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

Interacting quantum plasmons in metal-dielectric structures

Tigran V. Shahbazyan Department of Physics, Jackson State University, Jackson, Mississippi 39217 USA
Abstract

We develop a consistent quantum description of surface plasmons interacting with quantum emitters and external electromagnetic field. Within the framework of macroscopic electrodynamics in dispersive and absorptive medium, we derive, in the Markov approximation, the canonical Hamiltonian, commutation relations, and coupling parameters for the plasmon modes in metal-dielectric structures of an arbitrary shape whose characteristic size is well below the diffraction limit. We then develop a new quantum approach bridging the macroscopic and canonical schemes which describes the interacting plasmons in terms of bosonic modes with linear dispersion whose coupling to the electromagnetic field and quantum emitters is mediated by the classical plasmons. By accurately accounting for medium optical dispersion and losses in the interactions of surface plasmons with light and localized electron excitations, this approach can serve as a framework for studying non-Markovian effects in plasmonics.

I Introduction

Over the past decade, quantum plasmonics maier-np13 underwent a rapid development fueled by a host of recently discovered phenomena such as strong exciton-plasmon coupling effects bellessa-prl04 ; sugawara-prl06 ; wurtz-nl07 ; fofang-nl08 ; hakala-prl09 ; gomez-nl10 , plasmon-assisted hot carrier generation park-nl11 ; melosh-nl11 ; halas-science11 , plasmonic laser (spaser) bergman-prl03 ; stockman-natphot08 ; noginov-nature09 ; zhang-nature09 ; gwo-science12 , plasmon tunneling baumberg-nature12 ; aizpurua-nl12 ; dionne-nl13 ; tan-science14 and more, along with a growing number of applications. Surface plasmons are collective electron excitations living at the metal-dielectric interfaces which can interact strongly with light and localized electron excitations such as excitons in molecules or semiconductors stockman-review . Although classical description of many experiments in terms of local field enhancement has largely been successful, a growing number of topics and applications require a more rigorous quantum approach hohenester-prb08 ; nordlander-nl09 ; stockman-jo10 ; waks-pra10 ; garcia-prl13 ; garcia-optica17 ; aizpurua-nanophot20 .

In nanoscale systems, the local fields can change strongly over the length scale well below the diffraction limit, and so the plasmon interactions with the electromagnetic (EM) field and excitons depend sensitively on the system parameters such as the geometry of a metal-dielectric structure and the exciton position relative to it. While the coupling parameters, characterizing these interactions, have been suggested in several forms by using analogy with the cavity modes khitrova-nphys06 ; andreani-prb12 ; shahbazyan-nl19 ; mortensen-rpp20 , these parameters have yet to emerge within a consistent quantization approach for interacting plasmons. Yet another longstanding challenge for quantum plasmonics is to account properly for strong optical dispersion and losses in metals that give rise to non-Markovian effects nori-prb09 ; tejedor-prb10 ; thanopulos-prb17 ; moradi-sr18 ; molmer-acsph19 .

Within the canonical quantization scheme, localized plasmon modes with discrete frequency spectrum ωm\omega_{m} are described by the Hamiltonian

H^pl=mωma^ma^m,\hat{H}_{\rm pl}=\sum_{m}\hbar\omega_{m}\hat{a}^{\dagger}_{m}\hat{a}_{m}, (1)

where a^m\hat{a}^{\dagger}_{m} and a^m\hat{a}_{m} are, respectively, the plasmon creation and annihilation operators obeying the canonical commutation relations [a^m,a^n]=δmn[\hat{a}_{m},\hat{a}^{\dagger}_{n}]=\delta_{mn}. Plasmon interactions with quantum emitters (QEs), modeled hereafter by two-level systems, are usually described by the Jaynes-Cummings interaction Hamiltonian

H^plqe=imgim(σ^ia^m+a^mσ^i),\hat{H}_{\rm pl-qe}=\sum_{im}\hbar g_{im}(\hat{\sigma}^{\dagger}_{i}\hat{a}_{m}+\hat{a}^{\dagger}_{m}\hat{\sigma}_{i}), (2)

where σ^i\hat{\sigma}^{\dagger}_{i} and σ^i\hat{\sigma}_{i} are, respectively, the raising and lowering operators for the iith QE (related to the Pauli matrices for Fermionic systems), while gimg_{im} is the QE-plasmon coupling, which, within the canonical approach, is an ad hoc parameter. Although widely employed, the canonical scheme has significant limitations when used in metal-dielectric structures characterized by a complex dielectric function ε(ω,𝒓)=ε(ω,𝒓)+iε′′(ω,𝒓)\varepsilon(\omega,\bm{r})=\varepsilon^{\prime}(\omega,\bm{r})+i\varepsilon^{\prime\prime}(\omega,\bm{r}) since it ignores the medium optical dispersion and, hence, is unsuitable for describing non-Markovian effects in plasmonics.

On the other hand, the medium optical dispersion is accurately accounted for in the macroscopic electrodynamics approach based upon the fluctuation-dissipation (FD) theorem welsch-pra98 ; welsch-p00 ; philbin-njp10 . In this framework, the EM fields are quantized in terms of reservoir noise operators 𝒇^(ω,𝒓)\hat{\bm{f}}(\omega,\bm{r}) driven by the Hamiltonian

H^N=0𝑑ω𝑑Vω𝒇^(ω,𝒓)𝒇^(ω,𝒓)\hat{H}_{N}=\!\int_{0}^{\infty}\!d\omega\!\int\!dV\,\hbar\omega\hat{\bm{f}}^{\dagger}(\omega,\bm{r})\!\cdot\!\hat{\bm{f}}(\omega,\bm{r}) (3)

and obeying the commutation relations

[𝒇^(ω,𝒓),𝒇^(ω,𝒓)]=𝑰δ(ωω)δ(𝒓𝒓),[\hat{\bm{f}}(\omega,\bm{r}),\hat{\bm{f}}^{\dagger}(\omega^{\prime},\bm{r}^{\prime})]=\bm{I}\delta(\omega-\omega^{\prime})\delta(\bm{r}-\bm{r}^{\prime}), (4)

where 𝑰\bm{I} is the unit tensor. Interactions with QEs are described by the Hamiltonian term H^int=i𝒑^i𝑬^(𝒓i)\hat{H}_{\rm int}=-\sum_{i}\hat{\bm{p}}_{i}\cdot\hat{\bm{E}}(\bm{r}_{i}), where 𝒑^i\hat{\bm{p}}_{i} and 𝑬^(𝒓i)\hat{\bm{E}}(\bm{r}_{i}) are, respectively, the QE dipole moment and electric field operators. The latter is given by

𝑬^(𝒓)=0𝑑ω𝑑V𝑫(ω;𝒓,𝒓)𝑷^N(ω,𝒓)+H.c.,\hat{\bm{E}}(\bm{r})=\!\int_{0}^{\infty}\!d\omega\!\!\int\!dV^{\prime}\!\bm{D}(\omega;\bm{r},\bm{r}^{\prime})\hat{\bm{P}}_{N}(\omega,\bm{r}^{\prime})+\text{H.c.}, (5)

where 𝑷^N(ω,𝒓)=(i/2π)ε′′(ω,𝒓)𝒇^(ω,𝒓)\hat{\bm{P}}_{N}(\omega,\bm{r})=(i/2\pi)\sqrt{\hbar\varepsilon^{\prime\prime}(\omega,\bm{r})}\hat{\bm{f}}(\omega,\bm{r}) is the noise polarization vector operator and 𝑫(ω;𝒓,𝒓)\bm{D}(\omega;\bm{r},\bm{r}^{\prime}) is the EM dyadic Green function defined as

××𝑫(ω;𝒓,𝒓)ω2c2ε(ω,𝒓)𝑫(ω;𝒓,𝒓)\displaystyle\bm{\nabla}\!\times\!\bm{\nabla}\!\times\!\bm{D}(\omega;\bm{r},\bm{r}^{\prime})-\frac{\omega^{2}}{c^{2}}\varepsilon(\omega,\bm{r})\bm{D}(\omega;\bm{r},\bm{r}^{\prime})~{}~{}~{}~{}~{}~{}
=4πω2c2𝑰δ(𝒓𝒓).\displaystyle=\frac{4\pi\omega^{2}}{c^{2}}\bm{I}\delta(\bm{r}-\bm{r}^{\prime}). (6)

The macroscopic FD approach has been extensively used to model spontaneous emission, strong coupling effects and non-Markovian dynamics in metal-dielectric structures welsch-pra00 ; dzsotjan-prb10 ; hughes-prb12 ; andreani-prb12 ; hughes-prb13 ; zubairy-prb14 ; garcia-prl14 ; rousseaux-prb18 ; sivan-prb19 ; nori-prb09 ; tejedor-prb10 ; thanopulos-prb17 ; moradi-sr18 ; molmer-acsph19 . Its major drawback in relation to plasmonics is that, even though surface plasmons reside primarily at the metal-dielectric interfaces, the eigenstates of H^N\hat{H}_{N} extend over the entire system reservoir, i.e., the Hilbert space, spanned by the operators 𝒇^(ω,𝒓)\hat{\bm{f}}(\omega,\bm{r}), is excessively large. Furthermore, the plasmons only appear as resonances in the classical EM Green function 𝑫\bm{D}, so that, in practical terms, this approach is limited to systems that allow 𝑫\bm{D} to be evaluated analytically or numerically.

In principle, the Hamiltonians (1) and (2), along with the canonical commutations relations and QE-plasmon coupling, should be obtained within the macroscopic FD framework starting with a suitable mode expansion for the EM Green function that would define the basis set for normal mode expansion of the electric field operator (5) varguet-ol16 ; dzsotjan-pra16 ; hughes-prl19 . However, for systems of general shape, a straightforward expansion of 𝑫\bm{D} over a discrete set of EM modes that include radiation, such as quasinormal modes hughes-njp14 ; hughes-acsphot14 ; lalanne-pra14 ; hughes-pra15 ; muljarov-prb16 ; lalanne-prb18 ; lalanne-lpr18 , gives rise to dissipation-induced coupling between the modes which compicates the commutations relations hughes-prl19 . On the other hand, on the length scale well below the diffraction limit, the plasmon interactions with QEs take place primarily via the near field coupling rather than photon exchange, while the rate of nonradiative losses, determined by ε′′(ω,𝒓)\varepsilon^{\prime\prime}(\omega,\bm{r}), by far exceeds the plasmon radiative decay rate. This implies that in nanoscale systems, the plasmons should be treated as primarily electronic excitations interacting with the EM field and QEs, and so the quantization of plasmons should be carried on its own, i.e., separate from the radiation field. Furthermore, since the electron motion in metals is unretarded, the plasmon modes in such systems can be described within quasistatic approximation stockman-review , which allows, as we show below, one to define the orthogonal basis set that is free of dissipation-induced coupling.

In this paper, we develop a consistent quantum approach to localized plasmons interacting with quantum emitters and the electromagnetic field which accurately accounts for medium optical dispersion and losses. First, starting within the macroscopic FD framework, we employ the near-field plasmon Green function shahbazyan-prl16 ; shahbazyan-prb18 do define the normal mode expansion for the plasmon field operators free of dissipation-induced coupling. Using this basis set, we explicitly obtain the plasmon Hamiltonian (1), along with the equal-time canonical commutations relations, and show that the canonical quantization scheme is valid only in the Markov approximation, i.e., if the dielectric function ε(ω,𝒓)\varepsilon(\omega,\bm{r}) is replaced by its value at the plasmon frequency (i.e., ω=ωm\omega=\omega_{m}). In this way, we also obtain the microscopic coupling parameters that define the plasmon interactions with the EM field and QEs in terms of the plasmon local fields, system geometry and QE positions relative to the plasmonic structure. Second, moving beyond the Markov approximation, we present a new approach that bridges the macroscopic and canonical quantization schemes while accounting accurately for the medium optical dispersion and losses. In this approach, quantum plasmons are described in terms of a discrete set of bosonic modes with linear dispersion which reside at the metal-dielelectric interfaces, and whose operators span a reduced Hilbert space obtained by projecting the full reservoir states upon localized plasmon modes. The interactions of such projected reservoir modes with the EM field and QEs are mediated by classical plasmons, so that the classical plasmon enhancement effects are encoded in the Hamiltonian coupling parameters.

II From macroscopic to canonical quantization of surface plasmons

In this section, starting within the macroscopic quantization scheme welsch-pra98 ; welsch-p00 ; philbin-njp10 , we obtain the canonical Hamiltonian (1) and the equal-time commutation relations for localized surface plasmon in metal-dielectric structures of arbitrary shape.

II.1 Quasistatic modes and plasmon Green function

We consider a metal-dielectric structure characterized by dielectric function of the form ε(ω,𝒓)=iθi(𝒓)εi(ω)\varepsilon(\omega,\bm{r})=\sum_{i}\theta_{i}(\bm{r})\varepsilon_{i}(\omega), where θi(𝒓)\theta_{i}(\bm{r}) is unit step function that vanishes outside the connected region, metal or dielectric, of volume ViV_{i} with uniform dielectric function εi(ω)\varepsilon_{i}(\omega). For unretarded electron motion in metals, the potentials Φm(𝒓)\Phi_{m}(\bm{r}) and frequencies ωm\omega_{m} of plasmon modes are determined by the quasistatic Gauss law as stockman-review

[ε(ωm,𝒓)Φm(𝒓)]=0.\bm{\nabla}\cdot\left[\varepsilon^{\prime}(\omega_{m},\bm{r})\bm{\nabla}\Phi_{m}(\bm{r})\right]=0. (7)

Accordingly, the plasmon mode fields, which we choose to be real, are defined as 𝑬m(𝒓)=Φm(𝒓)\bm{E}_{m}(\bm{r})=-\bm{\nabla}\Phi_{m}(\bm{r}). Importantly, the plasmon mode fields are orthogonal in each region ViV_{i},

𝑑Vi𝑬m(𝒓)𝑬n(𝒓)=δmn𝑑Vi𝑬m2(𝒓),\int\!dV_{i}\bm{E}_{m}(\bm{r})\!\cdot\!\bm{E}_{n}(\bm{r})=\delta_{mn}\int\!dV_{i}\bm{E}_{m}^{2}(\bm{r}), (8)

so that 𝑑Vε′′(ω,𝒓)𝑬m(𝒓)𝑬n(𝒓)=0\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{E}_{m}(\bm{r})\bm{E}_{n}(\bm{r})=0 for mnm\neq n, implying no dissipation-induced coupling (see Appendix).

The near-field Green function that defines the field operator (5) can be split into free-space and plasmon parts as shahbazyan-prl16 ; shahbazyan-prb18 𝑫=𝑫0+𝑫pl\bm{D}=\bm{D}_{0}+\bm{D}_{\rm pl}. When inserted into Eq. (5), the first term yields the electric field due to noise fluctuations, while the second term defines the normal mode expansion of the plasmon field operator. In this paper, we focus only on the plasmonic sector of the Hilbert space. In the absence of dissipation-induced coupling, the plasmon Green function can be derived exactly in the following form (see Appendix):

𝑫pl(ω;𝒓,𝒓)=mDm(ω)𝑬m(𝒓)𝑬m(𝒓)\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=\sum_{m}D_{m}(\omega)\bm{E}_{m}(\bm{r})\bm{E}_{m}(\bm{r}^{\prime}) (9)

where

Dm(ω)=4π𝑑V𝑬m2(𝒓)4π𝑑Vε(ω,𝒓)𝑬m2(𝒓).\displaystyle D_{m}(\omega)=\dfrac{4\pi}{\int\!dV\bm{E}_{m}^{2}(\bm{r})}-\dfrac{4\pi}{\int\!dV\varepsilon(\omega,\bm{r})\bm{E}_{m}^{2}(\bm{r})}. (10)

The first term ensures that 𝑫pl=0\bm{D}_{\rm pl}=0 for ε=1\varepsilon=1 (or, in the limit ω\omega\rightarrow\infty). The plasmon Green functions exhibits poles in the complex frequency plane defined by the condition 𝑑Vε(ω,𝒓)𝑬m2(𝒓)=0\int\!dV\varepsilon(\omega,\bm{r})\bm{E}_{m}^{2}(\bm{r})=0. In the following, we restrict ourselves to the plasmonics frequency domain ε′′(ω)/ε(ω)1\varepsilon^{\prime\prime}(\omega)/\varepsilon^{\prime}(\omega)\ll 1. Since 𝑑Vε(ωm,𝒓)𝑬m2(𝒓)=0\int\!dV\varepsilon^{\prime}(\omega_{m},\bm{r})\bm{E}_{m}^{2}(\bm{r})=0 due to the Gauss law, we can expand ε(ω,𝒓)\varepsilon^{\prime}(\omega,\bm{r}) in Eq. (10) near ωm\omega_{m} to present the plasmon Green function as a sum over the plasmon poles shahbazyan-prb18 (see Appendix)

𝑫pl(ω;𝒓,𝒓)=mωm4Um𝑬m(𝒓)𝑬m(𝒓)ωmωi2γm(ω),\displaystyle\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=\sum_{m}\frac{\omega_{m}}{4U_{m}}\frac{\bm{E}_{m}(\bm{r})\bm{E}_{m}(\bm{r}^{\prime})}{\omega_{m}-\omega-\frac{i}{2}\gamma_{m}(\omega)}, (11)

where UmU_{m} is the plasmon mode energy landau ,

Um=116π𝑑V[ωmε(ωm,𝒓)]ωm𝑬m2(𝒓),\displaystyle U_{m}=\frac{1}{16\pi}\!\int\!dV\dfrac{\partial[\omega_{m}\varepsilon^{\prime}(\omega_{m},\bm{r})]}{\partial\omega_{m}}\,\bm{E}_{m}^{2}(\bm{r}), (12)

and γm(ω)\gamma_{m}(\omega) is the frequency-dependent decay rate

γm(ω)=2𝑑Vε′′(ω,𝒓)𝑬m2(𝒓)𝑑V[ε(ωm,𝒓)/ωm]𝑬m2(𝒓).\gamma_{m}(\omega)=\dfrac{2\!\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{E}_{m}^{2}(\bm{r})}{\!\int\!dV[\partial\varepsilon^{\prime}(\omega_{m},\bm{r})/\partial\omega_{m}]\bm{E}_{m}^{2}(\bm{r})}. (13)

In structures with a single metallic component, the decay rate takes the form stockman-review γm(ω)=2ε′′(ω)/[ε(ωm)/ωm]\gamma_{m}(\omega)=2\varepsilon^{\prime\prime}(\omega)/[\partial\varepsilon^{\prime}(\omega_{m})/\partial\omega_{m}]. Finally, with help of Eqs. (8) and (13), we obtain the following relation

𝑑Vε′′(ω,𝒓)𝑫pl(ω;𝒓,𝒓)𝑫pl(ω;𝒓,𝒓′′)\displaystyle\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{D}_{\rm pl}^{*}(\omega;\bm{r},\bm{r}^{\prime})\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime\prime})~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}
=4πIm𝑫pl(ω;𝒓,𝒓′′),\displaystyle=4\pi\text{Im}\bm{D}_{\rm pl}(\omega;\bm{r}^{\prime},\bm{r}^{\prime\prime}), (14)

which ensures consistency with the FD theorem welsch-pra98 ; welsch-p00 ; philbin-njp10 .

II.2 Plasmon Hamiltonian and canonical commutation relations

The normal mode expansion of the plasmon field operator is obtained upon inserting the plasmon Green function (11) into Eq. (5): 𝑬^pl(𝒓)=m𝑬^m(𝒓)\hat{\bm{E}}_{\rm pl}(\bm{r})=\sum_{m}\hat{\bm{E}}_{m}(\bm{r}), where

𝑬^m(𝒓)=ωm4Um𝑬m(𝒓)(a^m+a^m),\displaystyle\hat{\bm{E}}_{m}(\bm{r})=\sqrt{\frac{\hbar\omega_{m}}{4U_{m}}}\,\bm{E}_{m}(\bm{r})(\hat{a}_{m}+\hat{a}^{\dagger}_{m}), (15)

is the field operator for an individual mode. Here, a^m\hat{a}_{m} is the plasmon annihilation operator defined as

a^m=i0dω2πf^m(ω)ωωm+i2γm(ω),\hat{a}_{m}=i\!\int_{0}^{\infty}\!\frac{d\omega}{\sqrt{2\pi}}\frac{\hat{f}_{m}(\omega)}{\omega-\omega_{m}+\frac{i}{2}\gamma_{m}(\omega)}, (16)

where f^m\hat{f}_{m} is noise operator projected on plasmon mode:

f^m(ω)\displaystyle\hat{f}_{m}(\omega) =iπωm2Um𝑑V𝑬m(𝒓)𝑷^N(ω,𝒓)\displaystyle=i\sqrt{\frac{\pi\omega_{m}}{2\hbar U_{m}}}\!\int\!dV\bm{E}_{m}(\bm{r})\!\cdot\!\hat{\bm{P}}_{N}(\omega,\bm{r}) (17)
=ωm8πUm𝑑Vε′′(ω,𝒓)𝑬m(𝒓)𝒇^(ω,𝒓).\displaystyle=-\sqrt{\frac{\omega_{m}}{8\pi U_{m}}}\!\int\!dV\!\sqrt{\varepsilon^{\prime\prime}(\omega,\bm{r})}\,\bm{E}_{m}(\bm{r})\!\cdot\!\hat{\bm{f}}(\omega,\bm{r}).

Commutation relations for f^m\hat{f}_{m} follow from those for 𝒇^\hat{\bm{f}} and from Eqs. (8) and (13),

[f^m(ω),f^n(ω)]=δmnδ(ωω)γm(ω),[\hat{f}_{m}(\omega),\hat{f}_{n}^{\dagger}(\omega^{\prime})]=\delta_{mn}\delta(\omega-\omega^{\prime})\gamma_{m}(\omega), (18)

where commutativity of operators at mnm\neq n is insured by the absence of dissipation-induced coupling. Now, using Eqs. (16) and (18), we obtain the commutation relations for plasmon operators as

[a^m,a^n]=δmn0dω2πγm(ω)(ωmω)2+γm2(ω)/4.[\hat{a}_{m},\hat{a}_{n}^{\dagger}]=\delta_{mn}\!\int_{0}^{\infty}\!\frac{d\omega}{2\pi}\frac{\gamma_{m}(\omega)}{(\omega_{m}-\omega)^{2}+\gamma_{m}^{2}(\omega)/4}. (19)

In the Markov approximation, by replacing γm(ω)\gamma_{m}(\omega) with γmγm(ωm)\gamma_{m}\equiv\gamma_{m}(\omega_{m}) and extending the integral to negative frequencies, we obtain the canonical commutation relations [a^m,a^n]=δmn[\hat{a}_{m},\hat{a}_{n}^{\dagger}]=\delta_{mn}. The plasmon Hamiltonian (1) follows from the normal mode expansion (15) by verifying that the normal-ordered Hamiltonian for individual modes has the form

H^m=18π𝑑V(ωmε)ωm𝑬^m2=ωma^ma^m,\hat{H}_{m}=\frac{1}{8\pi}\!\int\!dV\dfrac{\partial(\omega_{m}\varepsilon^{\prime})}{\partial\omega_{m}}\hat{\bm{E}}_{m}^{2}=\hbar\omega_{m}\hat{a}_{m}^{\dagger}\hat{a}_{m}, (20)

where we dropped the terms a^ma^m\hat{a}_{m}\hat{a}_{m} and a^ma^m\hat{a}_{m}^{\dagger}\hat{a}^{\dagger}_{m} and disregarded the zero-point energy. The factor 1/21/2 difference between Eqs. (12) and (20) reflects the presence of both positive and negative frequency terms in 𝑬^m(𝒓)\hat{\bm{E}}_{m}(\bm{r}) landau . We stress that, by using the plasmon Green function (11), both the Hamiltonian (1) and canonical commutation relations have been explicitly obtained.

Turning to the plasmon dynamics, the time-evolution of projected noise operators (17) is determined by the Heisenberg equations,

f^˙m(ω)=(i/)[f^m(ω),H^N]=iωf^m(ω),\dot{\hat{f}}_{m}(\omega)=-(i/\hbar)[\hat{f}_{m}(\omega),\hat{H}_{N}]=-i\omega\hat{f}_{m}(\omega), (21)

where the dot stands for time derivative. From this relation and Eq. (16), the Heisenberg equations for plasmon operators readily follow (in the Markov approximation),

a^˙m(t)=(γm/2+iωm)a^m(t)+f^m(t),\dot{\hat{a}}_{m}(t)=-(\gamma_{m}/2+i\omega_{m})\hat{a}_{m}(t)+\hat{f}_{m}(t), (22)

where f^m(t)=(2π)1/20𝑑ωf^m(ω)eiωt\hat{f}_{m}(t)\!=\!(2\pi)^{-1/2}\!\int_{0}^{\infty}\!d\omega\hat{f}_{m}(\omega)e^{-i\omega t} is time-domain projected noise operator. The commutation relations for f^m(t)\hat{f}_{m}(t) are obtained from Eq. (18) as

[f^m(t),f^n(t)]=δmnγmδ(tt),[\hat{f}_{m}(t),\hat{f}_{n}^{\dagger}(t^{\prime})]=\delta_{mn}\gamma_{m}\delta(t-t^{\prime}), (23)

where the Markov approximation was used again. Thus, the Markovian dynamics of plasmon operators a^m(t)\hat{a}_{m}(t) is described by quantum Langevin equation (22) with white-noise source f^m(t)\hat{f}_{m}(t), which guarantees scully-book the equal-time commutation relations: [a^m(t),a^n(t)]=δmn[\hat{a}_{m}(t),\hat{a}_{n}^{\dagger}(t)]=\delta_{mn}.

III Plasmon interactions with quantum emitters and the electromagnetic field

Consider now the plasmon coupling to QEs and the EM field. In contrast to cavity modes, the plasmons are localized at the scale well below the diffraction limit and, therefore, interact with the EM field 𝓔(t)\bm{\mathcal{E}}(t) in a way similar to point dipoles. The interaction Hamiltonian has the form Hplem=m𝒑^m𝓔(t)H_{\rm pl-em}=-\sum_{m}\hat{\bm{p}}_{m}\cdot\bm{\mathcal{E}}(t), where 𝒑^m=𝑑V𝑷^m(𝒓)\hat{\bm{p}}_{m}=\!\int\!dV\hat{\bm{P}}_{m}(\bm{r}) is the plasmon dipole moment and 𝑷^m(𝒓)\hat{\bm{P}}_{m}(\bm{r}) is the polarization vector operators. To determine 𝑷^m(𝒓)\hat{\bm{P}}_{m}(\bm{r}), we present the Gauss law (7) as 𝑬m(𝒓)+4π𝑷m(𝒓)=0\bm{\nabla}\!\cdot\!\bm{E}_{m}(\bm{r})+4\pi\bm{\nabla}\!\cdot\!\bm{P}_{m}(\bm{r})=0, where 𝑷m(𝒓)=χ(ωm,𝒓)𝑬m(𝒓)\bm{P}_{m}(\bm{r})=\chi^{\prime}(\omega_{m},\bm{r})\bm{E}_{m}(\bm{r}) is the mode polarization vector and χ(ω,𝒓)\chi(\omega,\bm{r}) is the system susceptibility. In the Markov approximation, converting this relation to operator form as 𝑷^m(𝒓)=χ(ωm,𝒓)𝑬^m(𝒓)\hat{\bm{P}}_{m}(\bm{r})=\chi^{\prime}(\omega_{m},\bm{r})\hat{\bm{E}}_{m}(\bm{r}) and using Eq. (15), for the EM field of the form 𝓔(t)=𝓔eiωLt+𝓔eiωLt\bm{\mathcal{E}}(t)=\bm{\mathcal{E}}e^{-i\omega_{L}t}+\bm{\mathcal{E}}^{*}e^{i\omega_{L}t} that is uniform on the system scale, we obtain in the rotating wave approximation (RWA)

Hplem=m(𝝁m𝓔eiωLta^m+H.c.),H_{\rm pl-em}=-\sum_{m}\left(\bm{\mu}_{m}\cdot\bm{\mathcal{E}}e^{-i\omega_{L}t}\,\hat{a}_{m}^{\dagger}+{\rm H.c.}\right), (24)

where 𝝁m𝝁m(ωm)\bm{\mu}_{m}\equiv\bm{\mu}_{m}(\omega_{m}), and we introduced the frequency-dependent transition matrix element,

𝝁m(ω)=12ωmUm𝑑Vχ(ω,𝒓)𝑬m(𝒓).\bm{\mu}_{m}(\omega)=\frac{1}{2}\sqrt{\frac{\hbar\omega_{m}}{U_{m}}}\!\int\!dV\chi^{\prime}(\omega,\bm{r})\bm{E}_{m}(\bm{r}). (25)

to be used in the following section. The scaling factor ωm/Um\sqrt{\hbar\omega_{m}/U_{m}} in Eq. (25) converts the plasmon energy UmU_{m} to ωm\hbar\omega_{m} in order to match the energy of the EM field. With matrix element (25), the plasmon radiative decay rate is given by the standard expression

γmrad=Wmrad/Um=(4ωm3μm2)/(3c3),\gamma_{m}^{rad}=W_{m}^{rad}/U_{m}=(4\omega_{m}^{3}\mu_{m}^{2})/(3\hbar c^{3}), (26)

where Wmrad=pm2ωm4/3c3W_{m}^{rad}=p_{m}^{2}\omega_{m}^{4}/3c^{3} is the power radiated by dipole 𝒑m\bm{p}_{m} scully-book .

We now turn to the interactions between plasmons and QEs situated at positions 𝒓i\bm{r}_{i} with excitation frequency ωe\omega_{e} and dipole moments 𝒑^i=𝝁i(σ^i+σ^i)\hat{\bm{p}}_{i}=\bm{\mu}_{i}(\hat{\sigma}^{\dagger}_{i}+\hat{\sigma}_{i}), where 𝝁i=μe𝒏i\bm{\mu}_{i}=\mu_{e}\bm{n}_{i} is the transition matrix element (𝒏i\bm{n}_{i} is the dipole orientation). Employing the mode expansion (15) in the interaction Hamiltonian H^plqe=i𝒑^i𝑬^pl(𝒓i)\hat{H}_{\rm pl-qe}=-\sum_{i}\hat{\bm{p}}_{i}\cdot\hat{\bm{E}}_{\rm pl}(\bm{r}_{i}), we obtain, in RWA, the coupling Hamiltonian (2) with gimg_{im} given by

gim=ωm4Um𝝁i𝑬m(𝒓i).\hbar g_{im}=-\sqrt{\frac{\hbar\omega_{m}}{4U_{m}}}\,\bm{\mu}_{i}\!\cdot\!\bm{E}_{m}(\bm{r}_{i}). (27)

Using Eq. (12), the QE-plasmon coupling gimg_{im} can be recast in the form similar to cavities

gim2=2πμe2ωm𝒱m(i),1𝒱m(i)=2[𝒏i𝑬m(𝒓i)]2𝑑V[(ωmε)/ωm]𝑬m2,g^{2}_{im}=\frac{2\pi\mu_{e}^{2}\omega_{m}}{\hbar{\cal V}^{(i)}_{m}},~{}~{}\frac{1}{{\cal V}^{(i)}_{m}}=\frac{2[\bm{n}_{i}\!\cdot\!\bm{E}_{m}(\bm{r}_{i})]^{2}}{\int\!dV[\partial(\omega_{m}\varepsilon^{\prime})/\partial\omega_{m}]\bm{E}_{m}^{2}}, (28)

where 𝒱m(i){\cal V}^{(i)}_{m} is the projected plasmon mode volume shahbazyan-acsphot17 ; shahbazyan-prb18 , which characterizes the plasmon field confinement at a point 𝒓i\bm{r}_{i} in the direction 𝒏i\bm{n}_{i}. Since the Gauss equation (7) is scale-invariant stockman-review , the coupling parameters (25) and (27) are independent of the overall field normalization. By rescaling the fields as 𝑬~m(𝒓)=ωm/4Um𝑬m(𝒓)\tilde{\bm{E}}_{m}(\bm{r})=\sqrt{\hbar\omega_{m}/4U_{m}}\bm{E}_{m}(\bm{r}), these parameters are brought to a more familiar form

gim=𝝁i𝑬~m(𝒓i)/,𝝁m=𝑑Vχ(ωm,𝒓)𝑬~m(𝒓).g_{im}=-\bm{\mu}_{i}\!\cdot\!\tilde{\bm{E}}_{m}(\bm{r}_{i})/\hbar,~{}~{}~{}\bm{\mu}_{m}=\!\int\!dV\chi^{\prime}(\omega_{m},\bm{r})\tilde{\bm{E}}_{m}(\bm{r}). (29)

To illustrate the sensitivity of QE-plasmon coupling to the system parameters, we present the results of numerical calculations of the coupling parameter gemg_{em}, given by Eq. (28), for a single QE situated at a distance dd from the tip of an Au nanorod in water (see Fig. 1). The nanorod was modeled by a prolate spheroid with semi-major and semi-minor axes aa and bb, respectively, at aspect ratio a/b=3.0a/b=3.0, the standard spherical harmonics were used for calculating the longitudinal plasmon fields, the QE dipole orientation was chosen along the nanorod symmetry axis, and the Au experimental dielectric function was used in all calculations. The calculated QE-plasmon coupling is normalized by full plasmon decay rate γm\gamma_{m}, which includes the nonradiative decay rate (13) and radiative decay rate (26) taken at plasmon frequency ω=ωm\omega=\omega_{m}. The coupling parameter behaves as gem1/𝒱mg_{em}\propto 1/\sqrt{{\cal V}_{m}}, where, for small nanostructures, the plasmon mode volume 𝒱m{\cal V}_{m} scales as the metal volume shahbazyan-prb18 ; shahbazyan-nl19 , resulting in the enhanced coupling for smaller nanorods. The coupling sharply increases as QE approaches hot spot at the nanorod tip characterized by very large plasmon field confinement (small mode volume).

Refer to caption
Figure 1: The QE-plasmon coupling gemg_{em} for a QE near the tip of an Au nanorod in water is plotted against the distance to the tip for several values of the nanorod overall size. Inset: Schematics of a QE near the Au nanorod tip.

Finally, the full canonical Hamiltonian for plasmons interacting with the EM field and QEs has the form H=Hpl+Hplqe+Hplem+Hqe+HqeemH=H_{\rm pl}+H_{\rm pl-qe}+H_{\rm pl-em}+H_{\rm qe}+H_{\rm qe-em}, where we included the standard QE Hamiltonian Hqe=ωeiσ^iσ^iH_{\rm qe}=\hbar\omega_{e}\!\sum_{i}\hat{\sigma}_{i}^{\dagger}\hat{\sigma}_{i} and the coupling term Hqeem=i(𝝁i𝓔eiωLtσ^i+H.c.)H_{\rm qe-em}=-\sum_{i}(\bm{\mu}_{i}\!\cdot\!\bm{\mathcal{E}}e^{-i\omega_{L}t}\hat{\sigma}_{i}^{\dagger}+{\rm H.c.}) describing the QEs’ interactions with the EM field. We stress that for plasmons, the canonical quantization scheme is valid only in the Markov approximation that ignores the dielectric function dispersion. However, in metal-dielectric structures, the role of materials optical dispersion can be very significant, and so a quantum description that includes such effects is necessary, as we discuss in the following section.

IV Beyond Markov approximation: Intermediate quantization scheme

In this section, we present yet another plasmon quantization approach that bridges the macroscopic and canonical schemes while combining the advantages of both. In this ”intermediate” scheme, the Hilbert space is restricted to plasmonic degrees of freedom, in contrast to macroscopic approach, while the medium optical dispersion is still accurately accounted for, in contrast to the canonical scheme.

IV.1 Recasting quantum plasmons via projected reservoir modes

The Langevin equation (22) with the source (17) imply that each plasmon mode is driven by a fraction of the reservoir that overlaps with the plasmon electric field. These projected reservoir modes (PRM) form a discrete subspace of the full reservoir Hilbert space spanned by the operators b^m(ω)=f^m(ω)/γm(ω)\hat{b}_{m}(\omega)=\hat{f}_{m}(\omega)/\sqrt{\gamma_{m}(\omega)} or, using Eqs. (13) and (17),

b^m(ω)=𝑑Vε′′(ω,𝒓)𝑬m(𝒓)𝒇^(ω,𝒓)[𝑑Vε′′(ω,𝒓)𝑬m2(𝒓)]1/2,\hat{b}_{m}(\omega)=-\frac{\int\!dV\sqrt{\varepsilon^{\prime\prime}(\omega,\bm{r})}\,\bm{E}_{m}(\bm{r})\!\cdot\!\hat{\bm{f}}(\omega,\bm{r})}{\left[\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\,\bm{E}_{m}^{2}(\bm{r})\right]^{1/2}}, (30)

satisfying the commutation relations

[b^m(ω),b^n(ω)]=δmnδ(ωω).[\hat{b}_{m}(\omega),\hat{b}_{n}^{\dagger}(\omega^{\prime})]=\delta_{mn}\delta(\omega-\omega^{\prime}). (31)

Time-evolution in the reduced Hilbert space is driven by the PRM Hamiltonian with linear dispersion

H^b=m0𝑑ωωb^m(ω)b^m(ω),\hat{H}_{\rm b}=\sum_{m}\!\int_{0}^{\infty}\!d\omega\,\hbar\omega\,\hat{b}^{\dagger}_{m}(\omega)\hat{b}_{m}(\omega), (32)

which leads to accurate Heisenberg equations of the form b^˙m(ω)=(i/)[b^m(ω),H^b]=iωb^m(ω)\dot{\hat{b}}_{m}(\omega)=-(i/\hbar)[\hat{b}_{m}(\omega),\hat{H}_{\rm b}]=-i\omega\hat{b}_{m}(\omega) [compare to Eq. (21)]. The PRMs and plasmons can be set as independent variables by adding the coupling Hamiltonian term: H^plb=im0𝑑ωκm(ω)[a^mb^m(ω)b^m(ω)a^m]\hat{H}_{\rm pl-b}=i\hbar\sum_{m}\!\int_{0}^{\infty}\!\!d\omega\kappa_{m}(\omega)[\hat{a}^{\dagger}_{m}\hat{b}_{m}(\omega)-\hat{b}_{m}^{\dagger}(\omega)\hat{a}_{m}], where κm(ω)=γm(ω)/2π\kappa_{m}(\omega)=\sqrt{\gamma_{m}(\omega)/2\pi} is the plasmon-PRM coupling. Then, upon tracing the PRMs out, one would arrive, in the standard way, at the master equation for density matrix scully-book . Here, we chose a different approach and instead describe the system directly in terms of PRMs.

The interaction Hamiltonian between PRMs and QEs is obtained from the QE-plasmon coupling term (2) by using the relation (16) between the plasmon and PRM operators, with f^m(ω)=b^m(ω)γm(ω)\hat{f}_{m}(\omega)=\hat{b}_{m}(\omega)\sqrt{\gamma_{m}(\omega)}. We then obtain

H^bqe=im0𝑑ω[qim(ω)σ^ib^m(ω)+H.c.],\hat{H}_{\rm b-qe}=\sum_{im}\int_{0}^{\infty}\!d\omega\left[\hbar q_{im}(\omega)\,\hat{\sigma}^{\dagger}_{i}\,\hat{b}_{m}(\omega)+\text{H.c.}\right], (33)

where qim(ω)q_{im}(\omega) is the QE-PRM coupling,

qim(ω)=γm(ω)2πigimωωm+i2γm(ω),\displaystyle q_{im}(\omega)=\sqrt{\frac{\gamma_{m}(\omega)}{2\pi}}\frac{ig_{im}}{\omega-\omega_{m}+\frac{i}{2}\gamma_{m}(\omega)}, (34)

with gimg_{im} given by Eq. (27).

The PRM coupling to the EM field 𝓔(t)\bm{\mathcal{E}}(t) that is uniform on the system scale is described by the Hamiltonian Hint=Re𝑑V𝑬^pl(𝒓)𝑷(t,𝒓)H_{\rm int}=-\text{Re}\int dV\hat{\bm{E}}_{\rm pl}(\bm{r})\cdot\bm{P}(t,\bm{r}), where 𝑷=χ^𝓔\bm{P}=\hat{\chi}\bm{\mathcal{E}} is the induced polarization vector. For a monochromatic field, using Eqs. (15) and (16), we obtain

H^bem=m0𝑑ω[𝒅m(ωL,ω)𝓔eiωLtb^m(ω)+H.c.],\hat{H}_{\rm b-em}=-\!\!\sum_{m}\!\int_{0}^{\infty}\!\!\!d\omega\!\left[\bm{d}_{m}^{*}(\omega_{L},\omega)\!\cdot\!\bm{\mathcal{E}}e^{-i\omega_{L}t}\,\hat{b}_{m}^{\dagger}(\omega)+\text{H.c.}\right], (35)

where 𝒅m(ωL,ω)\bm{d}_{m}(\omega_{L},\omega) is the optical transition matrix element for PRMs [compare to Eq. (34)],

𝒅m(ωL,ω)=γm(ω)2πi𝝁m(ωL)ωωm+i2γm(ω),\displaystyle\bm{d}_{m}(\omega_{L},\omega)=\sqrt{\frac{\gamma_{m}(\omega)}{2\pi}}\frac{i\bm{\mu}_{m}(\omega_{L})}{\omega-\omega_{m}+\frac{i}{2}\gamma_{m}(\omega)}, (36)

and the plasmon transition matrix element 𝝁m(ω)\bm{\mu}_{m}(\omega) is given by Eq. (25).

IV.2 Classical plasmonic enhancement effects and first-order transition probability rates

To elucidate the mechanism behind the QE-PRM interaction, let us evaluate the first-order transition probability rate for PRM excitation by a QE. For an excited QE with energy ω\hbar\omega, the transition rate for an individual PRM has the form

Γim(ω)=2π0𝑑ω|qim(ω)|2δ(ωω),\Gamma_{im}(\omega)=\frac{2\pi}{\hbar}\int_{0}^{\infty}\!d\omega^{\prime}\left|\hbar q_{im}(\omega^{\prime})\right|^{2}\delta(\hbar\omega^{\prime}-\hbar\omega), (37)

where the integration is taken over the PRM’s final states. Evaluating the integral, we obtain

Γim(ω)=2π|qim(ω)|2=gim2γm(ω)(ωmω)2+14γm2(ω),\Gamma_{im}(\omega)=2\pi\left|q_{im}(\omega)\right|^{2}=\frac{g_{im}^{2}\gamma_{m}(\omega)}{(\omega_{m}-\omega)^{2}+\frac{1}{4}\gamma_{m}^{2}(\omega)}, (38)

where we used Eq. (34). The full transition rate is obtained by summing Eq. (38) over all PRMs: Γi(ω)=mΓim(ω)\Gamma_{i}(\omega)=\sum_{m}\Gamma_{im}(\omega). In fact, the above expression represents the rate of energy transfer (ET) from a QE to plasmons evaluated, in a standard way, using the classical plasmon Green function (11) with help of Eq. (27) shahbazyan-prl16 ,

Γi(ω)=2Im[𝝁i𝑫pl(ω;𝒓i,𝒓i)𝝁i]=mΓim(ω),\Gamma_{i}(\omega)=\dfrac{2}{\hbar}\,\text{Im}\left[\bm{\mu}_{i}\bm{D}_{\rm pl}(\omega;\bm{r}_{i},\bm{r}_{i})\bm{\mu}_{i}\right]=\sum_{m}\Gamma_{im}(\omega), (39)

which indicates that QE-PRM interactions are mediated by classical plasmons absorbing the QE energy. Thus, the classical effect of resonance ET from a QE to plasmons emerges from the Hamiltonian (33) in the lowest order.

Turning to PRM interactions with the EM field, the transition probability rate for excitation of a PRM by the incident monochromatic light of frequency ω\omega is

Γm(ω)=2π0𝑑ω|𝒅m(ω,ω)𝓔|2δ(ωω),\Gamma_{m}(\omega)=\frac{2\pi}{\hbar}\!\int_{0}^{\infty}\!d\omega^{\prime}\left|\bm{d}_{m}(\omega,\omega^{\prime})\!\cdot\!\bm{\mathcal{E}}\right|^{2}\delta(\hbar\omega^{\prime}-\hbar\omega), (40)

which, after evaluating the integral, can be presented as [compare to Eqs. (39) and (38)]

Γm(ω)=2π2|𝒅m(ω,ω)𝓔|2=2Im[𝓔𝜶m(ω)𝓔].\Gamma_{m}(\omega)=\dfrac{2\pi}{\hbar^{2}}\left|\bm{d}_{m}(\omega,\omega)\!\cdot\!\bm{\mathcal{E}}\right|^{2}=\dfrac{2}{\hbar}\,\text{Im}\left[\bm{\mathcal{E}}^{*}\bm{\alpha}_{m}(\omega)\bm{\mathcal{E}}\right]. (41)

Here, 𝜶m(ω)\bm{\alpha}_{m}(\omega) is the optical polarizability tensor of a plasmon mode that defines its response to an external field shahbazyan-prb18 (see Appendix):

𝜶m(ω)=1𝝁m(ω)𝝁m(ω)ωmωi2γm(ω).\bm{\alpha}_{m}(\omega)=\frac{1}{\hbar}\frac{\bm{\mu}_{m}(\omega)\bm{\mu}_{m}(\omega)}{\omega_{m}-\omega-\frac{i}{2}\gamma_{m}(\omega)}. (42)

The full absorption rate is obtained by summing over all PRM modes as Γpl(ω)=(2/)Im[𝓔𝜶pl(ω)𝓔]\Gamma_{\rm pl}(\omega)=(2/\hbar)\text{Im}\left[\bm{\mathcal{E}}^{*}\bm{\alpha}_{\rm pl}(\omega)\bm{\mathcal{E}}\right], where 𝜶pl(ω)=m𝜶m(ω)\bm{\alpha}_{\rm pl}(\omega)=\sum_{m}\bm{\alpha}_{m}(\omega) is the full optical polarizability of plasmonic structure. Within RWA, the absorbed power is given by P(ω)=ωΓpl(ω)/4=(ω/2)Im[𝓔𝜶pl(ω)𝓔]P(\omega)=\hbar\omega\Gamma_{\rm pl}(\omega)/4=(\omega/2)\text{Im}\left[\bm{\mathcal{E}}^{*}\bm{\alpha}_{\rm pl}(\omega)\bm{\mathcal{E}}\right]. Upon normalizing it by the incident flux S0=(c/8π)2S_{0}=(c/8\pi)\mathcal{E}^{2}, we obtain the absorption cross-section as σplabs=mσmabs\sigma_{\rm pl}^{\rm abs}=\sum_{m}\sigma_{m}^{\rm abs}, where σmabs(ω)\sigma_{m}^{\rm abs}(\omega) is the absorption cross-section for an individual mode,

σmabs=4πωcIm[𝜺𝜶m(ω)𝜺]=2πωc|𝜺𝝁m(ω)|2γm(ω)(ωωm)2+14γm2(ω),\sigma_{m}^{\rm abs}=\dfrac{4\pi\omega}{c}\text{Im}\left[\bm{\varepsilon}\bm{\alpha}_{m}(\omega)\bm{\varepsilon}\right]=\dfrac{2\pi\omega}{\hbar c}\frac{\bm{|\varepsilon}\!\cdot\!\bm{\mu}_{m}(\omega)|^{2}\gamma_{m}(\omega)}{(\omega-\omega_{m})^{2}+\frac{1}{4}\gamma_{m}^{2}(\omega)}, (43)

and 𝜺\bm{\varepsilon} is the incident light polarization.Thus, the PRM-EM interaction is mediated by plasmon absorption of the incident light, while the classical absorption cross-section is obtained from the interaction Hamiltonian (36) in the lowest order. Obviously, if the incident light frequency ω\omega is close to a plasmon mode frequency ωm\omega_{m}, this interaction is enhanced due to resonant plasmon absorption.

To illustrate the role of optical dispersion in the above first-order effects, in Fig. 2 we compare the normalized absorption cross-section for an Au nanorod in water, calculated using Eq. (43), with the result of Markov approximation obtained by setting ω=ωm\omega=\omega_{m} in the coupling parameters 𝝁m(ω)\bm{\mu}_{m}(\omega) and γm(ω)\gamma_{m}(\omega). With optical dispersion included, the plasmon resonance spectral shape deviates slightly from the Lorentzian by exhibiting a red shift of the peak position along with overall enhancement of the lower frequency range. Although, in this case, the effect appear to be relatively small, much stronger non-Markovian effects are expected in higher orders, including in strongly-coupled QE-plasmon systems to be discussed elsewhere.

Refer to caption
Figure 2: The calculated absorption cross-section of an Au nanorod in water, given by the full expression (43), is compared to the Markov approximation result obtained from the same expression but with ω=ωm\omega=\omega_{m} in the Au dielectric function. Inset: Schematics of the Au nanorod.

The Hamiltonian H^=H^b+H^bqe+H^bem+H^qe+H^qeem\hat{H}=\hat{H}_{\rm b}+\hat{H}_{\rm b-qe}+\hat{H}_{\rm b-em}+\hat{H}_{\rm qe}+\hat{H}_{\rm qe-em} provides a starting point for studying quantum correlations and non-Markovian dynamics in hybrid plasmonic systems involving localized plasmons interacting with QEs and the EM field. Within this framework, classical plasmons are encoded in the coupling parameters (36) and (34), respectively, to mediate resonant coupling between the system components. Namely, the classical effects of resonant plasmon excitation by the EM field and resonance ET between the QEs and plasmons, which underpin most of the plasmon-enhanced spectroscopy phenomena, now emerge in the lowest order of quantum perturbation theory. In higher orders, these classical effects will modulate quantum correlations and non-Markovian dynamics in hybrid plasmonic systems.

V Conclusions

In summary, we developed a quantization approach for surface plasmons in metal-dielectric structures that accounts for optical dispersion of the medium complex dielectric function and, hence, is suitable for describing non-Markovian effects in quantum plasmonics. Starting within the macroscopic quantization framework based on the dissipation-fluctuation theorem, we derived, in the Markov approximation, the canonical quantization scheme for interacting surface plasmons and obtained the relevant coupling parameters in terms of local fields and system geometry. Beyond Markov approximation, we developed a quantum approach in terms of discrete set of bosonic modes with linear dispersion whose interactions with the electromagnetic field and quantum emitters are mediated by the classical plasmons, providing resonant coupling between the system components. We have shown that, within this approach, the classical plasmonic enhancement effects, such as resonant plasmon excitation by incident light and resonance energy transfer from a quantum emitter to plasmonic structure, are obtained in the first-order of perturbation theory. The quantum dynamics of bosonic modes is restricted to the reduced Hilbert space of reservoir states which overlaps with the electric fields of plasmon modes.

Acknowledgements.
This work was supported in part by the National Science Foundation Grants Nos. DMR-2000170, DMR-1856515, DMR-1826886 and HRD-1547754.

Appendix A Plasmon modes

We consider a metal-dielectric structure supporting surface plasmons that are localized at the length scale much smaller than the radiation wavelength. In the absence of retardation effects, each connected volume ViV_{i} of the structure, metallic or dielectric, is characterized by a uniform dielectric function εi(ω)\varepsilon_{i}(\omega) so that the full dielectric function has the form ε(ω,𝒓)=iθi(𝒓)εi(ω)\varepsilon(\omega,\bm{r})=\sum_{i}\theta_{i}(\bm{r})\varepsilon_{i}(\omega), where θi(𝒓)\theta_{i}(\bm{r}) is unit step function that vanishes outside ViV_{i}. The system eigenmodes are determined by the quasistatic Gauss law stockman-review ,

[ε(ωm,𝒓)Φm(𝒓)]=0,\bm{\nabla}\!\cdot\!\left[\varepsilon^{\prime}(\omega_{m},\bm{r})\bm{\nabla}\Phi_{m}(\bm{r})\right]=0, (44)

where Φm(𝒓)\Phi_{m}(\bm{r}) and ωm\omega_{m} are the mode potentials and frequencies, respectively, and the mode electric fields, which can be chosen real, are defined as 𝑬m(𝒓)=Φm(𝒓)\bm{E}_{m}(\bm{r})=-\bm{\nabla}\Phi_{m}(\bm{r}). In the plasmon frequency region, where ε′′(ω)/ε(ω)1\varepsilon^{\prime\prime}(\omega)/\varepsilon^{\prime}(\omega)\ll 1, the mode frequencies are defined by the real part of dielectric function, while its imaginary part defines the mode decay rates.

Let us show that the eigenmodes of Eq. (44) are orthogonal in each connected volume ViV_{i}:

𝑑Vi𝑬m(𝒓)𝑬n(𝒓)=δmn𝑑Vi𝑬m2(𝒓).\int\!dV_{i}\bm{E}_{m}(\bm{r})\!\cdot\!\bm{E}_{n}(\bm{r})=\delta_{mn}\int\!dV_{i}\bm{E}_{m}^{2}(\bm{r}). (45)

Using ε(ω,𝒓)=1+4πχ(ω,𝒓)=1+4πiχi(ω)θi(𝒓)\varepsilon(\omega,\bm{r})=1+4\pi\chi(\omega,\bm{r})=1+4\pi\sum_{i}\chi_{i}(\omega)\theta_{i}(\bm{r}), where χ\chi is the susceptibility, we multiply Eq. (44) by Φn(𝒓)\Phi_{n}(\bm{r}) and integrate over the system volume to obtain

𝑑V𝑬m𝑬n+4πiχi(ωm)𝑑Vi𝑬m𝑬n=0\int\!dV\bm{E}_{m}\!\cdot\!\bm{E}_{n}+4\pi\sum_{i}\chi^{\prime}_{i}(\omega_{m})\!\int\!dV_{i}\bm{E}_{m}\!\cdot\!\bm{E}_{n}=0 (46)

Making a replacement mnm\leftrightarrow n in Eq. (46) and subtracting the result from Eq. (46), we arrive at overcomplete system

i[χi(ωm)χi(ωn)]𝑑Vi𝑬m𝑬n=0,\sum_{i}[\chi^{\prime}_{i}(\omega_{m})-\chi^{\prime}_{i}(\omega_{n})]\!\int\!dV_{i}\bm{E}_{m}\!\cdot\!\bm{E}_{n}=0, (47)

and the orthogonality relation Eq. (45) readily follows. An important consequence of Eq. (45) is the absence of dissipation-induced coupling between the modes, i.e., for mnm\neq n,

𝑑Vε′′(ω,𝒓)𝑬m(𝒓)𝑬n(𝒓)=iεi′′𝑑Vi𝑬m𝑬n=0,\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{E}_{m}(\bm{r})\bm{E}_{n}(\bm{r})=\sum_{i}\varepsilon_{i}^{\prime\prime}\int\!dV_{i}\bm{E}_{m}\bm{E}_{n}=0, (48)

which allows one to obtain the exact plasmon Green function in the presence of losses.

Appendix B Plasmon Green function

The EM dyadic Green function for Maxwell equations in the presence of inhomogeneous medium satisfies

[××ω2c2ε(ω,𝒓)]𝑫(ω;𝒓,𝒓)=4πω2c2𝑰δ(𝒓𝒓)\left[\bm{\nabla}\!\times\!\bm{\nabla}\!\times-\frac{\omega^{2}}{c^{2}}\varepsilon(\omega,\bm{r})\right]\bm{D}(\omega;\bm{r},\bm{r}^{\prime})=\frac{4\pi\omega^{2}}{c^{2}}\bm{I}\delta(\bm{r}-\bm{r}^{\prime}) (49)

where we adopted normalization convenient in the near field limit. Applying \bm{\nabla} to both sides, one finds equation for the longitudinal part of the Green function

[ε(ω,𝒓)𝑫(ω;𝒓,𝒓)]=4π𝑰δ(𝒓𝒓).\bm{\nabla}[\varepsilon(\omega,\bm{r})\bm{D}(\omega;\bm{r},\bm{r}^{\prime})]=-4\pi\bm{\nabla}\bm{I}\delta(\bm{r}-\bm{r}^{\prime}). (50)

In the near field, it is convenient to switch to the Green function for the potentials D(ω;𝒓,𝒓)D(\omega;\bm{r},\bm{r}^{\prime}), defined as 𝑫(ω;𝒓,𝒓)=D(ω;𝒓,𝒓)\bm{D}(\omega;\bm{r},\bm{r}^{\prime})=\bm{\nabla}\bm{\nabla}^{\prime}D(\omega;\bm{r},\bm{r}^{\prime}), which satisfies

[ε(ω,𝒓)D(ω;𝒓,𝒓)]=4πδ(𝒓𝒓).\bm{\nabla}\!\cdot\!\left[\varepsilon(\omega,\bm{r})\bm{\nabla}D(\omega;\bm{r},\bm{r}^{\prime})\right]=4\pi\delta(\bm{r}-\bm{r}^{\prime}). (51)

In free space (ε=1\varepsilon=1), the near-field Green’s function has the form D0(𝒓𝒓)=1/|𝒓𝒓|D_{0}(\bm{r}-\bm{r}^{\prime})=-1/|\bm{r}-\bm{r}^{\prime}|. For arbitrary ε(ω,𝒓)\varepsilon(\omega,\bm{r}), we separate out the free-space and plasmon parts as D=D0+DplD=D_{0}+D_{\rm pl} to obtain the equation for DplD_{\rm pl}:

[ε(ω,𝒓)\displaystyle\bm{\nabla}\!\cdot\!\bigl{[}\varepsilon(\omega,\bm{r})\bm{\nabla} Dpl(ω;𝒓,𝒓)]\displaystyle D_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})\bigr{]}
=[[ε(ω,𝒓)1]D0(ω;𝒓,𝒓)].\displaystyle=-\bm{\nabla}\!\cdot\!\bigl{[}[\varepsilon(\omega,\bm{r})-1]\bm{\nabla}D_{0}(\omega;\bm{r},\bm{r}^{\prime})\bigr{]}. (52)

Assume, for a moment, that the dielectric function ε(ω,𝒓)\varepsilon(\omega,\bm{r}) is real (ε′′=0\varepsilon^{\prime\prime}=0) and expand the plasmon Green’s function in terms of eigenmodes of Eq. (44) as

Dpl(ω;𝒓,𝒓)=mDm(ω)Φm(𝒓)Φm(𝒓),D_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=\sum_{m}D_{m}(\omega)\Phi_{m}(\bm{r})\Phi_{m}(\bm{r}^{\prime}), (53)

with real coefficients Dm(ω)D_{m}(\omega). Let us apply to both sides of Eq. (B) the integral operator 𝑑VΦm(𝒓)Δ\int\!dV^{\prime}\Phi_{m}(\bm{r}^{\prime})\Delta^{\prime}. Using the mode orthogonality, it is easy to prove the relation

𝑑VΦm(𝒓)ΔDpl(ω;𝒓,𝒓)=DmΦm(𝒓)𝑑V𝑬m2(𝒓)\int\!dV^{\prime}\Phi_{m}(\bm{r}^{\prime})\Delta^{\prime}D_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=-D_{m}\Phi_{m}(\bm{r})\!\int\!dV\bm{E}_{m}^{2}(\bm{r}) (54)

to use in the left-hand side, and the relation

𝑑VΦm(𝒓)ΔD0(ω;𝒓,𝒓)=4πΦm(𝒓)\int\!dV^{\prime}\Phi_{m}(\bm{r}^{\prime})\Delta^{\prime}D_{0}(\omega;\bm{r},\bm{r}^{\prime})=4\pi\Phi_{m}(\bm{r}) (55)

to use in the right-hand side. Then, we obtain

Dm[ε(ω,𝒓)Φm(𝒓)]=4π[[ε(ω,𝒓)1]Φm(𝒓)]𝑑V𝑬m2(𝒓).D_{m}\bm{\nabla}\!\cdot\!\bigl{[}\varepsilon(\omega,\bm{r})\bm{\nabla}\Phi_{m}(\bm{r})\bigr{]}=4\pi\dfrac{\bm{\nabla}\!\cdot\!\bigl{[}[\varepsilon(\omega,\bm{r})-1]\bm{\nabla}\Phi_{m}(\bm{r})\bigr{]}}{\int\!dV\bm{E}_{m}^{2}(\bm{r})}. (56)

Finally, multiplying Eq. (56) by Φm(𝒓)\Phi_{m}(\bm{r}) and integrating the result over the system volume, we obtain shahbazyan-prb18

Dm(ω)=4π𝑑V𝑬m2(𝒓)4π𝑑Vε(ω,𝒓)𝑬m2(𝒓),D_{m}(\omega)=\dfrac{4\pi}{\int\!dV\bm{E}_{m}^{2}(\bm{r})}-\dfrac{4\pi}{\int\!dV\varepsilon(\omega,\bm{r})\bm{E}_{m}^{2}(\bm{r})}, (57)

and the plasmon Green function takes the form

𝑫(ω;𝒓,𝒓)=mDm(ω)𝑬m(𝒓)𝑬m(𝒓).\bm{D}(\omega;\bm{r},\bm{r}^{\prime})=\sum_{m}D_{m}(\omega)\bm{E}_{m}(\bm{r})\bm{E}_{m}(\bm{r}^{\prime}). (58)

The first term in Eq. (57) ensures that Dm=0D_{m}=0 in the limit ω\omega\rightarrow\infty (or, in free space with ε=1\varepsilon=1).

To incorporate the losses, we note that in Eq. (B) with complex dielectric function ε(ω,𝒓)=ε(ω,𝒓)+iε′′(ω,𝒓)\varepsilon(\omega,\bm{r})=\varepsilon^{\prime}(\omega,\bm{r})+i\varepsilon^{\prime\prime}(\omega,\bm{r}), the imaginary part can be considered as a perturbation. In the first order, according to the standard perturbation theory, the diagonal matrix element 𝑑Vε′′(ω,𝒓)𝑬m2(𝒓)\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{E}_{m}^{2}(\bm{r}) affects the spectrum but leaves the eigenmodes unchanged, which is equivalent to having full complex dielectric function ε(ω,𝒓)\varepsilon(\omega,\bm{r}) in Eq. (57). In higher orders, both the spectrum and the eigenmodes should change as the perturbation causes transitions between the basis states via non-diagonal terms 𝑑Vε′′(ω,𝒓)𝑬m(𝒓)𝑬n(𝒓)\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{E}_{m}(\bm{r})\bm{E}_{n}(\bm{r}) with mnm\neq n. However, for quasistatic modes, all non-diagonal matrix elements vanish [see Eq. (48)], implying that the plasmon Green function Eq. (58) with complex coefficients (57) is exact in all orders.

Appendix C Plasmon pole expansion

For real ε(ω,𝒓)\varepsilon(\omega,\bm{r}), due to the Gauss law (44), the Green function (58) with coefficients (57) develops a pole as |ω||\omega| approaches ωm\omega_{m}. For a complex dielectric function, the plasmon poles move to the lower half of the complex-frequency plane, and so the Green’s function, being analytic in the entire complex-frequency plane except those poles, can be presented as a sum over all plasmon poles. For ω\omega approaching ωm\omega_{m}, we expand ε(ω,𝒓)\varepsilon^{\prime}(\omega,\bm{r}) near ωm\omega_{m}

ε(ω,𝒓)ε(ωm,𝒓)+ε(ωm,𝒓)ωm2(ω2ωm2),\varepsilon^{\prime}(\omega,\bm{r})\approx\varepsilon^{\prime}(\omega_{m},\bm{r})+\dfrac{\partial\varepsilon^{\prime}(\omega_{m},\bm{r})}{\partial\omega_{m}^{2}}\left(\omega^{2}-\omega_{m}^{2}\right), (59)

where we used ε(ω,𝒓)=ε(ω,𝒓)\varepsilon^{\prime}(\omega,\bm{r})=\varepsilon^{\prime}(-\omega,\bm{r}), and so the coefficient (57), after omitting the non-resonant term, becomes

Dm(ω)=ωm4Um2ωmωm2ω2iωmγm(ω).D_{m}(\omega)=\frac{\omega_{m}}{4U_{m}}\dfrac{2\omega_{m}}{\omega_{m}^{2}-\omega^{2}-i\omega_{m}\gamma_{m}(\omega)}. (60)

Here, we introduced the plasmon mode energy landau

Um=116π𝑑V[ωmε(ωm,𝒓)]ωm𝑬m2(𝒓),\displaystyle U_{m}=\frac{1}{16\pi}\!\int\!dV\dfrac{\partial[\omega_{m}\varepsilon^{\prime}(\omega_{m},\bm{r})]}{\partial\omega_{m}}\,\bm{E}_{m}^{2}(\bm{r}), (61)

and the frequency-dependent decay rate shahbazyan-prb18 ,

γm(ω)=2𝑑Vε′′(ω,𝒓)𝑬m2(𝒓)𝑑V[ε(ωm,𝒓)/ωm]𝑬m2(𝒓),\gamma_{m}(\omega)=\dfrac{2\!\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{E}_{m}^{2}(\bm{r})}{\!\int\!dV[\partial\varepsilon^{\prime}(\omega_{m},\bm{r})/\partial\omega_{m}]\bm{E}_{m}^{2}(\bm{r})}, (62)

where γm(ω)=γm(ω)\gamma_{m}(\omega)=-\gamma_{m}(-\omega). Note that Eq. (60) is valid in the frequency region ε′′(ω)/ε(ω)1\varepsilon^{\prime\prime}(\omega)/\varepsilon^{\prime}(\omega)\ll 1 or, equivalently, ωm/γm1\omega_{m}/\gamma_{m}\gg 1.

The plasmon dyadic Green’s function is given by 𝑫pl(ω;𝒓,𝒓)=Dpl(ω;𝒓,𝒓)\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=\bm{\nabla}\bm{\nabla}^{\prime}D_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime}), where Dpl(ω;𝒓,𝒓)D_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime}) is defined by Eqs. (53) and (60),

𝑫pl(ω;𝒓,𝒓)=mωm22Um𝑬m(𝒓)𝑬m(𝒓)ωm2ω2iωmγm(ω).\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=\sum_{m}\frac{\omega_{m}^{2}}{2U_{m}}\frac{\bm{E}_{m}(\bm{r})\bm{E}_{m}(\bm{r}^{\prime})}{\omega_{m}^{2}-\omega^{2}-i\omega_{m}\gamma_{m}(\omega)}. (63)

Using Eqs. (48) and (62), it is easy to check that the plasmon Green function (63) satisfies the relation

𝑑Vε′′(ω,𝒓)𝑫pl(ω;𝒓,𝒓)\displaystyle\int\!dV\varepsilon^{\prime\prime}(\omega,\bm{r})\bm{D}_{\rm pl}^{*}(\omega;\bm{r},\bm{r}^{\prime}) 𝑫pl(ω;𝒓,𝒓′′)\displaystyle\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime\prime})
=4πIm𝑫pl(ω;𝒓,𝒓′′),\displaystyle=4\pi\text{Im}\bm{D}_{\rm pl}(\omega;\bm{r}^{\prime},\bm{r}^{\prime\prime}), (64)

which is essential in the FD quantization approach.

For ω>0\omega>0, nonresonant contributions to 𝑫pl\bm{D}_{\rm pl} can be disregarded and the Green function takes the form

𝑫pl(ω;𝒓,𝒓)=mωm4Um𝑬m(𝒓)𝑬m(𝒓)ωmωi2γm(ω),\displaystyle\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})=\sum_{m}\frac{\omega_{m}}{4U_{m}}\frac{\bm{E}_{m}(\bm{r})\bm{E}_{m}(\bm{r}^{\prime})}{\omega_{m}-\omega-\frac{i}{2}\gamma_{m}(\omega)}, (65)

which satisfies the relation (C) as well. In the Markov approximation, i.e., γm(ω)γm(ωm)γm\gamma_{m}(\omega)\rightarrow\gamma_{m}(\omega_{m})\equiv\gamma_{m}, the full Green functions (63) or (65) no longer satisfy the relation (C) but, near the resonance, the individual terms do.

Appendix D Optical polarizability

Consider a plasmonic system subjected to an incident monochromatic field 𝓔ieiωt\bm{\mathcal{E}}_{i}e^{-i\omega t} that is uniform on the system scale. The near field generated by the plasmonic system in response to the incident has the form shahbazyan-prb18

𝓔(ω,𝒓)=𝑑Vχ(ω,𝒓)𝑫pl(ω;𝒓,𝒓)𝓔i.\bm{\mathcal{E}}(\omega,\bm{r})=\int\!dV^{\prime}\chi^{\prime}(\omega,\bm{r}^{\prime})\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})\bm{\mathcal{E}}_{i}. (66)

Multiplying by Eq. (66) by χ(ω,𝒓)\chi^{\prime}(\omega,\bm{r}) and integrating over the system volume, we obtain the system induced dipole moment, 𝓟=𝑑Vχ𝓔\bm{\mathcal{P}}=\int dV\chi^{\prime}\bm{\mathcal{E}}, as

𝓟(ω)=𝑑V𝑑Vχ(ω,𝒓)χ(ω,𝒓)𝑫pl(ω;𝒓,𝒓)𝓔i.\bm{\mathcal{P}}(\omega)=\int\!dVdV^{\prime}\chi^{\prime}(\omega,\bm{r})\chi^{\prime}(\omega,\bm{r}^{\prime})\bm{D}_{\rm pl}(\omega;\bm{r},\bm{r}^{\prime})\!\cdot\!\bm{\mathcal{E}}_{i}. (67)

Inserting the plasmon Green function Eq. (63) into Eq. (67), we obtain

𝓟(ω)=𝜶pl(ω)𝓔i\bm{\mathcal{P}}(\omega)=\bm{\alpha}_{\rm pl}(\omega)\bm{\mathcal{E}}_{i} (68)

where 𝜶pl(ω)=m𝜶m(ω)\bm{\alpha}_{\rm pl}(\omega)=\sum_{m}\bm{\alpha}_{m}(\omega) is the plasmon polarizability tensor shahbazyan-prb18 and

𝜶m(ω)=12ωm𝝁m(ω)𝝁m(ω)ωm2ω2iωmγm(ω),\bm{\alpha}_{m}(\omega)=\frac{1}{\hbar}\frac{2\omega_{m}\,\bm{\mu}_{m}(\omega)\bm{\mu}_{m}(\omega)}{\omega_{m}^{2}-\omega^{2}-i\omega_{m}\gamma_{m}(\omega)}, (69)

is the individual mode polarizability tensor while

𝝁m(ω)=ωm4Um𝑑Vχ(ω,𝒓)𝑬m(𝒓)\bm{\mu}_{m}(\omega)=\sqrt{\frac{\hbar\omega_{m}}{4U_{m}}}\!\int\!dV\chi^{\prime}(\omega,\bm{r})\bm{E}_{m}(\bm{r}) (70)

is the plasmon optical transition matrix element. Near the resonance, the mode polarizability simplifies to

𝜶m(ω)=1𝝁m(ω)𝝁m(ω)ωmωi2γm(ω).\bm{\alpha}_{m}(\omega)=\frac{1}{\hbar}\frac{\bm{\mu}_{m}(\omega)\bm{\mu}_{m}(\omega)}{\omega_{m}-\omega-\frac{i}{2}\gamma_{m}(\omega)}. (71)

Note that, in order to satisfy the optical theorem that guarantees energy flux conservation, the plasmon decay rate γm(ω)\gamma_{m}(\omega) should also include the radiative decay contribution shahbazyan-prb18 . The latter is given by a standard expression for a point-like dipole

γmr(ω)=4μm2ω33c3,\gamma_{m}^{r}(\omega)=\frac{4\mu_{m}^{2}\omega^{3}}{3\hbar c^{3}}, (72)

where ω\omega-dependence of μm\mu_{m} is implied. In the Markov approximation, one should set ω=ωm\omega=\omega_{m} in μm\mu_{m} and γm\gamma_{m}.

References

  • (1) M. S. Tame, K. R. McEnery, S. K. Özdemir, J. Lee, S. A. Maier, and M. S. Kim, Quantum plasmonics, Nat. Phys. 9, 329 (2013).
  • (2) J. Bellessa, C. Bonnand, J. C. Plenet, and J. Mugnier, Phys. Rev. Lett. 93, 036404 (2004).
  • (3) Y. Sugawara, T. A. Kelf, J. J. Baumberg, M. E. Abdelsalam, and P. N. Bartlett, Phys. Rev. Lett. 97, 266808 (2006).
  • (4) G. A. Wurtz, P. R. Evans, W. Hendren, R. Atkinson, W. Dickson, R. J. Pollard, A. V. Zayats, W. Harrison, and C. Bower, Nano Lett. 7, 1297 (2007).
  • (5) N. T. Fofang, T.-H. Park, O. Neumann, N. A. Mirin, P. Nordlander, and N. J. Halas, Nano Lett. 8, 3481 (2008).
  • (6) T. K. Hakala, J. J. Toppari, A. Kuzyk, M. Pettersson, H. Tikkanen, H. Kunttu, and P. Torma, Phys. Rev. Lett. 103, 053602 (2009).
  • (7) D. E. Gomez, K. C. Vernon, P. Mulvaney, and T. J. Davis, Nano Lett. 10, 274 (2010).
  • (8) Y. K. Lee, C. H. Jung, J.Park, H. Seo, G. A. Somorjai, and J. Y. Park, Nano Lett. 11, 4251 (2011).
  • (9) F. Wang and N. A. Melosh, Nano Lett. 11, 5426 (2011).
  • (10) M. W. Knight, H. Sobhani, P. Nordlander, and N. J. Halas, Science 332, 702 (2011).
  • (11) D. J. Bergman and M. I. Stockman, Phys. Rev. Lett., 90, 027402, (2003).
  • (12) M. I. Stockman, Nat. Photonics 2, 327, (2008).
  • (13) M. A. Noginov, G. Zhu, A. M. Belgrave, R. Bakker, V. M. Shalaev, E. E. Narimanov, S. Stout, E. Herz, T. Suteewong and U. Wiesner, Nature, 460, 1110, (2009).
  • (14) R. F. Oulton, V. J. Sorger, T. Zentgraf, R.-M. Ma, C. Gladden, L. Dai, G. Bartal, and X. Zhang, Nature 461, 629, (2009).
  • (15) Y.-J. Lu, J. Kim, H.-Y. Chen, C.i Wu, N. Dabidian, C. E. Sanders, C.-Y. Wang, M.-Y. Lu, B.-H. Li, X. Qiu, W.-H. Chang, L.-J. Chen, G. Shvets, C.-K. Shih, and S. Gwo, Science 337, 450 (2012).
  • (16) K. J. Savage, M. M. Hawkeye, R. Esteban, A. G. Borisov, J. Aizpurua, J. J. Baumberg, Nature 491, 574 (2012).
  • (17) D. C. Marinica, A. K. Kazansky, P. Nordlander, J. Aizpurua, and A. G. Borisov, Nano Lett. 12, 1333 (2012).
  • (18) J. A. Scholl, A. García-Etxarri, A. L. Koh, J. A. Dionne, Nano Lett. 13, 564 (2013).
  • (19) S. F. Tan, L. Wu, J. K. Yang, P. Bai, M. Bosman, and C. A. Nijhuis, Science 343, 1496 (2014).
  • (20) M. I. Stockman, in Plasmonics: Theory and Applications, edited by T. V. Shahbazyan and M. I. Stockman (Springer, New York, 2013).
  • (21) A. Trugler and U. Hohenester, Phys. Rev. B 77, 115403 (2008).
  • (22) J. Zuloaga, E. Prodan, and P. Nordlander, Nano Lett. 9, 887 (2009).
  • (23) M. I. Stockman, J. Opt. 12, 024004 (2010).
  • (24) E. Waks and D. Sridharan, Phys. Rev. A 82, 043845 (2010).
  • (25) A. Gonzalez-Tudela, P. A. Huidobro, L. Martin-Moreno, C. Tejedor, and F. J. Garcia-Vidal, Phys. Rev. Lett. 110, 126801 (2013).
  • (26) R. Saez-Blazquez, J. Feist, A. I. Fernandez-Dominguez, and F. J. Garcia-Vidal, Optica 4, 1363 (2017).
  • (27) T. Neuman, J. Aizpurua and R. Esteban, Nanophotonics 9, 295 (2020).
  • (28) G. Khitrova, H. M. Gibbs, M. Kira, S. W. Koch, and A. Scherer, Nature Phys. 2, 81 (2006).
  • (29) F. Alpeggiani, S. D’Agostino, and L. C. Andreani, Phys. Rev. B 86, 035421 (2012).
  • (30) T. V. Shahbazyan Nano Lett. 19, 3273 (2019).
  • (31) C. Tserkezis, A. I. Fernandez-Dominguez, P. A. D. Goncalves, F. Todisco, J. D. Cox, K. Busch, N. Stenger, S. I. Bozhevolnyi, N. A. Mortensen, and C. Wolff, Rep. Prog. Phys. 83, 082401 (2020).
  • (32) Y.-N. Chen, G.-Y. Chen, Y.-Y. Liao, N. Lambert, and F. Nori, Phys. Rev. B 79, 245312 (2009).
  • (33) A. Gonzalez-Tudela, F. J. Rodriguez, L. Quiroga, and C. Tejedor, Phys. Rev. B 82, 115334 (2010).
  • (34) I. Thanopulos, V. Yannopapas, and E. Paspalakis, Phys. Rev. B 95, 075412 (2017).
  • (35) T. Moradi, M. B. Harouni, and M. H. Naderi, Sci. Rep. 8, 12435 (2018).
  • (36) Y.-X. Zhang, Y. Zhang, and K. Molmer, ACS Phot. 6, 871 (2019).
  • (37) H. T. Dung, L. Knöll, and D.-G. Welsch, Phys. Rev. A 57, 3931 (1998).
  • (38) L. Knöll, S. Scheel, D.-G. Welsch, in Coherence and Statistics of Photons and Atoms, Ed. J. Perina (Wiley, New York, 2001), p. 1.
  • (39) T. G. Philbin, New J. Phys. 12 123008 (2010).
  • (40) H.T. Dung, L. Knöll and D.G. Welsch, Phys. Rev. A 62, 053804 (2000).
  • (41) C. Van Vlack, P. T. Kristensen, and S. Hughes, Phys. Rev. B 85, 075303 (2012).
  • (42) R.-C. Ge, C. Van Vlack, P. Yao, J. F. Young, and S. Hughes, Phys. Rev. B 87, 205425 (2013).
  • (43) J. Hakami, L. Wang, and M. S. Zubairy, Phys. Rev. A 89, 053835 (2014).
  • (44) A. Sivan and M. Orenstein, Phys. Rev. B 99, 115436 (2019).
  • (45) D Dzsotjan, A. S. Sorensen, and M. Fleischhauer, Phys. Rev. B 82, 075427 (2010).
  • (46) A. Delga, J. Feist, J. Bravo-Abad, and F. J. Garcia-Vidal, Phys. Rev. Lett. 112, 253601 (2014).
  • (47) B. Rousseaux, D. G. Baranov, M. Käll, T. Shegai, and G. Johansson, Phys. Rev. B 98, 045435 (2018).
  • (48) H. Varguet, B. Rousseaux, D. Dzsotjan, H.R. Jauslin, S. Guerin and G. Colas des Francs, Opt. Lett. 41, 4480 (2016).
  • (49) D. Dzsotjan, B. Rousseaux, H. R. Jauslin, G. Colas des Francs, C. Couteau, and S. Guerin, Phys. Rev. A 94, 023818 (2016).
  • (50) S. Franke, S. Hughes, M. K. Dezfouli, P. T. Kristensen, K. Busch, A. Knorr, and M. Richter, Phys. Rev. Lett. 122, 213901 (2019).
  • (51) R.-C. Ge, P. T. Kristensen, J. F. Young, and S. Hughes, New J. Phys. 16, 113048 (2014).
  • (52) P. T. Kristensen and S. Hughes, ACS Phot. 1, 2 (2014).
  • (53) C. Sauvan, J. P. Hugonin, R. Carminati, and P. Lalanne, Phys. Rev. A 89, 043825 (2014).
  • (54) P. T. Kristensen, R.-C. Ge, and S. Hughes, Phys. Rev. A 92, 053810 (2015).
  • (55) E. A. Muljarov and W. Langbein, Phys. Rev. B 94, 235438 (2016).
  • (56) W. Yan, R.i Faggiani, and P. Lalanne, Phys. Rev. B 97, 205422 (2018).
  • (57) P. Lalanne, W. Yan, K. Vynck, C. Sauvan, and J.‐P. Hugonin, Laser Photon. Rev. 12, 1700113 (2018).
  • (58) T. V. Shahbazyan, Phys. Rev. Lett. 117, 207401 (2016).
  • (59) T. V. Shahbazyan, Phys. Rev. B 98, 115401 (2018).
  • (60) L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media (Elsevier, Amsterdam, 2004).
  • (61) M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University Pres, 1997).
  • (62) T. V. Shahbazyan, ACS Photon. 4, 1003 (2017).