c_{ng}italic_c start_POSTSUBSCRIPT italic_n italic_g end_POSTSUBSCRIPT and cnmexsuperscriptsubscript𝑐𝑛𝑚exc_{nm}^{\text{ex}}italic_c start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ex end_POSTSUPERSCRIPT were obtained by diagonalizing the system Hamiltonian (describing only the excitons and photons without the phonons), defined by:

HSsubscript𝐻S\displaystyle H_{{\rm S}}italic_H start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT =Hcav+n𝒟,𝒜En|ψnψn|absentsubscript𝐻cav𝑛𝒟𝒜subscript𝐸𝑛ketsubscript𝜓𝑛brasubscript𝜓𝑛\displaystyle=H_{\text{cav}}+\underset{n\in\mathcal{D},\mathcal{A}}{\sum}E_{n}% \left|\psi_{n}\right\rangle\left\langle\psi_{n}\right|= italic_H start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT + start_UNDERACCENT italic_n ∈ caligraphic_D , caligraphic_A end_UNDERACCENT start_ARG ∑ end_ARG italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT |
+n𝒟,m𝒜Jnm(|ψnψm|+h.c.).\displaystyle+\underset{n\in\mathcal{D},m\in\mathcal{A}}{\sum}J_{nm}\left(% \left|\psi_{n}\right\rangle\left\langle\psi_{m}\right|+h.c.\right).+ start_UNDERACCENT italic_n ∈ caligraphic_D , italic_m ∈ caligraphic_A end_UNDERACCENT start_ARG ∑ end_ARG italic_J start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( | italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_ψ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT | + italic_h . italic_c . ) . (5)

In the above, we assumed that the ground state energy is zero, namely Eg=0subscript𝐸𝑔0E_{g}=0italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 0. Using the polaritonic basis obtained by diagonalizing the system Hamiltonian in Eq. (5), the total Hamiltonian (=S+B+IsubscriptSsubscriptBsubscriptI\mathcal{H}=\mathcal{H}_{{\rm S}}+\mathcal{H}_{\text{B}}+\mathcal{H}_{\text{I}}caligraphic_H = caligraphic_H start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT + caligraphic_H start_POSTSUBSCRIPT B end_POSTSUBSCRIPT + caligraphic_H start_POSTSUBSCRIPT I end_POSTSUBSCRIPT) given by H𝒟𝒜+Hcavsubscript𝐻𝒟𝒜subscript𝐻cavH_{{\rm\mathcal{DA}}}+H_{{\rm cav}}italic_H start_POSTSUBSCRIPT caligraphic_D caligraphic_A end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT roman_cav end_POSTSUBSCRIPT can be written as:

\displaystyle{\cal H}caligraphic_H =n(E~nλn)|φnφn|SabsentsubscriptSsubscript𝑛subscript~𝐸𝑛subscript𝜆𝑛ketsubscript𝜑𝑛brasubscript𝜑𝑛\displaystyle=\overset{{\scriptstyle\mathcal{H}_{{\rm S}}}}{\overbrace{\sum_{n% }\left(\tilde{E}_{n}-\lambda_{n}\right)\left|\varphi_{n}\right\rangle\left% \langle\varphi_{n}\right|}}= start_OVERACCENT caligraphic_H start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT end_OVERACCENT start_ARG over⏞ start_ARG ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) | italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | end_ARG end_ARG
+αωαbαbα+BnmWnm|φnφm|I.subscriptBlimit-fromsubscript𝛼Planck-constant-over-2-pisubscript𝜔𝛼superscriptsubscript𝑏𝛼subscript𝑏𝛼subscriptIsubscript𝑛𝑚subscript𝑊𝑛𝑚ketsubscript𝜑𝑛brasubscript𝜑𝑚\displaystyle+\overset{{\scriptstyle\mathcal{H}_{\text{B}}}}{\overbrace{\sum_{% \alpha}\hbar\omega_{\alpha}b_{\alpha}^{\text{$\dagger$}}b_{\alpha}}+}\overset{% {\scriptstyle\mathcal{H}_{\text{I}}}}{\overbrace{\sum_{n\neq m}W_{nm}\left|% \varphi_{n}\right\rangle\left\langle\varphi_{m}\right|}}.+ start_OVERACCENT caligraphic_H start_POSTSUBSCRIPT B end_POSTSUBSCRIPT end_OVERACCENT start_ARG over⏞ start_ARG ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT roman_ℏ italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG + end_ARG start_OVERACCENT caligraphic_H start_POSTSUBSCRIPT I end_POSTSUBSCRIPT end_OVERACCENT start_ARG over⏞ start_ARG ∑ start_POSTSUBSCRIPT italic_n ≠ italic_m end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT | italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_φ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT | end_ARG end_ARG . (6)

In the above, E~nsubscript~𝐸𝑛\tilde{E}_{n}over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the polaritonic energy of state |φnketsubscript𝜑𝑛\left|\varphi_{n}\right\rangle| italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ (an eigenvalue of HSsubscript𝐻SH_{{\rm S}}italic_H start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT in Eq. (5)) and λn=12α(V~nnα)2/ωα2subscript𝜆𝑛12subscript𝛼superscriptsuperscriptsubscript~𝑉𝑛𝑛𝛼2superscriptsubscript𝜔𝛼2\lambda_{n}=\frac{1}{2}\sum_{\alpha}\left(\tilde{V}_{nn}^{\alpha}\right)^{2}/% \omega_{\alpha}^{2}italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the corresponding reorganization energy (polaron shift). The dressed coupling between polaritons and phonons, Wnmsubscript𝑊𝑛𝑚W_{nm}italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT, is given by:

Wnmsubscript𝑊𝑛𝑚\displaystyle W_{nm}italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT =exp(iγpγV~nnγωγ2)×\displaystyle=\exp\left(-\frac{i}{\hbar}\sum_{\gamma}\frac{p_{\gamma}\tilde{V}% _{nn}^{\gamma}}{\omega_{\gamma}^{2}}\right)\times= roman_exp ( - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ∑ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT divide start_ARG italic_p start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ×
×[αV~nmαqα]exp(+iξpξV~mmξωξ2).absentdelimited-[]subscript𝛼superscriptsubscript~𝑉𝑛𝑚𝛼subscript𝑞𝛼𝑖Planck-constant-over-2-pisubscript𝜉subscript𝑝𝜉superscriptsubscript~𝑉𝑚𝑚𝜉superscriptsubscript𝜔𝜉2\displaystyle\times\left[\sum_{\alpha}\tilde{V}_{nm}^{\alpha}q_{\alpha}\right]% \exp\left(+\frac{i}{\hbar}\sum_{\xi}\frac{p_{\xi}\tilde{V}_{mm}^{\xi}}{\omega_% {\xi}^{2}}\right).× [ ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ] roman_exp ( + divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ∑ start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT divide start_ARG italic_p start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_m italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ξ end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) . (7)

To derive the Hamiltonian in Eq. (6) we also performed a small polaron transformation (in addition to diagonalizing the system Hamiltonian to form the polaritonic states), as detailed in Refs. 37 and 46. As a result of the polaron transformation, the dressed coupling between polaritons, Wnmsubscript𝑊𝑛𝑚W_{nm}italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT, depends exponentially on the momenta (pαsubscript𝑝𝛼p_{\alpha}italic_p start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT) and position (qαsubscript𝑞𝛼q_{\alpha}italic_q start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT) of all phonon modes, enabling the description of multi–phonon relaxation channels within the lowest order perturbation in IsubscriptI\mathcal{H}_{\text{I}}caligraphic_H start_POSTSUBSCRIPT I end_POSTSUBSCRIPT. In the above, the dressed coupling (and the reorganization energy) depend on V~nmαsuperscriptsubscript~𝑉𝑛𝑚𝛼\tilde{V}_{nm}^{\alpha}over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT, the transformed coupling between two polaritonic states |φnketsubscript𝜑𝑛\left|\varphi_{n}\right\rangle| italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ and |φmketsubscript𝜑𝑚\left|\varphi_{m}\right\rangle| italic_φ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ⟩ and phonon mode α𝛼\alphaitalic_α, obtained by applying the unitary transformation U𝑈Uitalic_U that diagonalizes HSsubscript𝐻SH_{{\rm S}}italic_H start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT to V~α=UVαUsuperscript~𝑉𝛼superscript𝑈superscript𝑉𝛼𝑈\tilde{V}^{\alpha}=U^{\dagger}V^{\alpha}Uover~ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT = italic_U start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_V start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_U, for each mode α𝛼\alphaitalic_α.

II.3 Dynamics

In this work, we leverage on the fact that the rescaled coupling between polaritons and phonon, IsubscriptI\mathcal{H}_{\text{I}}caligraphic_H start_POSTSUBSCRIPT I end_POSTSUBSCRIPT (Eq. (6)), can be treated perturbatively. Thus, to generate the dynamics within the Hilbert space of polaritons, we adopt the Redfield equations instead of using more advanced methods suitable for strong coupling.[49] The equation of motion for the reduced density matrix, σ(t)𝜎𝑡\sigma\left(t\right)italic_σ ( italic_t ), representing the polaritonic subspace can then be written as:[50]

dσ(t)dt𝑑𝜎𝑡𝑑𝑡\displaystyle\frac{d\sigma\left(t\right)}{dt}divide start_ARG italic_d italic_σ ( italic_t ) end_ARG start_ARG italic_d italic_t end_ARG =i[S+IB,σ(t)]absent𝑖Planck-constant-over-2-pisubscriptSsubscriptdelimited-⟨⟩subscriptIB𝜎𝑡\displaystyle=-\frac{i}{\hbar}\left[\mathcal{H}_{\text{S}}+\left\langle% \mathcal{H}_{\text{I}}\right\rangle_{\text{B}},\sigma\left(t\right)\right]= - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG [ caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT + ⟨ caligraphic_H start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT , italic_σ ( italic_t ) ]
\displaystyle-- 12nm,kl0tdτ{Cnm,kl(τ)[(SI)nm,ei/Sτ(SI)klσ(tτ)ei/Sτ]\displaystyle\frac{1}{\hbar^{2}}\sum_{nm,kl}\int_{0}^{t}d\tau\left\{C_{nm,kl}% \left(\tau\right)\left[\left(\mathcal{H}_{\text{S}}^{{\rm I}}\right)_{nm},e^{-% i/\hbar\mathcal{H}_{\text{S}}\tau}\left(\mathcal{H}_{\text{S}}^{{\rm I}}\right% )_{kl}\sigma\left(t-\tau\right)e^{i/\hbar\mathcal{H}_{\text{S}}\tau}\right]\right.divide start_ARG 1 end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_n italic_m , italic_k italic_l end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_τ { italic_C start_POSTSUBSCRIPT italic_n italic_m , italic_k italic_l end_POSTSUBSCRIPT ( italic_τ ) [ ( caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT , italic_e start_POSTSUPERSCRIPT - italic_i / roman_ℏ caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT ( caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT italic_σ ( italic_t - italic_τ ) italic_e start_POSTSUPERSCRIPT italic_i / roman_ℏ caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT ]
\displaystyle-- Cnm,kl(τ)[(SI)nm,ei/Sτσ(tτ)(SI)klei/Sτ]}+lossσ(t),\displaystyle\left.C_{nm,kl}^{*}\left(\tau\right)\left[\left(\mathcal{H}_{% \text{S}}^{{\rm I}}\right)_{nm},e^{-i/\hbar\mathcal{H}_{\text{S}}\tau}\sigma% \left(t-\tau\right)\left(\mathcal{H}_{\text{S}}^{{\rm I}}\right)_{kl}e^{i/% \hbar\mathcal{H}_{\text{S}}\tau}\right]\right\}+\mathcal{L_{\rm loss}}\sigma(t),italic_C start_POSTSUBSCRIPT italic_n italic_m , italic_k italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_τ ) [ ( caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT , italic_e start_POSTSUPERSCRIPT - italic_i / roman_ℏ caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT italic_σ ( italic_t - italic_τ ) ( caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i / roman_ℏ caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT italic_τ end_POSTSUPERSCRIPT ] } + caligraphic_L start_POSTSUBSCRIPT roman_loss end_POSTSUBSCRIPT italic_σ ( italic_t ) , (8)

where the polariton-phonon coupling is expressed as I=nmWnm(SI)nmsubscriptIsubscript𝑛𝑚subscript𝑊𝑛𝑚subscriptsuperscriptsubscriptSI𝑛𝑚\mathcal{H}_{\text{I}}=\sum_{nm}W_{nm}\left(\mathcal{H}_{\text{S}}^{{\rm I}}% \right)_{nm}caligraphic_H start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT with (SI)nm=(1δnm)|φnφm|subscriptsuperscriptsubscriptSI𝑛𝑚1subscript𝛿𝑛𝑚ketsubscript𝜑𝑛brasubscript𝜑𝑚\left(\mathcal{H}_{\text{S}}^{{\rm I}}\right)_{nm}=\left(1-\delta_{nm}\right)% \left|\varphi_{n}\right\rangle\left\langle\varphi_{m}\right|( caligraphic_H start_POSTSUBSCRIPT S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_I end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = ( 1 - italic_δ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ) | italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ ⟨ italic_φ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT |. In the above equation, Bsubscriptdelimited-⟨⟩B\left\langle\cdots\right\rangle_{\text{B}}⟨ ⋯ ⟩ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT denotes an equilibrium thermal average over the so called ”bath” degrees of freedom. The bath correlation function appearing above is given by:

Cnm,kl(t)subscript𝐶𝑛𝑚𝑘𝑙𝑡\displaystyle C_{nm,kl}\left(t\right)italic_C start_POSTSUBSCRIPT italic_n italic_m , italic_k italic_l end_POSTSUBSCRIPT ( italic_t ) =Wnm(t)Wkl(0)BWnmBWklB,absentsubscriptdelimited-⟨⟩subscript𝑊𝑛𝑚𝑡subscript𝑊𝑘𝑙0Bsubscriptdelimited-⟨⟩subscript𝑊𝑛𝑚Bsubscriptdelimited-⟨⟩subscript𝑊𝑘𝑙B\displaystyle=\left\langle W_{nm}\left(t\right)W_{kl}\left(0\right)\right% \rangle_{\text{B}}-\left\langle W_{nm}\right\rangle_{\text{B}}\left\langle W_{% kl}\right\rangle_{\text{B}},= ⟨ italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( italic_t ) italic_W start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT ( 0 ) ⟩ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT - ⟨ italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT ⟨ italic_W start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT , (9)

with Wnm(t)=ei/BtWnmei/Btsubscript𝑊𝑛𝑚𝑡superscript𝑒𝑖Planck-constant-over-2-pisubscriptB𝑡subscript𝑊𝑛𝑚superscript𝑒𝑖Planck-constant-over-2-pisubscriptB𝑡W_{nm}\left(t\right)=e^{i/\hbar\mathcal{H}_{\text{B}}t}W_{nm}e^{-i/\hbar% \mathcal{H}_{\text{B}}t}italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( italic_t ) = italic_e start_POSTSUPERSCRIPT italic_i / roman_ℏ caligraphic_H start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i / roman_ℏ caligraphic_H start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT. These thermal averages and thermal correlation functions were computed analytically for the above Hamiltonian as further described in the supporting information.

In addition, to describe cavity losses, we added a phenomenological loss term represented by the last term on the right hand side of Eq. (8), given by:

[lossσ(t)]nm=δnm(cng2τcav)σnm(t),subscriptdelimited-[]subscriptloss𝜎𝑡𝑛𝑚subscript𝛿𝑛𝑚superscriptsubscript𝑐𝑛𝑔2superscript𝜏cavsubscript𝜎𝑛𝑚𝑡\displaystyle\left[\mathcal{L_{\text{loss}}}\sigma\left(t\right)\right]_{nm}=-% \delta_{nm}\left(\frac{c_{ng}^{2}}{\tau^{\text{cav}}}\right)\sigma_{nm}\left(t% \right),[ caligraphic_L start_POSTSUBSCRIPT loss end_POSTSUBSCRIPT italic_σ ( italic_t ) ] start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = - italic_δ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( divide start_ARG italic_c start_POSTSUBSCRIPT italic_n italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_τ start_POSTSUPERSCRIPT cav end_POSTSUPERSCRIPT end_ARG ) italic_σ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( italic_t ) , (10)

where σnm(t)=φn|σ(t)|φmsubscript𝜎𝑛𝑚𝑡quantum-operator-productsubscript𝜑𝑛𝜎𝑡subscript𝜑𝑚\sigma_{nm}\left(t\right)=\left\langle\varphi_{n}\left|\sigma\left(t\right)% \right|\varphi_{m}\right\rangleitalic_σ start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( italic_t ) = ⟨ italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | italic_σ ( italic_t ) | italic_φ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ⟩, and the cavity loss rate from polaritonic state |φnketsubscript𝜑𝑛\left|\varphi_{n}\right\rangle| italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ is give by the ratio of the cavity fraction of the polaritonic state, cng2=|ψg,1|φn|2superscriptsubscript𝑐𝑛𝑔2superscriptinner-productsubscript𝜓𝑔1subscript𝜑𝑛2c_{ng}^{2}=\left|\left\langle\psi_{g},1|\varphi_{n}\right\rangle\right|^{2}italic_c start_POSTSUBSCRIPT italic_n italic_g end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = | ⟨ italic_ψ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , 1 | italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and the cavity lifetime, τcavsuperscript𝜏cav\tau^{\text{cav}}italic_τ start_POSTSUPERSCRIPT cav end_POSTSUPERSCRIPT.

III Exciton energy transfer between QDs

Refer to caption
Figure 2: The calculated linear absorption spectrum for CdSe/CdS core–shell NCs, with a CdSe core diameter of 3333 nm and 2222 monolayers of CdS shell. The vertical lines in upper subplot indicate the magnitude of the oscillator strength of the transition from the ground state to that excitonic state. The continuous absorption spectra (red curve) was obtained by broadening the individual transitions with a variance of 35 meV35 meV35\text{ meV}35 meV Gaussian function. The blue dots indicate the radiative lifetime of each excitonic state given by Eq. (11) below. The lower subplot in this figure shows the density of excitonic states.

In this section, we examine the excitonic structure of a typical 3333 nm CdSe core with a 2-monolayer CdS shell and compare the predicted radiative lifetimes to experimental data. Additionally, we demonstrate that the Redfield approach, which treats the dressed exciton–phonon coupling perturbatively but includes all orders of the hybridization Jnmsubscript𝐽𝑛𝑚J_{nm}italic_J start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT, aligns well with Förster theory in the absence of a cavity. The main purpose of this section is to provide further validation for the approach developed herein and in particular, for the calculated oscillator strengths and vibronic couplings, essential to accurately describe the energy transfer mechanisms and rates.

Refer to caption
Figure 3: Excitation energy transfer rates between two QDS, one with a 3.9nm CdSe core and 3 monolayers CdS shell, the other with a 3.9nm CdSe core 4 monolayers CdS shell, as a function of center-to-center distances between QDs, at 300K. The purple triangles represent the rates obtained from Eq. (8) (labeled as “Redfield”); the blue triangles depict the rates calculated directly from Fermi’s Golden Rule (Förster theory) as in Eq. (1) (labeled as “Förster”). The black star is an experimental measurement between two CdSe QDs of different sizes, 3.85nm and 6.2nm in diameter, respectively.[51] The dash line depicts a R6superscript𝑅6R^{-6}italic_R start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT relationship and is a guide to the eye.
Refer to caption
Figure 4: Population dynamics for different light–matter interaction strengths and different temperatures. The dashed and solid lines represent results from an ideal and lossy cavity, respectively. The donor–acceptor distance was fixed at 6nm6nm6\thinspace\text{nm}6 nm. Three different QD-cavity couplings were considered, 3meV3meV3\thinspace\text{meV}3 meV, 10meV10meV10\,\text{meV}10 meV, and 60meV60meV60\,\text{meV}60 meV, corresponding to effective coupling of 0.5meV0.5meV0.5\thinspace{\rm meV}0.5 roman_meV, 4meV4meV4\thinspace{\rm meV}4 roman_meV, and 20meV20meV20\thinspace{\rm meV}20 roman_meV, respectively.

Fig. 2 presents the calculated absorption spectrum of a typical 3333 nm CdSe core with a 2-monolayer CdS shell, alongside the density of excitonic states. The absorption onset is marked by several distinct “bright” excitonic transitions, highlighted by vertical lines representing the oscillator strength. As excitonic energy increases, the density of excitons rises rapidly, with the appearance of many “dim” transitions arising from the high density of holes.[52, 37] The bright excitons play a crucial role in the energy transfer between the donor and acceptor, whereas the dim excitons are essential to account for dephasing and relaxation. To accurately describe both energy transfer dynamics and the relaxation to equilibrium, we include the 30303030 lowest excitonic states for each QD, namely 30303030 states for the donor and the acceptor. Fig. 2 also shows the calculated radiative lifetime for each excitonic state, obtained from

τnex=3c34En3|𝝁nex|2.superscriptsubscript𝜏𝑛ex3superscript𝑐34superscriptsubscript𝐸𝑛3superscriptsuperscriptsubscript𝝁𝑛ex2\tau_{n}^{{\rm ex}}=\frac{3c^{3}}{4E_{n}^{3}\left|\boldsymbol{\mu}_{n}^{{\rm ex% }}\right|^{2}}.italic_τ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT = divide start_ARG 3 italic_c start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT | bold_italic_μ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_ex end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (11)

The thermal average radiative lifetime, τex=nPnτnexsuperscript𝜏exsubscript𝑛subscript𝑃𝑛superscriptsubscript𝜏𝑛ex\tau^{\text{ex}}=\sum_{n}P_{n}\tau_{n}^{\text{ex}}italic_τ start_POSTSUPERSCRIPT ex end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ex end_POSTSUPERSCRIPT (Pnsubscript𝑃𝑛P_{n}italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the Boltzmann population of exciton |ψnketsubscript𝜓𝑛|\psi_{n}\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩), is found to be 24ns24ns24\thinspace\text{ns}24 ns, in very good agreement with experimental measurements, ranging between 2128212821-2821 - 28 ns.[53, 54, 55, 56]

In Fig. 3, we plot the calculated energy transfer rates as a function of the distance between the two QDs in the absence of a cavity. This is used to assess the accuracy of the Redfield approach in describing the energy transfer dynamics in the FRET limit, which can be compared to Förster perturbation theory as well as experimental values. The Redfield rates were obtained by fitting the calculated longtime population dynamics, ρ𝒟/𝒜(t)=n𝒟/𝒜σnn(t)subscript𝜌𝒟𝒜𝑡subscript𝑛𝒟𝒜subscript𝜎𝑛𝑛𝑡\rho_{{\cal D}/{\cal A}}\left(t\right)=\sum_{n\in{\cal{\cal D}}/{\cal A}}% \sigma_{nn}\left(t\right)italic_ρ start_POSTSUBSCRIPT caligraphic_D / caligraphic_A end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_n ∈ caligraphic_D / caligraphic_A end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT ( italic_t ), to an exponential decay:

ρ𝒟(t)=(1P𝒟eq)exp[kETt]+P𝒟eq,subscript𝜌𝒟𝑡1superscriptsubscript𝑃𝒟eqsubscript𝑘ET𝑡superscriptsubscript𝑃𝒟eq\rho_{{\cal D}}\left(t\right)=\left(1-P_{{\cal D}}^{\text{eq}}\right)\exp\left% [-k_{\text{ET}}t\right]+P_{{\cal D}}^{\text{eq}},italic_ρ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT ( italic_t ) = ( 1 - italic_P start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT start_POSTSUPERSCRIPT eq end_POSTSUPERSCRIPT ) roman_exp [ - italic_k start_POSTSUBSCRIPT ET end_POSTSUBSCRIPT italic_t ] + italic_P start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT start_POSTSUPERSCRIPT eq end_POSTSUPERSCRIPT , (12)

where P𝒟eqsuperscriptsubscript𝑃𝒟eqP_{{\cal D}}^{\text{eq}}italic_P start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT start_POSTSUPERSCRIPT eq end_POSTSUPERSCRIPT is the donor total exciton population at equilibrium and kETsubscript𝑘ETk_{\text{ET}}italic_k start_POSTSUBSCRIPT ET end_POSTSUBSCRIPT is the energy transition rate. For implementing Förster theory (Eq. (1)), we have convoluted the absorption transition with a Gaussian broadening function, with a width (60meVabsent60meV\approx 60\thinspace{\rm meV}≈ 60 roman_meV) taken from the calculated photoluminescence spectrum.[57] The energy transfer rates calculated by both methods agree quite well with each other (note that the Förster theory rate depends on the choice of the broadening) and both recover the R6superscript𝑅6R^{-6}italic_R start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT distance dependence, as expected. The calculated results are within a reasonable range compared to experimental measurements,[51, 58] also shown in Fig. 3.

IV Polaritonic energy transfer between QDs

We now turn to discuss the role of the cavity mode in the exciton transfer dynamics. We consider the transfer between two CdSe/CdS core-shell QDs with 3333 nm core size and 2 layers of CdS shell and a core-to-core distance of 6666 nm. The photon energy was set to ωc=2.06eVPlanck-constant-over-2-pisubscript𝜔𝑐2.06eV\hbar\omega_{c}=2.06\thinspace\text{eV}roman_ℏ italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 2.06 eV, slightly above the absorption onset of the QDs and the orientation of the QDs was fixed with the dipole moment parallel to the cavity polarization vector. The cavity coupling strength to the lowest donor excited state g𝒟1gPlanck-constant-over-2-pisubscript𝑔subscript𝒟1Planck-constant-over-2-pi𝑔\hbar g_{{\cal D}_{1}}\equiv\hbar groman_ℏ italic_g start_POSTSUBSCRIPT caligraphic_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≡ roman_ℏ italic_g is used as a measure of the QD-cavity coupling strength. We considered two scenarios, an ideal cavity without energy loss and a more realistic cavity with a photon escape time of[59] τcav=0.5pssubscript𝜏cav0.5ps\tau_{\text{cav}}=0.5\,\text{ps}italic_τ start_POSTSUBSCRIPT cav end_POSTSUBSCRIPT = 0.5 ps. We prepare the system in an initial thermal population of the excitonic states on the donor and a canonical distribution of all phonon modes.

Fig. 4 shows representative population dynamics for an ideal cavity (dashed lines) and a “lossy” cavity (solid lines) at two different temperatures and varying cavity-QD coupling values. The behavior of the donor and acceptor populations is qualitatively similar for the ideal and lossy cavities on timescales relevant to energy transfer. At low QD-cavity couplings (g=3 meVPlanck-constant-over-2-pi𝑔3 meV\hbar g=3\text{\,meV}roman_ℏ italic_g = 3 meV), the donor population dynamics exhibit a single exponential decay with a rate approximated by Eq. (12). The coupling of both QDs to the same cavity mode results in an effective direct coupling between them (see supporting information), estimated by tunneling splitting Jeff=12(E~LP+1E~LP)subscript𝐽eff12subscript~𝐸LP1subscript~𝐸LPJ_{{\rm eff}}=\frac{1}{2}\left(\tilde{E}_{{\rm LP+1}}-\tilde{E}_{{\rm LP}}\right)italic_J start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT roman_LP + 1 end_POSTSUBSCRIPT - over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT roman_LP end_POSTSUBSCRIPT ),[60, 61, 62] where E~LPsubscript~𝐸LP\tilde{E}_{{\rm LP}}over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT roman_LP end_POSTSUBSCRIPT and E~LP+1subscript~𝐸LP1\tilde{E}_{{\rm LP+1}}over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT roman_LP + 1 end_POSTSUBSCRIPT are the two lowest polaritonic energies obtained by diagonalizing HSsubscript𝐻SH_{{\rm S}}italic_H start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT, as described above. This effective coupling is much larger than the direct dipole-dipole coupling given by Eq. (3). As gPlanck-constant-over-2-pi𝑔\hbar groman_ℏ italic_g increases above 60meV60meV60\thinspace\text{meV}60 meV, the population dynamics display pronounced oscillations with a frequency approximately equal to 2Jeff2subscript𝐽eff2J_{\text{eff}}2 italic_J start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT. In this case, the exciton transfer rate is governed by decoherence time, especially at lower temperatures. A similar transition from incoherent to coherent dynamics has recently been reported for electron transfer between fused NC QDs.[63]

To better understand the transition from the overdamped to the coherent regime and make connections with previous works, we define a nonadiabatic parameter γ𝛾\gammaitalic_γ, given by:[64, 65]

γ=|Jeff|2ωnuπ3(λ𝒟+λ𝒜)kBT,𝛾superscriptsubscript𝐽eff2Planck-constant-over-2-pisubscript𝜔nusuperscript𝜋3subscript𝜆𝒟subscript𝜆𝒜subscript𝑘𝐵𝑇\gamma=\frac{\left|J_{{\rm eff}}\right|^{2}}{\hbar\omega_{{\rm nu}}}\sqrt{% \frac{\pi^{3}}{\left(\lambda_{{\cal D}}+\lambda_{{\cal A}}\right)k_{B}T}},italic_γ = divide start_ARG | italic_J start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ italic_ω start_POSTSUBSCRIPT roman_nu end_POSTSUBSCRIPT end_ARG square-root start_ARG divide start_ARG italic_π start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_λ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT ) italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG end_ARG , (13)

where ωnusubscript𝜔nu\omega_{{\rm nu}}italic_ω start_POSTSUBSCRIPT roman_nu end_POSTSUBSCRIPT is the averaged effective phonon modes that contribute most to the system dynamics. The terms λ𝒟subscript𝜆𝒟\lambda_{{\cal D}}italic_λ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT and λ𝒜subscript𝜆𝒜\lambda_{{\cal A}}italic_λ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT represent the average reorganization energies for the donor and acceptor, respectively, defined as λ𝒟/𝒜=n𝒟/𝒜Pnλnsubscript𝜆𝒟𝒜subscript𝑛𝒟𝒜subscript𝑃𝑛subscript𝜆𝑛\lambda_{{\cal D}/{\cal A}}=\sum_{n\in{\cal D}/{\cal A}}P_{n}\lambda_{n}italic_λ start_POSTSUBSCRIPT caligraphic_D / caligraphic_A end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_n ∈ caligraphic_D / caligraphic_A end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, where Pnsubscript𝑃𝑛P_{n}italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the Boltzmann population distribution of exciton |ψnketsubscript𝜓𝑛|\psi_{n}\rangle| italic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ and, as before, λn=12α(V~nnα)2/ωα2subscript𝜆𝑛12subscript𝛼superscriptsuperscriptsubscript~𝑉𝑛𝑛𝛼2superscriptsubscript𝜔𝛼2\lambda_{n}=\frac{1}{2}\sum_{\alpha}\left(\tilde{V}_{nn}^{\alpha}\right)^{2}/% \omega_{\alpha}^{2}italic_λ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_n italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_ω start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In Fig. 5(a) we plot the nonadiabatic parameter γ𝛾\gammaitalic_γ as a function of the QD-cavity coupling strength gPlanck-constant-over-2-pi𝑔\hbar groman_ℏ italic_g. For small coupling to the cavity mode, γ1much-less-than𝛾1\gamma\ll 1italic_γ ≪ 1 and the system is in the nonadiabatic regime, where the dynamics are characterized by an exponential decay of populations. As the coupling to the cavity mode increases, γ𝛾\gammaitalic_γ also increases due to the increase in the effective coupling between the two QDs. This parameter reaches and surpasses γ=1𝛾1\gamma=1italic_γ = 1 when g=20Planck-constant-over-2-pi𝑔20\hbar g=20roman_ℏ italic_g = 20 meV, marking the transition to an adiabatic regime. In this regime, the effective coupling strength between the QDs becomes significantly larger than the couplings of the donor and acceptor polaritonic states to phonons, leading to more pronounced and sustained coherent dynamics.

Refer to caption
Figure 5: (a) Plot of the nonadiabatic parameter γ𝛾\gammaitalic_γ versus the QD-cavity coupling strength gPlanck-constant-over-2-pi𝑔\hbar groman_ℏ italic_g at 300K. The dashed line shows γ=1𝛾1\gamma=1italic_γ = 1. (b) Energy transfer rate kETsubscript𝑘ETk_{\text{ET}}italic_k start_POSTSUBSCRIPT ET end_POSTSUBSCRIPT as a function of effective coupling Jeffsubscript𝐽effJ_{\text{eff}}italic_J start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT for ideal cavity (orange dots) and lossy cavity (red hollow circles) at 300K. The black dashed lines show a quadratic and square root dependence and the gray dashed line corresponds to an effective coupling for γ=1𝛾1\gamma=1italic_γ = 1.

In Fig. 5(b), we present the exciton transfer rate as a function of the effective coupling for both ideal and lossy cavities. For small value of Jeffsubscript𝐽effJ_{\text{eff}}italic_J start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT, the transfer rate scales with Jeff2superscriptsubscript𝐽eff2J_{\text{eff}}^{2}italic_J start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, consistent with Fermi’s golden rule. As Jeffsubscript𝐽effJ_{\text{eff}}italic_J start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT increases, the system transitions to the adiabatic regime, and the transfer rate which is determined by decoherence times exhibits a more moderate dependence, fitting to a Jeffsubscript𝐽eff\sqrt{J_{\text{eff}}}square-root start_ARG italic_J start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG relationship. In this limit, the transfer rate was obtained by fitting the envelop function of ρ𝒟(t)subscript𝜌𝒟𝑡\rho_{{\cal D}}\left(t\right)italic_ρ start_POSTSUBSCRIPT caligraphic_D end_POSTSUBSCRIPT ( italic_t )/ρ𝒜(t)subscript𝜌𝒜𝑡\rho_{{\cal A}}\left(t\right)italic_ρ start_POSTSUBSCRIPT caligraphic_A end_POSTSUBSCRIPT ( italic_t ) to an exponential decay/rise. The gray dashed line in the figure marks the crossover coupling point, where the adiabatic parameter γ=1𝛾1\gamma=1italic_γ = 1. In the lossy cavity, the energy transfer rate is inherently slower when compared to an ideal, lossless cavity. However, as the coupling to the cavity mode is increased, this discrepancy becomes less significant. This is because the rate of energy transfer accelerates, eventually becoming so rapid that it surpasses the cavity’s inherent loss rate. As a result, the effects of the cavity’s losses are effectively mitigated, rendering them negligible in the context of the swift energy transfer times that are achieved.

Before we conclude, we would like to mention that our calculations have been performed for a single QD pair and such large couplings required to drive the system into the adiabatic limit are typically achieved in plasmonic nanocavities, with complications arising from the coupling of excitons to near-fields.[66] While our approach ignores the role of near fields, it provides the guidelines that would be useful in future studies of QDs in nanocavities.

V Conclusion

In this study, we investigated the dynamics of exciton energy transfer between two nanocrystal quantum dots positioned within an optical microcavity. We delved into this intricate phenomenon, examining how the cavity photon mode influences the energy transfer process between the quantum dots, essentially serving as a conduit. To achieve this, we developed a model Hamiltonian, parameterized by atomistic calculations of the excitonic level structure, phonon, exciton-phonon couplings, and transition dipole moments of the semiconductor NCs within the microcavity. To describe the dynamics comprehensively, we utilized a quantum master equation and approximated its kernel to second order in the exciton-phonon coupling (Redfield with memory), while also performing a small polaron transformation to account for multiphonon relaxation channels.

Our investigation revealed that by fine-tuning the coupling strength between the cavity photon mode and the NCs, we can effectively modulate the interaction between the donor and acceptor. Particularly noteworthy is the identification of a critical transition from a nonadiabatic to an adiabatic process, wherein the exciton energy transition rate is accelerated and governed by decoherence times. This transition holds significant implications for the controllability and efficiency of energy transfer processes at the nanoscale. Overall, our findings lay the groundwork for the design of innovative energy transfer systems and nanophotonic devices, promising enhanced functionality and performance in various applications.

Supplementary Material

The supplementary material comprises nanocrystal configurations, simplification of polaron transformation, time–local non-Markovian Redfield equation, simplification of correlation function Wnm(t)Wkl(0)Bsubscriptdelimited-⟨⟩subscript𝑊𝑛𝑚𝑡subscript𝑊𝑘𝑙0B\left\langle W_{nm}\left(t\right)W_{kl}\left(0\right)\right\rangle_{\text{B}}⟨ italic_W start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT ( italic_t ) italic_W start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT ( 0 ) ⟩ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT, and derivation of the effective coupling for a 3-level system.

Acknowledgements.
This work was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division, under Contract No. DE461AC02-05-CH11231 within the Fundamentals of Semiconductor Nanowire Program (KCPY23). Computational resources were provided in part by the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility operated under contract no. DEAC02-05CH11231.

Author Declarations

Conflict of Interest Statement

The authors have no conflicts to disclose.

Author Contributions

Kaiyue Peng: Data curation (lead); Formal analysis (lead); Software (lead), Methodology (equal); Project administration (equal); Conceptualization (equal); Validation (lead); Writing – original draft (equal). Eran Rabani: Funding Acquisition (lead), Resources (lead), Supervision (lead), Writing/Review and Editing (lead), Methodology (equal), Project Administration (equal), Conceptualization (equal); Writing – original draft (equal).

Data Availability Statement

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

References

  • Choi et al. [2012] S. Choi, H. Jin, J. Bang, and S. Kim, “Layer-by-layer quantum dot assemblies for the enhanced energy transfers and their applications toward efficient solar cells,” J. Phys. Chem. Lett. 3, 3442–3447 (2012).
  • Almutlaq et al. [2024] J. Almutlaq, Y. Liu, W. J. Mir, R. P. Sabatini, D. Englund, O. M. Bakr, and E. H. Sargent, “Engineering colloidal semiconductor nanocrystals for quantum information processing,” Nat. Nanotechnol. , 1–10 (2024).
  • Alivisatos [1996] A. P. Alivisatos, “Perspectives on the physical chemistry of semiconductor nanocrystals,” J. Phys. Chem. 100, 13226–13239 (1996).
  • Efros and Rosen [2000] A. L. Efros and M. Rosen, “The electronic structure of semiconductor nanocrystals,” Annu. Rev. Mater. Sci. 30, 475–521 (2000).
  • Gomez, Califano, and Mulvaney [2006] D. E. Gomez, M. Califano, and P. Mulvaney, “Optical properties of single semiconductor nanocrystals,” Phys. Chem. Chem. Phys. 8, 4989–5011 (2006).
  • Klimov [2007] V. I. Klimov, “Spectral and dynamical properties of multiexcitons in semiconductor nanocrystals,” Annu. Rev. Phys. Chem. 58, 635–673 (2007).
  • Kodaimati et al. [2018] M. S. Kodaimati, S. Lian, G. C. Schatz, and E. A. Weiss, “Energy transfer-enhanced photocatalytic reduction of protons within quantum dot light-harvesting–catalyst assemblies,” Proc. Natl. Acad. Sci. U. S. A. 115, 8290–8295 (2018).
  • Roy et al. [2020] P. Roy, G. Devatha, S. Roy, A. Rao, and P. P. Pillai, “Electrostatically driven resonance energy transfer in an all-quantum dot based donor–acceptor system,” J. Phys. Chem. Lett. 11, 5354–5360 (2020).
  • Kodaimati et al. [2017] M. S. Kodaimati, C. Wang, C. Chapman, G. C. Schatz, and E. A. Weiss, “Distance-dependence of interparticle energy transfer in the near-infrared within electrostatic assemblies of pbs quantum dots,” ACS nano 11, 5041–5050 (2017).
  • Scholes, Jordanides, and Fleming [2001] G. D. Scholes, X. J. Jordanides, and G. R. Fleming, “Adapting the förster theory of energy transfer for modeling dynamics in aggregated molecular assemblies,” J. Phys. Chem. B 105, 1640–1651 (2001).
  • Barth et al. [2022] A. Barth, O. Opanasyuk, T.-O. Peulen, S. Felekyan, S. Kalinin, H. Sanabria, and C. A. M. Seidel, “Unraveling multi-state molecular dynamics in single-molecule FRET experiments. I. Theory of FRET-lines,” J. Chem. Phys. 156, 141501 (2022).
  • Guo et al. [2018] J. Guo, X. Qiu, C. Mingoes, J. R. Deschamps, K. Susumu, I. L. Medintz, and N. Hildebrandt, “Conformational details of quantum dot-dna resolved by forster resonance energy transfer lifetime nanoruler,” ACS nano 13, 505–514 (2018).
  • Jares-Erijman and Jovin [2003] E. A. Jares-Erijman and T. M. Jovin, “Fret imaging,” Nat. Biotechnol. 21, 1387–1395 (2003).
  • Kim, Kim, and Lee [2015] J.-Y. Kim, C. Kim, and N. K. Lee, “Real-time submillisecond single-molecule fret dynamics of freely diffusing molecules with liposome tethering,” Nat. Commun. 6, 6992 (2015).
  • Sun et al. [2011] Y. Sun, H. Wallrabe, S.-A. Seo, and A. Periasamy, “Fret microscopy in 2010: the legacy of theodor förster on the 100th anniversary of his birth,” ChemPhysChem 12, 462–474 (2011).
  • Zhong et al. [2023] K. Zhong, H. L. Nguyen, T. N. Do, H.-S. Tan, J. Knoester, and T. L. C. Jansen, “An efficient time-domain implementation of the multichromophoric Förster resonant energy transfer method,” J. Chem. Phys. 158, 064103 (2023).
  • Baer and Rabani [2008] R. Baer and E. Rabani, “Theory of resonance energy transfer involving nanocrystals: The role of high multipoles,” J. Chem. Phys. 128, 184710 (2008).
  • Loiudice, Saris, and Buonsanti [2020] A. Loiudice, S. Saris, and R. Buonsanti, “Tunable metal oxide shell as a spacer to study energy transfer in semiconductor nanocrystals,” J. Phys. Chem. Lett. 11, 3430–3435 (2020).
  • Govorov, Lee, and Kotov [2007] A. O. Govorov, J. Lee, and N. A. Kotov, “Theory of plasmon-enhanced förster energy transfer in optically excited semiconductor and metal nanoparticles,” Phys. Rev. B 76, 125308 (2007).
  • Schachenmayer et al. [2015] J. Schachenmayer, C. Genes, E. Tignone, and G. Pupillo, “Cavity-enhanced transport of excitons,” Phys. Rev. Lett. 114, 196403 (2015).
  • Wang, Hertzog, and Börjesson [2021] M. Wang, M. Hertzog, and K. Börjesson, “Polariton-assisted excitation energy channeling in organic heterojunctions,” Nat. Commun. 12, 1874 (2021).
  • Chowdhury, Zhang, and Beratan [2022] S. N. Chowdhury, P. Zhang, and D. N. Beratan, “Interference between molecular and photon field-mediated electron transfer coupling pathways in cavities,” J. Phys. Chem. Lett. 13, 9822–9828 (2022).
  • Albinsson and Mårtensson [2010] B. Albinsson and J. Mårtensson, “Excitation energy transfer in donor–bridge–acceptor systems,” Phys. Chem. Chem. Phys. 12, 7338–7351 (2010).
  • Xu et al. [2023] D. Xu, A. Mandal, J. M. Baxter, S.-W. Cheng, I. Lee, H. Su, S. Liu, D. R. Reichman, and M. Delor, “Ultrafast imaging of polariton propagation and interactions,” Nat. Commun. 14, 3881 (2023).
  • Feist and Garcia-Vidal [2015] J. Feist and F. J. Garcia-Vidal, “Extraordinary exciton conductance induced by strong coupling,” Phys. Rev. Lett. 114, 196402 (2015).
  • Zhong et al. [2017] X. Zhong, T. Chervy, L. Zhang, A. Thomas, J. George, C. Genet, J. A. Hutchison, and T. W. Ebbesen, “Energy transfer between spatially separated entangled molecules,” Angew. Chem. 129, 9162–9166 (2017).
  • Weight, Krauss, and Huo [2023] B. M. Weight, T. D. Krauss, and P. Huo, “Investigating molecular exciton polaritons using ab initio cavity quantum electrodynamics,” J. Phys. Chem. Lett. 14, 5901–5913 (2023).
  • Du et al. [2018] M. Du, L. A. Martínez-Martínez, R. F. Ribeiro, Z. Hu, V. M. Menon, and J. Yuen-Zhou, “Theory for polariton-assisted remote energy transfer,” Chem. Sci. 9, 6659–6669 (2018).
  • Kimura [2009] A. Kimura, “General theory of excitation energy transfer in donor-mediator-acceptor systems,” J. Chem. Phys. 130, 154103 (2009).
  • Sáez-Blázquez et al. [2018] R. Sáez-Blázquez, J. Feist, A. I. Fernández-Domínguez, and F. García-Vidal, “Organic polaritons enable local vibrations to drive long-range energy transfer,” Phys. Rev. B 97, 241407 (2018).
  • Reitz, Mineo, and Genes [2018] M. Reitz, F. Mineo, and C. Genes, “Energy transfer and correlations in cavity-embedded donor-acceptor configurations,” Sci. Rep. 8, 9050 (2018).
  • Coles et al. [2014] D. M. Coles, N. Somaschi, P. Michetti, C. Clark, P. G. Lagoudakis, P. G. Savvidis, and D. G. Lidzey, “Polariton-mediated energy transfer between organic dyes in a strongly coupled optical microcavity,” Nat. Mater. 13, 712–719 (2014).
  • Xiang et al. [2020] B. Xiang, R. F. Ribeiro, M. Du, L. Chen, Z. Yang, J. Wang, J. Yuen-Zhou, and W. Xiong, “Intermolecular vibrational energy transfer enabled by microcavity strong light–matter coupling,” Science 368, 665–667 (2020).
  • Philbin and Rabani [2018] J. P. Philbin and E. Rabani, “Electron–hole correlations govern auger recombination in nanostructures,” Nano Lett. 18, 7889–7895 (2018).
  • Jasrasaria and Rabani [2021] D. Jasrasaria and E. Rabani, “Interplay of surface and interior modes in exciton–phonon coupling at the nanoscale,” Nano Lett. 21, 8741–8748 (2021).
  • Jasrasaria et al. [2022] D. Jasrasaria, D. Weinberg, J. P. Philbin, and E. Rabani, “Simulations of nonradiative processes in semiconductor nanocrystals,” J. Chem. Phys. 157, 020901 (2022).
  • Jasrasaria and Rabani [2023] D. Jasrasaria and E. Rabani, “Circumventing the phonon bottleneck by multiphonon-mediated hot exciton cooling at the nanoscale,” Npj Comput. Mater. 9, 1–8 (2023).
  • Wall and Neuhauser [1995] M. R. Wall and D. Neuhauser, “Extraction, through filter-diagonalization, of general quantum eigenvalues or classical normal mode frequencies from a small number of residues or a short-time segment of a signal. i. theory and application to a quantum-dynamics model,” J. Chem. Phys. 102, 8011–8022 (1995).
  • Toledo and Rabani [2002] S. Toledo and E. Rabani, “Very large electronic structure calculations using an out-of-core filter-diagonalization method,” J. Comput. Phys. 180, 256–269 (2002).
  • Rohlfing and Louie [2000] M. Rohlfing and S. G. Louie, “Electron-hole excitations and optical spectra from first principles,” Phys. Rev. B 62, 4927 (2000).
  • Wang and Zunger [1996] L.-W. Wang and A. Zunger, “Pseudopotential calculations of nanoscale cdse quantum dots,” Phys. Rev. B 53, 9579 (1996).
  • Latt, Cheung, and Blout [1965] S. Latt, H. Cheung, and E. Blout, “Energy transfer. a system with relatively fixed donor-acceptor separation,” J. Am. Chem. Soc. 87, 995–1003 (1965).
  • Fedchenia and Westlund [1994] I. Fedchenia and P.-O. Westlund, “Influence of molecular reorientation on electronic energy transfer between a pair of mobile chromophores: The stochastic liouville equation combined with brownian dynamic simulation techniques,” Phys. Rev. E. 50, 555 (1994).
  • Zhou et al. [2013] X. Zhou, D. Ward, J. Martin, F. Van Swol, J. Cruz-Campa, and D. Zubia, “Stillinger-weber potential for the ii-vi elements zn-cd-hg-s-se-te,” Phys. Rev. B 88, 085309 (2013).
  • Mandal, Krauss, and Huo [2020] A. Mandal, T. D. Krauss, and P. Huo, “Polariton-mediated electron transfer via cavity quantum electrodynamics,” J. Phys. Chem. B 124, 6321–6340 (2020).
  • Peng and Rabani [2023] K. Peng and E. Rabani, “Polaritonic bottleneck in colloidal quantum dots,” Nano Lett. 23, 10587–10593 (2023).
  • Rokaj et al. [2018] V. Rokaj, D. M. Welakuh, M. Ruggenthaler, and A. Rubio, “Light-matter interaction in the long-wavelength limit: No ground-statewithout dipole self-energy,” J. Phys. B: At., Mol. Opt. Phys. 51, 034005 (2018).
  • Sidler, Ruggenthaler, and Rubio [2023] D. Sidler, M. Ruggenthaler, and A. Rubio, “Numerically exact solution for a real polaritonic system under vibrational strong coupling in thermodynamic equilibrium: loss of light–matter entanglement and enhanced fluctuations,” J. Chem. Theory Comput. 19, 8801–8814 (2023).
  • Pérez-Sánchez et al. [2023] 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. 120, e2219223120 (2023).
  • Nitzan [2006] A. Nitzan, Chemical dynamics in condensed phases: relaxation, transfer and reactions in condensed molecular systems (Oxford university press, 2006).
  • Kagan et al. [1996] C. Kagan, C. Murray, M. Nirmal, and M. Bawendi, “Electronic energy transfer in cdse quantum dot solids,” Phys. Rev. Lett. 76, 1517 (1996).
  • Brosseau et al. [2023] P. J. Brosseau, J. J. Geuchies, D. Jasrasaria, A. J. Houtepen, E. Rabani, and P. Kambhampati, “Ultrafast hole relaxation dynamics in quantum dots revealed by two-dimensional electronic spectroscopy,” Commun. Phys. 6, 48 (2023).
  • Rabouw et al. [2015] F. T. Rabouw, M. Kamp, R. J. van Dijk-Moes, D. R. Gamelin, A. F. Koenderink, A. Meijerink, and D. Vanmaekelbergh, “Delayed exciton emission and its relation to blinking in cdse quantum dots,” Nano Lett. 15, 7718–7725 (2015).
  • Bae et al. [2013a] W. K. Bae, L. A. Padilha, Y.-S. Park, H. McDaniel, I. Robel, J. M. Pietryga, and V. I. Klimov, “Controlled alloying of the core–shell interface in cdse/cds quantum dots for suppression of auger recombination,” ACS nano 7, 3411–3419 (2013a).
  • Bae et al. [2013b] W. K. Bae, Y.-S. Park, J. Lim, D. Lee, L. A. Padilha, H. McDaniel, I. Robel, C. Lee, J. M. Pietryga, and V. I. Klimov, “Controlling the influence of auger recombination on the performance of quantum-dot light-emitting diodes,” Nat. Commun. 4, 2661 (2013b).
  • Crooker et al. [2002] S. Crooker, J. Hollingsworth, S. Tretiak, and V. I. Klimov, “Spectrally resolved dynamics of energy transfer in quantum-dot assemblies: towards engineered energy flows in artificial materials,” Phys. Rev. Lett. 89, 186802 (2002).
  • Lin et al. [2023] K. Lin, D. Jasrasaria, J. J. Yoo, M. Bawendi, H. Utzat, and E. Rabani, “Theory of photoluminescence spectral line shapes of semiconductor nanocrystals,” J. Phys. Chem. Lett. 14, 7241–7248 (2023).
  • Lingley, Lu, and Madhukar [2011] Z. Lingley, S. Lu, and A. Madhukar, “A high quantum efficiency preserving approach to ligand exchange on lead sulfide quantum dots and interdot resonant energy transfer,” Nano Lett. 11, 2887–2891 (2011).
  • le Feber et al. [2018] B. le Feber, F. Prins, E. De Leo, F. T. Rabouw, and D. J. Norris, “Colloidal-quantum-dot ring lasers with active color control,” Nano Lett. 18, 1028–1034 (2018).
  • Davis et al. [2009] D. Davis, M. C. Toroker, S. Speiser, and U. Peskin, “On the effect of nuclear bridge modes on donor–acceptor electronic coupling in donor–bridge–acceptor molecules,” Chem. Phys. 358, 45–51 (2009).
  • Skourtis, Archontis, and Xie [2001] S. S. Skourtis, G. Archontis, and Q. Xie, “Electron transfer through fluctuating bridges: On the validity of the superexchange mechanism and time-dependent tunneling matrix elements,” J. Chem. Phys. 115, 9444–9462 (2001).
  • Priyadarshy et al. [1996] S. Priyadarshy, S. S. Skourtis, S. M. Risser, and D. N. Beratan, “Bridge-mediated electronic interactions: Differences between hamiltonian and green function partitioning in a non-orthogonal basis,” J. Chem. Phys. 104, 9473–9481 (1996).
  • Hou et al. [2023] B. Hou, M. Thoss, U. Banin, and E. Rabani, “Incoherent nonadiabatic to coherent adiabatic transition of electron transfer in colloidal quantum dot molecules,” Nat. Commun. 14, 3073 (2023).
  • Newton and Sutin [1984] M. D. Newton and N. Sutin, “Electron transfer reactions in condensed phases,” Annu. Rev. Phys. Chem. 35, 437–480 (1984).
  • Balzani et al. [2001] V. Balzani, P. Piotrowiak, M. Rodgers, J. Mattay, D. Astruc, et al.Electron transfer in chemistry, Vol. 1 (Wiley-VCh Weinheim, 2001).
  • Chikkaraddy et al. [2016] R. Chikkaraddy, B. De Nijs, F. Benz, S. J. Barrow, O. A. Scherman, E. Rosta, A. Demetriadou, P. Fox, O. Hess, and J. J. Baumberg, “Single-molecule strong coupling at room temperature in plasmonic nanocavities,” Nature 535, 127–130 (2016).