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

Degenerate parametric down-conversion facilitated by exciton-plasmon polariton states in nonlinear plasmonic cavity

Andrei Piryatinski [email protected] Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA    Maxim Sukharev Department of Physics, Arizona State University, Tempe, AZ 85287, USA College of Integrative Sciences and Arts, Arizona State University, Mesa, AZ 85201, USA
Abstract

We study the effect of degenerate parametric down-conversion (DPDC) in an ensemble of two-level quantum emitters (QEs) coupled via near-field interactions to a single surface plasmon (SP) mode of a nonlinear plasmonic cavity. For this purpose, we develop a quantum driven-dissipative model capturing non-equilibrium dynamics of the system in which incoherently pumped QEs have transition frequency tuned near the second-harmonic response of the SPs. Considering the strong coupling regime, i.e., the SP-QE interaction rate exceeds system dissipation rates, we find a critical SP-QE coupling attributed to the phase transition between normal and lasing steady states. Examining fluctuations above the system’s steady states, we predict new elementary excitations, namely, the exciton-plasmon polaritons formed by the two-SP quanta and single-exciton states of QEs. The contribution of two-SP quanta results in the linear scaling of the SP-QE interaction rate with the number of QEs, 𝒩o{\cal N}_{o}, as opposed to the 𝒩o\sqrt{{\cal N}_{o}}-scaling known for the Dicke and Tavis-Cummings models. We further examine how SP-QE interaction scaling affects the polariton dispersions and power spectra in the vicinity of the critical coupling. For this purpose, we compare the calculation results assuming a finite ensemble of QEs and the model thermodynamic limit. The calculated power spectra predict an interplay of coherent photon emission by QEs near the second-harmonic frequency and correlated photon-pair emission at the fundamental frequency by the SPs (i.e., the photonic DPDC effect).

I Introduction

Plasmonics has been in the spotlight in engineering, GramotnevNatPhot:2010 chemistry, EbbesenAcCemRes:2016 ; FeistASCNano:2018 and biology HuangLaserSurgMed:2007 for more than a decade. StockmanOptEx:2011 During this time, fascinating progress in nanofabrication provided new tools to create metal-dielectric, i.e., plasmonic, interfaces with nanometer precision.PanoiuJ.Opt:2018 Even though the surface plasmon (SP) modes in such nanostructures show relatively low characteristic quality factors, large local field enhancements WuACSPhot:2021 make them very attractive for applications. EbbesenAcCemRes:2016 ; JiangChemRev:2017 On the other hand, ultrafast optical detection techniques have offered time resolution reaching a few femtosecond limit. VasaNatPhot:2013 These fabrication and characterization tools tremendously assist us in advancing our knowledge of the light-matter interactions at the nanoscale, where the quantum and classical properties of light and matter merge.TormaRepPrgoPhys:2014

Despite such progress, the nonlinear plasmonics relying on the strong local field enhancement GordonACSPhot:2018 and high nonlinear electron susceptibilities in metal KrasavinLaserPhotRev:2018 is still in its infancy.PanoiuJ.Opt:2018 Several recent reports address nonlinear phenomena such as second/third-harmonic generation GalantyLightSciTech:2018 ; WeissmanAdvOptMat:2017 ; YooOptExp:2019 , four-wave mixing AlmeidaSciRep:2015 ; BlechmanNanoLett:2018 , and optical bistability. WurtzPRL:2006 Moreover, the use of coherent control was reported to study the second-harmonic generation (SHG) by gold nanoparticles combining spatial localization of electromagnetic fields with laser pulse shaping.BaharPRB:2020

Quantum plasmonics interprets the metallic nanostructure as a plasmonic cavity where strong local electric fields enhance the spontaneous decay of organic dyes or semiconductor quantum dots representing quantum emitters (QEs).TameNatPhys:2013 This effect is known as the plasmonic Purcell effect. Advances beyond demonstration of the Purcell effect hinge on the use of 2D plasmonic lattices facilitating delocalized hybrid plasmonic-diffraction surface lattice resonance of relatively high quality factor (low energy losses).CherquiAccChemRes:2019 Use of these structures allowed researchers to reach a strong coupling regime between QE and SP in which the SP-QE quantum exchange rate exceeds plasmonic losses and results in the formation of mixed QE and SP exciton-plasmon polariton stats. RamezaniJOSAB:2019 Current theoretical and experimental studies of the exciton-polariton states focus on plasmonic lasing Pusch_ACSnano:2012 ; Guan_ASCnano:2020 ; Winkler_ACSnano:2020 , polariton Bose Einstein condensation, and polariton lasing. MartikainenPRA:2014 ; ZasterJPCS:2016 ; DeGiorgiPEP:2018 ; HakalaNP:2018 All these studies asume linear SP response with the QEs providing source of nonlinearity.

The strong SP-QE coupling regime has recently become a spotlight in nonlinear plasmonics, promising coherent photonic up/down-conversion and lasing applications. One of the authors contributed to a theoretical study demonstrating that strongly coupled nonlinear SP systems and QEs may exhibit unique second-order optical signals facilitated by mixed exciton-plasmon polariton states.DrobnyhJCP:2020 ; SukharevJCP:2021 Subsequent experimental studies confirmed the predictions. LiNanoLett:2020 In another theory report, we examined the effect of QE gain on the SHG by a plasmonic nanostructure, e.g., 2D plasmonic lattice, exhibiting nonlinear response.Sukharev_JCP:2021 For the QEs strongly coupled with fundamental SP mode, we found a lasing state contributing to efficient photon emission at the second-harmonic frequency. In that report, we also introduced a concept of the nonlinear plasmonic cavity in which nonlinear interaction occurs between quantized fundamental and second-harmonic plasmonic modes. Noteworthy that photonic cavities demonstrate nonlinearity due to embedded QEs only. RibeiroPRA:2021

In this paper, we use the concept of nonlinear SP cavity to investigate a reverse effect of the degenerate parametric down-conversion (DPDC) of excited QEs (excitons) strongly coupled to the second-harmonic of the cavity via near-field interactions. For a localized SP mode, the effects of SHG and parametric down-conversion have been studied experimentally in light of nonlinear photon-plasmon interactions.GrossePRL:2012 ; HeckmannOptExp:2013 The authors of Refs. LootJOpt:2018 ; LootOptik:2018 considered spontaneous parametric downconversion in nonlinear dielectric media interfacing with metal surfaces so that the linear SP response of the metal enhances the process via weak local field coupling. None of these reports considers the effects of QEs.

Therefore, we focus our paper on the case in which QEs forming gain medium exchange energy via the near-filed interactions with the second-harmonic of a nonlinear plasmonic cavity. As demonstrated below, the presence of QEs results in new exciton-plasmon polariton states mixing two-SP quanta with a single exciton state. Considering the strong coupling regime, we predict a non-trivial lasing phase supporting these polariton states as elementary excitations. Our analysis of the polariton power spectra predicts coherent photon emission by QEs around the second harmonic frequency and correlated photon pairs at the fundamental frequency by the SP mode. Given the SP photon emission rate significantly exceeds the same rate for QEs, we predict the photon DPDC effect in such a system.

We organize the paper as follows. Sec. II presents our quantum open system theory of the DPDC, identifying lasing steady states and providing expressions for the exciton-plasmon polariton dispersion and fluctuation auto-correlation functions. Sec. III presents numerical calculations and discussion of the polariton dispersion relations and the auto-correlation functions. Finally, we conclude in Sec. IV.

II Quantum driven-dissipative model

Refer to caption
Figure 1: Schematics of a plasmonic cavity formed by an ensemble of metal nanoparticles (MNPs) supporting a delocalized SP mode (red wave line). QEs whose energy ωo\hbar\omega_{o} is tuned near the second-harmonic response of the SP mode are coupled to the SP mode via the near-field interactions. The QEs are subject to incoherent pump providing energy for the DPDC which results in the formation of correlated two-SP states of energy ωsp\sim\hbar\omega_{sp}. The cavity can radiate correlated two-photon pairs of energy 2ω2\hbar\omega and photons of energy ω\hbar\omega due the two-SP and QE radiative decay, respectively.

As illustrated in Fig. 1, we introduce an ensemble of identical two-level QEs with the transition frequency ωo\omega_{o}. The QEs occupy sites n=1,𝒩o¯n=\overline{1,{\cal N}_{o}} and interact with a delocalized SP mode whose fundamental frequency is ωsp\omega_{sp}. The SP mode is assumed to have second-order nonlinear response allowing for the SHG, i.e., creation of a single SP quantum with frequency 2ωsp2\omega_{sp} or a down-conversion of a single SP quantum with frequency 2ωsp2\omega_{sp} to two quanta each having frequency ωsp\omega_{sp}. By tuning the QEs around the SP second-harmonic frequency, i.e., ωo2ωsp\omega_{o}\sim 2\omega_{sp}, we allow for the down-conversion of the SP quanta produced by QEs. The interaction between each QE and SP is characterized by the quantum exchange rate λ\lambda proportional to the SP second-order susceptibility χsp(2)\chi_{sp}^{(2)}.

For the introduced system, the Hamiltonian, assuming rotating wave approximation for the SP-QE interaction term, reads

^\displaystyle\hat{\cal H} =\displaystyle= ωspa^a^+ωon=1𝒩o(s^nz+12)\displaystyle\hbar\omega_{sp}\hat{a}^{\dagger}\hat{a}+\hbar\omega_{o}\sum_{n=1}^{{\cal N}_{o}}\left(\hat{s}_{n}^{z}+\frac{1}{2}\right)
+\displaystyle+ λn=1𝒩o(a^2s^n+s^n+a^2).\displaystyle\hbar\lambda\sum_{n=1}^{{\cal N}_{o}}\left(\hat{a}^{{\dagger}2}\hat{s}_{n}^{-}+\hat{s}_{n}^{+}\hat{a}^{2}\right).

Here, the SP mode is described by Bose creation (annihilation) operator a^\hat{a}^{\dagger} (a^\hat{a}), and the two-level QEs by spin operators s^n±=s^nx±is^ny\hat{s}_{n}^{\pm}=\hat{s}_{n}^{x}\pm i\hat{s}^{y}_{n}, snzs_{n}^{z}, related to the Pauli matrices as s^nj=12σ^nj\hat{s}_{n}^{j}=\frac{1}{2}\hat{\sigma}_{n}^{j} where j=x,y,zj=x,y,z.

Both the SPs and QEs are subject to energy supply originating form incoherent excitation of the QEs and dissipation processes due to environment fluctuations. Thus, the time-evolution of such an open quantum system characterized by an operator set 𝒪^={a^,s^n,s^nz}\hat{\cal O}=\{\hat{a},\hat{s}_{n}^{-},\hat{s}_{n}^{z}\} can be described by the Heisenberg equation of motion

t𝒪^\displaystyle\partial_{t}\hat{\cal O} =\displaystyle= i[^,𝒪^]+2γsp^a^[𝒪^]+γn=1𝒩o^s^n+[𝒪^]\displaystyle\frac{i}{\hbar}\left[\hat{\cal H},\hat{\cal O}\right]+2\gamma_{sp}\hat{\cal L}_{\hat{a}}[\hat{\cal O}]+\gamma_{\uparrow}\sum\limits_{n=1}^{{\cal N}_{o}}\hat{\cal L}_{\hat{s}^{+}_{n}}[\hat{\cal O}]
+\displaystyle+ γn=1𝒩o^s^n[𝒪^]+γϕn=1𝒩o^s^nz[𝒪^],\displaystyle\gamma_{\downarrow}\sum\limits_{n=1}^{{\cal N}_{o}}\hat{\cal L}_{\hat{s}^{-}_{n}}[\hat{\cal O}]+\gamma_{\phi}\sum\limits_{n=1}^{{\cal N}_{o}}\hat{\cal L}_{\hat{s}_{n}^{z}}[\hat{\cal O}],

where the first term in the right-hand side accounts for the coherent dynamics governed by the Hamiltonian ^\hat{\cal H}. The second term describes SP decay with the rate 2γsp2\gamma_{sp} and the last three terms account for incoherent pumping of QEs with the rate γ\gamma_{\uparrow}, QE decay with the rate γ\gamma_{\downarrow} and pure dephasing with the rate γϕ\gamma_{\phi}. Here, the Lindblad operator describing dissipation process in the Markovian approximation is

^C^[𝒪^]\displaystyle\hat{\cal L}_{\hat{C}}[\hat{\cal O}] =\displaystyle= 12(C^[𝒪^,C^]+[C^,𝒪^]C^).\displaystyle\frac{1}{2}\left(\hat{C}^{\dagger}\left[\hat{\cal O},\hat{C}\right]+\left[\hat{C}^{\dagger},\hat{\cal O}\right]\hat{C}\right). (3)

The substitution of the Hamiltonian (II) and 𝒪^={a^,s^n,s^nz}\hat{\cal O}=\{\hat{a},\hat{s}_{n}^{-},\hat{s}_{n}^{z}\} into Eqs. (II) – (3) results in the following set of coupled Heisenberg equations of motion for the SP and QE degrees of freedom

ta^=(iωsp+γsp)a^2iλa^n=1𝒩os^n,\displaystyle\partial_{t}\hat{a}~{}=-\left(i\omega_{sp}+\gamma_{sp}\right)\hat{a}-2i\lambda\hat{a}^{\dagger}\sum_{n=1}^{{\cal N}_{o}}\hat{s}_{n}^{-}, (4)
tsn=(iωo+γo)sn+2iλsnza^2,\displaystyle\partial_{t}s_{n}^{-}=-\left(i\omega_{o}+\gamma_{o}\right)s_{n}^{-}+2i\lambda s_{n}^{z}\hat{a}^{2}, (5)
tsnz=γt(snzdo2)+iλ(a^2snsn+a^2).\displaystyle\partial_{t}s_{n}^{z}=-\gamma_{t}\left(s_{n}^{z}-\frac{d_{o}}{2}\right)+i\lambda\left(\hat{a}^{{\dagger}2}s_{n}^{-}-s_{n}^{+}\hat{a}^{2}\right). (6)

Eqs. (5) and (6) contain pumping rate dependent QE dephasing, γo=γ/2+γ/2+γϕ\gamma_{o}=\gamma_{\downarrow}/2+\gamma_{\uparrow}/2+\gamma_{\phi} and population decay, γt=γ+γ\gamma_{t}=\gamma_{\uparrow}+\gamma_{\downarrow}, rates. Eq. (6) also contains the population inversion parameter

do=γγγ+γ,\displaystyle d_{o}=\frac{\gamma_{\uparrow}-\gamma_{\downarrow}}{\gamma_{\uparrow}+\gamma_{\downarrow}}, (7)

whose value varies in the interval 1do<0-1\leq d_{o}<0 (0<do10<d_{o}\leq 1) for γ<γ\gamma_{\uparrow}<\gamma_{\downarrow} (γ>γ\gamma_{\uparrow}>\gamma_{\downarrow}) indicating that QEs are pumped below (above) population inversion threshold.

Eq. (4) has a well familial form of equation of motion for the parametric down-conversion processes Loudon_book:10 . While conventional DPDC occurs due to a coherent optical pump tuned in resonance with the second-harmonic, our model replaces the laser pump with the QE coherences interacting with the counter-rotating SP coherence in the last term of Eq. (4).

II.1 Lasing phase transition

To get a better insight into this model, we first identify its nontrivial steady state appearing in the strong coupling regime. We start the analysis by deriving mean-field equations of motion for uncorrelated SP and QE degrees of freedom. Specifically, for the two-SP coherence, a^2\langle\hat{a}^{2}\rangle, and SP population, a^a^\langle\hat{a}^{\dagger}\hat{a}\rangle, and the following normalized coherence s^=n=1𝒩os^n/𝒩o\langle\hat{s}_{-}\rangle=\sum_{n=1}^{{\cal N}_{o}}\langle\hat{s}_{n}^{-}\rangle/{\cal N}_{o} and population inversion s^z=n=1𝒩os^nz/𝒩o\langle\hat{s}_{z}\rangle=\sum_{n=1}^{{\cal N}_{o}}\langle\hat{s}_{n}^{z}\rangle/{\cal N}_{o} of QEs. To find conditions for the lasing phase transition, we break U(1) gauge symmetry by transforming SP and QE coherences into rotating frame, i.e., replacing a^a^eiωLt/2\hat{a}\rightarrow\hat{a}e^{-i\omega_{L}t/2} and s^s^eiωLt\hat{s}_{-}\rightarrow\hat{s}_{-}e^{-i\omega_{L}t}. In these variables Eqs. (4)–(6), result in a set of equations of motion

ta^2\displaystyle\partial_{t}\langle\hat{a}^{2}\rangle =\displaystyle= 2(iδωsp+γsp)a^2\displaystyle-2\left(i\delta\omega_{sp}+\gamma_{sp}\right)\langle\hat{a}^{2}\rangle
2i𝒩oλ(2a^a^+1)s^,\displaystyle-2i{\cal N}_{o}\lambda\left(2\langle\hat{a}^{\dagger}\hat{a}\rangle+1\right)\langle\hat{s}_{-}\rangle,
ts^\displaystyle\partial_{t}\langle\hat{s}_{-}\rangle =\displaystyle= (iδωo+γo)s^+2iλs^za^2,\displaystyle-\left(i\delta\omega_{o}+\gamma_{o}\right)\langle\hat{s}_{-}\rangle+2i\lambda\langle\hat{s}_{z}\rangle\langle\hat{a}^{2}\rangle, (9)
ta^a^\displaystyle\partial_{t}\langle\hat{a}^{\dagger}\hat{a}\rangle =\displaystyle= 2γspa^a^\displaystyle-2\gamma_{sp}\langle\hat{a}^{\dagger}\hat{a}\rangle
+2iλ𝒩o[s^+a^2a^2s^]\displaystyle+2i\lambda{\cal N}_{o}\left[\langle\hat{s}_{+}\rangle\langle\hat{a}^{2}\rangle-\langle\hat{a}^{{\dagger}2}\rangle\langle\hat{s}_{-}\rangle\right]
ts^z\displaystyle\partial_{t}\langle\hat{s}_{z}\rangle =\displaystyle= γt(s^zdo2)\displaystyle-\gamma_{t}\left(\langle\hat{s}_{z}\rangle-\frac{d_{o}}{2}\right)
iλ[s^+a^2a^2s^],\displaystyle-i\lambda\left[\langle\hat{s}_{+}\rangle\langle\hat{a}^{2}\rangle-\langle\hat{a}^{{\dagger}2}\rangle\langle\hat{s}_{-}\rangle\right],

with the short hand notations δωsp=ωspωL/2\delta\omega_{sp}=\omega_{sp}-\omega_{L}/2, and δωo=ωoωL\delta\omega_{o}=\omega_{o}-\omega_{L} for the frequency offsets due to the lasing frequency, ωL\omega_{L}, determined below.

Denoting the steady states of the two-SP coherence as α¯2\bar{\alpha}_{2}, SP population as n¯sp\bar{n}_{sp}, QE coherence as s¯\bar{s}_{-}, and population inversion as s¯z\bar{s}_{z}, we present their non-trivial expressions

|α¯2|=n¯sp,\displaystyle|\bar{\alpha}_{2}|=\bar{n}_{sp}, (12)
n¯sp=𝒩odoγt4γsp(1±1[λc𝒩oλ]2),\displaystyle\bar{n}_{sp}={\cal N}_{o}\frac{d_{o}\gamma_{t}}{4\gamma_{sp}}\left(1\pm\sqrt{1-\left[\frac{\lambda_{c}}{{\cal N}_{o}\lambda}\right]^{2}}\right), (13)
s¯z=do4(11[λc𝒩oλ]2),\displaystyle\bar{s}_{z}=\frac{d_{o}}{4}\left(1\mp\sqrt{1-\left[\frac{\lambda_{c}}{{\cal N}_{o}\lambda}\right]^{2}}\right), (14)
|s¯|2=δωsp2+γsp2(2𝒩oλ)2,\displaystyle|\bar{s}_{-}|^{2}=\frac{\delta\omega_{sp}^{2}+\gamma_{sp}^{2}}{\left(2{\cal N}_{o}\lambda\right)^{2}}, (15)

satisfying Eqs. (II.1)–(II.1) with the time derivatives set to zero. This solution exists in the QE inversion regime, i.e., do>0d_{o}>0, and above the SP-QE critical coupling rate

λc2=4γsp2do2γoγt(δωo2+γo2),\displaystyle\lambda_{c}^{2}=\frac{4\gamma_{sp}^{2}}{d_{o}^{2}\gamma_{o}\gamma_{t}}\left(\delta\omega_{o}^{2}+\gamma_{o}^{2}\right), (16)

i.e., 𝒩oλ>λc{\cal N}_{o}\lambda>\lambda_{c}. Associated lasing frequency offsets are

δωo\displaystyle\delta\omega_{o} =\displaystyle= ωo2ωsp1+2γsp/γo,\displaystyle\frac{\omega_{o}-2\omega_{sp}}{1+2\gamma_{sp}/\gamma_{o}}, (17)
δωsp\displaystyle\delta\omega_{sp} =\displaystyle= ωspωo/21+γo/(2γsp).\displaystyle\frac{\omega_{sp}-\omega_{o}/2}{1+\gamma_{o}/(2\gamma_{sp})}. (18)

Below the critical coupling Eqs. (II.1)–(II.1) have a trivial steady state with s¯z=do/2\bar{s}_{z}=d_{o}/2, α¯2=n¯sp=s¯=0\bar{\alpha}_{2}=\bar{n}_{sp}=\bar{s}_{-}=0 and δωo=δωsp=0\delta\omega_{o}=\delta\omega_{sp}=0.

The dependence of our critical coupling, λc\lambda_{c}, on the dissipation rates (Eq. (16)) shows that for 𝒩oλ>λc{\cal N}_{o}\lambda>\lambda_{c} our system enters the strong coupling regime, i.e., the interaction rate exceeds any dissipation rate. In the case of Dicke and Tavis-Cummings models, strong coupling regime can be achieved by increasing the number of QEs, 𝒩o{\cal N}_{o}, since, effective coupling rate, 𝒩oλ\sqrt{{\cal N}_{o}}\lambda, scales as 𝒩o\sqrt{{\cal N}_{o}}.Kirton_AdvQT:2019 ; Piryatinski_PRR:2020 ; Sukharev_JCP:2021 Interestingly in our model, the effective coupling, 𝒩oλ{\cal N}_{o}\lambda, scales linearly with 𝒩o{\cal N}_{o} as can be noticed in Eqs. (13) and (14). This allows one to reach strong coupling regime for a smaller number of QEs compared to the Dicke and Tavis-Cummings models. The difference in the scaling reflects the fact that in our model single energy quantum due to a QE is coupled to two-SP quanta whereas in the aforementioned models a single cavity quantum couples a single QE excitation.

The mean field steady states (Eqs. (13)–(15)) are exact in the thermodynamic limit difined as 𝒩o{\cal N}_{o}\rightarrow\infty and λ0\lambda\rightarrow 0 preserving a finite value of their product 𝒩oλ{\cal N}_{o}\lambda. As we demonstrate in Sec. III, obtained steady state solution can be used as a good approximation to examine formation of the lasing phase in finite but large enough ensembles of QEs. Much better description of the finite size ensemble can be obtained by means of equations of motion which account for the correlations between QEs and SP degrees of freedom. Kirton_AdvQT:2019 ; Piryatinski_PRR:2020 ; Kirton_NJP:2018 Thus, we introduce correlated SP-QE coherence c+sp=ns^n+a^2/𝒩oc_{+sp}=\sum\limits_{n}\langle\hat{s}^{+}_{n}\hat{a}^{2}\rangle/{\cal N}_{o}, SP-SP long range coherence c+=nms^n+s^m/[𝒩o(𝒩o1)]c_{+-}=\sum\limits_{n\neq m}\langle\hat{s}^{+}_{n}\hat{s}^{-}_{m}\rangle/[{\cal N}_{o}({\cal N}_{o}-1)], and keep nsp=a^a^n_{sp}=\langle\hat{a}^{\dagger}\hat{a}\rangle, and sz=ns^nz/𝒩os_{z}=\sum_{n}\langle\hat{s}_{n}^{z}\rangle/{\cal N}_{o}, to form a new set of variables.

Following the methodology used in  Kirton_AdvQT:2019 ; Piryatinski_PRR:2020 ; Kirton_NJP:2018 , we derived the following closed set of equations of motion for these variables

tc+sp\displaystyle\partial_{t}c_{+sp} =\displaystyle= [i(2ωspωo)+2γsp+γo]c+sp\displaystyle-\left[i\left(2\omega_{sp}-\omega_{o}\right)+2\gamma_{sp}+\gamma_{o}\right]c_{+sp}
2iλ[2nsp2sz+(2nsp+1)\displaystyle-2i\lambda\left[2n^{2}_{sp}s_{z}+\left(2n_{sp}+1\right)\right.
×(sz+1/2+(𝒩o1)c+)],\displaystyle\left.\times\left(s_{z}+1/2+({\cal N}_{o}-1)c_{+-}\right)\right],
tc+\displaystyle\partial_{t}c_{+-} =\displaystyle= 2γoc++2iλsz(c+spc+sp),\displaystyle-2\gamma_{o}c_{+-}+2i\lambda s_{z}\left(c_{+sp}-c^{*}_{+sp}\right), (20)
tsz\displaystyle\partial_{t}s_{z} =\displaystyle= γt(szdo/2)iλ(c+spc+sp),\displaystyle-\gamma_{t}\left(s_{z}-d_{o}/2\right)-i\lambda\left(c_{+sp}-c^{*}_{+sp}\right), (21)
tnsp\displaystyle\partial_{t}n_{sp} =\displaystyle= 2γspnsp+2iλ𝒩o(c+spc+sp),\displaystyle-2\gamma_{sp}n_{sp}+2i\lambda{\cal N}_{o}\left(c_{+sp}-c^{*}_{+sp}\right), (22)

starting with the operator Eqs. (II.1)–(II.1). Steady states of Eqs. (II.1)–(22) will be numerically identified and results will be compared with the mean-filed theory in Sec. III.

II.2 Cavity polaritons and related auto-correlation functions

Now, we turn our attention to the two-SP, α2=a^2α¯2{\alpha}_{2}=\langle\hat{a}^{2}\rangle-\bar{\alpha}_{2}, and QEs, s=s^s¯s_{-}=\langle\hat{s}_{-}\rangle-\bar{s}_{-} coherence fluctuations. According to Eqs. (II.1) and (9) the fluctuations satisfy linearized equation of motion

t𝝃(t)=𝝃(t).\displaystyle\partial_{t}{\bm{\xi}}(t)={\cal M}{\bm{\xi}}(t). (23)

where, the fluctuation vector is 𝝃=(α2,s)T{\bm{\xi}}=({\alpha}_{2},s_{-})^{T} and stability matrix

=[2(iδωsp+γsp)2i𝒩oλ(2n¯sp+1)2iλs¯ziδωoγo],\displaystyle{\cal M}=\begin{bmatrix}-2\left(i\delta\omega_{sp}+\gamma_{sp}\right)&-2i{\cal N}_{o}\lambda(2\bar{n}_{sp}+1)\\ 2i\lambda\bar{s}_{z}&-i\delta\omega_{o}-\gamma_{o}\end{bmatrix}, (24)

depends on the steady state parameters s¯z\bar{s}_{z}, n¯sp\bar{n}_{sp}, δωo\delta\omega_{o}, δωsp\delta\omega_{sp}. The time-dependent and Fourier transformed solutions of linear Eq. (23) in terms of the initial condition vector, 𝝃(0){\bm{\xi}}(0), are given in Appendix A. As expected the solution depends on the eigenvalues of {\cal M},

Λ±=iδωsp+γsp+iδωo+γo2\displaystyle\Lambda_{\pm}=i\delta\omega_{sp}+\gamma_{sp}+\frac{i\delta\omega_{o}+\gamma_{o}}{2}
±{[iωsp+γspiωo+γo2]2+4𝒩oλ2s¯z[2n¯sp+1]}1/2,\displaystyle\pm\left\{\left[i\omega_{sp}+\gamma_{sp}-\frac{i\omega_{o}+\gamma_{o}}{2}\right]^{2}+4{\cal N}_{o}\lambda^{2}\bar{s}_{z}[2\bar{n}_{sp}+1]\right\}^{1/2},

as a function of the SP-QE coupling rate.

Refer to caption
Figure 2: Calculated polariton dispersion curves as a function of normalized QE-SP effective coupling rate for various values of the incoherent pumping rates γ\gamma_{\uparrow} indicated in each panel: (a) and (b) QE are pumped above population inversion with the inversion ratios (Eq. (7)) of do=0.33d_{o}=0.33 and do=0.046d_{o}=0.046, respectively. (c) QEs are pumped slightly below population inversion with do=0.02d_{o}=-0.02.

Next we introduce, the following auto-correlation functions associated with the two-SP and QE coherences

Csp(t)\displaystyle C_{sp}(t) =\displaystyle= a^2(t)a^2(0),\displaystyle\left\langle\hat{a}^{{\dagger}2}(t)\hat{a}^{2}(0)\right\rangle, (26)
Cqe(t)\displaystyle C_{qe}(t) =\displaystyle= s^+(t)s^(0),\displaystyle\left\langle\hat{s}_{+}(t)\hat{s}_{-}(0)\right\rangle, (27)

respectively. To account for the fluctuation dynamics in terms of these correlation functions, we use the quantum regression theorem.Loudon_book:10 ; Piryatinski_PRR:2020 According to this theorem, Csp(t)C_{sp}(t) complimented by the cross-correlation function CX1=s^+(t)a^2(0)C_{X1}=\left\langle\hat{s}_{+}(t)\hat{a}^{2}(0)\right\rangle form a vector 𝝃(t)=(Csp(t),CX1(t))T{\bm{\xi}}^{*}(t)=\left(C_{sp}(t),C_{X1}(t)\right)^{T} that satisfies complex conjugate Eq. (23). The same applies to the pair of Cqe(t)C_{qe}(t) and CX2=a^2(t)s^(0)C_{X2}=\left\langle\hat{a}^{{\dagger}2}(t)\hat{s}_{-}(0)\right\rangle in the vector representation 𝝃(t)=(CX2(t),Cqe(t))T{\bm{\xi}}^{*}(t)=\left(C_{X2}(t),C_{qe}(t)\right)^{T}. Using Fourier transformed solution of Eq. (23) given in Appendix A, we provide explicit form for the frequency domain representation of the auto-correlation functions

Csp(ω)\displaystyle C_{sp}(\omega) =\displaystyle= i(ωωo)+γo(iω+Λ+)(iω+Λ)a^2a^2\displaystyle\frac{i\left(\omega-\omega_{o}\right)+\gamma_{o}}{\left(i\omega+\Lambda^{*}_{+}\right)\left(i\omega+\Lambda^{*}_{-}\right)}\left\langle\hat{a}^{{\dagger}2}\hat{a}^{2}\right\rangle
+\displaystyle+ 2i𝒩oλ(2n¯sp+1)(iω+Λ+)(iω+Λ)s^+a^2,\displaystyle\frac{2i{\cal N}_{o}\lambda\left(2\bar{n}_{sp}+1\right)}{\left(i\omega+\Lambda^{*}_{+}\right)\left(i\omega+\Lambda^{*}_{-}\right)}\left\langle\hat{s}_{+}\hat{a}^{2}\right\rangle,
Cqe(ω)\displaystyle C_{qe}(\omega) =\displaystyle= 2iλs¯z(iω+Λ+)(iω+Λ)a^2s^\displaystyle\frac{2i\lambda\bar{s}_{z}}{\left(i\omega+\Lambda^{*}_{+}\right)\left(i\omega+\Lambda^{*}_{-}\right)}\left\langle\hat{a}^{{\dagger}2}\hat{s}_{-}\right\rangle
+\displaystyle+ i(ω2ωsp)+2γsp(iω+Λ+)(iω+Λ)s^+s^.\displaystyle\frac{i\left(\omega-2\omega_{sp}\right)+2\gamma_{sp}}{\left(i\omega+\Lambda^{*}_{+}\right)\left(i\omega+\Lambda^{*}_{-}\right)}\left\langle\hat{s}_{+}\hat{s}_{-}\right\rangle.

To extract the fluctuation dynamics, zero-time operator correlation functions entering into Eqs. (II.2) and (II.2) as initial conditions should be evaluated in the steady state. Adopting the the mean-field approximation, we factorize the mean operator products as

a^2a^2\displaystyle\left\langle\hat{a}^{{\dagger}2}\hat{a}^{2}\right\rangle =\displaystyle= 2n¯sp2,\displaystyle 2\bar{n}_{sp}^{2}, (30)
a^2s^\displaystyle\left\langle\hat{a}^{{\dagger}2}\hat{s}_{-}\right\rangle =\displaystyle= s^+a^2=n¯sp|s¯|eiφo,\displaystyle\left\langle\hat{s}_{+}\hat{a}^{2}\right\rangle=\bar{n}_{sp}|\bar{s}_{-}|e^{-i\varphi_{o}}, (31)
s^+s^\displaystyle\left\langle\hat{s}_{+}\hat{s}_{-}\right\rangle =\displaystyle= |s¯|2+s¯z+1/2.\displaystyle|\bar{s}_{-}|^{2}+\bar{s}_{z}+1/2. (32)

In Eq. (31), φo\varphi_{o} is initial phase of the QE spontaneous coherence, ss_{-}.

Accordingly below critical coupling, the two-SP auto-correlation function vanishes, since, no coherent energy exchange between SP-QE exists. In contrast, the QE auto-correlation function acquires a finite but trivial form

Cqe(ω)\displaystyle C_{qe}(\omega) =\displaystyle= do+1/2i(ωωo)+γo,\displaystyle\frac{d_{o}+1/2}{i\left(\omega-\omega_{o}\right)+\gamma_{o}}, (33)

whose real part describes photon emission lineshape for the non-interacting QEs. Non-vanishing prefactor do+1/2d_{o}+1/2 reflects incoherent pumping resulting in the energy emission. Above critical coupling, full set of Eqs. (II.2) and (II.2) should be used to evaluate the correlation functions. It should be complimented with Eq. (II.2) for polariton branches and Eqs. (13)–(18) for the lasing steady state parameters.

As stated above, the steady states of our model can also be calculated using a set of Eq. (II.1)-(22) directly accounting for the QE-SP correlations. In this case, the zero-time correlation functions entering Eqs. (II.2) and (II.2) can be factorized as

a^2a^2\displaystyle\left\langle\hat{a}^{{\dagger}2}\hat{a}^{2}\right\rangle =\displaystyle= 2n¯sp2,\displaystyle 2\bar{n}_{sp}^{2}, (34)
s^+a^2\displaystyle\left\langle\hat{s}_{+}\hat{a}^{2}\right\rangle =\displaystyle= a^2s^=c¯+sp,\displaystyle\left\langle\hat{a}^{{\dagger}2}\hat{s}_{-}\right\rangle^{*}=\bar{c}_{+sp}, (35)
s^+s^\displaystyle\left\langle\hat{s}_{+}\hat{s}_{-}\right\rangle =\displaystyle= c¯++s¯z+1/2.\displaystyle\bar{c}_{+-}+\bar{s}_{z}+1/2. (36)

Here, the right-hand side contains the steady state SP population, n¯sp\bar{n}_{sp}, and QE inversion, s¯z\bar{s}_{z}, as well as the steady state correlated two-SP-QE coherence, c¯+sp\bar{c}_{+sp}, and the long range QE-QE coherences, c¯+\bar{c}_{+-}.

III Numerical Results and Discussion

In this section, we calculate the polariton dispersion (Eq. (II.2)) as a function of the SP-QE coupling and power spectra of the SP and QE elementary excitations. Let us define the spectra as Ssp=Re[Csp(ω)]/𝒩o2S_{sp}=\text{Re}[C_{sp}(\omega)]/{\cal N}_{o}^{2} and Sqe=Re[Cqe(ω)]S_{qe}=\text{Re}[C_{qe}(\omega)] using Eqs. (26) and (27) for the auto-correlation functions. The correlation functions are calculated using both the mean-field (Eq. (12)–(18)) and correlated QE-SP (Eqs. (II.1)–(22)) models allowing for comparison between their predictions. Parameterization of the models involves setting SP and QE energies to ωsp=ωo/2=1.95eV\hbar\omega_{sp}=\hbar\omega_{o}/2=1.95\text{eV}, QE decay (dephasing) rate to γ=0.014eV\hbar\gamma_{\downarrow}=0.014\text{eV} (γϕ=0.01eV\hbar\gamma_{\phi}=0.01\text{eV}), and the SP dephasing rate to γsp=0.014eV\hbar\gamma_{sp}=0.014\text{eV}. These parameters are consistent with our previous calculations reported in Ref. Sukharev_JCP:2021 . Finally, for the correlated SP-QE model, Eqs. (II.1)–(22) were solved numerically by setting the number of QEs to 𝒩o=5000{\cal N}_{o}=5000.

Figure 2 shows the dispersion of polariton frequency Ω=Im[Λ±]\Omega=\text{Im}[\Lambda_{\pm}] (orange curves) and dissipation rate Γ=Re[Λ±]\Gamma=\text{Re}[\Lambda_{\pm}] (blue curves) as a function of normalized coupling parameter. Probing various values of the incoherent pumping rate and associated (Eq. (7)) inversion parameter, dod_{o}, reveals several distinct features of polariton states and limitations of the mean-field model. For a relatively high pumping rate, do=0.33d_{o}=0.33, results of both mean-field and correlated SP-QE calculations are the same (see panel a). In this case, the lower and upper polariton frequencies do not split as the system transitions from the normal (i.e., trivial ) steady state to the lasing state. In contrast, the polariton dissipation rate branches split above the critical coupling, and the lower branch reaches zero. This trend is due to the QE gain, which fully compensates for the losses indicating lasing regime. For a lower inversion characterized by do=0.46d_{o}=0.46, panel b shows no changes in the behavior of the mean-field solution (thin blue lines). However, the dissipation rate calculated using the correlated QE-SP model (blue dots) behaves differently. Specifically, the polariton frequency splitting associated with the strong coupling regime occurs below the mean-field threshold (0.75λc\sim 0.75\lambda_{c}). At the same time, the lower polariton dissipation rate never reaches zero, i.e., the QE gain does not fully compensate the losses.

Refer to caption
Figure 3: Calculated power spectra for (a), (b) ensemble of QEs normalized per 𝒩o{\cal N}_{o} and (c), (d) correlated two-SP states normalized per 𝒩o2{\cal N}_{o}^{2} in strong couplings regime characterized by λ=2.1λc\lambda=2.1\lambda_{c} and values of incoherent pumping rate γ\gamma_{\uparrow} same as in Fig. 2. The mean-field calculations are presented in the left column and the correlated QE-SP model is used in the right column.

These features indicate deviations in the properties of finite-size QE ensembles from the predictions of the mean-field theory. The latter, indeed, is exact only in the thermodynamic limit. For example, failure of the lower polariton dissipation branch to reach zero is a signature of cross-over behavior between normal and lasing states rather than the lasing phase transition. Furthermore, below the QE inversion (panel c), our mean-field theory is not defined, but the correlated QE-SP calculations produce a satisfactory result. They show a well-known behavior of the polariton branches while crossing a threshold between weak and strong coupling regimes at 𝒩oλ0.2λc{\cal N}_{o}\lambda\approx 0.2\lambda_{c}. At this point, the polariton frequency (orange dots) experiences the Rabi splitting, and the QE and SP dissipation rates (blue dots) merge into a single rate. The trend of lowering polariton strong coupling threshold, observed in panels b and c, with the pumping rate decrease, is likely related to the dependence of the dissipation rates on the pumping rate (see comments following Eqs. (4) and (6)).

Trends in the variations of the QE power spectra, calculated in the mean-field approximation (Fig. 3a) and using the QE-SP correlated model (Fig. 3b), are consistent with the discussed trends in polariton dispersion (Fig. 2). Calculated above critical coupling at 𝒩oλ=2.0λc{\cal N}_{o}\lambda=2.0\lambda_{c}, both panels of the plot show narrow lasing peaks at the pumping rate twice the decay rate (red curves). Differences appear when the pumping rate gets closer to the inversion threshold. In this situation (blue lines), the mean-field spectrum still has a sharp lasing peak, whereas the correlated QE-SP model predicts a broader peak. The broadening occurs because the lower polariton dissipative branch (blue dots in Fig. 2b) does not reach zero, i.e., the QE gain does not fully compensate for the losses.

Below the inversion threshold, our mean-field model predicts a single broad peak (black curve in panel a) originating from the trivial limit of the fluctuation correlation function (Eq. (33)) with the resonance at ωo\omega_{o}. In contrast, the correlated QE-SP model (black line in panel b) predicts two resonances due to the Rabi splitting of the polariton frequency (orange dots in Fig. 2c). The two-SP power spectra in panels c and d of Fig. 3 show the same features as panels a and b. The mean-field model predicts no DPDC effect below inversion (notice the absence of a black curve in pane c). To understand this, recall that the two-SP coherence directly couples to the QE steady state coherence s¯\bar{s}_{-} (see, e.g., Eq. (II)). The latter vanishes below the population inversion threshold and cancels the two-SP coherence.

As illustrated in Fig. 1, the radiative decay of QEs and two-SP states results in the photon emission around the second-harmonic frequency and correlated fundamental-frequency photon-pair production, respectively. The QE and SP spectra are appropriately normalized, so comparing peak values, we see an even quantum number distribution between the QEs and two-SP degrees of freedom. In contrast, the number of photons produced via the radiative decay of QEs and two-SP states depends on their material-specific spontaneous emission rates. This quantity in plasmonic nanostructures can be orders of magnitude higher than in organic chromophores or semiconductor QEs. As a result, we expext that the photon emission spectra should be dominated by the correlated two-photon pairs resulting in the photon DPDC effect.

IV Concluding Remarks

We have examined the mechanism of DPDC in nonlinear plasmonic cavities in which the incoherently pumped QEs interact with the SP second-harmonic mode via near-field coupling. As demonstrated, such a model allows for a lasing phase transition in the strong coupling regime. Interestingly, the effective coupling rate in our model scales as 𝒩oλ{\cal N}_{o}\lambda in contrast to conventional scaling of 𝒩oλ\sqrt{{\cal N}_{o}}\lambda known for the Dicke and Tavis-Cummings models facilitating the strong coupling regime for smaller ensembles of QEs. Comparing the finite-size ensemble of QEs with the model thermodynamic limit, we have found that in the former case, the strong coupling threshold lowers as the QE pumping rate decreases. Furthermore, the finite-size ensemble of QEs pumped below population inversion is predicted to demonstrate the polariton Rabi splitting and the DPDC effect. In the thermodynamic limit, no such effects take place. Our model also predicts that the lasing threshold in finite-size systems occurs for high pumping rate than in the thermodynamic limit.

We expect that 2D plasmonic lattices showing the lattice-plasmon resonances RamezaniJOSAB:2019 ; Sukharev_JCP:2021 could be a good material platform for observation predicted DPDC effect. However, our basic quantum optics model cannot incorporate some material-specific features of the lattice-plasmon resonances. This calls for numerical simulations SukharevJPCM:2017 accounting for the composition and geometry of metal nanoparticles, detailed mechanisms of nonlinearity, and QE electronic structure. In particular, the simulations should address the question of photon emission yield for the correlated two-SP states compared to the emission yield of the QEs. Finally, reported progress on developing new plasmonic nanomaterials, e.g., doped semiconductors, AgrawalChenRev:2018 ; Dolgopolova_NanoHorizons:2022 with SP resonances in the near-infrared, will pave the way for practical applications of the proposed effect in on-chip photonic devices.

V Acknowledgements

This work was performed, in part, at the Center for Integrated Nanotechnologies, an Office of Science User Facility operated for the U.S. Department of Energy (DOE) Office of Science by Los Alamos National Laboratory (Contract 89233218CNA000001) and Sandia National Laboratories (Contract DE-NA-0003525) (user project No. 2021BC0087). A.P. would like to acknowledge support via LDRD exploratory research grant No. 20200407ER. M.S. is grateful for the financial support by the Air Force Office of Scientific Research under Grant No. FA9550-22-1-0175. We would like to thank Professor Joseph Zyss for stimulating discussions.

Appendix A Solution of Eqs. (23) and (24)

Solution of Eqs. (23) and (24) is

𝝃(t)\displaystyle\bm{\xi}(t) =\displaystyle= 𝒢(t)𝝃(0)\displaystyle{\cal G}(t)\bm{\xi}(0) (37)

where 𝝃(0)\bm{\xi}(0) is the vector of initial conditions and the Green function, 𝒢(t){\cal G}(t), matrix elements are

g11(t)\displaystyle g_{11}(t) =\displaystyle= 12[1+iωo+γo2(iωsp+γsp)Λ+Λ]eΛt\displaystyle\frac{1}{2}\left[1+\frac{i\omega_{o}+\gamma_{o}-2\left(i\omega_{sp}+\gamma_{sp}\right)}{\Lambda_{+}-\Lambda_{-}}\right]e^{-\Lambda_{-}t}\;\;\;\;
+\displaystyle+ 12[1iωo+γo2(iωsp+γsp)Λ+Λ]eΛ+t,\displaystyle\frac{1}{2}\left[1-\frac{i\omega_{o}+\gamma_{o}-2\left(i\omega_{sp}+\gamma_{sp}\right)}{\Lambda_{+}-\Lambda_{-}}\right]e^{-\Lambda_{+}t},
g12(t)\displaystyle g_{12}(t) =\displaystyle= 2i𝒩oλ(2n¯sp+1)Λ+Λ(eΛ+teΛt),\displaystyle\frac{2i{\cal N}_{o}\lambda\left(2\bar{n}_{sp}+1\right)}{\Lambda_{+}-\Lambda_{-}}\left(e^{-\Lambda_{+}t}-e^{-\Lambda_{-}t}\right), (39)
g21(t)\displaystyle g_{21}(t) =\displaystyle= 2iλs¯zΛ+Λ(eΛteΛ+t),\displaystyle\frac{2i\lambda\bar{s}_{z}}{\Lambda_{+}-\Lambda_{-}}\left(e^{-\Lambda_{-}t}-e^{-\Lambda_{+}t}\right), (40)
g22(t)\displaystyle g_{22}(t) =\displaystyle= 12[1+iωo+γo2(iωsp+γsp)Λ+Λ]eΛ+t\displaystyle\frac{1}{2}\left[1+\frac{i\omega_{o}+\gamma_{o}-2\left(i\omega_{sp}+\gamma_{sp}\right)}{\Lambda_{+}-\Lambda_{-}}\right]e^{-\Lambda_{+}t}
+\displaystyle+ 12[1iωo+γo2(iωsp+γsp)Λ+Λ]eΛt.\displaystyle\frac{1}{2}\left[1-\frac{i\omega_{o}+\gamma_{o}-2\left(i\omega_{sp}+\gamma_{sp}\right)}{\Lambda_{+}-\Lambda_{-}}\right]e^{-\Lambda_{-}t}.

where, Λ±\Lambda_{\pm} given by Eq. (24).

Next, we perform the Fourier transform of the Green function

𝒢(ω)=𝒢(t)eiδωt𝑑t\displaystyle{\cal G}(\omega)=\int\limits_{-\infty}^{\infty}{\cal G}(t)e^{i\delta\omega t}dt (42)

where δω=ωωL\delta\omega=\omega-\omega_{L} and the lower integration limit assumes that 𝒢(|t|)=0{\cal G}(-|t|)=0. This allows us to represent the spectra in terms of the initial condition vector ξ(0)\xi(0). The result of the calculation is

𝝃(ω)\displaystyle\bm{\xi}(\omega) =\displaystyle= 𝒢(ω)𝝃(0),\displaystyle{\cal G}(\omega)\bm{\xi}(0), (43)

with the Green function matrix elements

g11(ω)\displaystyle g_{11}(\omega) =\displaystyle= i(ωωo+iγo)(ω+iΛ+)(ω+iΛ),\displaystyle\frac{i\left(\omega-\omega_{o}+i\gamma_{o}\right)}{\left(\omega+i\Lambda_{+}\right)\left(\omega+i\Lambda_{-}\right)}, (44)
g12(ω)\displaystyle g_{12}(\omega) =\displaystyle= 2i𝒩oλ(2n¯sp+1)(ω+iΛ+)(ω+iΛ),\displaystyle\frac{2i{\cal N}_{o}\lambda\left(2\bar{n}_{sp}+1\right)}{\left(\omega+i\Lambda_{+}\right)\left(\omega+i\Lambda_{-}\right)}, (45)
g21(ω)\displaystyle g_{21}(\omega) =\displaystyle= 2iλs¯z(ω+iΛ+)(ω+iΛ),\displaystyle-\frac{2i\lambda\bar{s}_{z}}{\left(\omega+i\Lambda_{+}\right)\left(\omega+i\Lambda_{-}\right)}, (46)
g22(ω)\displaystyle g_{22}(\omega) =\displaystyle= i(ω2ωsp+2iγsp)(ω+iΛ+)(ω+iΛ).\displaystyle\frac{i\left(\omega-2\omega_{sp}+2i\gamma_{sp}\right)}{\left(\omega+i\Lambda_{+}\right)\left(\omega+i\Lambda_{-}\right)}. (47)

References

  • (1) Gramotnev D K and Bozhevolnyi S I 2010 Nat. Photonics 4 83–91
  • (2) Ebbesen T W 2016 Acc. Chem. Res. 49 2403–2412
  • (3) Feist J, Galego J and Garcia-Vidal F J 2018 ACS Photonics 5 205–216
  • (4) Huang X, Qian W, El-Sayed I H and El-Sayed M A 2007 Lasers Surg. Med. 39 747–753
  • (5) Stockman M I 2011 Opt. Express 19 22029–22106
  • (6) Panoiu N, Sha W, Lei D and Li G 2018 J. Opt-Uk 20 083001
  • (7) Wu T, Gurioli M and Lalanne P 2021 ACS Photonics 8 1522–1538
  • (8) Jiang N, Zhuo X and Wang J 2017 Chem. Rev. 118 3054–3099
  • (9) Vasa P, Wang W, Pomraenke R, Lammers M, Maiuri M, Manzoni C, Cerullo G and Lienau C 2013 Nat. Photonics 7 128–132
  • (10) Törmä P and Barnes W L 2014 Rep. Prog. Phys. 78 013901
  • (11) Gordon R and Ahmed A 2018 ACS Photonics 5 4222–4228
  • (12) Krasavin A V, Ginzburg P and Zayats A V 2018 Laser Photonics Rev. 12 1700082
  • (13) Galanty M, Shavit O, Weissman A, Aharon H, Gachet D, Segal E and Salomon A 2018 Light: Science & Applications 7 1–8
  • (14) Weissman A, Galanty M, Gachet D, Segal E, Shavit O and Salomon A 2017 Adv. Opt. Mat. 5 1700097
  • (15) Yoo K, Becker S F, Silies M, Yu S, Lienau C and Park N 2019 Opt. Express 27 18246–18261
  • (16) Almeida E and Prior Y 2015 Sci. Rep. 5 1–10
  • (17) Blechman Y, Almeida E, Sain B and Prior Y 2018 Nano Lett. 19 261–268
  • (18) Wurtz G, Pollard R and Zayats A V 2006 Phys. Rev. Lett 97 057402
  • (19) Bahar E, Arieli U, Mrejen M and Suchowski H 2020 Phys. Rev. B 101 035141
  • (20) Tame M, McEnery K, Özdemir Ş, Lee J, Maier S and Kim M 2013 Nature Physics 9 329–340
  • (21) Cherqui C R, Bourgeois M R, Wang D and Schatz G C 2019 Acc. Chem. Res. 52 2548–2558
  • (22) Ramezani M, Berghuis M and Rivas J G 2019 J. Opt. Soc. Am. B 36 E88
  • (23) Pusch A, Wuestner S, Hamm J M, Tsakmakidis K L and Hess O 2012 ACS Nano 6 2420–2431
  • (24) Guan J, Sagar L K, Li R, Wang D, Bappi G, Wang W, Watkins N, Bourgeois M R, Levina L, Fan F et al. 2020 ACS Nano 14 3426–3433
  • (25) Winkler J M, Ruckriegel M J, Rojo H, Keitel R C, De Leo E, Rabouw F T and Norris D J 2020 ACS Nano 14 5223–5232
  • (26) Martikainen J P, Heikkinen M O J and Törmä P 2014 Phys. Rev. A 90 053604
  • (27) Zaster S, Bittner E R and Piryatinski A 2016 J. Phys. Chem. A 120 3109–3116 ISSN 1520-5215 URL http://dx.doi.org/10.1021/acs.jpca.5b10726
  • (28) Giorgi M D, Ramezani M, Todisco F, Caputo D, Halpin A, Fieramosca A, Sanvitto D and Rivas J G 2017 arXiv:1709.04803
  • (29) Hakala T K, Moilanen A J, Väkeväinen A I, Guo R, Martikainen J P, Daskalakis K S, Rekola H T, Julku A and Törmä P 2018 Nature Phys. 14 739–744 URL https://doi.org/10.1038/s41567-018-0109-9Download
  • (30) Drobnyh E and Sukharev M 2020 J. Chem. Phys. 152 094706
  • (31) Sukharev M, Salomon A and Zyss J 2021 J. Chem. Phys. 154 244701
  • (32) Li C, Lu X, Srivastava A, Storm S D, Gelfand R, Pelton M, Sukharev M and Harutyunyan H 2020 Nano Lett. 21 1599–1605
  • (33) Sukharev M, Roslyak O and Piryatinski A 2021 J. Chem. Phys. 154 084703 URL https://doi.org/10.1063%2F5.0037453
  • (34) Ribeiro R F, Campos-Gonzalez-Angulo J A, Giebink N C, Xiong W and Yuen-Zhou J 2021 Phys. Rev. A 103 063111
  • (35) Grosse N B, Heckmann J and Woggon U 2012 Phys. Rev. Lett. 108 136802
  • (36) Heckmann J, Kleemann M E, Grosse N B and Woggon U 2013 Opt. Express 21 28856–28861
  • (37) Loot A and Hizhnyakov V 2018 J. Opt. 20 055502 URL https://doi.org/10.1088%2F2040-8986%2Faab6c0
  • (38) Loot A, Sildos I, Kiisk V, Romann T and Hizhnyakov V 2018 Optik 171 557–564
  • (39) Loudon R 2010 The quantum theory of light 4th ed (New York: Oxford University Press)
  • (40) Kirton P, Roses M M, Keeling J and Dalla Torre E G 2019 Adv. Quantum Technol. 2 1800043
  • (41) Piryatinski A, Roslyak O, Li H and Bittner E R 2020 Phys. Rev. Research 2(1) 013141 URL https://link.aps.org/doi/10.1103/PhysRevResearch.2.013141
  • (42) Kirton P and Keeling J 2018 New J. Phys. 20 015009
  • (43) Sukharev M and Nitzan A 2017 J. Phys.: Condens. Matter 29 443003
  • (44) Agrawal A, Cho S H, Zandi O, Ghosh S, Johns R W and Milliron D J 2018 Chem. Rev. 118 3121–3207
  • (45) Dolgopolova E A, Li D, Hartman S T, Watt J, Ríos C, Hu J, Kukkadapu R, Casson J, Bose R, Malko A V et al. 2022 Nanoscale Horizons 7 267–275