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

Comprehensive Microscopic Theory for Coupling of Longitudinal–Transverse Fields and Individual–Collective Excitations

Tomohiro Yokoyama [email protected] Department of Material Engineering Science, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan    Masayuki Iio Department of Material Engineering Science, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan Department of Physics and Electronics, Graduate School of Engineering, Osaka Prefecture University, 1-1 Gakuen-cho, Naka-ku, Sakai, Osaka 599-8531, Japan    Takashi Kinoshita Department of Physics and Electronics, Graduate School of Engineering, Osaka Prefecture University, 1-1 Gakuen-cho, Naka-ku, Sakai, Osaka 599-8531, Japan    Takeshi Inaoka Department of Physics and Earth Sciences, Faculty of Science, University of the Ryukyus, 1 Senbaru, Nishihara-cho, Okinawa 903-0213, Japan    Hajime Ishihara Department of Material Engineering Science, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan Department of Physics and Electronics, Graduate School of Engineering, Osaka Prefecture University, 1-1 Gakuen-cho, Naka-ku, Sakai, Osaka 599-8531, Japan Center for Quantum Information and Quantum Biology, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan
Abstract

A plasmon is a collective excitation of electrons due to the Coulomb interaction. Both plasmons and single-particle excitations (SPEs) are eigenstates of bulk metallic systems and they are orthogonal to each other. In non-translationally symmetric systems such as nanostructures, plasmons, and SPEs coherently interact. It has been well discussed that the plasmons and SPEs, respectively, can couple with transverse (T) electric fields in such systems, and also that they are coupled with each other via longitudinal (L) field. However, there has been a missing link in the previous studies: the coherent coupling between the plasmons and SPEs mediated by the T field. Herein, we develop a theoretical framework to describe the self-consistent relationship between the plasmons and SPEs through both the L and T fields. The excitations are described in terms of the charge and current densities in a constitutive equation with a nonlocal susceptibility, where the densities include the L and T components. The electromagnetic fields originating from the densities are described in terms of the Green’s function in the Maxwell’s equations. The T field is generated from both densities, whereas the L component is attributed to the charge density only. We introduce a four-vector representation incorporating the vector and scalar potentials in the Coulomb gauge, in which the T and L fields are separated explicitly. The eigenvalues of the matrix for the self-consistent equations appear as the poles of the system excitations. Numerical demonstration of the excitation spectrum is performed for a rectangular nanorod. It indicates a non-negligible shift of the collective excitation and an enhancement of the energy transfer between the excitations by the T-field-mediated interaction. The developed formulation enables to approach unknown mechanisms for enhancement of the coherent coupling between the plasmons and the hot carriers generated by radiative fields.

I INTRODUCTION

Elucidation of the light–matter interaction is one of the core research topics of contemporary physics. Recently, the behaviors of light-induced plasmons in metals have attracted considerable interest Tame13 . A plasmon is a quantum of collective electron motion due to the electron–electron interaction, and has been clearly observed in experiment by Powell and Swan Powell59 . In bulk metals, a plasmon has a longitudinal (L) component only and, hence, cannot be excited by transverse (T) electromagnetic (EM) fields. To induce a plasmon using light, the L component of the EM field, e.g., the evanescent mode on a surface, is required book:NovotnyHecht . Cardinal methods to excite surface plasmon-polaritons (SPPs) on a 2D metal surface have been developed by Otto Otto68 and Kretschmann Kretschmann68 . The induced SPPs cause a charge–density spatial distribution on the surface, which in turn generates L fields. Therefore, the EM fields and induced plasmons must be determined self-consistently Feibelman82 ; Gerhardts84 . In metal nanostructures, the SPPs are localized and enhance the L fields significantly. Thus, SPPs are sensitive to surface states and are, therefore, applied in sensors for gases, molecules, and biological matter Garcia11 ; Zeng11 ; Gandhi19 .

One of the fascinating applications of light-induced plasmons is hot-carrier generation  Govorov14 ; Brongersma15 and injection to combined materials White12 ; Govorov13 ; Kumar19 , which can be utilized for photocatalysis Ueno16 , photodetection Li17 , photocarrier injection Tatsuma17 , and photovoltaics Clavero14 . The hot-carrier generation processes involving localized SPPs are categorized into two types: those achieved through plasmon relaxation or through coherent coupling between plasmons and individual single-particle excitations (SPEs). Some theoretical studies have elucidated the relevant mechanisms of the plasmon relaxation approach. The first-principles calculation has been examined for hot-carrier generation on 2D metal surfaces Sundararaman14 ; Bernardi15 , where the interband and intraband excitations make dominant and tunable contributions to the generation, respectively. Hot-carrier generation for metallic nanostructures has also been investigated based on a phenomenological model with several relaxation times Zhang14 ; Ross15 ; Besteiro17 and using density functional theory Manjavacas14 . It was found that the nanostructure sizes and shapes influence the plasmon via the electronic wavefunction and the enhanced electric field due to the hot spots. In those studies, a unidirectional energy transfer from the plasmon(-like) state to the hot carriers by the electron SPEs was discussed Zhang14 ; Ross15 ; Besteiro17 ; Manjavacas14 ; Chapkin18 . As noted above, the second type of hot-carrier generation is due to coherent coupling between the plasmons and individual SPEs. Recently, Ma et al. explicitly discussed the interplay between the plasmons and SPEs in a nanocluster of Ag atoms using a real-time simulation based on time-dependent density functional theory Ma15 . The on- and off-resonant conditions between the plasmon and SPE were found to change the energy transfer processes. For the off-resonant condition, a coherent Rabi oscillation between the plasmon and SPE was found. You et al. also elucidated the bidirectional energy transfer between the plasmon and SPE based on a model Hamiltonian with the Coulomb interaction You18 , considering excited, injected, and extracted electrons.

In the present work, we focus on the fact that the coupling between the plasmons and SPEs via the T field, as well as the treatment of the microscopic nonlocal response involving the T field, has been missed in the previous frameworks, which is particularly related with the latter case in the above discussion. Recent ab initio calculations successfully implemented coherent coupling or hybridization between plasmon-like and single-particle-like excitations Zhang17 ; Chapkin18 ; Senanayake19 ; Fitzgerald17 ; Ma15 ; You18 ; Schira19 . In those studies, the Coulomb interaction corresponding to part of the induced L field was considered. However, the T field was not precisely considered despite the light-induced excitations. As a phenomenological and semi-classical treatment of the microscopic nonlocal response involving the T field, the hydrodynamic model considering the “pressure” on the conduction electrons was applied to the relation between the polarization (or current) and the electric field Pitarke07 ; McMahan09 ; Raza11 ; Toscano12 ; Moreau13 ; Wang13 ; Mortensen14 ; Christensen14 . In such studies, aspects such as the plasmon peak shift, the presence of additional resonance above the plasma frequency, and the size effect were well discussed Raza11 ; Toscano12 ; Mortensen14 ; Christensen14 . The importance of the nonlocal effect for a dimer nanostructure with a gap much narrower than the light wavelength was noted Toscano12 ; Mortensen14 . However, depending on the sample structure or the situation, the hydrodynamic model requires an additional boundary condition or additional treatments from outside the microscopic model Svendsen20 . Importantly, coherent coupling between the plasmons and SPEs has not yet been fully discussed with this phenomenological model as a basis. For a small nanostructure, the plasmon spectrum becomes indistinguishable from the SPE spectrum Inaoka96 . Therefore, for the nonlocal effect in nanostructures with radiative fields, a microscopic formulation must be developed.

Motivated by the above-mentioned situation, in this study, we develop a quantum mechanical framework for mesoscopic plasmonics, focusing on the L and T components of EM fields and induced electronic responses in nanostructures that describe light-induced plasmons (collective excitations), SPEs (individual excitations), and their coherent coupling via both the L and T fields. An EM field induces charge polarization, charge current, spin fluctuation, and so on, and such electronic responses are described by a constitutive equation with a nonlocal susceptibility. The responses generate both L and T field components, and both field components must be comprehensively considered in a self-consistent treatment of the Maxwell’s and constitutive equations book:cho1 . Our formulation can be applied to arbitrary models of electronic states, such as those based on the Drude model Besteiro17 , jellium model with random phase approximation (RPA) Prodan01 , first-principles calculation Ma15 , or specific exotic materials, e.g., graphene Wang13 ; Koppens11 ; Lu19 . Such applications are expected to facilitate theoretical understanding of the interplay between the individual and collective excitations in nanoscale systems. Very recently, to understand the quantum effect in the light-matter interaction, a combination of the density functional theory and the macroscopic Maxwell’s equation has been discussed Svendsen21 . However, a discussion on the coupling of excitations caused by the L and T field components has not been developed yet. In this study, we develop an understanding of the coherent coupling between the individual and collective excitations in terms of the L and T fields and the charge and current densities. To examine the feasibility of our formulation, we numerically demonstrate the excitation spectrum for a rectangular nanorod. Then, we determine a non-negligible contribution of the T field component for a coherent coupling between the individual and collective excitations and energy dissipation. Further, our formulation provides a platform for the study of unrevealed physics concerning not only plasmons, but also excitons and/or other quasiparticles within nanostructures.

The remainder of this article is structured as follows. In Section II, we describe the basic quantities used in our framework and the nanostructure Hamiltonian. The Hamiltonian is separated into a nonperturbative term and an interaction with the field-induced potentials. This separation determines the treatment of the optical response and the electronic states of the nanostructures. In Section III, we introduce the four-vector representation and deduce the nonlocal susceptibility. In Section IV, we present a formal solution of the Maxwell’s equations with vector and scalar potentials, which is given by the Green’s function describing the propagated fields from the excited current and charge densities. In Section V, we explain the self-consistent treatment of the constitutive and Maxwell’s equations demonstrated in Secs. III and IV. In Section VI, we show the result of the numerical calculation of individual and collective excitations for a rectangular nanorod to demonstrate practicality of our formulation. In Section VII, we discuss possible applications and advantaged developments of our LT formulation for plasmonics and photonics. Finally, we present the summary in Section VIII.

II Basic quantities and Hamiltonian of electrons with electromagnetic field

In the present formulation, the fields are described by the vector and scalar potentials, 𝑨(𝒓,t)\bm{A}(\bm{r},t) and ϕ(𝒓,t)\phi(\bm{r},t), respectively. In the quantum mechanics, they are essential rather than the electric and magnetic fields, 𝑬\bm{E} and 𝑩\bm{B}. It has been experimentally proven by the Aharonov–Bohm experiment AB59 ; Tonomura86 . To divide the fields into the L and T components, we apply the Coulomb gauge, div𝑨(𝒓,t)=0{\rm div}\bm{A}(\bm{r},t)=0, where the vector potential describes the T field only. In the constitutive equation, the electronic responses corresponding to the plasmon and SPE are described by the charge and current densities, ρ(𝒓,t)\rho(\bm{r},t) and 𝒋(𝒓,t)\bm{j}(\bm{r},t), respectively, which are directly coupled with ϕ\phi and 𝑨\bm{A} in the interaction Hamiltonian. The nonlocal susceptibility is deduced from the Hamiltonian in accordance with quantum linear response theory Kubo57 ; book:cho1 , where the electron eigenstates determine the susceptibility. Therefore, the nanostructure sizes, shapes, and internal structures strongly affect the optical response via the electron states.

To clearly define the light–matter interaction, we first discuss the matter and interaction Hamiltonians in terms of the potentials. Within the nanostructures, mixing of the L and T field components and the densities occurs spontaneously. To formulate the L and T components explicitly, we introduce a four-vector representation for the fields 𝓐=(𝑨,ϕ/c)\bm{\mathcal{A}}=(\bm{A},-\phi/c) and densities 𝓙=(𝒋,cρ)\bm{\mathcal{J}}=(\bm{j},c\rho). Hence, the nonlocal susceptibility is expressed as a 4×44\times 4 matrix. This formulation enables us to exhibit the roles of the LT components of the fields and densities in enhancing the optical response and the interplay between the plasmons and SPEs.

The Hamiltonian for the electrons in the EM fields is

H^=j[{𝒑^je𝑑𝒓𝑨(𝒓,t)δ(𝒓𝒓^j)}22mεF]+ej𝑑𝒓ϕ(𝒓,t)δ(𝒓𝒓^j)+12ije24πε0|𝒓^i𝒓^j|,\hat{H}=\sum_{j}\left[\frac{\left\{\hat{\bm{p}}_{j}-e\int d\bm{r}\bm{A}(\bm{r},t)\delta(\bm{r}-\hat{\bm{r}}_{j})\right\}^{2}}{2m^{*}}-\varepsilon_{\rm F}\right]+e\sum_{j}\int d\bm{r}\phi(\bm{r},t)\delta(\bm{r}-\hat{\bm{r}}_{j})+\frac{1}{2}\sum_{i\neq j}\frac{e^{2}}{4\pi\varepsilon_{0}|\hat{\bm{r}}_{i}-\hat{\bm{r}}_{j}|}, (1)

where ee is the charge of electron and ε0\varepsilon_{0} is the dielectric constant in vacuum. εF\varepsilon_{\rm F} is the Fermi energy and mm is the (effective) mass for the conduction electrons. The ^\hat{\ } symbol indicates the electron operator. In the case of the Coulomb gauge, div𝑨=0{\rm div}\bm{A}=0, and ϕ(𝒓,t)=ϕncl(𝒓,t)+ϕext(𝒓,t)\phi(\bm{r},t)=\phi_{\rm ncl}(\bm{r},t)+\phi_{\rm ext}(\bm{r},t) in the second term describes the L field due to the nuclei and external origins. The third term is the electron–electron (Coulomb) interaction, which is also longitudinal.

The first term is expanded as

{𝒑^je𝑑𝒓𝑨(𝒓,t)δ(𝒓𝒓^j)}22m\displaystyle\frac{\left\{\hat{\bm{p}}_{j}-e\int d\bm{r}\bm{A}(\bm{r},t)\delta(\bm{r}-\hat{\bm{r}}_{j})\right\}^{2}}{2m^{*}} =\displaystyle= 𝒑^j22me2m𝑑𝒓(𝒑^jδ(𝒓𝒓^j)+δ(𝒓𝒓^j)𝒑^j)𝑨(𝒓,t)\displaystyle\frac{\hat{\bm{p}}_{j}^{2}}{2m^{*}}-\frac{e}{2m^{*}}\int d\bm{r}\left(\hat{\bm{p}}_{j}\delta(\bm{r}-\hat{\bm{r}}_{j})+\delta(\bm{r}-\hat{\bm{r}}_{j})\hat{\bm{p}}_{j}\right)\cdot\bm{A}(\bm{r},t) (2)
+e22m𝑑𝒓δ(𝒓𝒓^j)𝑨(𝒓^j,t)𝑨(𝒓,t).\displaystyle\hskip 85.35826pt+\frac{e^{2}}{2m^{*}}\int d\bm{r}\delta(\bm{r}-\hat{\bm{r}}_{j})\bm{A}(\hat{\bm{r}}_{j},t)\cdot\bm{A}(\bm{r},t).

With the introduction of a current operator

𝑰^(𝒓)e2mj(𝒑^jδ(𝒓𝒓^j)+δ(𝒓𝒓^j)𝒑^j),\hat{\bm{I}}(\bm{r})\equiv\frac{e}{2m^{*}}\sum_{j}\left(\hat{\bm{p}}_{j}\delta(\bm{r}-\hat{\bm{r}}_{j})+\delta(\bm{r}-\hat{\bm{r}}_{j})\hat{\bm{p}}_{j}\right), (3)

the second term on the right-hand side (r.h.s.) of Eq. (2) gives the light–matter interaction

H^IA=𝑑𝒓𝑰^(𝒓)𝑨(𝒓,t).\hat{H}_{IA}=-\int d\bm{r}\hat{\bm{I}}(\bm{r})\cdot\bm{A}(\bm{r},t). (4)

Then, the Hamiltonian can be separated as follows:

H^=H^0+(H^IA+H^A2+H^ext),\hat{H}=\hat{H}_{0}+\left(\hat{H}_{IA}+\hat{H}_{A^{2}}+\hat{H}_{\rm ext}\right), (5)

with

H^0=j[𝒑^j22mεF]+ej𝑑𝒓ϕncl(𝒓,t)δ(𝒓𝒓^j)+12e24πε0ij1|𝒓^i𝒓^j|.\hat{H}_{0}=\sum_{j}\left[\frac{\hat{\bm{p}}_{j}^{2}}{2m^{*}}-\varepsilon_{\rm F}\right]+e\sum_{j}\int d\bm{r}\phi_{\rm ncl}(\bm{r},t)\delta(\bm{r}-\hat{\bm{r}}_{j})+\frac{1}{2}\frac{e^{2}}{4\pi\varepsilon_{0}}\sum_{i\neq j}\frac{1}{|\hat{\bm{r}}_{i}-\hat{\bm{r}}_{j}|}. (6)

Note that the summations over i,ji,j include the electron spin degrees of freedom.

The second quantization of these Hamiltonians is

H^0\displaystyle\hat{H}_{0} =\displaystyle= 𝑑𝒙Ψ^(𝒙)(2x22mεF)Ψ^(𝒙)+e𝑑𝒙Ψ^(𝒙)𝑑𝒓ϕncl(𝒓,t)δ(𝒓𝒙)Ψ^(𝒙)\displaystyle\int d\bm{x}\hat{\Psi}^{\dagger}(\bm{x})\left(-\frac{\hbar^{2}\bm{\nabla}_{x}^{2}}{2m^{*}}-\varepsilon_{\rm F}\right)\hat{\Psi}(\bm{x})+e\int d\bm{x}\hat{\Psi}^{\dagger}(\bm{x})\int d\bm{r}\phi_{\rm ncl}(\bm{r},t)\delta(\bm{r}-\bm{x})\hat{\Psi}(\bm{x}) (7)
+12e24πε0𝑑𝒙𝑑𝒙Ψ^(𝒙)Ψ^(𝒙)1|𝒙𝒙|Ψ^(𝒙)Ψ^(𝒙)\displaystyle\hskip 56.9055pt+\frac{1}{2}\frac{e^{2}}{4\pi\varepsilon_{0}}\int d\bm{x}\int d\bm{x}^{\prime}\hat{\Psi}^{\dagger}(\bm{x})\hat{\Psi}^{\dagger}(\bm{x}^{\prime})\frac{1}{|\bm{x}-\bm{x}^{\prime}|}\hat{\Psi}(\bm{x}^{\prime})\hat{\Psi}(\bm{x})
=\displaystyle= n,n𝑑𝒙ψn(𝒙)(2x22mεF)ψn(𝒙)a^na^n+n,n𝑑𝒙ρnn(𝒙)ϕncl(𝒙,t)a^na^n\displaystyle\sum_{n,n^{\prime}}\int d\bm{x}\psi_{n^{\prime}}^{*}(\bm{x})\left(-\frac{\hbar^{2}\bm{\nabla}_{x}^{2}}{2m^{*}}-\varepsilon_{\rm F}\right)\psi_{n}(\bm{x})\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{n}+\sum_{n,n^{\prime}}\int d\bm{x}\rho_{n^{\prime}n}(\bm{x})\phi_{\rm ncl}(\bm{x},t)\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{n}
+12n,n,m,m𝑑𝒙𝑑𝒙ρnn(𝒙)ρmm(𝒙)4πε0|𝒙𝒙|a^na^ma^ma^n,\displaystyle\hskip 113.81102pt+\frac{1}{2}\sum_{n,n^{\prime},m,m^{\prime}}\int d\bm{x}\int d\bm{x}^{\prime}\frac{\rho_{n^{\prime}n}(\bm{x})\rho_{m^{\prime}m}(\bm{x}^{\prime})}{4\pi\varepsilon_{0}|\bm{x}-\bm{x}^{\prime}|}\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{m^{\prime}}^{\dagger}\hat{a}_{m}\hat{a}_{n},
H^IA\displaystyle\hat{H}_{IA} =\displaystyle= n,n𝑑𝒙𝑰nn(𝒙)𝑨(𝒙,t)a^na^n,\displaystyle-\sum_{n,n^{\prime}}\int d\bm{x}\bm{I}_{n^{\prime}n}(\bm{x})\cdot\bm{A}(\bm{x},t)\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{n}, (8)
H^A2\displaystyle\hat{H}_{A^{2}} =\displaystyle= e22mn,n𝑑𝒙ψn(𝒙)𝑨(𝒙,t)𝑨(𝒙,t)ψn(𝒙)a^na^n,\displaystyle\frac{e^{2}}{2m^{*}}\sum_{n,n^{\prime}}\int d\bm{x}\psi_{n^{\prime}}^{*}(\bm{x})\bm{A}(\bm{x},t)\cdot\bm{A}(\bm{x},t)\psi_{n}(\bm{x})\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{n}, (9)
H^ext\displaystyle{\hat{H}_{\rm ext}} =\displaystyle= n,n𝑑𝒙ρnn(𝒙)ϕext(𝒙,t)a^na^n,\displaystyle{\sum_{n,n^{\prime}}\int d\bm{x}\rho_{n^{\prime}n}(\bm{x})\phi_{\rm ext}(\bm{x},t)\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{n},} (10)

where Ψ^(𝒙)=nψn(𝒙)a^n\hat{\Psi}(\bm{x})=\sum_{n}\psi_{n}(\bm{x})\hat{a}_{n} is the field operator. Note that the nn state includes the spin degrees of freedom. Here,

ρnn(𝒙)=eψn(𝒙)ψn(𝒙)\rho_{n^{\prime}n}(\bm{x})=e\psi_{n^{\prime}}^{*}(\bm{x})\psi_{n}(\bm{x}) (11)

indicates an element of charge density and

𝑰nn(𝒙)=e2im[{xψn(𝒙)}ψn(𝒙)ψn(𝒙){xψn(𝒙)}]\bm{I}_{n^{\prime}n}(\bm{x})=-\frac{e\hbar}{2im^{*}}\Big{[}\left\{\bm{\nabla}_{x}\psi_{n^{\prime}}^{*}(\bm{x})\right\}\psi_{n}(\bm{x})-\psi_{n^{\prime}}^{*}(\bm{x})\left\{\bm{\nabla}_{x}\psi_{n}(\bm{x})\right\}\Big{]} (12)

is derived from the current operator in Eq. (3). Note that ρnn(𝒙)\rho_{n^{\prime}n}(\bm{x}) has to follow the continuous relation with the charge current density. Further, the single-electron wave function follows the time-dependent Schrödinger equation,

iψ(𝒙,t)t=[22m(xie𝑨(𝒙,t))2+eϕncl(𝒙,t)]ψ(𝒙,t).i\hbar\frac{\partial\psi(\bm{x},t)}{\partial t}=\left[-\frac{\hbar^{2}}{2m^{*}}\left(\bm{\nabla}_{x}-\frac{ie}{\hbar}\bm{A}(\bm{x},t)\right)^{2}+e\phi_{\rm ncl}(\bm{x},t)\right]\psi(\bm{x},t). (13)

Then, ρnn(𝒙)\rho_{n^{\prime}n}(\bm{x}) satisfies

ρnn(𝒙,t)t=x𝒋nn(𝒙,t),\frac{\partial\rho_{n^{\prime}n}(\bm{x},t)}{\partial t}=-\bm{\nabla}_{x}\cdot\bm{j}_{n^{\prime}n}(\bm{x},t), (14)

where

𝒋nn(𝒙,t)=𝑰nn(𝒙)e2mψn(𝒙)𝑨(𝒙,t)ψn(𝒙)\bm{j}_{n^{\prime}n}(\bm{x},t)=\bm{I}_{n^{\prime}n}(\bm{x})-\frac{e^{2}}{m^{*}}\psi_{n^{\prime}}^{*}(\bm{x})\bm{A}(\bm{x},t)\psi_{n}(\bm{x}) (15)

is the modified current density in the EM field. The second term contributes to the O(𝑨2)O(\bm{A}^{2}) term. The light–matter interaction component with the vector potential in Eq. (5) is approximately

H^IA+H^A2\displaystyle\hat{H}_{IA}+\hat{H}_{A^{2}} \displaystyle\simeq n,n𝑑𝒙𝒋nn(𝒙,t)𝑨(𝒙,t)a^na^n\displaystyle-\sum_{n,n^{\prime}}\int d\bm{x}\bm{j}_{n^{\prime}n}(\bm{x},t)\cdot\bm{A}(\bm{x},t)\hat{a}_{n^{\prime}}^{\dagger}\hat{a}_{n} (16)
=\displaystyle= 𝑑𝒙𝒋^(𝒙,t)𝑨(𝒙,t)H^int.\displaystyle-\int d\bm{x}\hat{\bm{j}}(\bm{x},t)\cdot\bm{A}(\bm{x},t)\equiv\hat{H}_{\rm int}.

The second and third terms on the r.h.s. of Eq. (7) can be interpreted as the L component of the electric fields due to internal sources. In a following formulation of the constitutive equation, the T and L fields are explicitly distinguished by the vector and scalar potentials, respectively, with the Coulomb gauge. To describe the susceptibility for the vector and scalar potentials, we treat the many-electron system with H^0\hat{H}_{0} on a one-electron basis and rewrite the second and third terms as

𝑑𝒙n,na^nρnn(𝒙,t)[ϕncl(𝒙,t)+e4πε0𝑑𝒙Ψ^(𝒙,t)Ψ^(𝒙,t)|𝒙𝒙|]a^n\displaystyle\int d\bm{x}\sum_{n,n^{\prime}}\hat{a}_{n^{\prime}}^{\dagger}\rho_{n^{\prime}n}(\bm{x},t)\left[\phi_{\rm ncl}(\bm{x},t)+\frac{e}{4\pi\varepsilon_{0}}\int d\bm{x}^{\prime}\frac{\hat{\Psi}^{\dagger}(\bm{x}^{\prime},t)\hat{\Psi}(\bm{x}^{\prime},t)}{|\bm{x}-\bm{x}^{\prime}|}\right]\hat{a}_{n}
=𝑑𝒙n,na^nρnn(𝒙,t)[ϕ^mat(𝒙,t)]a^n,\displaystyle=\int d\bm{x}\sum_{n,n^{\prime}}\hat{a}_{n^{\prime}}^{\dagger}\rho_{n^{\prime}n}(\bm{x},t)\left[\hat{\phi}_{\rm{mat}}(\bm{x},t)\right]\hat{a}_{n}, (17)

with an inherent scalar potential operator for matter,

ϕ^mat(𝒙,t)\displaystyle\hat{\phi}_{\rm{mat}}(\bm{x},t) =\displaystyle= ϕncl(𝒙,t)+ϕ^ee(𝒙,t),\displaystyle\phi_{\rm ncl}(\bm{x},t)+\hat{\phi}_{\rm e-e}(\bm{x},t), (18)
ϕ^ee(𝒙,t)\displaystyle{\hat{\phi}_{\rm e-e}(\bm{x},t)} \displaystyle\equiv e4πε0𝑑𝒙Ψ^(𝒙,t)Ψ^(𝒙,t)|𝒙𝒙|.\displaystyle{\frac{e}{4\pi\varepsilon_{0}}\int d\bm{x}^{\prime}\frac{\hat{\Psi}^{\dagger}(\bm{x}^{\prime},t)\hat{\Psi}(\bm{x}^{\prime},t)}{|\bm{x}-\bm{x}^{\prime}|}.} (19)

Note that, in the absence of 𝑨\bm{A} and ϕext\phi_{\rm ext}, a material is electrically neutral, and ϕ^mat(𝒙,t)0=0\left\langle\hat{\phi}_{\rm{mat}}(\bm{x},t)\right\rangle_{0}=0. Here, we take the basis for the static state without external fields and consider H^0\hat{H}_{0} in Eq. (7) as a non-perturbative Hamiltonian. When one applies the electromagnetic field of 𝑨(𝒙,t)\bm{A}(\bm{x},t) and ϕext(𝒙,t)\phi_{\rm ext}(\bm{x},t) to the material, a polarized charge is induced. It gives an additional interaction between the induced polarized charges,

H^pp\displaystyle\hat{H}_{\rm p-p} \displaystyle\simeq 𝑑𝒙ρ^(𝒙,t)ρ0(𝒙)4πε0𝑑𝒙ρ^(𝒙,t)ρ0(𝒙)|𝒙𝒙|\displaystyle\int d\bm{x}\frac{\hat{\rho}(\bm{x},t)-\rho_{0}(\bm{x})}{4\pi\varepsilon_{0}}\int d\bm{x}^{\prime}\frac{\hat{\rho}(\bm{x}^{\prime},t)-\rho_{0}(\bm{x}^{\prime})}{|\bm{x}-\bm{x}^{\prime}|} (20)
=\displaystyle= 𝑑𝒙δρ^(𝒙,t)ϕ^pol(𝒙,t).\displaystyle{\int d\bm{x}\delta\hat{\rho}(\bm{x},t)\hat{\phi}_{\rm pol}(\bm{x},t).}

Here, ρ0(𝒙)=ρ^(𝒙,t)0\rho_{0}(\bm{x})=\left\langle\hat{\rho}(\bm{x},t)\right\rangle_{0} is the average charge density from the electrons [ρ^(𝒙,t)eΨ^(𝒙,t)Ψ^(𝒙,t)\hat{\rho}(\bm{x},t)\equiv e\hat{\Psi}^{\dagger}(\bm{x},t)\hat{\Psi}(\bm{x},t)]. Thus, δρ^(𝒙,t)=ρ^(𝒙,t)ρ0(𝒙)\delta\hat{\rho}(\bm{x},t)=\hat{\rho}(\bm{x},t)-\rho_{0}(\bm{x}) indicates a deviation from the static distribution. The scalar potential due to the induced polarized charge is

ϕ^pol(𝒙,t)=𝑑𝒙δρ^(𝒙,t)4πε0|𝒙𝒙|.{\hat{\phi}_{\rm pol}(\bm{x},t)=\int d\bm{x}^{\prime}\frac{\delta\hat{\rho}(\bm{x}^{\prime},t)}{4\pi\varepsilon_{0}|\bm{x}-\bm{x}^{\prime}|}.} (21)

Therefore, the L field consists of three elements: ϕ^mat\hat{\phi}_{\rm{mat}}, ϕ^pol\hat{\phi}_{\rm{pol}}, and ϕext\phi_{\rm{ext}}, which are associated with the electron–electron interaction in the static case, the induced charge densities, and the external field, respectively.

In this study, we divide the total Hamiltonian as H^=H^0+H^ind\hat{H}=\hat{H}_{0}+\hat{H}_{\rm ind} with

H^ind=H^int+H^ext+H^pp.\hat{H}_{\rm ind}=\hat{H}_{\rm int}+\hat{H}_{\rm ext}+\hat{H}_{\rm p-p}. (22)

H^0\hat{H}_{0} describes the static system without any field irradiation, whereas all induced effects by the external electric and magnetic fields are in H^ind\hat{H}_{\rm ind}. Then, the incident and induced components are included in the perturbative term. For the scalar potential of the induced polarized charge, we apply the mean field approximation, Ψ^(𝒙,t)ϕ^pol(𝒙,t)Ψ^(𝒙,t)Ψ^(𝒙,t)Ψ^(𝒙,t)ϕpol(𝒙,t)\hat{\Psi}^{\dagger}(\bm{x},t)\hat{\phi}_{\rm pol}(\bm{x},t)\hat{\Psi}(\bm{x},t)\simeq\hat{\Psi}^{\dagger}(\bm{x},t)\hat{\Psi}(\bm{x},t)\phi_{\rm pol}(\bm{x},t). For the external potential, 𝑑𝒙ρ0(𝒙)ϕext(𝒙,t)\int d\bm{x}\rho_{0}(\bm{x})\phi_{\rm ext}(\bm{x},t) corresponds to the system energy shifts without excitation, as it is constant. Subtracting this term from the Hamiltonian, we obtain

H^ind\displaystyle\hat{H}_{\rm ind}^{\prime} =\displaystyle= H^ind(𝑑𝒙ρ0(𝒙)ϕext(𝒙,t))\displaystyle\hat{H}_{\rm ind}-\left(\int d\bm{x}\rho_{0}(\bm{x})\phi_{\rm ext}(\bm{x},t)\right) (23)
\displaystyle\simeq 𝑑𝒙[𝒋^(𝒙,t)𝑨(𝒙,t)δρ^(𝒙,t)(ϕext(𝒙,t)+ϕpol(𝒙,t))],\displaystyle-\int d\bm{x}\left[\hat{\bm{j}}(\bm{x},t)\cdot\bm{A}(\bm{x},t)-\delta\hat{\rho}(\bm{x},t)\Big{(}\phi_{\rm ext}(\bm{x},t)+\phi_{\rm pol}(\bm{x},t)\Big{)}\right],

which describes the interaction between the matter and applied and induced EM fields for both the L and T components. The L field is included in both the constitutive and Maxwell’s equations. However, we do not need to take care of a double count of the electron–electron interaction (see Appendix A).

The susceptibility is obtained from the linear response theory Kubo57 . The susceptibility, field-induced charge density, and induced current density formulations depend on the treatment of the non-perturbative and perturbative Hamiltonians. Note that H^0\hat{H}_{0} includes neither a net charge nor current. Therefore, for the formulation of the susceptibility based on H^=H^0+H^ind\hat{H}=\hat{H}_{0}+\hat{H}_{\rm ind}^{\prime} is evaluated by the static states |μ|\mu\rangle for H^0\hat{H}_{0}. As this susceptibility is irrelevant to the induced fields, we can analyze the T and L components of the field-induced charge and current in terms of 𝑨\bm{A} (T field) and ϕ\phi (L field). A self-consistent relation between the induced charge and ϕpol\phi_{\rm{pol}} can describe a plasmon (collective excitation) within the RPA Prodan01 , as discussed in the following section.

We can omit δ\delta of δρ^(𝒙)\delta\hat{\rho}(\bm{x}) in Eq. (23) for simplicity. In the following discussion, the charge density ρ^(𝒙)\hat{\rho}(\bm{x}) indicates the induced charge only. This abbreviation does not change the continuous relation, as ρ0/t=0\partial\rho_{0}/\partial t=0.

III Four-vector representation for nonlocal susceptibility

The light–matter interaction H^ind(t)\hat{H}_{\rm ind}^{\prime}(t) in Eq. (23) gives a susceptibility to the external fields, 𝑨(𝒙,t)\bm{A}(\bm{x},t) and ϕind(𝒙,t)=ϕext(𝒙,t)+ϕpol(𝒙,t)\phi_{\rm ind}(\bm{x},t)=\phi_{\rm ext}(\bm{x},t)+\phi_{\rm pol}(\bm{x},t) in linear response theory Kubo57 . Here, ϕind\phi_{\rm ind} indicates a scalar potential caused by the field irradiation. In the following, let us omit the subscript “ind{\rm ind}” on H^ind\hat{H}_{\rm ind}^{\prime} and ϕind\phi_{\rm ind} for simplicity. For the interaction Hamiltonian H^(t)\hat{H}^{\prime}(t), let us introduce the following four-vector representation:

𝓐(𝒙,t)\displaystyle{\bm{\mathcal{A}}(\bm{x},t)} =\displaystyle= (𝑨(𝒙,t)1cϕ(𝒙,t)),\displaystyle{\left(\begin{matrix}\bm{A}(\bm{x},t)\\ -\frac{1}{c}\phi(\bm{x},t)\end{matrix}\right),} (24)
𝓙^(𝒙,t)\displaystyle\hat{\bm{\mathcal{J}}}(\bm{x},t) =\displaystyle= (𝒋^(𝒙,t)cρ^(𝒙,t)),\displaystyle\left(\begin{matrix}\hat{\bm{j}}(\bm{x},t)\\ c\hat{\rho}(\bm{x},t)\end{matrix}\right), (25)
H^(t)\displaystyle\hat{H}^{\prime}(t) =\displaystyle= 𝑑𝒙𝓙^(𝒙,t)𝓐(𝒙,t).\displaystyle-\int d\bm{x}\hat{\bm{\mathcal{J}}}(\bm{x},t)\cdot\bm{\mathcal{A}}(\bm{x},t). (26)

The time-dependence of H^(t)\hat{H}^{\prime}(t) in Eq. (26) is attributed to the external field 𝓐(𝒙,t)\bm{\mathcal{A}}(\bm{x},t) only, and cc is the light velocity in vacuum. Note that we define the sign for the scalar potential component as negative in Eq. (24) for a simpler formulation.

We assume a monochromatic field, 𝓐(𝒙,t)=𝓐(𝒙;ω)eiωt\bm{\mathcal{A}}(\bm{x},t)=\bm{\mathcal{A}}(\bm{x};\omega)e^{-i\omega t}. The statistical average of the four-vector current 𝓙^(𝒙)\hat{\bm{\mathcal{J}}}(\bm{x}) gives the susceptibility by the field 𝓐(𝒙;ω)\bm{\mathcal{A}}(\bm{x};\omega):

𝓙^(𝒙)(t)\displaystyle\left\langle\hat{\bm{\mathcal{J}}}(\bm{x})\right\rangle(t) =\displaystyle= 𝓙(𝒙;ω)eiωt\displaystyle{\bm{\mathcal{J}}(\bm{x};\omega)e^{-i\omega t}} (27)
=\displaystyle= 𝓙0(𝒙;ω)eiωt+𝑑𝒙𝒳¯(𝒙,𝒙;ω)𝓐(𝒙;ω)eiωt,\displaystyle{\bm{\mathcal{J}}_{0}(\bm{x};\omega)e^{-i\omega t}}+\int d\bm{x}^{\prime}\bar{\mathcal{X}}(\bm{x},\bm{x}^{\prime};\omega)\cdot\bm{\mathcal{A}}(\bm{x}^{\prime};\omega)e^{-i\omega t},

with a nonlocal susceptibility book:cho1

𝒳¯(𝒙,𝒙;ω)\displaystyle\bar{\mathcal{X}}(\bm{x},\bm{x}^{\prime};\omega) \displaystyle\equiv μ,ν[fνμ𝓙μν(𝒙)(𝓙νμ(𝒙))t+hνμ𝓙νμ(𝒙)(𝓙μν(𝒙))t]\displaystyle\sum_{\mu,\nu}\left[f_{\nu\mu}\bm{\mathcal{J}}_{\mu\nu}(\bm{x})\left(\bm{\mathcal{J}}_{\nu\mu}(\bm{x}^{\prime})\right)^{\rm t}+h_{\nu\mu}\bm{\mathcal{J}}_{\nu\mu}(\bm{x})\left(\bm{\mathcal{J}}_{\mu\nu}(\bm{x}^{\prime})\right)^{\rm t}\right] (28)
=\displaystyle= (χ¯jj(𝒙,𝒙;ω)𝝌jρ(𝒙,𝒙;ω)𝝌ρjt(𝒙,𝒙;ω)χρρ(𝒙,𝒙;ω)),\displaystyle\left(\begin{matrix}\bar{\chi}_{jj}(\bm{x},\bm{x}^{\prime};\omega)&\bm{\chi}_{j\rho}(\bm{x},\bm{x}^{\prime};\omega)\\ \bm{\chi}_{\rho j}^{\rm\ t}(\bm{x},\bm{x}^{\prime};\omega)&\chi_{\rho\rho}(\bm{x},\bm{x}^{\prime};\omega)\end{matrix}\right),

and

𝓙νμ(𝒙)\displaystyle\bm{\mathcal{J}}_{\nu\mu}(\bm{x}) =\displaystyle= ν|𝓙^(𝒙)|μ,\displaystyle\left\langle\nu\left|\hat{\bm{\mathcal{J}}}(\bm{x})\right|\mu\right\rangle, (29)
fνμ\displaystyle f_{\nu\mu} =\displaystyle= ρ0,μωνμωiγ,\displaystyle\frac{\rho_{0,\mu}}{\hbar\omega_{\nu\mu}-\hbar\omega-i\gamma}, (30)
hνμ\displaystyle h_{\nu\mu} =\displaystyle= ρ0,μωνμ+ω+iγ.\displaystyle\frac{\rho_{0,\mu}}{\hbar\omega_{\nu\mu}+\hbar\omega+i\gamma}. (31)

The nonlocal susceptibility is a 4×44\times 4 matrix in the four-vector space and |μ|\mu\rangle represents the electronic eigenstates of H^0\hat{H}_{0}. Thus, the susceptibility is affected by the nanostructure geometry via |μ|\mu\rangle. If the electronic system has translational symmetry, e.g., in the case of a bulk metal, the nonlocality in the susceptibility is given only by a relative position, 𝒳¯(𝒙𝒙;ω)\bar{\mathcal{X}}(\bm{x}-\bm{x}^{\prime};\omega). Further, ωνμ=ενεμ\hbar\omega_{\nu\mu}=\varepsilon_{\nu}-\varepsilon_{\mu} is the energy difference between the eigenenergies for H^0\hat{H}_{0}, γ\gamma is an infinitesimal positive value for the causality, and ρ0,μ\rho_{0,\mu} in the numerators is an element of the density matrix at equilibrium, where

ρ0,μ=μ|ρ^0|μ=1Z0μ|eβH^0|μ\rho_{0,\mu}=\langle\mu|\hat{\rho}_{0}|\mu\rangle=\frac{1}{Z_{0}}\langle\mu|e^{-\beta\hat{H}_{0}}|\mu\rangle (32)

with the partition function Z0=Tr{eβH^0}Z_{0}={\rm Tr}\{e^{-\beta\hat{H}_{0}}\}. The first term in the r.h.s. of Eq. (27) is

𝓙0(𝒙;ω)=(𝒋^(𝒙;ω)0c(ρ^(𝒙)0ρ0(𝒙)))=(emρ0(𝒙)𝑨(𝒙;ω)0)\bm{\mathcal{J}}_{0}(\bm{x};\omega)=\left(\begin{matrix}\langle\hat{\bm{j}}(\bm{x};\omega)\rangle_{0}\\ c\left(\langle\hat{\rho}(\bm{x})\rangle_{0}-\rho_{0}(\bm{x})\right)\end{matrix}\right)=\left(\begin{matrix}-\frac{e}{m^{*}}\rho_{0}(\bm{x})\bm{A}(\bm{x};\omega)\\ 0\end{matrix}\right) (33)

at equilibrium. The elements of susceptibility in Eq. (28) are related to each other and satisfy the continuous relation x𝒋^(𝒙)iωρ^(𝒙)=0\bm{\nabla}_{x}\cdot\left\langle\hat{\bm{j}}(\bm{x})\right\rangle-i\omega\left\langle\hat{\rho}(\bm{x})\right\rangle=0, where

xχ¯jj(𝒙,𝒙;ω)\displaystyle\bm{\nabla}_{x}\cdot\bar{\chi}_{jj}(\bm{x},\bm{x}^{\prime};\omega) =\displaystyle= i(ω/c)𝝌ρjt(𝒙,𝒙;ω),\displaystyle i(\omega/c)\bm{\chi}_{\rho j}^{\rm\ t}(\bm{x},\bm{x}^{\prime};\omega), (34)
x𝝌jρ(𝒙,𝒙;ω)\displaystyle\bm{\nabla}_{x}\cdot\bm{\chi}_{j\rho}(\bm{x},\bm{x}^{\prime};\omega) =\displaystyle= i(ω/c)χρρ(𝒙,𝒙;ω).\displaystyle i(\omega/c)\chi_{\rho\rho}(\bm{x},\bm{x}^{\prime};\omega). (35)

In the linear response, the 𝓐\bm{\mathcal{A}}-dependence of the susceptibility 𝒳¯(𝒙,𝒙;ω)\bar{\mathcal{X}}(\bm{x},\bm{x}^{\prime};\omega) cannot be discussed. To evaluate the effect of the 𝑨(𝒙,t)\bm{A}(\bm{x},t) term in Eq. (28), the higher-order terms must be considered.

IV Maxwell’s equations for Coulomb gauge

Taking the Coulomb gauge, the vector and scalar potentials describe only the T and L components, respectively. For the Maxwell’s equations, we consider the potentials 𝓐(𝒙,t)\bm{\mathcal{A}}(\bm{x},t) induced by the densities 𝓙(𝒙,t)\bm{\mathcal{J}}(\bm{x},t) via the Green’s function.

The Maxwell’s equations for the potentials, ϕ(𝒙,t)\phi(\bm{x},t) and 𝑨(𝒙,t)\bm{A}(\bm{x},t), are expressed as

2𝑨(𝒙,t)1c22𝑨(𝒙,t)t21c2tϕ(𝒙,t)\displaystyle\bm{\nabla}^{2}\bm{A}(\bm{x},t)-\frac{1}{c^{2}}\frac{\partial^{2}\bm{A}(\bm{x},t)}{\partial t^{2}}-\frac{1}{c^{2}}\frac{\partial}{\partial t}\bm{\nabla}\phi(\bm{x},t) =\displaystyle= μ0𝒋(𝒙,t),\displaystyle-\mu_{0}\bm{j}(\bm{x},t), (36)
2ϕ(𝒙,t)\displaystyle\bm{\nabla}^{2}\phi(\bm{x},t) =\displaystyle= ρ(𝒙,t)ε0.\displaystyle-\frac{\rho(\bm{x},t)}{\varepsilon_{0}}. (37)

Note that the L component of the field is relevant only to ρ(𝒙,t)\rho(\bm{x},t), whereas 𝒋(𝒙,t)\bm{j}(\bm{x},t) generates both of the L and T components. Another description may also be considered (see Appendix B). Equations (36) and (37) are unified in the four-vector representation

(22(ct)2(ct)02)𝓐(𝒙,t)=μ0𝓙(𝒙,t),\left(\begin{matrix}\bm{\nabla}^{2}-\frac{\partial^{2}}{\partial(ct)^{2}}&{\frac{\partial}{\partial(ct)}\bm{\nabla}}\\ 0&{-\bm{\nabla}^{2}}\end{matrix}\right)\bm{\mathcal{A}}(\bm{x},t)=-\mu_{0}\bm{\mathcal{J}}(\bm{x},t), (38)

with the statistical average of the densities, 𝓙(𝒙,t)=𝓙^(𝒙)(t)\bm{\mathcal{J}}(\bm{x},t)=\left\langle\hat{\bm{\mathcal{J}}}(\bm{x})\right\rangle(t). For the Fourier component, 𝓞(𝒙,t)=𝓞(𝒙;ω)eiωt\bm{\mathcal{O}}(\bm{x},t)=\bm{\mathcal{O}}(\bm{x};\omega)e^{-i\omega t} (𝓞=𝓐,𝓙\bm{\mathcal{O}}=\bm{\mathcal{A}},\bm{\mathcal{J}}), the Maxwell’s equation is written as

𝒟¯(𝒙;ω)𝓐(𝒙;ω)=μ0𝓙(𝒙;ω),\bar{\mathcal{D}}(\bm{x};\omega)\bm{\mathcal{A}}(\bm{x};\omega)=-\mu_{0}\bm{\mathcal{J}}(\bm{x};\omega), (39)

with the 4×44\times 4 differential matrix

𝒟¯(𝒙;ω)(2+ω2c2iωc02).\bar{\mathcal{D}}(\bm{x};\omega)\equiv\left(\begin{matrix}\bm{\nabla}^{2}+\frac{\omega^{2}}{c^{2}}&{-i\frac{\omega}{c}\bm{\nabla}}\\ 0&{-\bm{\nabla}^{2}}\end{matrix}\right). (40)

From Eqs. (27) and (39), we can construct a self-consistent relation between 𝓙\bm{\mathcal{J}} and 𝓐\bm{\mathcal{A}}. For the self-consistent equation, let us obtain the formal solution of the Maxwell’s equation (39) using the Green’s function

𝓐(𝒙;ω)=𝓐0(𝒙;ω)μ0𝑑𝒙𝒢¯(𝒙,𝒙;ω)𝓙(𝒙;ω).\bm{\mathcal{A}}(\bm{x};\omega)=\bm{\mathcal{A}}_{0}(\bm{x};\omega)-\mu_{0}\int d\bm{x}^{\prime}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}(\bm{x}^{\prime};\omega). (41)

The first term in Eq. (41) corresponds to an “incident” field satisfying

𝒟¯(𝒙;ω)𝓐0(𝒙;ω)=0.\bar{\mathcal{D}}(\bm{x};\omega)\bm{\mathcal{A}}_{0}(\bm{x};\omega)=0. (42)

The Green’s function is given as

𝒢¯(𝒙,𝒙;ω)=14π(eiωc|𝒙𝒙||𝒙𝒙|1¯iω4πc𝑑𝒙′′eiωc|𝒙𝒙′′||𝒙𝒙′′|𝒙′′𝒙|𝒙′′𝒙|301|𝒙𝒙|).\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)=-\frac{1}{4\pi}\left(\begin{matrix}\frac{\ e^{i\frac{\omega}{c}|\bm{x}-\bm{x}^{\prime}|}}{|\bm{x}-\bm{x}^{\prime}|}\bar{1}\ &\frac{-i\omega}{4\pi c}\int d\bm{x}^{\prime\prime}\frac{e^{i\frac{\omega}{c}|\bm{x}-\bm{x}^{\prime\prime}|}}{|\bm{x}-\bm{x}^{\prime\prime}|}\frac{\bm{x}^{\prime\prime}-\bm{x}^{\prime}}{|\bm{x}^{\prime\prime}-\bm{x}^{\prime}|^{3}}\\ 0&{-\frac{1}{|\bm{x}-\bm{x}^{\prime}|}}\end{matrix}\right). (43)

A detailed derivation of 𝒢¯\bar{\mathcal{G}} in Eq. (43) is presented in Appendix C. This Green’s function provides a scheme for field (potential) generation due to the electronic excitations (current and charge densities).

V Self-consistent equation

In previous sections, we derived the constitutive equation (27) with the nonlocal susceptibility (28) and the solution of the Maxwell’s equations (41) with Green’s function (43) in a four-vector representation. These expressions form a self-consistent relation.

V.1 𝒙\bm{x}-space representation

Let us first formulate the self-consistent equation in terms of 𝓐(𝒙;ω)\bm{\mathcal{A}}(\bm{x};\omega) and 𝓙(𝒙;ω)\bm{\mathcal{J}}(\bm{x};\omega). Substitution of the constitutive equation into the solution of Maxwell’s equation yields the self-consistent equation for the four-vector potential 𝓐(𝒙;ω)\bm{\mathcal{A}}(\bm{x};\omega),

𝓐(𝒙;ω)\displaystyle\bm{\mathcal{A}}(\bm{x};\omega) =\displaystyle= 𝓐0(𝒙;ω)μ0𝑑𝒙𝒢¯(𝒙,𝒙;ω)𝓙0(𝒙;ω)\displaystyle\bm{\mathcal{A}}_{0}(\bm{x};\omega)-\mu_{0}\int d\bm{x}^{\prime}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{0}(\bm{x}^{\prime};\omega) (44)
μ0𝑑𝒙𝑑𝒙′′𝒢¯(𝒙,𝒙;ω)𝒳¯(𝒙,𝒙′′;ω)𝓐(𝒙′′;ω).\displaystyle\hskip 56.9055pt-\mu_{0}\int d\bm{x}^{\prime}\int d\bm{x}^{\prime\prime}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bar{\mathcal{X}}(\bm{x}^{\prime},\bm{x}^{\prime\prime};\omega)\bm{\mathcal{A}}(\bm{x}^{\prime\prime};\omega).

By applying the separable integral kernel δ(𝒙𝒙)=mφm(𝒙)φm(𝒙)\delta(\bm{x}-\bm{x}^{\prime})=\sum_{m}\varphi_{m}^{\ast}(\bm{x})\varphi_{m}(\bm{x}^{\prime}) for the vector potential term (33) in the current, we expand the delta function as

(μ0)𝓙0(𝒙;ω)\displaystyle(-\mu_{0})\bm{\mathcal{J}}_{0}(\bm{x};\omega) =\displaystyle= μ0e2n0mV𝑑𝒙δ(𝒙𝒙)(1¯0)𝓐(𝒙;ω)\displaystyle\frac{\mu_{0}e^{2}n_{0}}{m^{*}}\int_{V}d\bm{x}^{\prime}\delta(\bm{x}-\bm{x}^{\prime})\left(\begin{matrix}\bar{1}&&\\ &&0\end{matrix}\right)\bm{\mathcal{A}}(\bm{x}^{\prime};\omega) (45)
=\displaystyle= (ωpc)2mα=x,y,zV𝑑𝒙(φm(𝒙)𝒆α)(φm(𝒙)𝒆α)t𝓐(𝒙;ω).\displaystyle\left(\frac{\omega_{\rm p}}{c}\right)^{2}\sum_{m}\sum_{\alpha=x,y,z}\int_{V}d\bm{x}^{\prime}\left(\varphi_{m}^{\ast}(\bm{x})\bm{e}_{\alpha}\right)\left(\varphi_{m}(\bm{x}^{\prime})\bm{e}_{\alpha}\right)^{\rm t}\bm{\mathcal{A}}(\bm{x}^{\prime};\omega).

Here, n0=ρ0/en_{0}=\rho_{0}/e is the electron density and ωp=e2n0/(ε0m)\omega_{\rm p}=\sqrt{e^{2}n_{0}/(\varepsilon_{0}m^{*})} is the plasma frequency in bulk. 𝒆α\bm{e}_{\alpha} means a unit vector in the α\alpha direction. The integral is applied only inside the nanostructure.

Since the nonlocal susceptibility 𝒳¯(𝒙,𝒙′′;ω)\bar{\mathcal{X}}(\bm{x}^{\prime},\bm{x}^{\prime\prime};\omega) shown in Eq. (28) is a separable integral kernel, by multiplying Eq. (44) by (𝓙νμ(𝒙))t\left(\bm{\mathcal{J}}_{\nu^{\prime}\mu^{\prime}}(\bm{x})\right)^{\rm t}, (𝓙μν(𝒙))t\left(\bm{\mathcal{J}}_{\mu^{\prime}\nu^{\prime}}(\bm{x})\right)^{\rm t}, and (φm(𝒙)𝒆βt)\left(\varphi_{m^{\prime}}(\bm{x})\bm{e}_{\beta}^{\rm t}\right) from the left and integrating with respect to 𝒙\bm{x}, we formulate the matrix form of the self-consistent equation as

[Ξ¯(ω)](𝑿()(ω)𝑿(+)(ω))=(𝒀(0,)(ω)𝒀(0,+)(ω))\left[\bar{\Xi}(\omega)\right]\left(\begin{matrix}\bm{X}^{(-)}(\omega)\\ \bm{X}^{(+)}(\omega)\end{matrix}\right)=\left(\begin{matrix}\bm{Y}^{\prime(0,-)}(\omega)\\ \bm{Y}^{\prime(0,+)}(\omega)\end{matrix}\right) (46)

with

Ξ¯(ω)=(Ω¯(ω+iγ)1¯Ω¯+(ω+iγ)1¯)+(K¯(ω)L¯(ω)M¯(ω)N¯(ω)).\bar{\Xi}(\omega)=\left(\begin{matrix}\hbar\bar{\Omega}-(\hbar\omega+i\gamma)\bar{1}&\\ &\hbar\bar{\Omega}+(\hbar\omega+i\gamma)\bar{1}\\ \end{matrix}\right)+\left(\begin{matrix}\bar{K}^{\prime}(\omega)&\bar{L}^{\prime}(\omega)\\ \bar{M}^{\prime}(\omega)&\bar{N}^{\prime}(\omega)\end{matrix}\right). (47)

The vectors

(𝒀(0,)(ω)𝒀(0,+)(ω))=(𝒀(0,)(ω)𝒀(0,+)(ω))+(U¯(ω)11¯R¯(ω)𝒀(A)(ω)V¯(ω)11¯R¯(ω)𝒀(A)(ω)),\left(\begin{matrix}\bm{Y}^{\prime(0,-)}(\omega)\\ \bm{Y}^{\prime(0,+)}(\omega)\end{matrix}\right)=\left(\begin{matrix}\bm{Y}^{(0,-)}(\omega)\\ \bm{Y}^{(0,+)}(\omega)\end{matrix}\right)+\left(\begin{matrix}\bar{U}(\omega)\frac{1}{\bar{1}-\bar{R}(\omega)}\bm{Y}^{(A)}(\omega)\\ \bar{V}(\omega)\frac{1}{\bar{1}-\bar{R}(\omega)}\bm{Y}^{(A)}(\omega)\end{matrix}\right), (48)

correspond to an incident field. The detail derivation and formulation are described in Appendix D. Ω¯=diag(Ω¯0,Ω¯1,,Ω¯N)\bar{\Omega}={\rm diag}(\bar{\Omega}_{0},\bar{\Omega}_{1},\cdots,\bar{\Omega}_{N}) is a diagonal matrix with Ω¯μ=diag(ω0μ,ω1μ,ω2μ,ωNμ)\bar{\Omega}_{\mu}={\rm diag}(\omega_{0\mu},\omega_{1\mu},\omega_{2\mu}\cdots,\omega_{N\mu}) for the excitation energy ωνμ=ενεμ\hbar\omega_{\nu\mu}=\varepsilon_{\nu}-\varepsilon_{\mu}. γ\gamma is an infinitesimal value for the causality. The matrices

K¯(ω)\displaystyle\bar{K}^{\prime}(\omega) =\displaystyle= K¯(ω)+U¯(ω)11¯R¯(ω)S¯(ω),\displaystyle\bar{K}(\omega)+\bar{U}(\omega)\frac{1}{\bar{1}-\bar{R}(\omega)}\bar{S}(\omega), (49)
L¯(ω)\displaystyle\bar{L}^{\prime}(\omega) =\displaystyle= L¯(ω)+U¯(ω)11¯R¯(ω)T¯(ω),\displaystyle\bar{L}(\omega)+\bar{U}(\omega)\frac{1}{\bar{1}-\bar{R}(\omega)}\bar{T}(\omega), (50)
M¯(ω)\displaystyle\bar{M}^{\prime}(\omega) =\displaystyle= M¯(ω)+V¯(ω)11¯R¯(ω)S¯(ω),\displaystyle\bar{M}(\omega)+\bar{V}(\omega)\frac{1}{\bar{1}-\bar{R}(\omega)}\bar{S}(\omega), (51)
N¯(ω)\displaystyle\bar{N}^{\prime}(\omega) =\displaystyle= N¯(ω)+V¯(ω)11¯R¯(ω)T¯(ω).\displaystyle\bar{N}(\omega)+\bar{V}(\omega)\frac{1}{\bar{1}-\bar{R}(\omega)}\bar{T}(\omega). (52)

describe the radiative correction or the correlation between the excited charge and current densities by the longitudinal and transverse fields. Their matrix and vector components are defined as

Kνμ,μν(ω)\displaystyle K_{\nu^{\prime}\mu^{\prime},\mu\nu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒙𝑑𝒙(𝓙νμ(𝒙))t𝒢¯(𝒙,𝒙;ω)𝓙μν(𝒙),\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{x}\int d\bm{x}^{\prime}\left(\bm{\mathcal{J}}_{\nu^{\prime}\mu^{\prime}}(\bm{x})\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{\mu\nu}(\bm{x}^{\prime}), (53)
Lνμ,νμ(ω)\displaystyle L_{\nu^{\prime}\mu^{\prime},\nu\mu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒙𝑑𝒙(𝓙νμ(𝒙))t𝒢¯(𝒙,𝒙;ω)𝓙νμ(𝒙),\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{x}\int d\bm{x}^{\prime}\left(\bm{\mathcal{J}}_{\nu^{\prime}\mu^{\prime}}(\bm{x})\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{\nu\mu}(\bm{x}^{\prime}), (54)
Mμν,μν(ω)\displaystyle M_{\mu^{\prime}\nu^{\prime},\mu\nu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒙𝑑𝒙(𝓙μν(𝒙))t𝒢¯(𝒙,𝒙;ω)𝓙μν(𝒙),\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{x}\int d\bm{x}^{\prime}\left(\bm{\mathcal{J}}_{\mu^{\prime}\nu^{\prime}}(\bm{x})\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{\mu\nu}(\bm{x}^{\prime}), (55)
Nμν,νμ(ω)\displaystyle N_{\mu^{\prime}\nu^{\prime},\nu\mu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒙𝑑𝒙(𝓙μν(𝒙))t𝒢¯(𝒙,𝒙;ω)𝓙νμ(𝒙),\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{x}\int d\bm{x}^{\prime}\left(\bm{\mathcal{J}}_{\mu^{\prime}\nu^{\prime}}(\bm{x})\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{\nu\mu}(\bm{x}^{\prime}), (56)
Rmβ,mα(ω)\displaystyle R_{m^{\prime}\beta,m\alpha}(\omega) =\displaystyle= (ωpc)2𝑑𝒙𝑑𝒙(φm(𝒙)𝒆β)t𝒢¯(𝒙,𝒙;ω)(φm(𝒙)𝒆α),\displaystyle\left(\frac{\omega_{\rm p}}{c}\right)^{2}\int d\bm{x}\int d\bm{x}^{\prime}\left(\varphi_{m^{\prime}}(\bm{x})\bm{e}_{\beta}\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\left(\varphi_{m}^{\ast}(\bm{x}^{\prime})\bm{e}_{\alpha}\right), (57)
Uνμ,mα(ω)\displaystyle U_{\nu^{\prime}\mu^{\prime},m\alpha}(\omega) =\displaystyle= (ωpc)2𝑑𝒙𝑑𝒙(𝓙νμ(𝒙))t𝒢¯(𝒙,𝒙;ω)(φm(𝒙)𝒆α),\displaystyle\left(\frac{\omega_{\rm p}}{c}\right)^{2}\int d\bm{x}\int d\bm{x}^{\prime}\left(\bm{\mathcal{J}}_{\nu^{\prime}\mu^{\prime}}(\bm{x})\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\left(\varphi_{m}^{\ast}(\bm{x}^{\prime})\bm{e}_{\alpha}\right), (58)
Vμν,mα(ω)\displaystyle V_{\mu^{\prime}\nu^{\prime},m\alpha}(\omega) =\displaystyle= (ωpc)2𝑑𝒙𝑑𝒙(𝓙μν(𝒙))t𝒢¯(𝒙,𝒙;ω)(φm(𝒙)𝒆α),\displaystyle\left(\frac{\omega_{\rm p}}{c}\right)^{2}\int d\bm{x}\int d\bm{x}^{\prime}\left(\bm{\mathcal{J}}_{\mu^{\prime}\nu^{\prime}}(\bm{x})\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\left(\varphi_{m}^{\ast}(\bm{x}^{\prime})\bm{e}_{\alpha}\right), (59)
Smβ,μν(ω)\displaystyle S_{m^{\prime}\beta,\mu\nu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒙𝑑𝒙(φm(𝒙)𝒆β)t𝒢¯(𝒙,𝒙;ω)𝓙μν(𝒙),\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{x}\int d\bm{x}^{\prime}\left(\varphi_{m^{\prime}}(\bm{x})\bm{e}_{\beta}\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{\mu\nu}(\bm{x}^{\prime}), (60)
Tmβ,νμ(ω)\displaystyle T_{m^{\prime}\beta,\nu\mu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒙𝑑𝒙(φm(𝒙)𝒆β)t𝒢¯(𝒙,𝒙;ω)𝓙νμ(𝒙)\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{x}\int d\bm{x}^{\prime}\left(\varphi_{m^{\prime}}(\bm{x})\bm{e}_{\beta}\right)^{\rm t}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{\nu\mu}(\bm{x}^{\prime}) (61)

and

Xνμ()(ω)\displaystyle X_{\nu\mu}^{(-)}(\omega) =\displaystyle= 1ωνμωiγ𝑑𝒙(𝓙νμ(𝒙))t𝓐(𝒙;ω),\displaystyle\frac{1}{\hbar\omega_{\nu\mu}-\hbar\omega-i\gamma}\int d\bm{x}\left(\bm{\mathcal{J}}_{\nu\mu}(\bm{x})\right)^{\rm t}\bm{\mathcal{A}}(\bm{x};\omega), (62)
Xμν(+)(ω)\displaystyle X_{\mu\nu}^{(+)}(\omega) =\displaystyle= 1ωνμ+ω+iγ𝑑𝒙(𝓙μν(𝒙))t𝓐(𝒙;ω),\displaystyle\frac{1}{\hbar\omega_{\nu\mu}+\hbar\omega+i\gamma}\int d\bm{x}\left(\bm{\mathcal{J}}_{\mu\nu}(\bm{x})\right)^{\rm t}\bm{\mathcal{A}}(\bm{x};\omega), (63)
Xmα(A)(ω)\displaystyle X_{m\alpha}^{(A)}(\omega) =\displaystyle= 𝑑𝒙(φm(𝒙)𝒆α)t𝓐(𝒙;ω),\displaystyle\int d\bm{x}\left(\varphi_{m}(\bm{x})\bm{e}_{\alpha}\right)^{\rm t}\bm{\mathcal{A}}(\bm{x};\omega), (64)
Yνμ(0)(ω)\displaystyle Y^{\rm(0)}_{\nu^{\prime}\mu^{\prime}}(\omega) =\displaystyle= 𝑑𝒙(𝓙νμ(𝒙))t𝓐0(𝒙;ω),\displaystyle\int d\bm{x}\left(\bm{\mathcal{J}}_{\nu^{\prime}\mu^{\prime}}(\bm{x})\right)^{\rm t}\bm{\mathcal{A}}_{0}(\bm{x};\omega), (65)
Ymα(A)(ω)\displaystyle Y_{m\alpha}^{(A)}(\omega) =\displaystyle= 𝑑𝒙(φm(𝒙)𝒆α)t𝓐0(𝒙;ω).\displaystyle\int d\bm{x}\left(\varphi_{m}(\bm{x})\bm{e}_{\alpha}\right)^{\rm t}\bm{\mathcal{A}}_{0}(\bm{x};\omega). (66)

The matrix Ξ¯(ω)\bar{\Xi}(\omega) describes coupling between the electron–hole excitations mediated by both longitudinal and transverse electromagnetic fields, which forms the collective excitation. The individual electron–hole excitations are followed by Ω¯\bar{\Omega}, whereas the field-mediated coupling, namely the collective excitation, is in K¯(ω)\bar{K}^{\prime}(\omega), L¯(ω)\bar{L}^{\prime}(\omega), M¯(ω)\bar{M}^{\prime}(\omega), and N¯(ω)\bar{N}^{\prime}(\omega). Due to coupling, the eigenvalues of Ξ¯(ω)\bar{\Xi}(\omega) become complex and provide zero points in the complex ω=ωr+iωi\omega=\omega_{\rm r}+i\omega_{\rm i} plane. Its real components provide the excitation spectrum, and the imaginary components means the radiative width. The vectors 𝑿()(ω)\bm{X}^{(\mp)}(\omega) and 𝒀(0,)(ω){\bm{Y}^{\prime}}^{(0,\mp)}(\omega) in Eq. (46) mean the coefficients of induced four-vector current and incident four-vector field, respectively. Therefore, the matrix Ξ¯(ω)\bar{\Xi}(\omega) exhibits the electronic properties of nanostructures, whereas the vector 𝑿()(ω)\bm{X}^{(\mp)}(\omega) provides the output charge and current densities.

At zero temperature, i.e., T=0T=0, ρ0,μ=δ0,μ\rho_{0,\mu}=\delta_{0,\mu} in fνμf_{\nu\mu} and hνμh_{\nu\mu} defined in Eqs. (30) and (31), respectively. Then, the energy differences ων0\hbar\omega_{\nu 0} in the denominators of Xν0()X_{\nu 0}^{(-)} and X0ν(+)X_{0\nu}^{(+)} are positive. For the matrix form of the self-consistent equation (46), only 𝓙ν0\bm{\mathcal{J}}_{\nu 0} and 𝓙0ν\bm{\mathcal{J}}_{0\nu} should be considered. Therefore, the matrix elements are reduced for Kν0,0νK_{\nu^{\prime}0,0\nu}, Lν0,ν0L_{\nu^{\prime}0,\nu 0}, M0ν,0νM_{0\nu^{\prime},0\nu}, and N0ν,ν0N_{0\nu^{\prime},\nu 0}.

The matrix form of the self-consistent equation (46) has an important advantage. The matrix Ξ¯(ω)\bar{\Xi}(\omega) with the matrices K¯\bar{K}, L¯\bar{L}, M¯\bar{M}, and N¯\bar{N} consisting of 𝒢¯\bar{\mathcal{G}} and ν|𝓙^|μ\left\langle\nu\left|\hat{\bm{\mathcal{J}}}\right|\mu\right\rangle gives the eigenmodes of the system coupled with the radiative field. The L and T fields are in 𝒢¯\bar{\mathcal{G}} and the properties of the nonlocal response are given by |μ|\mu\rangle. From the real and imaginary components of the eigenvalues of Ξ¯(ω)\bar{\Xi}(\omega), the properties of the collective and individual excitations are evaluated. Then, the current and charge densities are discussed in terms of the plasmons and SPEs.

V.2 𝒌\bm{k}-space representation

For a metallic structure having some symmetry, the Fourier transformation of the fields and densities is useful for the evaluation. In that case, we consider the 𝒌\bm{k}-space representation for the self-consistent equation. The fields and densities in the 𝒌\bm{k}-representation are given as

𝓞~(𝒌;ω)=1(2π)32𝑑𝒙𝓞(𝒙;ω)ei𝒌𝒙,(𝓞=𝓐,𝓙).\tilde{\bm{\mathcal{O}}}(\bm{k};\omega)=\frac{1}{(2\pi)^{\frac{3}{2}}}\int d\bm{x}\bm{\mathcal{O}}(\bm{x};\omega)e^{-i\bm{k}\cdot\bm{x}},\hskip 14.22636pt(\bm{\mathcal{O}}=\bm{\mathcal{A}},\bm{\mathcal{J}}). (67)

The Maxwell’s equations (39) and their solutions (41) are also transferred to the 𝒌\bm{k}-representation, with the latter being expressed as

𝓐~(𝒌;ω)=𝓐~0(𝒌;ω)μ0𝒢¯𝒌(ω)𝓙~(𝒌;ω).\tilde{\bm{\mathcal{A}}}(\bm{k};\omega)=\tilde{\bm{\mathcal{A}}}_{0}(\bm{k};\omega)-\mu_{0}\bar{\mathcal{G}}_{\bm{k}}(\omega)\tilde{\bm{\mathcal{J}}}(\bm{k};\omega). (68)

Here, the Green’s function is given as

𝒢¯𝒌(ω)=(1𝒌2+ω2/c21¯1𝒌21𝒌2+ω2/c2ωc𝒌01𝒌2).\bar{\mathcal{G}}_{\bm{k}}(\omega)=\left(\begin{matrix}\frac{1}{-\bm{k}^{2}+\omega^{2}/c^{2}}\bar{1}&-\frac{1}{\bm{k}^{2}}\frac{1}{-\bm{k}^{2}+\omega^{2}/c^{2}}\frac{\omega}{c}\bm{k}\\ 0&{\frac{1}{\bm{k}^{2}}}\end{matrix}\right). (69)

The incident term satisfies

((𝒌2+ω2c2)1¯ωc𝒌0𝒌2)𝓐~0(𝒌;ω)=0.\left(\begin{matrix}\left(-\bm{k}^{2}+\frac{\omega^{2}}{c^{2}}\right)\bar{1}&{\frac{\omega}{c}\bm{k}}\\ 0&{\bm{k}^{2}}\end{matrix}\right)\tilde{\bm{\mathcal{A}}}_{0}(\bm{k};\omega)=0. (70)

Note that the solution of the Maxwell’s equations contains no integral.

The constitutive equation becomes

𝓙~(𝒌;ω)=𝓙~0(𝒌;ω)+𝑑𝒌𝒳¯𝒌,𝒌(ω)𝓐~(𝒌;ω),\tilde{\bm{\mathcal{J}}}(\bm{k};\omega)=\tilde{\bm{\mathcal{J}}}_{0}(\bm{k};\omega)+\int d\bm{k}^{\prime}\bar{\mathcal{X}}_{\bm{k},\bm{k}^{\prime}}(\omega)\tilde{\bm{\mathcal{A}}}(\bm{k}^{\prime};\omega), (71)

with the nonlocal susceptibility

𝒳¯𝒌,𝒌(ω)=μ,ν[fνμ𝓙~μν(𝒌)(𝓙~νμ(𝒌))t+hνμ𝓙~νμ(𝒌)(𝓙~μν(𝒌))t].\bar{\mathcal{X}}_{\bm{k},\bm{k}^{\prime}}(\omega)=\sum_{\mu,\nu}\left[f_{\nu\mu}\tilde{\bm{\mathcal{J}}}_{\mu\nu}(\bm{k})\left(\tilde{\bm{\mathcal{J}}}_{\nu\mu}(-\bm{k}^{\prime})\right)^{\rm t}+h_{\nu\mu}\tilde{\bm{\mathcal{J}}}_{\nu\mu}(\bm{k})\left(\tilde{\bm{\mathcal{J}}}_{\mu\nu}(-\bm{k}^{\prime})\right)^{\rm t}\right]. (72)

Equations (68) and (71) form the self-consistent equation:

𝓐~(𝒌;ω)=𝓐~0(𝒌;ω)μ0𝒢¯𝒌(ω)𝓙~0(𝒌;ω)μ0𝑑𝒌𝒢¯𝒌(ω)𝒳¯𝒌,𝒌(ω)𝓐~(𝒌;ω).\tilde{\bm{\mathcal{A}}}(\bm{k};\omega)=\tilde{\bm{\mathcal{A}}}_{0}(\bm{k};\omega)-\mu_{0}\bar{\mathcal{G}}_{\bm{k}}(\omega)\tilde{\bm{\mathcal{J}}}_{0}(\bm{k};\omega)-\mu_{0}\int d\bm{k}^{\prime}\bar{\mathcal{G}}_{\bm{k}}(\omega)\bar{\mathcal{X}}_{\bm{k},\bm{k}^{\prime}}(\omega)\tilde{\bm{\mathcal{A}}}(\bm{k}^{\prime};\omega). (73)

The matrix form of the self-consistent equation in the 𝒌\bm{k}-representation is obtained in the same manner as that of the previous subsection. It is exactly identical with Eq. (46), see Appendix D. The matrix elements in the 𝒌\bm{k}-representation are , e.g., as

K~νμ,μν(ω)\displaystyle\tilde{K}_{\nu^{\prime}\mu^{\prime},\mu\nu}(\omega) =\displaystyle= μ0ρ0,μ𝑑𝒌(𝓙~νμ(𝒌))t𝒢¯𝒌(ω)𝓙~μν(𝒌),\displaystyle\mu_{0}\rho_{0,\mu}\int d\bm{k}\left(\tilde{\bm{\mathcal{J}}}_{\nu^{\prime}\mu^{\prime}}(-\bm{k})\right)^{\rm t}\bar{\mathcal{G}}_{\bm{k}}(\omega)\tilde{\bm{\mathcal{J}}}_{\mu\nu}(\bm{k}), (74)
X~νμ()(ω)\displaystyle\tilde{X}_{\nu\mu}^{(-)}(\omega) =\displaystyle= 1ωνμωiγ𝑑𝒌(𝓙~νμ(𝒌))t𝓐~(𝒌;ω).\displaystyle\frac{1}{\hbar\omega_{\nu\mu}-\hbar\omega-i\gamma}\int d\bm{k}\left(\tilde{\bm{\mathcal{J}}}_{\nu\mu}(-\bm{k})\right)^{\rm t}\tilde{\bm{\mathcal{A}}}(\bm{k};\omega). (75)

The Fourier transformation demonstrates that these elements are equivalent to those of the 𝒙\bm{x}-representation, K~νμ,μν(ω)=Kνμ,μν(ω)\tilde{K}_{\nu^{\prime}\mu^{\prime},\mu\nu}(\omega)=K_{\nu^{\prime}\mu^{\prime},\mu\nu}(\omega), X~νμ()(ω)=Xνμ()(ω)\tilde{X}_{\nu\mu}^{(-)}(\omega)=X_{\nu\mu}^{(-)}(\omega), etc. Therefore, the matrix element evaluation can be performed in either the 𝒙\bm{x}- or 𝒌\bm{k}-representations depending on the case.

It is worth noting that, in (𝓙~νμ(𝒌))t𝒢¯𝒌(ω)𝓙~μν(𝒌)\left(\tilde{\bm{\mathcal{J}}}_{\nu^{\prime}\mu^{\prime}}(-\bm{k})\right)^{\rm t}\bar{\mathcal{G}}_{\bm{k}}(\omega)\tilde{\bm{\mathcal{J}}}_{\mu\nu}(\bm{k}), a component between the induced charge densities,

K0,νμ,μν=μ0ρ0,μ𝑑𝒌cρ~νμ(𝒌)1𝒌2cρ~μν(𝒌),K_{0,\nu^{\prime}\mu^{\prime},\mu\nu}=\mu_{0}\rho_{0,\mu}\int d\bm{k}c\tilde{\rho}_{\nu^{\prime}\mu^{\prime}}(-\bm{k})\frac{1}{\bm{k}^{2}}c\tilde{\rho}_{\mu\nu}(\bm{k}), (76)

describes the Coulomb interaction, which arises when an external field is applied. If the contribution of the current densities is negligible, the self-consistent equation (46) is reduced as

[Ξ¯0(ω)](𝑿()(ω)𝑿(+)(ω))=(𝒀(0,)(ω)𝒀(0,+)(ω))\left[\bar{\Xi}_{0}(\omega)\right]\left(\begin{matrix}\bm{X}^{(-)}(\omega)\\ \bm{X}^{(+)}(\omega)\end{matrix}\right)=\left(\begin{matrix}\bm{Y}^{(0,-)}(\omega)\\ \bm{Y}^{(0,+)}(\omega)\end{matrix}\right) (77)

with

Ξ¯0(ω)=(Ω¯(ω+iγ)1¯Ω¯+(ω+iγ)1¯)+(K¯0K¯0K¯0K¯0).\bar{\Xi}_{0}(\omega)=\left(\begin{matrix}\hbar\bar{\Omega}-(\hbar\omega+i\gamma)\bar{1}&\\ &\hbar\bar{\Omega}+(\hbar\omega+i\gamma)\bar{1}\\ \end{matrix}\right)+\left(\begin{matrix}\bar{K}_{0}&\bar{K}_{0}\\ \bar{K}_{0}&\bar{K}_{0}\end{matrix}\right). (78)

The eigenvalues of the matrix Ξ¯0(ω)\bar{\Xi}_{0}(\omega) generate the spectrum of the individual (electron-hole) excitation and the collective (plasmon) excitation. It follows

det{2Ω¯2(ω+iγ)2+K¯0Ω¯+Ω¯K¯0}=0.{\rm det}\left\{\hbar^{2}\bar{\Omega}^{2}-(\hbar\omega+i\gamma)^{2}+\hbar\bar{K}_{0}\bar{\Omega}+\hbar\bar{\Omega}\bar{K}_{0}\right\}=0. (79)

A rough estimate of the spectrum is

1K0ωνμ(ω+iγ)2(ωνμ)2.1\sim K_{0}\frac{\hbar\omega_{\nu\mu}}{(\hbar\omega+i\gamma)^{2}-(\hbar\omega_{\nu\mu})^{2}}. (80)

Hence, we find the spectrum at ωωνμ\omega\sim\omega_{\nu\mu} for the electron-hole excitation and ωωνμ\omega\gg\omega_{\nu\mu} for the plasmon excitation In a free electron model for the basis |μ|\mu\rangle of H^0\hat{H}_{0}, the plasmon spectrum should correspond to that in the RPA.

In addition to the ρ~νμ[𝒢¯𝒌]ρρρ~μν\tilde{\rho}_{\nu^{\prime}\mu^{\prime}}\left[\bar{\mathcal{G}}_{\bm{k}}\right]_{\rho\rho}\tilde{\rho}_{\mu\nu} term, K~νμ,μν\tilde{K}_{\nu^{\prime}\mu^{\prime},\mu\nu} and the other matrix components have 𝒋~νμ[𝒢¯𝒌]jρρ~μν\tilde{\bm{j}}_{\nu^{\prime}\mu^{\prime}}\left[\bar{\mathcal{G}}_{\bm{k}}\right]_{j\rho}\tilde{\rho}_{\mu\nu} and 𝒋~νμ[𝒢¯𝒌]jj𝒋~μν\tilde{\bm{j}}_{\nu^{\prime}\mu^{\prime}}\left[\bar{\mathcal{G}}_{\bm{k}}\right]_{jj}\tilde{\bm{j}}_{\mu\nu} terms, which have been neglected in many previous studies. However, the modifications of the eigenstates and spectrum of the electronic system due to the coupled radiative field are given by these components, as noted in the previous subsection. The nanostructure shape determines the spectrum shift via the |μ|\mu\rangle in ρ~μν\tilde{\rho}_{\mu\nu} and 𝒋~μν\tilde{\bm{j}}_{\mu\nu}. Therefore, analysis and exploration of an enhancement condition based on K~\tilde{K}, L~\tilde{L}, M~\tilde{M}, and N~\tilde{N}, with ρ~νμ[𝒢¯𝒌]ρρρ~μν\tilde{\rho}_{\nu^{\prime}\mu^{\prime}}\left[\bar{\mathcal{G}}_{\bm{k}}\right]_{\rho\rho}\tilde{\rho}_{\mu\nu}, 𝒋~νμ[𝒢¯𝒌]jρρ~μν\tilde{\bm{j}}_{\nu^{\prime}\mu^{\prime}}\left[\bar{\mathcal{G}}_{\bm{k}}\right]_{j\rho}\tilde{\rho}_{\mu\nu}, and 𝒋~νμ[𝒢¯𝒌]jj𝒋~μν\tilde{\bm{j}}_{\nu^{\prime}\mu^{\prime}}\left[\bar{\mathcal{G}}_{\bm{k}}\right]_{jj}\tilde{\bm{j}}_{\mu\nu} are important.

VI Application example: Rectangular Nanorod

To verify the formulation of self-consistent matrix equation (46) with Eq. (47) and the feasibility of calculation, we apply the formulation to a single rectangular nanorod.

VI.1 Radiative correction matrix for rectangular nanorod

Electron and hole wavefunctions in a rectangular nanorod are given as

ψ(eμ)=(nx,ny,nz)(𝒙)\displaystyle\psi_{({\rm e}\mu)=(n_{x},n_{y},n_{z})}(\bm{x}) =\displaystyle= 2Lxsin(nxπLxx)2Lysin(nyπLyy)2Lzsin(nzπLzz),\displaystyle\sqrt{\frac{2}{L_{x}}}\sin\left(\frac{n_{x}\pi}{L_{x}}x\right)\sqrt{\frac{2}{L_{y}}}\sin\left(\frac{n_{y}\pi}{L_{y}}y\right)\sqrt{\frac{2}{L_{z}}}\sin\left(\frac{n_{z}\pi}{L_{z}}z\right), (81)
ψ(hμ¯)=(n¯x,n¯y,n¯z)(𝒙)\displaystyle\psi_{({\rm h}\bar{\mu})=(\bar{n}_{x},\bar{n}_{y},\bar{n}_{z})}(\bm{x}) =\displaystyle= 2Lxsin(n¯xπLxx)2Lysin(n¯yπLyy)2Lzsin(n¯zπLzz)\displaystyle\sqrt{\frac{2}{L_{x}}}\sin\left(\frac{\bar{n}_{x}\pi}{L_{x}}x\right)\sqrt{\frac{2}{L_{y}}}\sin\left(\frac{\bar{n}_{y}\pi}{L_{y}}y\right)\sqrt{\frac{2}{L_{z}}}\sin\left(\frac{\bar{n}_{z}\pi}{L_{z}}z\right) (82)

with LxL_{x}, LyL_{y}, and LzL_{z} being the length of nanorod in the xx, yy, and zz directions, respectively. This section aims to verify calculation feasibility. We suppose the basis |μ=(eμ,hμ¯)|\mu=(\rm{e}\mu,\rm{h}\bar{\mu})\rangle for the Hamiltonian H^0\hat{H}_{0} are given by Eqs. (81) and (82). Note that the electron and hole energies are εeμ>εF\varepsilon_{{\rm e}\mu}>\varepsilon_{\rm F} and εhμ¯εF\varepsilon_{{\rm h}\bar{\mu}}\leq\varepsilon_{\rm F}, respectively. The excited charge and current densities at zero temperature are evaluated by the wavefunctions as

0|ρ^(𝒙)|μ\displaystyle\langle 0|\hat{\rho}(\bm{x})|\mu\rangle =\displaystyle= ρ0μ(𝒙)=eψ(hμ¯)(𝒙)ψ(eμ)(𝒙),\displaystyle\rho_{0\mu}(\bm{x})=e\psi_{(\rm{h}\bar{\mu})}(\bm{x})\psi_{({\rm e}\mu)}(\bm{x}), (83)
0|𝒋^(𝒙)|μ\displaystyle\langle 0|\hat{\bm{j}}(\bm{x})|\mu\rangle =\displaystyle= 𝒋0μ(𝒙)=e2im[(ψ(hμ¯)(𝒙))ψ(eμ)(𝒙)ψ(hμ¯)(𝒙)(ψ(eμ)(𝒙))].\displaystyle\bm{j}_{0\mu}(\bm{x})=-\frac{e\hbar}{2im^{*}}\left[\left(\bm{\nabla}\psi_{(\rm{h}\bar{\mu})}(\bm{x})\right)\psi_{({\rm e}\mu)}(\bm{x})-\psi_{(\rm{h}\bar{\mu})}(\bm{x})\left(\bm{\nabla}\psi_{({\rm e}\mu)}(\bm{x})\right)\right]. (84)

Since the wavefunctions are real, we find ρμ0(𝒙)=ρ0μ(𝒙)\rho_{\mu 0}(\bm{x})=\rho_{0\mu}(\bm{x}) and 𝒋μ0(𝒙)=𝒋μ0(𝒙)\bm{j}_{\mu 0}(\bm{x})=-\bm{j}_{\mu 0}(\bm{x}).

By applying the Fourier transformation of the densities, the matrix elements of K¯\bar{K}, L¯\bar{L}, M¯\bar{M}, and N¯\bar{N} are evaluated as

K~μ0,0μ\displaystyle\tilde{K}_{\mu^{\prime}0,0\mu} =\displaystyle= μ0d𝒌[(𝒋~μ0(𝒌))t(1𝒌2+(nbgω/c)2)𝒋~0μ(𝒌)\displaystyle\mu_{0}\int d\bm{k}\left[\left(\tilde{\bm{j}}_{\mu^{\prime}0}(-\bm{k})\right)^{\rm t}\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{\bm{j}}_{0\mu}(\bm{k})\right. (85)
+(𝒋~μ0(𝒌))t(1𝒌2nbgω/c𝒌2+(nbgω/c)2𝒌)(c/nbg)ρ~0μ(𝒌)\displaystyle\hskip 56.9055pt+\left(\tilde{\bm{j}}_{\mu^{\prime}0}(-\bm{k})\right)^{\rm t}\left(-\frac{1}{\bm{k}^{2}}\frac{n_{\rm bg}\omega/c}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\bm{k}\right)(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k})
+(c/nbg)ρ~μ0(𝒌)(1𝒌2)(c/nbg)ρ~0μ(𝒌)]\displaystyle\hskip 56.9055pt\left.+(c/n_{\rm bg})\tilde{\rho}_{\mu^{\prime}0}(-\bm{k})\left(\frac{1}{\bm{k}^{2}}\right)(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k})\right]
=\displaystyle= Aμ,μ(2)+Aμ,μ(1)+Aμ,μ(0),\displaystyle A_{\mu^{\prime},\mu}^{(2)}+A_{\mu^{\prime},\mu}^{(1)}+A_{\mu^{\prime},\mu}^{(0)},
L~μ0,μ0\displaystyle\tilde{L}_{\mu^{\prime}0,\mu 0} =\displaystyle= Aμ,μ(2)+Aμ,μ(1)+Aμ,μ(0),\displaystyle-A_{\mu^{\prime},\mu}^{(2)}+A_{\mu^{\prime},\mu}^{(1)}+A_{\mu^{\prime},\mu}^{(0)}, (86)
M~0μ,0μ\displaystyle\tilde{M}_{0\mu^{\prime},0\mu} =\displaystyle= Aμ,μ(2)Aμ,μ(1)+Aμ,μ(0),\displaystyle-A_{\mu^{\prime},\mu}^{(2)}-A_{\mu^{\prime},\mu}^{(1)}+A_{\mu^{\prime},\mu}^{(0)}, (87)
N~0μ,μ0\displaystyle\tilde{N}_{0\mu^{\prime},\mu 0} =\displaystyle= Aμ,μ(2)Aμ,μ(1)+Aμ,μ(0).\displaystyle A_{\mu^{\prime},\mu}^{(2)}-A_{\mu^{\prime},\mu}^{(1)}+A_{\mu^{\prime},\mu}^{(0)}. (88)

Here, nbgω/cn_{\rm bg}\omega/c means the wavenumber of light in the nanostructure and environment with the refractive index nbgn_{\rm bg}, for which the background dielectric constant εbg\varepsilon_{\rm bg} is introduced phenomenologically, nbg=εbg/ε0n_{\rm bg}=\sqrt{\varepsilon_{\rm bg}/\varepsilon_{0}}. The light velocity cc in the Maxwell’s equations (36) and (37) [or Eqs. (38) with the four-vector definition in Eqs. (24) and (25)] is replaced by c/nbgc/n_{\rm bg}. The back ground refractive index nbgn_{\rm bg} modulates the wavenumber of light, which enlarges the current–current interaction Aμ,μ(2)A_{\mu^{\prime},\mu}^{(2)} and the current–charge interaction Aμ,μ(1)A_{\mu^{\prime},\mu}^{(1)} and reduces effectively the charge–charge interaction Aμ,μ(0)A_{\mu^{\prime},\mu}^{(0)} by the screening effect.

For the separable integral kernel in the elements of matrices S¯\bar{S}, T¯\bar{T}, U¯\bar{U}, V¯\bar{V}, and R¯\bar{R}, we use

φm(𝒙)=2Lxsin(mxπLxx)2Lysin(myπLyy)2Lzsin(mzπLzz)\varphi_{m}(\bm{x})=\sqrt{\frac{2}{L_{x}}}\sin\left(\frac{m_{x}\pi}{L_{x}}x\right)\sqrt{\frac{2}{L_{y}}}\sin\left(\frac{m_{y}\pi}{L_{y}}y\right)\sqrt{\frac{2}{L_{z}}}\sin\left(\frac{m_{z}\pi}{L_{z}}z\right) (89)

with m=(mx,my,mz)m=(m_{x},m_{y},m_{z}). Its Fourier transferred function satisfies mφ~m(𝒌)φ~m(𝒌)=δ(𝒌𝒌)\sum_{m}\tilde{\varphi}_{m}^{*}(\bm{k})\tilde{\varphi}_{m}(-\bm{k}^{\prime})=\delta(\bm{k}-\bm{k}^{\prime}). Then, the matrix elements are

S~mβ,0μ\displaystyle\tilde{S}_{m^{\prime}\beta,0\mu} =\displaystyle= μ0𝑑𝒌φ~m(𝒌)(1𝒌2+(nbgω/c)2)j~0μ(β)(𝒌)\displaystyle\mu_{0}\int d\bm{k}\tilde{\varphi}_{m^{\prime}}(-\bm{k})\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{j}^{(\beta)}_{0\mu}(\bm{k}) (90)
+μ0𝑑𝒌φ~m(𝒌)(1𝒌2nbgω/c𝒌2+(nbgω/c)2kβ)(c/nbg)ρ~0μ(𝒌)\displaystyle\hskip 42.67912pt+\mu_{0}\int d\bm{k}\tilde{\varphi}_{m^{\prime}}(-\bm{k})\left(-\frac{1}{\bm{k}^{2}}\frac{n_{\rm bg}\omega/c}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}k_{\beta}\right)(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k})
=\displaystyle= Bmβ,μ(2)+Bmβ,μ(1),\displaystyle B_{m^{\prime}\beta,\mu}^{(2)}+B_{m^{\prime}\beta,\mu}^{(1)},
T~mβ,μ0\displaystyle\tilde{T}_{m^{\prime}\beta,\mu 0} =\displaystyle= Bmβ,μ(2)+Bmβ,μ(1),\displaystyle-B_{m^{\prime}\beta,\mu}^{(2)}+B_{m^{\prime}\beta,\mu}^{(1)}, (91)
U~μ0,mα\displaystyle\tilde{U}_{\mu^{\prime}0,m\alpha} =\displaystyle= (ωpc/nbg)2𝑑𝒌j~μ0(α)(𝒌)(1𝒌2+(nbgω/c)2)φ~m(𝒌)\displaystyle\left(\frac{\omega_{\rm p}^{\prime}}{c/n_{\rm bg}}\right)^{2}\int d\bm{k}\tilde{j}^{(\alpha)}_{\mu^{\prime}0}(-\bm{k})\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{\varphi}_{m}^{\ast}(\bm{k}) (92)
=\displaystyle= 1μ0(ωpc/nbg)2Bmα,μ(2),\displaystyle-\frac{1}{\mu_{0}}\left(\frac{\omega_{\rm p}^{\prime}}{c/n_{\rm bg}}\right)^{2}B_{m\alpha,\mu^{\prime}}^{(2)},
V~0μ,mα\displaystyle\tilde{V}_{0\mu^{\prime},m\alpha} =\displaystyle= 1μ0(ωpc/nbg)2Bmα,μ(2),\displaystyle\frac{1}{\mu_{0}}\left(\frac{\omega_{\rm p}^{\prime}}{c/n_{\rm bg}}\right)^{2}B_{m\alpha,\mu^{\prime}}^{(2)}, (93)
Rmβ,mα\displaystyle R_{m^{\prime}\beta,m\alpha} =\displaystyle= (ωpc/nbg)2d3𝒌φ~m(𝒌)(1𝒌2+(nbgω/c)2)φ~m(𝒌)δαβ\displaystyle\left(\frac{\omega_{\rm p}^{\prime}}{c/n_{\rm bg}}\right)^{2}\int d^{3}\bm{k}\tilde{\varphi}_{m^{\prime}}(-\bm{k})\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{\varphi}_{m}^{\ast}(\bm{k})\delta_{\alpha\beta} (94)
=\displaystyle= Cm,mδαβ.\displaystyle C_{m^{\prime},m}\delta_{\alpha\beta}.

Note that ωp=ωp/nbg\omega_{\rm p}^{\prime}=\omega_{\rm p}/n_{\rm bg} is the (bulk) plasma frequency in a material with the refractive index nbgn_{\rm bg}. The detail expression for Aμ,μ(0,1,2)A_{\mu^{\prime},\mu}^{(0,1,2)}, Bmα,μ(1,2)B_{m^{\prime}\alpha,\mu}^{(1,2)}, and Cm,mC_{m^{\prime},m} are summarized in Appendix E.

VI.2 Exciation spectrum

The matrix Ξ¯(ω)\bar{\Xi}(\omega) of the self-consistent formulation in Eq. (47) describes the spectrum for both the individual-like electron–hole excitations and the field-mediated collective-like excitation. Then, we demonstrate the eigenvalues and determinant of the matrix Ξ¯(ω)\bar{\Xi}(\omega) for the nanorod as functions of the real component of frequency, ω=ωr+iωi\omega=\omega_{\rm r}+i\omega_{\rm i}. By modulating the field-mediated couplings to reduce the T field contribution as shown in the following numerical demonstrations, we find that the T field contributes significantly to form the collective-like excitation and enlarges the radiative width related to the energy transfer between the excitations.

We consider Lx=10nmL_{x}=10\,\mathrm{nm}, Ly=15nmL_{y}=15\,\mathrm{nm}, and Lz=200nmL_{z}=200\,\mathrm{nm} rectangular nanorod. The Fermi energy of conduction electron is set at εF=0.5eV\varepsilon_{\rm F}=0.5\,\mathrm{eV}. The effective mass is m=0.02mem^{*}=0.02m_{\rm e} with mem_{\rm e} being the electron mass in vacuum. Such an effective mass is obtained for InSb. The background refractive index is taken as nbg=5n_{\rm bg}=5. For a typical size scale L0=100nmL_{0}=100\,\mathrm{nm}, an order of the confinement energy is E0=2π2/(2mL02)1.88meVE_{0}=\hbar^{2}\pi^{2}/(2m^{*}L_{0}^{2})\simeq 1.88\,\mathrm{meV}. Then, the electron–hole excitation energy is ωμ0=E0{(nx2+ny2+nz2)(n¯x2+n¯y2+n¯z2)}\hbar\omega_{\mu 0}=E_{0}\left\{(n_{x}^{2}+n_{y}^{2}+n_{z}^{2})-(\bar{n}_{x}^{2}+\bar{n}_{y}^{2}+\bar{n}_{z}^{2})\right\} with μ=(eμ,hμ¯)=({nx,ny,nz},{n¯x,n¯y,n¯z})\mu=({\rm e}\mu,{\rm h}\bar{\mu})=(\{n_{x},n_{y},n_{z}\},\{\bar{n}_{x},\bar{n}_{y},\bar{n}_{z}\}) satisfying E0(n¯x2+n¯y2+n¯z2)εF<E0(nx2+ny2+nz2)E_{0}(\bar{n}_{x}^{2}+\bar{n}_{y}^{2}+\bar{n}_{z}^{2})\leq\varepsilon_{\rm F}<E_{0}(n_{x}^{2}+n_{y}^{2}+n_{z}^{2}). We set γ=0.1×E0\gamma=0.1\times E_{0} as an infinitesimal value. In bulk, the electron density is evaluated as n0=(2mεF/2)3/2/(3π2)n_{0}=(2m^{*}\varepsilon_{\rm F}/\hbar^{2})^{3/2}/(3\pi^{2}). Then, the bulk plasma frequency for above parameters is ωp=e2n0/(εbgm)0.112eV\hbar\omega_{\rm p}^{\prime}=\hbar\sqrt{e^{2}n_{0}/(\varepsilon_{\rm bg}m^{*})}\simeq 0.112\,\mathrm{eV}.

Refer to caption
Figure 1: (a) Numerical results of det(Ξ¯(ω)){\rm det}\left(\bar{\Xi}(\omega)\right) as a function of ωr\omega_{\rm r} with ω=ωr+iωi\omega=\omega_{\rm r}+i\omega_{\rm i} when the T field contribution is reduced (ζ=0\zeta=0). A parameter ζ\zeta tunes the T components of the Green’s function 𝒢¯𝒌\bar{\mathcal{G}}_{\bm{k}}. ζ=1(0)\zeta=1(0) corresponds to a full (no transverse field) calculation. The energy dimension of Ξ¯(ω)\bar{\Xi}(\omega) is normalized. The imaginary part of frequency is introduced to prevent mathematical divergence and set at ωi=0.002eV\hbar\omega_{\rm i}=0.002\,\mathrm{eV}. (b) Modified plot of log|det(Ξ¯(ω))|\log\left|{\rm det}\left(\bar{\Xi}(\omega)\right)\right| in (a) to see clearly the collective excitation. The determinant is subtracted by a harmonic function f(x)6580.9077x2+5034.8828x1256.5965f(x)\simeq-6580.9077x^{2}+5034.8828x-1256.5965 with x=ω/(1eV)x=\hbar\omega/(1\,\mathrm{eV}), which is deduced from the values at x=0.95x=0.95, 1.051.05, and 1.151.15. Lines indicate log|det(Ξ¯(ω))|\log\left|{\rm det}\left(\bar{\Xi}(\omega)\right)\right| from ζ=1\zeta=1 to 0 by 0.10.1. (c) Color plot of modified log|det(Ξ¯(ω))|\log\left|{\rm det}\left(\bar{\Xi}(\omega)\right)\right| in (b). A line indicates the position of ωr\omega_{\rm r} satisfying a real part of the eigenvalues of Ξ¯(ω)\bar{\Xi}(\omega) being zero, Re[ξj(ω)]=0{\rm Re}\left[\xi_{j}(\omega)\right]=0.

We focus on the excitation with nx=n¯xn_{x}=\bar{n}_{x}, ny=n¯yn_{y}=\bar{n}_{y}, and nz=n¯z+1n_{z}=\bar{n}_{z}+1 to consider the individual-like and collective-like excitation spectrum at a small wavenumber |𝒒|=|𝒌eμ𝒌hμ¯|=π/Lz|\bm{q}|=|\bm{k}_{{\rm e}\mu}-\bm{k}_{{\rm h}\bar{\mu}}|=\pi/L_{z}. Because of stronger confinements in the xx- and yy-directions than that in the zz-direction, a subband structure characterized by index nx,yn_{x,y} is formed; hence, our consideration corresponds to only the intra-subband excitation. The excitation spectrum should be evaluated from the zero points of the eigenvalues of Ξ¯(ω)\bar{\Xi}(\omega) in the complex plane of ω=ωr+iωi\omega=\omega_{\rm r}+i\omega_{\rm i}. In this demonstration, however, we fix the imaginary part of frequency at ωi=0.002eV\hbar\omega_{\rm i}=0.002\,\mathrm{eV} for the visibility and discuss the eigenvalues and determinant of Ξ¯(ω)\bar{\Xi}(\omega) when ωr\omega_{\rm r} is swept. First, we consider the determinant when the T field contribution is reduced from the matrix Ξ¯(ω)\bar{\Xi}(\omega) to see clearly the individual-like and collective-like excitations since the T field generally causes the radiative correction and enlarges the imaginary components of the eigenvalues. Then, we introduce a parameter ζ\zeta to tune the T component in the Green’s function 𝒢¯𝒌(ω)\bar{\mathcal{G}}_{\bm{k}}(\omega), i.e., we set K~μ0,0μ=ζ2Aμ,μ(2)+ζAμ,μ(1)+Aμ,μ(0)\tilde{K}_{\mu^{\prime}0,0\mu}=\zeta^{2}A_{\mu^{\prime},\mu}^{(2)}+\zeta A_{\mu^{\prime},\mu}^{(1)}+A_{\mu^{\prime},\mu}^{(0)}, where the order of ζ\zeta corresponds to that of current density in the matrix Ξ¯(ω)\bar{\Xi}(\omega). Namely, ζ=1\zeta=1 and 0 signify a fully consideration and only the Coulomb interaction, respectively. Figure 1(a) represents log|det(Ξ¯(ω))|\log\left|{\rm det}\left(\bar{\Xi}(\omega)\right)\right| when ζ=0\zeta=0. The energy levels of the electron–hole excitations at |𝒒|=π/Lz|\bm{q}|=\pi/L_{z} are distributed in ωμ0<0.22eV\hbar\omega_{\mu 0}<0.22\,\mathrm{eV}. The eigenvalues of Ξ¯(ω)\bar{\Xi}(\omega) indicates zero points at ωrωμ0\omega_{\rm r}\approx\omega_{\mu 0}, which suppresses strongly the determinant and shows jagged behavior at ωr<0.22eV\hbar\omega_{\rm r}<0.22\,\mathrm{eV} in Fig. 1(a). In the present calculation, the imaginary frequency and the field-mediated couplings broaden the determinant dips. Hence, it is difficult to find respective dip structures due to the respective electron–hole excitations.

Above the level distribution, the determinant increases monotonically with ωr\omega_{\rm r}, where the signature of a dip structure due to the collective excitation is difficult to be found from the determinant in this scale. Then, we modulate the plot to see a signature of the collective-like excitation clearly by subtracting a harmonic smooth function deduced from the plot in Fig. 1(a). Figures 1(b) and (c) exhibit the modulated plot of log|det(Ξ¯(ω))|\log\left|{\rm det}\left(\bar{\Xi}(\omega)\right)\right|, which shows a shift of dip structure when the tuning parameter is tuned from ζ=0\zeta=0 to 11. By tuning ζ\zeta, we can discuss a contribution of the T field to the construction of collective-like excitation. Note that in our formulation, the existence of collective (plasmon) mode is not supposed in the Hamiltonian in an empirical way. If the plasmon mode was assumed as an empirical model, the radiative T field would simply contribute to the radiative shift of the assumed plasmon energy. On the contrary, in our result, the collective(-like) mode appears in the deductive process from a cooperation of the electron–hole excitations via both L and T fields. Then, the obtained shift of collective-like excitation spectrum in Fig. 1(b) contains not only the radiative shift but also contributions by an additional mechanism to form the collective mode by the L and T fields. If only the Coulomb interaction is considered (ζ=0\zeta=0), the modulated plot shows the dip structure due to the collective-like excitation at ωr0.284eV\hbar\omega_{\rm r}\approx 0.284\,\mathrm{eV} although its depth is much smaller than the determinant in Fig. 1(a). Note that generally the plasmon excitation energy in the nanostructure is largely blue-shifted from ωp\omega_{\rm p} in bulk. When ζ=0\zeta=0, the dip width is attributed to the imaginary part of the frequency. If ωi\omega_{\rm i} is reduced, the dip structure becomes sharp and shows divergence at ωi=0\omega_{\rm i}=0 (not shown).

With a tuning of ζ\zeta, in addition to the slight shift of dip position, the depth decreases and the width broadens. It signifies that the T-field-mediated interaction between the excitations induces an energy dissipation. When ζ=1\zeta=1, however, the dip structure is not distinguished in Fig. 1(b); hence, the energy of collective-like excitation cannot be evaluated from the determinant. Then, the zero position of (real part of) the eigenvalue of Ξ¯\bar{\Xi} corresponding to the collective-like excitation, Re[(Ξ¯(ω))j]=0{\rm Re}\left[\left(\bar{\Xi}(\omega)\right)_{j}\right]=0, is examined in Fig. 1(c). The position also shifts with ζ\zeta from ωr0.284eV\hbar\omega_{\rm r}\approx 0.284\,\mathrm{eV} to 0.260eV\approx 0.260\,\mathrm{eV}. The collective-like excitation is well defined even at ζ=1\zeta=1 although the determinant does not indicate the distinct signature. The shift of collective-like excitation with the increase of ζ\zeta is harmonic behavior. Thus, the current–current interaction Aμ,μ(2)A_{\mu^{\prime},\mu}^{(2)} is dominant than the current–charge interaction Aμ,μ(1)A_{\mu^{\prime},\mu}^{(1)}. The contribution of the T component to the collective-like excitation is not negligible, especially for the hybridization and the energy transfer between the individual-like and collective-like excitations. A generation of the hot electron and hole should be described by the individual excitations. Thus, the hybridization by the T field would be an important ingredient in the hot carrier generation from the plasmon excitation.

VII Discussion

In the self-consistent formulation in Eq. (44) and its matrix form (46) with Eq. (47), the optical response of the nanostructures is described in terms of 𝓙νμ(𝒙)\bm{\mathcal{J}}_{\nu\mu}(\bm{x}). The electron eigenstates |μ|\mu\rangle contribute to the optical response via 𝓙νμ(𝒙)\bm{\mathcal{J}}_{\nu\mu}(\bm{x}) in the susceptibility. When one prepares the |μ|\mu\rangle of H^0\hat{H}_{0}, including the Coulomb interaction at equilibrium, the nanostructure optical responses, e.g., the plasmon–polariton and SPE spectra, are obtained within a theoretical framework for the prepared eigenstates. The first-principles calculation for a nanostructure or a nanoscale cluster of atoms was developed Kumar19 ; Zhang17 ; Chapkin18 ; Senanayake19 ; Ma15 ; Schira19 ; Lu19 ; Svendsen21 , where the electronic states with the electron–electron interaction (the L component of the EM field) are evaluated precisely. In many previous studies, however, the T field was not self-consistently considered. In our formulation, both components of the EM fields are determined self-consistently on the mesoscopic scale. Further, if one employs the equilibrium states evaluated in the first-principles calculation in our formulation, the T fields generated by the electronic responses and the electron–electron interactions are taken into account com1 . Hence, our approach provides an important and convenient framework for investigations of mesoscopic plasmonics and photonics.

The present formulation is based on the classical Maxwell’s equations for the incident and induced EM fields. As the induced L field describes part of the electron–electron interaction, the quantum Maxwell’s equations for the EM fields are required for the higher-order electron correlations. For the linear response framework discussed in this paper, however, the classical treatment of the fields is equivalent to the quantum treatment.

Let us emphasize the difference in the four-vector representation of Eq. (46) compared to the conventional nonlocal and self-consistent theory book:cho1 . In our formulation, both the vector and scalar potentials are fully considered. The Coulomb gauge separates the L and T components in 𝓐=(𝑨,ϕ/c)\bm{\mathcal{A}}=(\bm{A},-\phi/c), which directly provides a picture of the L and T component mixing due to the nanostructure in the self-consistent relation in Eq. (46). For the densities 𝓙=(𝒋,cρ)\bm{\mathcal{J}}=(\bm{j},c\rho) in the constitutive equation (27), the current is induced not only by the T field, but also by the L field though H^(𝒋^𝑨ρ^ϕ)\hat{H}^{\prime}\sim-\left(\hat{\bm{j}}\cdot\bm{A}-\hat{\rho}\phi\right). The charge density is also attributed to both the L and T fields. These characteristics are due to the off-diagonal elements in the 4×44\times 4 susceptibility 𝒳¯\bar{\mathcal{X}}. In the Maxwell’s equations (41), the L field ϕ\phi is generated by the charge only, whereas both the current and charge generate the T field 𝑨\bm{A}. This “cross generation” of LT components is essential physics for the nanostructure and is enlarged when the nanostructure enhances its 𝝌jρ\bm{\chi}_{j\rho} and 𝝌ρj\bm{\chi}_{\rho j} due to the nonlocality. The present formulation might provide a guideline for an enhancement of the plasmon resonance since the generation of the current and charge densities is related to the collective-like and individual-like excitations.

We have demonstrated a single nanorod as an example of the applications in Sec. VI.2. For rectangular nanorods, the charge and current densities in the radiative correlation matrices are analytically obtained. Such analytical expressions present a clear understanding of the relation between the excitations and the induced densities. The matrix components were evaluated numerically, which exhibits the practicality of our formulation. The numerical results reveal the shift of collective-like excitation spectrum and enhanced radiative width due to the T field contribution. In our formulation and model, the collective-like excitation appears in the deductive process from the field-mediated interaction between the electron–hole excitations. Then, our results suggest a new mechanism for the collective(-like) excitation.

In the present demonstration, we consider the individual excitations with only small wavenumber |𝒒|=|𝒌eμ𝒌hμ¯|=π/Lz|\bm{q}|=|\bm{k}_{{\rm e}\mu}-\bm{k}_{{\rm h}\bar{\mu}}|=\pi/L_{z}. This corresponds to the intra-subband excitation for strong confinement in the xx- and yy-directions. However, if the system is larger, or there are several interacting nanostructures in the xyxy-plane, the charge and current deviations might be important. In such situations, the inter-subband excitations becomes essential. Moreover, the T field contributes to the coherent coupling between the collective-like and individual-like excitations. In most previous studies that describe hot carrier generation caused by the plasmon excitation, the energy transfer is unidirectional Zhang14 ; Ross15 ; Besteiro17 ; Manjavacas14 ; Chapkin18 . However, if the coherent coupling between the collective and individual excitations is large, a bidirectional energy transfer Ma15 ; You18 can be effective in the nanostructures. To discuss such coherent coupling mechanism, we will extend the model with large wavenumber for the individual excitations considering the energy closing with the collective-like excitation in our future study.

As a further discussion, based on our formulation, the optical responses of electrons are described in terms of the current and charge densities. This representation is useful for considering the responses to the L and T field components. In many previous studies, the optical properties were discussed in terms of the plasmons and SPEs. In nanostructures featuring nonlocality, the plasmons and SPEs are coupled with each other. Hence, the relation between the densities (𝒋,cρ\bm{j},c\rho) and the collective and individual excitations is complicated. The excitation energy spectrum is evaluated from the eigenvalues of Ξ¯(ω)\bar{\Xi}(\omega) in Eq. (47) as the excitations correspond to the poles of the system with the radiative field. The densities 𝓙\bm{\mathcal{J}} and induced fields 𝓐\bm{\mathcal{A}} are described in terms of the excitations. This is an advantage over another existing treatment of the nonlocal effect Pitarke07 . Moreover, from the poles, coherent coupling between the plasmon (collective-like) excitation and the carrier generation due to the SPE can be investigated.

The coherent coupling of the collective-like and individual-like excitations in the nanostructures gives rise to a bidirectional energy transfer Ma15 ; You18 . The plasmon and individual excitations form discrete and continuous spectra, respectively. Therefore, their coupling may demonstrate the Fano resonance as evidence of coherent coupling. In such a scenario, the self-consistency of the L and T total fields and the induced charge and current densities are important. In addition to the energy, the electron wavefunction in the nanostructures may significantly affect the coupling. Therefore, our microscopic-theory-based formulation reveals such light–nanostructure interaction and its enhancement. As a multi-dimensional integral is included in Eqs. (53)–(66), the presence of numerous electrons, even in submicro-scale metallic structures, would make a feasibility of numerical calculations difficult. The approach based on the first-principles calculation Ma15 ; You18 has more limited applications than the analytical approach based on our formulation. Note also that we can overcome this difficulty for several nanostructures, e.g., a 2D sheet (including graphene) and a rectangular rod.

The development based on our microscopic approach are not limited to several adaptable nanostructures, but can be applied to arbitrary nanostructures by considering appropriate approximations to restrict the bases. Our approach can also be applied to an array of two or more nanostructure units, in which the nonlocal effect is enlarged by the nanogap structure. The collective(-like) excitations in the respective nanostructures are coupled with each other by the T and L fields. Then, our formulation will contribute to reveal a coherent energy transfer between the nanostructures and its enhancement.

VIII Summary

To correctly describe the coherent coupling between the collective plasmon excitations and the individual single-particle excitations via a transverse electric field, which is an aspect that has been neglected in previous studies, we developed a self-consistent and microscopic nonlocal formulation for the light–matter interaction in plasmonic nanostructures. Our formulation is based on linear response theory with a nonlocal susceptibility and the classical Maxwell’s equations describing the transverse and longitudinal electromagnetic fields. The nonlocal susceptibility 𝒳¯(𝒙,𝒙)\bar{\mathcal{X}}(\bm{x},\bm{x}^{\prime}) is obtained from the interaction Hamiltonian and the eigenstates of the nanostructure. The Coulomb interaction is included not only in the non-perturbative Hamiltonian H^0\hat{H}_{0}, but also in the longitudinal electric fields caused by the charge density related to the collective and individual excitations. In the formulation, the longitudinal and transverse components of the fields and optically induced responses are described in terms of the four-vector potential 𝓐(𝒙,t)\bm{\mathcal{A}}(\bm{x},t) and density 𝓙(𝒙,t)\bm{\mathcal{J}}(\bm{x},t) with the Green’s function 𝒢¯(𝒙,𝒙)\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime}) in the four-vector space. The self-consistent equation is rewritten in matrix form with the incident and induced fields, Yμν(0,±)Y^{(0,\pm)}_{\mu\nu} and Xμν(±)X^{(\pm)}_{\mu\nu}. From the poles of this formulated matrix Ξ¯(ω)\bar{\Xi}(\omega) for the radiative correction, the excitation spectrum can be examined. The developed formulation can be applied to various frameworks by utilizing the electronic states obtained according to those respective frameworks. We examine the excitation spectrum for a rectangular nanorod to demonstrate the numerical feasibility of our formulation. The solutions of formulated matrix Ξ¯(ω)\bar{\Xi}(\omega) provides the collective-like excitation spectrum, where the transverse-field-mediated interaction contributes non-negligiblly to the coherent coupling between the individual-like and collective-like excitations.

Our formulation could be combined with real-time density functional theory based on the first-principles calculation, in which the electron–electron interaction is considered accurately. The four-vector formulation can describe the effects of transverse fields on the eigenstates given by the first-principles calculation at each step in a real-time simulation. In this treatment, the exchange correlation and quantum fluctuation can be partially considered via the excited states. Application of the present formulation to various types of metallic meso- and nanostructures will facilitate theoretical understanding of the interplay between the collective(-like) and individual(-like) excitations in nanoscale systems, which will aid the design of metallic systems to control photo-induced hot carrier generation.

Acknowledgements.
T.Y. and H.I. are supported by JSPS KAKENHI JP16H06504 in Scientific Research on Innovative Areas, “Nano-Material Optical-Manipulation.” T.Y. is supported by JSPS KAKENHI 18K13484. H.I. is supported by JSPS KAKENHI 18H01151.

Appendix A Another treatment of Hamiltonian

In Sec. II, we separate the Hamiltonian into a static component including a part of the Coulomb interaction, H^0\hat{H}_{0}, and a light-induced component, H^ind\hat{H}^{\prime}_{\rm ind}. In this description, the Coulomb interaction is included in the non-perturbative and perturbative Hamiltonian. In this appendix, we consider another treatment for the Coulomb interaction and compare the two descriptions.

The total Hamiltonian H^\hat{H} subtracting a constant value U0=𝑑𝒙ρ0ϕextU_{0}=\int d\bm{x}\rho_{0}\phi_{\rm ext} can be read as following two descriptions,

H^\displaystyle\hat{H} =\displaystyle= (H^0+H^pp)+(H^int+H^ext)=H^0+H^int,\displaystyle\left(\hat{H}_{0}+\hat{H}_{\rm p-p}^{\prime}\right)+\left(\hat{H}_{\rm int}+\hat{H}_{\rm ext}\right)=\hat{H}_{0}^{\prime}+\hat{H}_{\rm int}^{\prime}, (95)
=\displaystyle= H^0+(H^int+H^ext+H^pp)=H^0+H^ind.\displaystyle\hat{H}_{0}+\left(\hat{H}_{\rm int}+\hat{H}_{\rm ext}+\hat{H}_{\rm p-p}^{\prime}\right)=\hat{H}_{0}+\hat{H}_{\rm ind}^{\prime}. (96)

Here, H^pp=H^ppU0\hat{H}_{\rm p-p}^{\prime}=\hat{H}_{\rm p-p}-U_{0}.

In the former description of Eq. (95), the electron–electron interaction due to the L field caused by the light-induced charge density is fully included in the non-perturbative Hamiltonian, H^0H^0+H^pp\hat{H}_{0}^{\prime}\equiv\hat{H}_{0}+\hat{H}_{\rm p-p}^{\prime}. The interaction with an external L field is in the perturbative one,

H^int\displaystyle\hat{H}_{\rm int}^{\prime} \displaystyle\equiv H^int+H^ext\displaystyle\hat{H}_{\rm int}+\hat{H}_{\rm ext} (97)
=\displaystyle= 𝑑𝒙[𝒋^(𝒙,t)𝑨(𝒙,t)ρ^(𝒙,t)ϕext(𝒙,t)].\displaystyle-\int d\bm{x}\left[\hat{\bm{j}}(\bm{x},t)\cdot\bm{A}(\bm{x},t)-\hat{\rho}(\bm{x},t)\phi_{\rm ext}(\bm{x},t)\right].

Here, both the incident and induced T fields are coupled with the charge current density. Because the perturbative Hamiltonian is time-dependent via H^pp\hat{H}_{\rm p-p} in Eq. (20), one must prepare a quite complex and time-dependent basis |μ(t)|\mu(t)\rangle. Thus, the evaluation of susceptibility by the time-dependent basis becomes complicated. The induced Coulomb interaction contributes to the induced charge and current densities via the susceptibility. This approach is reasonable if the self-consistent relation between the constitutive and Maxwell’s equations is not considered. However, for the self-consistent equation with this description, the constitutive equation includes ϕ^mat\hat{\phi}_{\rm{mat}} and ϕ^pol\hat{\phi}_{\rm{pol}}, simultaneously, the interaction between the induced charges is also considered in the Maxwell’s equations. Then, a double count problem of the electron–electron interaction must also be handled. Therefore, because of the basis complexity and the double count problem, this approach is not useful to analyze the relation between the L and T components of the induced fields and densities.

For the latter description of Eq. (96), however, the non-perturbative Hamiltonian H^0\hat{H}_{0} and its basis |μ|\mu\rangle are static. All induced effects by the light irradiation are included in the perturbative one,

H^ind=𝑑𝒙[𝒋^(𝒙,t)𝑨(𝒙,t)δρ^(𝒙,t)(ϕext(𝒙,t)+ϕpol(𝒙,t))].\hat{H}_{\rm ind}^{\prime}=-\int d\bm{x}\left[\hat{\bm{j}}(\bm{x},t)\cdot\bm{A}(\bm{x},t)-\delta\hat{\rho}(\bm{x},t)\Big{(}\phi_{\rm ext}(\bm{x},t)+\phi_{\rm pol}(\bm{x},t)\Big{)}\right]. (98)

Note that although ϕpol\phi_{\rm pol} is attributed to the Coulomb interaction between the internal charges of the matter, it arises only when the external fields are applied. Because of the electric neutrality of the matter, ϕ^mat\hat{\phi}_{\rm{mat}} is not incorporated in the Maxwell’s equations. Therefore, no treatment of the double counting of the electron–electron interaction is required as the induced Coulomb interaction is irrelevant to the susceptibility in the constitutive equation and considered only in the Maxwell’s equations.

non-perturb. H perturb. H basis |μ|\mu\rangle Coulomb int. L field T field susceptibility
H^0+H^pp\hat{H}_{0}+\hat{H}_{\rm{p-p}}^{\prime} H^int\hat{H}_{\rm{int}}^{\prime} time-dependent fully in |μ|\mu\rangle ϕext\phi_{\rm{ext}} 𝑨\bm{A} time-dep. via |μ|\mu\rangle
H^0\hat{H}_{0} H^ind\hat{H}_{\rm{ind}}^{\prime} time-independent partially in |μ|\mu\rangle ϕpol+ϕext\phi_{\rm{pol}}+\phi_{\rm{ext}} 𝑨\bm{A} static
Table 1: Treatment of Hamiltonian in Eqs. (95) and (96). Here, ϕext\phi_{\rm{ext}} describes an external (applied) L field, ϕpol\phi_{\rm{pol}} is an induced L field, and 𝑨\bm{A} consists of the T field of both the applied and induced components.

The above discussion and classification are summarized in TABLE I and Figure 1.

Refer to caption
Figure 2: Schematic summary of two treatments of the Hamiltonian. In the upper case, the electronic state of non-perturbative Hamiltonian has induced polarized charges. The Coulomb interaction of the induced charges is included in both the constitutive and Maxwell’s equations. Then, one has to take care of the double count of Coulomb interaction. The lower case is our scheme, where the Coulomb interaction of the induced charges is taken into account only in the Maxwell’s equations. Therefore, one does not need to consider the double count.

Appendix B L-T separation of Maxwell’s equations

In the main text, we formulated the self-consistent equation in terms of {(𝑨,ϕ/c),(𝒋,cρ)}\{(\bm{A},-\phi/c),(\bm{j},c\rho)\} with the Coulomb gauge. Let us consider another description in terms of {(𝑨(T),ϕ(L)/c),(𝒋(T),𝒋(L))}\{(\bm{A}^{\rm(T)},-\phi^{\rm(L)}/c),(\bm{j}^{\rm(T)},\bm{j}^{\rm(L)})\}. Here, to emphasize the separation of the L and T components in the fields and densities, we present the superscripts explicitly. Equation (36) can be expressed as

(21c22t2)𝑨(T)(𝒙,t)\displaystyle\left(\bm{\nabla}^{2}-\frac{1}{c^{2}}\frac{\partial^{2}}{\partial t^{2}}\right)\bm{A}^{\rm(T)}(\bm{x},t) =\displaystyle= μ0𝒋(T)(𝒙,t),\displaystyle-\mu_{0}\bm{j}^{\rm(T)}(\bm{x},t), (99)
1c2tϕ(L)(𝒙,t)\displaystyle-\frac{1}{c^{2}}\frac{\partial}{\partial t}\bm{\nabla}\phi^{\rm(L)}(\bm{x},t) =\displaystyle= μ0𝒋(L)(𝒙,t),\displaystyle-\mu_{0}\bm{j}^{\rm(L)}(\bm{x},t), (100)

where 𝒋(𝒙,t)=𝒋(T)(𝒙,t)+𝒋(L)(𝒙,t)\bm{j}(\bm{x},t)=\bm{j}^{\rm(T)}(\bm{x},t)+\bm{j}^{\rm(L)}(\bm{x},t). The scalar potential ϕ(L)\phi^{\rm(L)} generates the L component only. Equation (37) also describes the L only, where

2ϕ(L)(𝒙,t)=ρ(L)(𝒙,t)ε0.\bm{\nabla}^{2}\phi^{\rm(L)}(\bm{x},t)=-\frac{\rho^{\rm(L)}(\bm{x},t)}{\varepsilon_{0}}. (101)

Equations (100) and (101) satisfy the continuous relation for the L component,

𝒋(L)(𝒙,t)+ρ(L)(𝒙,t)t=0.\bm{\nabla}\cdot\bm{j}^{\rm(L)}(\bm{x},t)+\frac{\partial\rho^{\rm(L)}(\bm{x},t)}{\partial t}=0. (102)

Let us refer to the T component in the scalar potential as ϕ(T)\phi^{\rm(T)}. The T component should be 𝑬=(ϕ(T))=0\bm{\nabla}\cdot\bm{E}=\bm{\nabla}\cdot\left(-\bm{\nabla}\phi^{\rm(T)}\right)=0. Thus, it must have a linear dependence up to ϕ(T)(𝒙,t)ax+b\phi^{\rm(T)}(\bm{x},t)\sim ax+b. These terms cannot be induced from the internal charge density of the system. We incorporate the external scalar potential ϕext(𝒙,t)\phi_{\rm ext}(\bm{x},t) in the interaction Hamiltonian (23). The T component in the external field, however, is included in the vector potential. Therefore, the scalar potential induces the L component only, and we find 𝑬(L)=ϕ(L)\bm{E}^{\rm(L)}=-\bm{\nabla}\phi^{\rm(L)} and 𝑬(T)=t𝑨(T)\bm{E}^{\rm(T)}=-\partial_{t}\bm{A}^{\rm(T)} for 𝑬=𝑬(L)+𝑬(T)\bm{E}=\bm{E}^{\rm(L)}+\bm{E}^{\rm(T)}.

Such L and T field separation has been also discussed by Cho book:cho2 , where the scalar potential is reduced from the Hamiltonian (or Lagrangian) and only the T components, 𝑨(T)\bm{A}^{\rm(T)} and 𝒋(T)\bm{j}^{\rm(T)}, are included in the Maxwell’s equations. The L field, 𝑬(L)\bm{E}^{\rm(L)}, is treated as the external field only.

The representation of (𝒋(T),𝒋(L)\bm{j}^{\rm(T)},\bm{j}^{\rm(L)}) is a modification of (𝒋,cρ\bm{j},c\rho). It may be useful to separate the L and T components clearly in the source terms. Moreover, the Green’s function for Eqs. (99) and (100) is simpler than Eq. (43), as the off-diagonal components are absent. However, for this representation, the constitutive equation (27) [and the perturbative Hamiltonian (23)] should be rewritten in terms of (𝒋(T),𝒋(L)\bm{j}^{\rm(T)},\bm{j}^{\rm(L)}). Then, the susceptibility 𝒳¯\bar{\mathcal{X}} must be a 6×46\times 4 matrix, which is not desirable in terms of the mathematics.

Appendix C Derivation of Green’s function

The Green’s function for the Maxwell’s equations is defined according to the differential operators in the equation, which corresponds to the matrix 𝒟¯(𝒙;ω)\bar{\mathcal{D}}(\bm{x};\omega) in Eq. (39). Therefore, the Green’s function 𝒢¯(𝒙,𝒙;ω)\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega) is the 4×44\times 4 matrix,

𝒟¯(𝒙;ω)𝒢¯(𝒙,𝒙;ω)=1¯δ(𝒙𝒙).\bar{\mathcal{D}}(\bm{x};\omega)\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)=\bar{1}\delta(\bm{x}-\bm{x}^{\prime}). (103)

Note that 𝒟¯(𝒙;ω)\bar{\mathcal{D}}(\bm{x};\omega) is defined as Eq. (40). The formal solution of the Maxwell’s equation is

𝓐(𝒙;ω)=𝓐0(𝒙;ω)μ0𝑑𝒙𝒢¯(𝒙,𝒙;ω)𝓙(𝒙;ω).\bm{\mathcal{A}}(\bm{x};\omega)=\bm{\mathcal{A}}_{0}(\bm{x};\omega)-\mu_{0}\int d\bm{x}^{\prime}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}(\bm{x}^{\prime};\omega). (104)

The first term in Eq. (104) satisfies

𝒟¯(𝒙;ω)𝓐0(𝒙;ω)=0.\bar{\mathcal{D}}(\bm{x};\omega)\bm{\mathcal{A}}_{0}(\bm{x};\omega)=0. (105)

To obtain an explicit form, 𝒢¯\bar{\mathcal{G}} is divided into the following matrix elements:

𝒢¯(𝒙,𝒙;ω)=(g¯AA(𝒙,𝒙;ω)𝒈Aϕ(𝒙,𝒙;ω)𝒈ϕAt(𝒙,𝒙;ω)gϕϕ(𝒙,𝒙;ω)),\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)=\left(\begin{matrix}\bar{g}_{AA}(\bm{x},\bm{x}^{\prime};\omega)&\bm{g}_{A\phi}(\bm{x},\bm{x}^{\prime};\omega)\\ \bm{g}_{\phi A}^{\ {\rm t}}(\bm{x},\bm{x}^{\prime};\omega)&g_{\phi\phi}(\bm{x},\bm{x}^{\prime};\omega)\end{matrix}\right), (106)

where g¯AA\bar{g}_{AA} is a 3×33\times 3 matrix, g¯ϕϕ\bar{g}_{\phi\phi} is scalar, and the other elements are vectors.

We have a mathematical preparation. The solution of

(2+k2)gk(𝒙,𝒙)=δ(𝒙𝒙)\left(\bm{\nabla}^{2}+k^{2}\right)g_{k}(\bm{x},\bm{x}^{\prime})=\delta(\bm{x}-\bm{x}^{\prime}) (107)

is given as

gk(𝒙,𝒙)=eik|𝒙𝒙|4π|𝒙𝒙|.g_{k}(\bm{x},\bm{x}^{\prime})=-\frac{e^{ik|\bm{x}-\bm{x}^{\prime}|}}{4\pi|\bm{x}-\bm{x}^{\prime}|}. (108)

Here, we avoid constant and linear terms, α0+𝜶1𝒙\alpha_{0}+\bm{\alpha}_{1}\cdot\bm{x}, in Eq. (108), because of the boundary condition g(|𝒙|)=0g(|\bm{x}|\to\infty)=0.

The scalar potential component of the Green’s function satisfies

2gϕϕ(𝒙,𝒙;ω)=δ(𝒙𝒙),-\bm{\nabla}^{2}g_{\phi\phi}(\bm{x},\bm{x}^{\prime};\omega)=\delta(\bm{x}-\bm{x}^{\prime}), (109)

the solution of which is given by Eq. (108) as follows:

gϕϕ(𝒙,𝒙;ω)=g0(𝒙,𝒙)=14π|𝒙𝒙|.g_{\phi\phi}(\bm{x},\bm{x}^{\prime};\omega)=-g_{0}(\bm{x},\bm{x}^{\prime})=\frac{1}{4\pi|\bm{x}-\bm{x}^{\prime}|}. (110)

Further, gϕϕg_{\phi\phi} gives another matrix element 𝒈Aϕ\bm{g}_{A\phi}. From Eq. (103),

(2+k2)𝒈Aϕ(𝒙,𝒙;ω)ikgϕϕ(𝒙,𝒙;ω)=0.\left(\bm{\nabla}^{2}+k^{2}\right)\bm{g}_{A\phi}(\bm{x},\bm{x}^{\prime};\omega)-ik\bm{\nabla}g_{\phi\phi}(\bm{x},\bm{x}^{\prime};\omega)=0. (111)

Here, k=ω/ck=\omega/c. By substituting Eq. (110),

(2+k2)𝒈Aϕ(𝒙,𝒙;ω)=ikgϕϕ(𝒙,𝒙;ω)=ik4π𝒙𝒙|𝒙𝒙|3.\left(\bm{\nabla}^{2}+k^{2}\right)\bm{g}_{A\phi}(\bm{x},\bm{x}^{\prime};\omega)=ik\bm{\nabla}g_{\phi\phi}(\bm{x},\bm{x}^{\prime};\omega)=-\frac{ik}{4\pi}\frac{\bm{x}-\bm{x}^{\prime}}{|\bm{x}-\bm{x}^{\prime}|^{3}}. (112)

This is an inhomogeneous equation, which is related to Eq. (107). The solution of 𝒈Aϕ\bm{g}_{A\phi} is also obtained using Eq. (108), with

𝒈Aϕ(𝒙,𝒙;ω)=𝒈k(𝒙)+𝑑𝒙′′gk(𝒙,𝒙′′)(ik4π𝒙′′𝒙|𝒙′′𝒙|3).\bm{g}_{A\phi}(\bm{x},\bm{x}^{\prime};\omega)=\bm{g}_{k}(\bm{x})+\int d\bm{x}^{\prime\prime}g_{k}(\bm{x},\bm{x}^{\prime\prime})\left(-\frac{ik}{4\pi}\frac{\bm{x}^{\prime\prime}-\bm{x}^{\prime}}{|\bm{x}^{\prime\prime}-\bm{x}^{\prime}|^{3}}\right). (113)

The first term satisfies (2+k2)𝒈k(𝒙)=0\left(\bm{\nabla}^{2}+k^{2}\right)\bm{g}_{k}(\bm{x})=0. It gives 𝒈k(𝒙)=A0ei𝒌(𝒙𝒙0)\bm{g}_{k}(\bm{x})=A_{0}e^{i\bm{k}\cdot(\bm{x}-\bm{x}_{0})} with |𝒌|=ω/c|\bm{k}|=\omega/c. Here, A0A_{0} and 𝒙0\bm{x}_{0} are constant. However, this term should be zero because of the boundary condition at |𝒙||\bm{x}|\to\infty. Then, we obtain

𝒈Aϕ(𝒙,𝒙;ω)=ik(4π)2𝑑𝒙′′eik|𝒙𝒙′′||𝒙𝒙′′|𝒙′′𝒙|𝒙′′𝒙|3.\bm{g}_{A\phi}(\bm{x},\bm{x}^{\prime};\omega)=\frac{ik}{(4\pi)^{2}}\int d\bm{x}^{\prime\prime}\frac{e^{ik|\bm{x}-\bm{x}^{\prime\prime}|}}{|\bm{x}-\bm{x}^{\prime\prime}|}\frac{\bm{x}^{\prime\prime}-\bm{x}^{\prime}}{|\bm{x}^{\prime\prime}-\bm{x}^{\prime}|^{3}}. (114)

Next, we consider g¯AA\bar{g}_{AA} and 𝒈ϕAt\bm{g}_{\phi A}^{\ {\rm t}}. From Eq. (103), 𝒈ϕAt\bm{g}_{\phi A}^{\ {\rm t}} satisfies

2𝒈ϕAt(𝒙,𝒙;ω)=0.\bm{\nabla}^{2}\bm{g}_{\phi A}^{\ {\rm t}}(\bm{x},\bm{x}^{\prime};\omega)=0. (115)

Note that 𝒈ϕAt\bm{g}_{\phi A}^{\ {\rm t}} is a 1×31\times 3 row vector. The solution is 𝒈ϕAt(𝒙,𝒙;ω)={β¯1𝒙+𝜷0}t\bm{g}_{\phi A}^{\ {\rm t}}(\bm{x},\bm{x}^{\prime};\omega)=\left\{\bar{\beta}_{1}\bm{x}+\bm{\beta}_{0}\right\}^{\rm t}, where β¯1\bar{\beta}_{1} and 𝜷0\bm{\beta}_{0} are a constant matrix and vector, respectively. For the boundary condition, 𝜷0=0\bm{\beta}_{0}=0 and β¯1=0\bar{\beta}_{1}=0, and

𝒈ϕAt(𝒙,𝒙;ω)=0.\bm{g}_{\phi A}^{\ {\rm t}}(\bm{x},\bm{x}^{\prime};\omega)=0. (116)

The last element of the Green’s function, g¯AA\bar{g}_{AA}, satisfies

(2+k2)g¯AA(𝒙,𝒙;ω)=1¯δ(𝒙𝒙),\left(\bm{\nabla}^{2}+k^{2}\right)\bar{g}_{AA}(\bm{x},\bm{x}^{\prime};\omega)=\bar{1}\delta(\bm{x}-\bm{x}^{\prime}), (117)

which is equivalent to Eq. (107), and the solution is

g¯AA(𝒙,𝒙;ω)=1¯gk(𝒙,𝒙)=eik|𝒙𝒙|4π|𝒙𝒙|1¯.\bar{g}_{AA}(\bm{x},\bm{x}^{\prime};\omega)=\bar{1}g_{k}(\bm{x},\bm{x}^{\prime})=-\frac{e^{ik|\bm{x}-\bm{x}^{\prime}|}}{4\pi|\bm{x}-\bm{x}^{\prime}|}\bar{1}. (118)

Finally, Eqs. (110), (116), (114), and (118) give the Green’s function in matrix form:

𝒢¯(𝒙,𝒙;ω)=14π(eik|𝒙𝒙||𝒙𝒙|1¯ik4π𝑑𝒙′′eik|𝒙𝒙′′||𝒙𝒙′′|𝒙′′𝒙|𝒙′′𝒙|301|𝒙𝒙|).\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)=-\frac{1}{4\pi}\left(\begin{matrix}\frac{\ e^{ik|\bm{x}-\bm{x}^{\prime}|}}{|\bm{x}-\bm{x}^{\prime}|}\bar{1}\ &\frac{-ik}{4\pi}\int d\bm{x}^{\prime\prime}\frac{e^{ik|\bm{x}-\bm{x}^{\prime\prime}|}}{|\bm{x}-\bm{x}^{\prime\prime}|}\frac{\bm{x}^{\prime\prime}-\bm{x}^{\prime}}{|\bm{x}^{\prime\prime}-\bm{x}^{\prime}|^{3}}\\ 0&{-\frac{1}{|\bm{x}-\bm{x}^{\prime}|}}\end{matrix}\right). (119)

This Green’s function provides a picture of the field (potential) generation due to the electronic excitations (current and charge densities).

The Green’s function in the 𝒌\bm{k}-space representation (69) is obtained via a conventional Fourier transformation of Eq. (39). By applying

𝓞(𝒙;ω)=1(2π)32𝑑𝒌𝓞~(𝒌;ω)ei𝒌𝒙\bm{\mathcal{O}}(\bm{x};\omega)=\frac{1}{(2\pi)^{\frac{3}{2}}}\int d\bm{k}\tilde{\bm{\mathcal{O}}}(\bm{k};\omega)e^{i\bm{k}\cdot\bm{x}} (120)

for 𝓞=𝓐,𝓙\bm{\mathcal{O}}=\bm{\mathcal{A}},\bm{\mathcal{J}}, we find

𝒟¯(𝒙;ω){1(2π)32𝑑𝒌𝓐~(𝒌;ω)ei𝒌𝒙}\displaystyle\bar{\mathcal{D}}(\bm{x};\omega)\left\{\frac{1}{(2\pi)^{\frac{3}{2}}}\int d\bm{k}\tilde{\bm{\mathcal{A}}}(\bm{k};\omega)e^{i\bm{k}\cdot\bm{x}}\right\} =\displaystyle= 1(2π)32𝑑𝒌((𝒌2+ω2c2)1¯ωc𝒌0𝒌2)𝓐~(𝒌;ω)ei𝒌𝒙\displaystyle\frac{1}{(2\pi)^{\frac{3}{2}}}\int d\bm{k}\left(\begin{matrix}\left(-\bm{k}^{2}+\frac{\omega^{2}}{c^{2}}\right)\bar{1}&{\frac{\omega}{c}\bm{k}}\\ 0&{\bm{k}^{2}}\end{matrix}\right)\tilde{\bm{\mathcal{A}}}(\bm{k};\omega)e^{i\bm{k}\cdot\bm{x}} (121)
=\displaystyle= 1(2π)32𝑑𝒌(μ0)𝓙~(𝒌;ω)ei𝒌𝒙.\displaystyle\frac{1}{(2\pi)^{\frac{3}{2}}}\int d\bm{k}(-\mu_{0})\tilde{\bm{\mathcal{J}}}(\bm{k};\omega)e^{i\bm{k}\cdot\bm{x}}.

Hence, the Green’s function is

𝒢¯𝒌(ω)=((𝒌2+ω2c2)1¯ωc𝒌0𝒌2)1=(1𝒌2+ω2/c21¯1𝒌21𝒌2+ω2/c2ωc𝒌01𝒌2).\bar{\mathcal{G}}_{\bm{k}}(\omega)=\left(\begin{matrix}\left(-\bm{k}^{2}+\frac{\omega^{2}}{c^{2}}\right)\bar{1}&{\frac{\omega}{c}\bm{k}}\\ 0&{\bm{k}^{2}}\end{matrix}\right)^{-1}=\left(\begin{matrix}\frac{1}{-\bm{k}^{2}+\omega^{2}/c^{2}}\bar{1}&-\frac{1}{\bm{k}^{2}}\frac{1}{-\bm{k}^{2}+\omega^{2}/c^{2}}\frac{\omega}{c}\bm{k}\\ 0&{\frac{1}{\bm{k}^{2}}}\end{matrix}\right). (122)

The formal solutions of 𝓐=(𝑨,ϕ/c)\bm{\mathcal{A}}=(\bm{A},-\phi/c) can also be obtained from the separated Maxwell’s equations in Eqs. (36) and (37). The latter equation gives

ϕ(𝒙;ω)=ϕ0(𝒙;ω)+1ε0𝑑𝒙gϕ(𝒙,𝒙)ρ(𝒙;ω).\phi(\bm{x};\omega)=\phi_{0}(\bm{x};\omega)+\frac{1}{\varepsilon_{0}}\int d\bm{x}^{\prime}g_{\phi}(\bm{x},\bm{x}^{\prime})\rho(\bm{x}^{\prime};\omega). (123)

The scalar Green’s function gϕ(𝒙,𝒙)g_{\phi}(\bm{x},\bm{x}^{\prime}) is equivalent to gϕϕ(𝒙,𝒙)g_{\phi\phi}(\bm{x},\bm{x}^{\prime}) in Eq. (110). The first term satisfies 2ϕ0(𝒙;ω)=0\bm{\nabla}^{2}\phi_{0}(\bm{x};\omega)=0. In the Coulomb gauge, however, the T field is described by 𝑨(𝒙;ω)\bm{A}(\bm{x};\omega) only and, hence, ϕ0=0\phi_{0}=0. The solution of Eq. (36) is

𝑨(𝒙;ω)=𝑨0(𝒙;ω)μ0𝑑𝒙G¯𝑨(𝒙,𝒙;ω)𝒋(𝒙;ω)\bm{A}(\bm{x};\omega)=\bm{A}_{0}(\bm{x};\omega)-\mu_{0}\int d\bm{x}^{\prime}\bar{G}_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\bm{j}(\bm{x}^{\prime};\omega) (124)

with (2+ω2c2)𝑨0(𝒙;ω)=0\left(\bm{\nabla}^{2}+\frac{\omega^{2}}{c^{2}}\right)\bm{A}_{0}(\bm{x};\omega)=0. Here, the dyadic Green’s function G¯𝑨\bar{G}_{\bm{A}} for the vector potential is expressed as

G¯𝑨(𝒙,𝒙;ω)\displaystyle\bar{G}_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega) =\displaystyle= g𝑨(𝒙,𝒙;ω)1¯+d𝒙′′g𝑨(𝒙,𝒙′′;ω)𝒙′′gϕ(𝒙′′,𝒙)𝒙\displaystyle g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\bar{1}+\int d\bm{x}^{\prime\prime}g_{\bm{A}}(\bm{x},\bm{x}^{\prime\prime};\omega)\bm{\nabla}_{\bm{x}^{\prime\prime}}g_{\phi}(\bm{x}^{\prime\prime},\bm{x}^{\prime})\bm{\nabla}_{\bm{x}^{\prime}}\cdot (125)
=\displaystyle= g𝑨(𝒙,𝒙;ω)1¯+d𝒙′′[𝒙′′g𝑨(𝒙,𝒙′′;ω)][𝒙gϕ(𝒙′′,𝒙)]\displaystyle g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\bar{1}+\int d\bm{x}^{\prime\prime}\left[\bm{\nabla}_{\bm{x}^{\prime\prime}}g_{\bm{A}}(\bm{x},\bm{x}^{\prime\prime};\omega)\right]\left[\bm{\nabla}_{\bm{x}^{\prime}}g_{\phi}(\bm{x}^{\prime\prime},\bm{x}^{\prime})\right]\cdot

with

g𝑨(𝒙,𝒙;ω)=eiωc|𝒙𝒙|4π|𝒙𝒙|.g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)=-\frac{e^{i\frac{\omega}{c}|\bm{x}-\bm{x}^{\prime}|}}{4\pi|\bm{x}-\bm{x}^{\prime}|}. (126)

For the derivation, Eq. (123) is substituted and the continuous equation 𝒋(𝒙;ω)iωρ(𝒙;ω)=0\bm{\nabla}\cdot\bm{j}(\bm{x};\omega)-i\omega\rho(\bm{x};\omega)=0 is used, as the source term is described by 𝒋(𝒙;ω)\bm{j}(\bm{x};\omega) only in this representation. In the dyadic Green’s function G¯𝑨(𝒙,𝒙;ω)\bar{G}_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega) in Eq. (125), gϕg_{\phi} is included, which implies that the L component is related to the vector potential. This seems strange at first glance, as the vector potential in Eq. (124) must satisfy x𝑨=0\bm{\nabla}_{x}\cdot\bm{A}=0.

To see it, we separate the T and L components of the current density, 𝒋=𝒋(T)+𝒋(L)\bm{j}=\bm{j}^{\rm(T)}+\bm{j}^{\rm(L)}. This is a description in terms of {(𝑨(T),ϕ(L)/c),(𝒋(T),𝒋(L))}\{(\bm{A}^{\rm(T)},-\phi^{\rm(L)}/c),(\bm{j}^{\rm(T)},\bm{j}^{\rm(L)})\}, which is discussed in Appendix B. The formal solution for Eq. (99) is given by g𝑨(𝒙,𝒙;ω)g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega),

𝑨(T)(𝒙;ω)=𝑨0(T)(𝒙;ω)μ0𝑑𝒙g𝑨(𝒙,𝒙;ω)𝒋(T)(𝒙;ω).\bm{A}^{\rm(T)}(\bm{x};\omega)=\bm{A}_{0}^{\rm(T)}(\bm{x};\omega)-\mu_{0}\int d\bm{x}^{\prime}g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\bm{j}^{\rm(T)}(\bm{x}^{\prime};\omega). (127)

Note that Eq. (127) differs slightly from Eq. (124) in its second term on the r.h.s. For the scalar potential, we apply the continuous relation, iωρ=𝒋(L)i\omega\rho=\bm{\nabla}\cdot\bm{j}^{\rm(L)}, to Eq. (123),

ϕ(L)(𝒙;ω)\displaystyle\phi^{\rm(L)}(\bm{x};\omega) =\displaystyle= 1iωε0𝑑𝒙gϕ(𝒙,𝒙)x𝒋(L)(𝒙;ω).\displaystyle\frac{1}{i\omega\varepsilon_{0}}\int d\bm{x}^{\prime}g_{\phi}(\bm{x},\bm{x}^{\prime})\bm{\nabla}_{x^{\prime}}\cdot\bm{j}^{\rm(L)}(\bm{x}^{\prime};\omega). (128)

The set of Eqs. (127) and (128) give the solutions of the Maxwell’s equation for the L–T separation. For the source term in Eq. (127), we apply Eqs. (100), (128), and the continuous relation,

𝒋(T)(𝒙;ω)\displaystyle\bm{j}^{\rm(T)}(\bm{x}^{\prime};\omega) =\displaystyle= 𝒋(𝒙;ω)𝒋(L)(𝒙;ω)\displaystyle\bm{j}(\bm{x}^{\prime};\omega)-\bm{j}^{\rm(L)}(\bm{x}^{\prime};\omega) (129)
=\displaystyle= 𝒋(𝒙;ω)iωμ0c2xϕ(L)(𝒙;ω)\displaystyle\bm{j}(\bm{x}^{\prime};\omega)-\frac{-i\omega}{\mu_{0}c^{2}}\bm{\nabla}_{x^{\prime}}\phi^{\rm(L)}(\bm{x}^{\prime};\omega)
=\displaystyle= 𝒋(𝒙;ω)+x𝑑𝒙′′gϕ(𝒙,𝒙′′)x′′𝒋(L)(𝒙′′;ω)\displaystyle\bm{j}(\bm{x}^{\prime};\omega)+\bm{\nabla}_{x^{\prime}}\int d\bm{x}^{\prime\prime}g_{\phi}(\bm{x}^{\prime},\bm{x}^{\prime\prime})\bm{\nabla}_{x^{\prime\prime}}\cdot\bm{j}^{\rm(L)}(\bm{x}^{\prime\prime};\omega)
=\displaystyle= d𝒙′′{1¯δ(𝒙𝒙′′)+xgϕ(𝒙,𝒙′′)x′′}𝒋(𝒙′′;ω).\displaystyle\int d\bm{x}^{\prime\prime}\left\{\bar{1}\delta(\bm{x}^{\prime}-\bm{x}^{\prime\prime})+\bm{\nabla}_{x^{\prime}}g_{\phi}(\bm{x}^{\prime},\bm{x}^{\prime\prime})\bm{\nabla}_{x^{\prime\prime}}\cdot\right\}\bm{j}(\bm{x}^{\prime\prime};\omega).

Then, Eqs. (127) and (124) are equivalent. Note that the last term can be rewritten as a dyadic function,

xgϕ(𝒙,𝒙′′)x′′=(xx)gϕ(𝒙,𝒙′′).\bm{\nabla}_{x^{\prime}}g_{\phi}(\bm{x}^{\prime},\bm{x}^{\prime\prime})\bm{\nabla}_{x^{\prime\prime}}\cdot=\overleftrightarrow{\left(\bm{\nabla}_{x^{\prime}}\bm{\nabla}_{x^{\prime}}\cdot\right)}g_{\phi}(\bm{x}^{\prime},\bm{x}^{\prime\prime}). (130)

As g𝑨(𝒙,𝒙;ω)g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega) depends only on |𝒙𝒙||\bm{x}-\bm{x}^{\prime}|, we find

x𝑑𝒙g𝑨(𝒙,𝒙;ω)𝒋(T)(𝒙;ω),\displaystyle\bm{\nabla}_{x}\cdot\int d\bm{x}^{\prime}g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\bm{j}^{\rm(T)}(\bm{x}^{\prime};\omega), =\displaystyle= 𝑑𝒙(xg𝑨(𝒙,𝒙;ω))𝒋(T)(𝒙;ω),\displaystyle-\int d\bm{x}^{\prime}\left(\bm{\nabla}_{x^{\prime}}\cdot g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\right)\bm{j}^{\rm(T)}(\bm{x}^{\prime};\omega),
=\displaystyle= 𝑑𝒙g𝑨(𝒙,𝒙;ω)(x𝒋(T)(𝒙;ω))\displaystyle\int d\bm{x}^{\prime}g_{\bm{A}}(\bm{x},\bm{x}^{\prime};\omega)\left(\bm{\nabla}_{x^{\prime}}\cdot\bm{j}^{\rm(T)}(\bm{x}^{\prime};\omega)\right)
=\displaystyle= 0.\displaystyle 0.

This follows the transverse vector potential, x𝑨=0\bm{\nabla}_{x}\cdot\bm{A}=0. Therefore, the solutions (123) and (124) obtained from the separated Maxwell’s equations have no inquiries.

These treatments of the source term 𝒋\bm{j} and the L and T field components complicate the theoretical framework. In our four-vector representation, such complexity becomes clear as the two components (𝑨,ϕ/c)(\bm{A},-\phi/c) of the fields and the two source components (𝒋,cρ)(\bm{j},c\rho) are related to each other in matrix form.

Appendix D Derivation of the matrix form of self-consistent equation

In this appendix, we describe the detail derivation and formulation of the matrix form of self-consistent equation from Eq. (44),

𝓐(𝒙;ω)\displaystyle\bm{\mathcal{A}}(\bm{x};\omega) =\displaystyle= 𝓐0(𝒙;ω)μ0𝑑𝒙𝒢¯(𝒙,𝒙;ω)𝓙0(𝒙;ω)\displaystyle\bm{\mathcal{A}}_{0}(\bm{x};\omega)-\mu_{0}\int d\bm{x}^{\prime}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bm{\mathcal{J}}_{0}(\bm{x}^{\prime};\omega) (131)
μ0𝑑𝒙𝑑𝒙′′𝒢¯(𝒙,𝒙;ω)𝒳¯(𝒙,𝒙′′;ω)𝓐(𝒙′′;ω).\displaystyle\hskip 56.9055pt-\mu_{0}\int d\bm{x}^{\prime}\int d\bm{x}^{\prime\prime}\bar{\mathcal{G}}(\bm{x},\bm{x}^{\prime};\omega)\bar{\mathcal{X}}(\bm{x}^{\prime},\bm{x}^{\prime\prime};\omega)\bm{\mathcal{A}}(\bm{x}^{\prime\prime};\omega).

For the third term in the r.h.s., one substitutes the nonlocal susceptibility 𝒳¯(𝒙,𝒙′′;ω)\bar{\mathcal{X}}(\bm{x}^{\prime},\bm{x}^{\prime\prime};\omega) in Eq. (28), and for the second term, one uses Eq. (45). By multiplying with (𝓙νμ(𝒙))t\left(\bm{\mathcal{J}}_{\nu^{\prime}\mu^{\prime}}(\bm{x})\right)^{\rm t}, (𝓙μν(𝒙))t\left(\bm{\mathcal{J}}_{\mu^{\prime}\nu^{\prime}}(\bm{x})\right)^{\rm t}, and (φi(𝒙)𝒆βt)\left(\varphi_{i}(\bm{x})\bm{e}_{\beta}^{\rm t}\right) from the left and integrating by 𝒙\bm{x}, one obtains

(ωνμωiγ)Xνμ()\displaystyle\left(\hbar\omega_{\nu^{\prime}\mu^{\prime}}-\hbar\omega-i\gamma\right)X_{\nu^{\prime}\mu^{\prime}}^{(-)} =\displaystyle= Yνμ(0)+j,αUνμ,jαXjα(A)μ,ν[Kνμ,μνXνμ()+Lνμ,νμXμν(+)],\displaystyle Y_{\nu^{\prime}\mu^{\prime}}^{(0)}+\sum_{j,\alpha}U_{\nu^{\prime}\mu^{\prime},j\alpha}X_{j\alpha}^{(A)}-\sum_{\mu,\nu}\left[K_{\nu^{\prime}\mu^{\prime},\mu\nu}X_{\nu\mu}^{(-)}+L_{\nu^{\prime}\mu^{\prime},\nu\mu}X_{\mu\nu}^{(+)}\right], (132)
(ωνμ+ω+iγ)Xμν(+)\displaystyle\left(\hbar\omega_{\nu^{\prime}\mu^{\prime}}+\hbar\omega+i\gamma\right)X_{\mu^{\prime}\nu^{\prime}}^{(+)} =\displaystyle= Yμν(0)+j,αVμν,jαXjα(A)μ,ν[Mμν,μνXνμ()+Nμν,νμXμν(+)],\displaystyle Y_{\mu^{\prime}\nu^{\prime}}^{(0)}+\sum_{j,\alpha}V_{\mu^{\prime}\nu^{\prime},j\alpha}X_{j\alpha}^{(A)}-\sum_{\mu,\nu}\left[M_{\mu^{\prime}\nu^{\prime},\mu\nu}X_{\nu\mu}^{(-)}+N_{\mu^{\prime}\nu^{\prime},\nu\mu}X_{\mu\nu}^{(+)}\right], (133)
Xiβ(A)\displaystyle X_{i\beta}^{(A)} =\displaystyle= Yiβ(A)+j,αRiβ,jαXjα(A)μ,ν[Siβ,μνXνμ()+Tiβ,νμXμν(+)].\displaystyle Y_{i\beta}^{(A)}+\sum_{j,\alpha}R_{i\beta,j\alpha}X_{j\alpha}^{(A)}-\sum_{\mu,\nu}\left[S_{i\beta,\mu\nu}X_{\nu\mu}^{(-)}+T_{i\beta,\nu\mu}X_{\mu\nu}^{(+)}\right]. (134)

Here, the factors are defined in Eqs. (53)–(66) in Sec. V.1. Note again that the excited states μ\mu indicating an electron–hole pair include their spin degrees of freedom. α,β\alpha,\beta takes only x,y,zx,y,z not ϕ\phi. The factors, Kνμ,μνK_{\nu^{\prime}\mu^{\prime},\mu\nu}, Lνμ,νμL_{\nu^{\prime}\mu^{\prime},\nu\mu}, Mμν,μνM_{\mu^{\prime}\nu^{\prime},\mu\nu}, and Nμν,νμN_{\mu^{\prime}\nu^{\prime},\nu\mu} are mathematically related. The matrix form of Eqs. (132)-(134) is written as

[(Ω¯(ω+iγ)1¯)+K¯]𝑿()+L¯𝑿(+)\displaystyle\left[\left(\hbar\bar{\Omega}-(\hbar\omega+i\gamma)\bar{1}\right)+\bar{K}\right]\bm{X}^{(-)}+\bar{L}\bm{X}^{(+)} =\displaystyle= 𝒀(0,)+U¯𝑿(A),\displaystyle\bm{Y}^{(0,-)}+\bar{U}\bm{X}^{(A)}, (135)
[(Ω¯+(ω+iγ)1¯)+N¯]𝑿(+)+M¯𝑿()\displaystyle\left[\left(\hbar\bar{\Omega}+(\hbar\omega+i\gamma)\bar{1}\right)+\bar{N}\right]\bm{X}^{(+)}+\bar{M}\bm{X}^{(-)} =\displaystyle= 𝒀(0,+)+V¯𝑿(A),\displaystyle\bm{Y}^{(0,+)}+\bar{V}\bm{X}^{(A)}, (136)
[1¯R¯]𝑿(A)\displaystyle\left[\bar{1}-\bar{R}\right]\bm{X}^{(A)} =\displaystyle= 𝒀(A)[S¯𝑿()+T¯𝑿(+)].\displaystyle\bm{Y}^{(A)}-\left[\bar{S}\bm{X}^{(-)}+\bar{T}\bm{X}^{(+)}\right]. (137)

K¯\bar{K} and N¯\bar{N} indicate the resonant and anti-resonant radiative correlations, respectively, and L¯\bar{L} and M¯\bar{M} describes their couplings. Here, one introduces an integer NN being a cut-off number of the states, and defines the vectors of Xνμ(±)X_{\nu\mu}^{(\pm)} for the fields as

𝑿(±)=(𝑿0(±)𝑿1(±)𝑿N(±))with𝑿μ()=(X0μ()X1μ()XNμ()),𝑿μ(+)=(Xμ0(+)Xμ1(+)XμN(+)).\bm{X}^{(\pm)}=\left(\begin{matrix}\bm{X}^{(\pm)}_{0}\\ \bm{X}^{(\pm)}_{1}\\ \vdots\\ \bm{X}^{(\pm)}_{N}\end{matrix}\right)\ \ \ {\rm with}\ \ \ \bm{X}^{(-)}_{\mu}=\left(\begin{matrix}X^{(-)}_{0\mu}\\ X^{(-)}_{1\mu}\\ \vdots\\ X^{(-)}_{N\mu}\end{matrix}\right)\ \ ,\ \ \bm{X}^{(+)}_{\mu}=\left(\begin{matrix}X^{(+)}_{\mu 0}\\ X^{(+)}_{\mu 1}\\ \vdots\\ X^{(+)}_{\mu N}\end{matrix}\right). (138)

For the incident field Yνμ(0)Y^{(0)}_{\nu\mu},

𝒀(0,±)=(𝒀0(0,±)𝒀1(0,±)𝒀N(0,±))with𝒀μ(0,)=(Y0μ(0)Y1μ(0)YNμ(0)),𝒀μ(0,+)=(Yμ0(0)Yμ1(0)YμN(0)).\bm{Y}^{(0,\pm)}=\left(\begin{matrix}\bm{Y}^{(0,\pm)}_{0}\\ \bm{Y}^{(0,\pm)}_{1}\\ \vdots\\ \bm{Y}^{(0,\pm)}_{N}\end{matrix}\right)\ \ \ {\rm with}\ \ \ \bm{Y}^{(0,-)}_{\mu}=\left(\begin{matrix}Y^{(0)}_{0\mu}\\ Y^{(0)}_{1\mu}\\ \vdots\\ Y^{(0)}_{N\mu}\end{matrix}\right)\ \ ,\ \ \bm{Y}^{(0,+)}_{\mu}=\left(\begin{matrix}Y^{(0)}_{\mu 0}\\ Y^{(0)}_{\mu 1}\\ \vdots\\ Y^{(0)}_{\mu N}\end{matrix}\right). (139)

Here, 𝒀(0,)\bm{Y}^{(0,-)} and 𝒀(0,+)\bm{Y}^{(0,+)} correspond to each other by a permutation of the elements. For the vector potential term in the current Xjα(A)X_{j\alpha}^{(A)},

𝑿(A)=(𝑿0(A)𝑿1(A)𝑿M(A))with𝑿j(A)=(Xjx(A)Xjy(A)Xjz(A)).\bm{X}^{(A)}=\left(\begin{matrix}\bm{X}^{(A)}_{0}\\ \bm{X}^{(A)}_{1}\\ \vdots\\ \bm{X}^{(A)}_{M}\end{matrix}\right)\ \ \ {\rm with}\ \ \ \bm{X}^{(A)}_{j}=\left(\begin{matrix}X^{(A)}_{jx}\\ X^{(A)}_{jy}\\ X^{(A)}_{jz}\end{matrix}\right). (140)

For the incident field Yjα(A)Y_{j\alpha}^{(A)} ,

𝒀(A)=(𝒀0(A)𝒀1(A)𝒀M(A))with𝒀j(A)=(Yjx(A)Yjy(A)Yjz(A)).\bm{Y}^{(A)}=\left(\begin{matrix}\bm{Y}^{(A)}_{0}\\ \bm{Y}^{(A)}_{1}\\ \vdots\\ \bm{Y}^{(A)}_{M}\end{matrix}\right)\ \ \ {\rm with}\ \ \ \bm{Y}^{(A)}_{j}=\left(\begin{matrix}Y^{(A)}_{jx}\\ Y^{(A)}_{jy}\\ Y^{(A)}_{jz}\end{matrix}\right). (141)

Next, the matrices in Eqs. (135)-(137) are defined as follows. For the energy differences ωνμ\omega_{\nu\mu},

Ω¯=(Ω¯0Ω¯1Ω¯N)\bar{\Omega}=\left(\begin{matrix}\bar{\Omega}_{0}&&&\\ &\bar{\Omega}_{1}&&\\ &&\ddots&\\ &&&\bar{\Omega}_{N}\end{matrix}\right) (142)

is a diagonal matrix with Ω¯μ=diag(ω0μ,ω1μ,ω2μ,ωNμ)\bar{\Omega}_{\mu}={\rm diag}(\omega_{0\mu},\omega_{1\mu},\omega_{2\mu}\cdots,\omega_{N\mu}). For the matrices K¯\bar{K}, L¯\bar{L}, M¯\bar{M}, and N¯\bar{N},

Q¯=(Q¯00Q¯01Q¯0NQ¯10Q¯11Q¯N0Q¯NN)(Q=K,L,M,N)\bar{Q}=\left(\begin{matrix}\bar{Q}_{00}&\bar{Q}_{01}&\cdots&\bar{Q}_{0N}\\ \bar{Q}_{10}&\bar{Q}_{11}&&\\ \vdots&&\ddots&\\ \bar{Q}_{N0}&&&\bar{Q}_{NN}\end{matrix}\right)\ \ (Q=K,L,M,N) (143)

with different subscript rules for (N+1)×(N+1)(N+1)\times(N+1) block matrices

K¯μμ\displaystyle\bar{K}_{\mu^{\prime}\mu} =\displaystyle= (K0μ,μ0K0μ,μ1K0μ,μNK1μ,μ0K1μ,μ1KNμ,μ0KNμ,μN),\displaystyle\left(\begin{matrix}K_{0\mu^{\prime},\mu 0}&K_{0\mu^{\prime},\mu 1}&\cdots&K_{0\mu^{\prime},\mu N}\\ K_{1\mu^{\prime},\mu 0}&K_{1\mu^{\prime},\mu 1}&&\\ \vdots&&\ddots&\\ K_{N\mu^{\prime},\mu 0}&&&K_{N\mu^{\prime},\mu N}\end{matrix}\right), (144)
L¯μμ\displaystyle\bar{L}_{\mu^{\prime}\mu} =\displaystyle= (L0μ,0μL0μ,1μL0μ,NμL1μ,0μL1μ,1μLNμ,0μLNμ,Nμ),\displaystyle\left(\begin{matrix}L_{0\mu^{\prime},0\mu}&L_{0\mu^{\prime},1\mu}&\cdots&L_{0\mu^{\prime},N\mu}\\ L_{1\mu^{\prime},0\mu}&L_{1\mu^{\prime},1\mu}&&\\ \vdots&&\ddots&\\ L_{N\mu^{\prime},0\mu}&&&L_{N\mu^{\prime},N\mu}\end{matrix}\right), (145)
M¯μμ\displaystyle\bar{M}_{\mu^{\prime}\mu} =\displaystyle= (Mμ0,μ0Mμ0,μ1Mμ0,μNMμ1,μ0Mμ1,μ1MμN,μ0MμN,μN),\displaystyle\left(\begin{matrix}M_{\mu^{\prime}0,\mu 0}&M_{\mu^{\prime}0,\mu 1}&\cdots&M_{\mu^{\prime}0,\mu N}\\ M_{\mu^{\prime}1,\mu 0}&M_{\mu^{\prime}1,\mu 1}&&\\ \vdots&&\ddots&\\ M_{\mu^{\prime}N,\mu 0}&&&M_{\mu^{\prime}N,\mu N}\end{matrix}\right), (146)
N¯μμ\displaystyle\bar{N}_{\mu^{\prime}\mu} =\displaystyle= (Nμ0,0μNμ0,1μNμ0,NμNμ1,0μNμ1,1μNμN,0μNμN,Nμ).\displaystyle\left(\begin{matrix}N_{\mu^{\prime}0,0\mu}&N_{\mu^{\prime}0,1\mu}&\cdots&N_{\mu^{\prime}0,N\mu}\\ N_{\mu^{\prime}1,0\mu}&N_{\mu^{\prime}1,1\mu}&&\\ \vdots&&\ddots&\\ N_{\mu^{\prime}N,0\mu}&&&N_{\mu^{\prime}N,N\mu}\end{matrix}\right). (147)

For the matrix R¯\bar{R} related to the vector potential term in the current,

R¯=(R¯11R¯12R¯1MR¯21R¯22R¯M1R¯MM)\bar{R}=\left(\begin{matrix}\bar{R}_{11}&\bar{R}_{12}&\cdots&\bar{R}_{1M}\\ \bar{R}_{21}&\bar{R}_{22}&&\\ \vdots&&\ddots&\\ \bar{R}_{M1}&&&\bar{R}_{MM}\end{matrix}\right) (148)

with 3×33\times 3 block matrices

R¯ij=(Rix,jxRix,jyRix,jzRiy,jxRiy,jyRiy,jzRiz,jxRiz,jyRiz,jz).\bar{R}_{ij}=\left(\begin{matrix}R_{ix,jx}&R_{ix,jy}&R_{ix,jz}\\ R_{iy,jx}&R_{iy,jy}&R_{iy,jz}\\ R_{iz,jx}&R_{iz,jy}&R_{iz,jz}\end{matrix}\right). (149)

For the off-diagonal matrices U¯\bar{U} and V¯\bar{V},

Q¯=(Q¯01Q¯02Q¯0MQ¯11Q¯12Q¯N1Q¯NM)(Q=U,V)\bar{Q}=\left(\begin{matrix}\bar{Q}_{01}&\bar{Q}_{02}&\cdots&\bar{Q}_{0M}\\ \bar{Q}_{11}&\bar{Q}_{12}&&\\ \vdots&&\ddots&\\ \bar{Q}_{N1}&&&\bar{Q}_{NM}\end{matrix}\right)\ \ (Q=U,V) (150)

with (N+1)×3(N+1)\times 3 block matrices

U¯μj\displaystyle\bar{U}_{\mu^{\prime}j} =\displaystyle= (U0μ,jxU0μ,jyU0μ,jzU1μ,jxU1μ,jyU1μ,jzUNμ,jxUNμ,jyUNμ,jz),\displaystyle\left(\begin{matrix}U_{0\mu^{\prime},jx}&U_{0\mu^{\prime},jy}&U_{0\mu^{\prime},jz}\\ U_{1\mu^{\prime},jx}&U_{1\mu^{\prime},jy}&U_{1\mu^{\prime},jz}\\ &\vdots&\\ U_{N\mu^{\prime},jx}&U_{N\mu^{\prime},jy}&U_{N\mu^{\prime},jz}\end{matrix}\right), (151)
V¯μj\displaystyle\bar{V}_{\mu^{\prime}j} =\displaystyle= (Vμ0,jxVμ0,jyVμ0,jzVμ1,jxVμ1,jyVμ1,jzVμN,jxVμN,jyVμN,jz)\displaystyle\left(\begin{matrix}V_{\mu^{\prime}0,jx}&V_{\mu^{\prime}0,jy}&V_{\mu^{\prime}0,jz}\\ V_{\mu^{\prime}1,jx}&V_{\mu^{\prime}1,jy}&V_{\mu^{\prime}1,jz}\\ &\vdots&\\ V_{\mu^{\prime}N,jx}&V_{\mu^{\prime}N,jy}&V_{\mu^{\prime}N,jz}\end{matrix}\right) (152)

and for S¯\bar{S} and T¯\bar{T},

Q¯=(Q¯10Q¯11Q¯1NQ¯20Q¯21Q¯M0Q¯MN)(Q=S,T)\bar{Q}=\left(\begin{matrix}\bar{Q}_{10}&\bar{Q}_{11}&\cdots&\bar{Q}_{1N}\\ \bar{Q}_{20}&\bar{Q}_{21}&&\\ \vdots&&\ddots&\\ \bar{Q}_{M0}&&&\bar{Q}_{MN}\end{matrix}\right)\ \ (Q=S,T) (153)

with 3×(N+1)3\times(N+1) block matrices

S¯iμ\displaystyle\bar{S}_{i\mu} =\displaystyle= (Six,μ0Six,μ1Six,μNSiy,μ0Siy,μ1Siy,μNSiz,μ0Siz,μ1Siz,μN),\displaystyle\left(\begin{matrix}S_{ix,\mu 0}&S_{ix,\mu 1}&&S_{ix,\mu N}\\ S_{iy,\mu 0}&S_{iy,\mu 1}&\cdots&S_{iy,\mu N}\\ S_{iz,\mu 0}&S_{iz,\mu 1}&&S_{iz,\mu N}\end{matrix}\right), (154)
T¯iμ\displaystyle\bar{T}_{i\mu} =\displaystyle= (Tix,0μTix,1μTix,NμTiy,0μTiy,1μTiy,NμTiz,0μTiz,1μTiz,Nμ).\displaystyle\left(\begin{matrix}T_{ix,0\mu}&T_{ix,1\mu}&&T_{ix,N\mu}\\ T_{iy,0\mu}&T_{iy,1\mu}&\cdots&T_{iy,N\mu}\\ T_{iz,0\mu}&T_{iz,1\mu}&&T_{iz,N\mu}\end{matrix}\right). (155)

For the many electrons state (μ,ν=0,1,2,3,,N\mu,\nu=0,1,2,3,\cdots,N), the size of K¯\bar{K}, L¯\bar{L}, M¯\bar{M}, and N¯\bar{N} is (N+1)2×(N+1)2(N+1)^{2}\times(N+1)^{2}. In case of zero temperature, the consideration can be reduced due to ρ0,μ=δμ,0\rho_{0,\mu}=\delta_{\mu,0}; hence, the matrix size is (N+1)×(N+1)(N+1)\times(N+1). For the separable kernel of delta function, contrarily, if one considers MM bases (i,j=1,2,3,,Mi,j=1,2,3,\cdots,M), R¯\bar{R} is 3M×3M3M\times 3M. Here, the factor 33 comes from α=x,y,z\alpha=x,y,z. In addition, U¯\bar{U} and V¯\bar{V} are (N+1)2×3M(N+1)^{2}\times 3M (or (N+1)×3M(N+1)\times 3M), and S¯\bar{S} and T¯\bar{T} are 3M×(N+1)23M\times(N+1)^{2} (or 3M×(N+1)3M\times(N+1)). Note that the physical dimension of the matrices is not identical. Hence, one has to modify the equations (135)-(137).

Equation (137) is deformed as

𝑿(A)=11¯R¯[𝒀(A)S¯𝑿()T¯𝑿(+)].\bm{X}^{(A)}=\frac{1}{\bar{1}-\bar{R}}\left[\bm{Y}^{(A)}-\bar{S}\bm{X}^{(-)}-\bar{T}\bm{X}^{(+)}\right]. (156)

A substitution to Eqs. (135) and (136) gives

[(Ω¯(ω+iγ)1¯)+K¯+U¯11¯R¯S¯]𝑿()+[L¯+U¯11¯R¯T¯]𝑿(+)\displaystyle\left[\left(\hbar\bar{\Omega}-(\hbar\omega+i\gamma)\bar{1}\right)+\bar{K}+\bar{U}\frac{1}{\bar{1}-\bar{R}}\bar{S}\right]\bm{X}^{(-)}+\left[\bar{L}+\bar{U}\frac{1}{\bar{1}-\bar{R}}\bar{T}\right]\bm{X}^{(+)} =\displaystyle= 𝒀(0,)+U¯11¯R¯𝒀(A),\displaystyle\bm{Y}^{(0,-)}+\bar{U}\frac{1}{\bar{1}-\bar{R}}\bm{Y}^{(A)}, (157)
[M¯+V¯11¯R¯S¯]𝑿()+[(Ω¯+(ω+iγ)1¯)+N¯+V¯11¯R¯T¯]𝑿(+)\displaystyle\left[\bar{M}+\bar{V}\frac{1}{\bar{1}-\bar{R}}\bar{S}\right]\bm{X}^{(-)}+\left[\left(\hbar\bar{\Omega}+(\hbar\omega+i\gamma)\bar{1}\right)+\bar{N}+\bar{V}\frac{1}{\bar{1}-\bar{R}}\bar{T}\right]\bm{X}^{(+)} =\displaystyle= 𝒀(0,+)+V¯11¯R¯𝒀(A).\displaystyle\bm{Y}^{(0,+)}+\bar{V}\frac{1}{\bar{1}-\bar{R}}\bm{Y}^{(A)}. (158)

As a result, one obtains Eq. (46) with Eq. (47) and (48):

[Ξ¯(ω)](𝑿()𝑿(+))=(𝒁(0,)𝒁(0,+))\left[\bar{\Xi}(\omega)\right]\left(\begin{matrix}\bm{X}^{(-)}\\ \bm{X}^{(+)}\end{matrix}\right)=\left(\begin{matrix}\bm{Z}^{(0,-)}\\ \bm{Z}^{(0,+)}\end{matrix}\right) (159)

with

Ξ¯=[(Ω¯(ω+iγ)1¯Ω¯+(ω+iγ)1¯)+(K¯L¯M¯N¯)+(U¯11¯R¯S¯U¯11¯R¯T¯V¯11¯R¯S¯V¯11¯R¯T¯)]\bar{\Xi}=\left[\left(\begin{matrix}\hbar\bar{\Omega}-(\hbar\omega+i\gamma)\bar{1}&\\ &\hbar\bar{\Omega}+(\hbar\omega+i\gamma)\bar{1}\\ \end{matrix}\right)+\left(\begin{matrix}\bar{K}&\bar{L}\\ \bar{M}&\bar{N}\end{matrix}\right)+\left(\begin{matrix}\bar{U}\frac{1}{\bar{1}-\bar{R}}\bar{S}&\bar{U}\frac{1}{\bar{1}-\bar{R}}\bar{T}\\ \bar{V}\frac{1}{\bar{1}-\bar{R}}\bar{S}&\bar{V}\frac{1}{\bar{1}-\bar{R}}\bar{T}\end{matrix}\right)\right] (160)

and

(𝒁(0,)𝒁(0,+))=(𝒀(0,)𝒀(0,+))+(U¯11¯R¯𝒀(A)V¯11¯R¯𝒀(A)).\left(\begin{matrix}\bm{Z}^{(0,-)}\\ \bm{Z}^{(0,+)}\end{matrix}\right)=\left(\begin{matrix}\bm{Y}^{(0,-)}\\ \bm{Y}^{(0,+)}\end{matrix}\right)+\left(\begin{matrix}\bar{U}\frac{1}{\bar{1}-\bar{R}}\bm{Y}^{(A)}\\ \bar{V}\frac{1}{\bar{1}-\bar{R}}\bm{Y}^{(A)}\end{matrix}\right). (161)

Appendix E Radiative correction for rectangular nanorod

In this appendix, we summarize the detailed calculations for the rectangular nanorod discussed in Sec. VI.1.

Single particle wavefunction for the electron and hole are given by Eqs. (81) and (82).

ψeμ(𝒙)\displaystyle\psi_{{\rm e}\mu}(\bm{x}) =\displaystyle= 2Lxsin(nxπLxx)2Lysin(nyπLyy)2Lzsin(nzπLzz),\displaystyle\sqrt{\frac{2}{L_{x}}}\sin\left(\frac{n_{x}\pi}{L_{x}}x\right)\sqrt{\frac{2}{L_{y}}}\sin\left(\frac{n_{y}\pi}{L_{y}}y\right)\sqrt{\frac{2}{L_{z}}}\sin\left(\frac{n_{z}\pi}{L_{z}}z\right), (162)
ψhμ¯(𝒙)\displaystyle\psi_{{\rm h}\bar{\mu}}(\bm{x}) =\displaystyle= 2Lxsin(n¯xπLxx)2Lysin(n¯yπLyy)2Lzsin(n¯zπLzz).\displaystyle\sqrt{\frac{2}{L_{x}}}\sin\left(\frac{\bar{n}_{x}\pi}{L_{x}}x\right)\sqrt{\frac{2}{L_{y}}}\sin\left(\frac{\bar{n}_{y}\pi}{L_{y}}y\right)\sqrt{\frac{2}{L_{z}}}\sin\left(\frac{\bar{n}_{z}\pi}{L_{z}}z\right). (163)

For these wavefunctions, the excited charge and current densities at zero temperature, ρ0μ(𝒙)\rho_{0\mu}(\bm{x}) and 𝒋0μ(𝒙)\bm{j}_{0\mu}(\bm{x}), are obtained by Eqs. (83) and (84) with μ=(eμ,hμ¯)\mu=(\rm{e}\mu,\rm{h}\bar{\mu}):

(c/nbg)ρ0μ(𝒙)\displaystyle(c/n_{\rm bg})\rho_{0\mu}(\bm{x}) =\displaystyle= (c/nbg)eLxLyLz{cos((qn¯xqnx)x)cos((qn¯x+qnx)x)}\displaystyle\frac{(c/n_{\rm bg})e}{L_{x}L_{y}L_{z}}\left\{\cos\Big{(}(q_{\bar{n}_{x}}-q_{n_{x}})x\Big{)}-\cos\Big{(}(q_{\bar{n}_{x}}+q_{n_{x}})x\Big{)}\right\} (164)
×{cos((qn¯yqny)y)cos((qn¯y+qny)y)}\displaystyle\hskip 28.45274pt\times\left\{\cos\Big{(}(q_{\bar{n}_{y}}-q_{n_{y}})y\Big{)}-\cos\Big{(}(q_{\bar{n}_{y}}+q_{n_{y}})y\Big{)}\right\}
×{cos((qn¯zqnz)z)cos((qn¯z+qnz)z)},\displaystyle\hskip 28.45274pt\times\left\{\cos\Big{(}(q_{\bar{n}_{z}}-q_{n_{z}})z\Big{)}-\cos\Big{(}(q_{\bar{n}_{z}}+q_{n_{z}})z\Big{)}\right\},
j0μ(x)(𝒙)\displaystyle j_{0\mu}^{(x)}(\bm{x}) =\displaystyle= e2ime1LxLyLz[qn¯x{sin((qn¯x+qnx)x)sin((qn¯xqnx)x)}\displaystyle-\frac{e\hbar}{2im_{\rm e}}\frac{1}{L_{x}L_{y}L_{z}}\left[q_{\bar{n}_{x}}\left\{\sin\Big{(}(q_{\bar{n}_{x}}+q_{n_{x}})x\Big{)}-\sin\Big{(}(q_{\bar{n}_{x}}-q_{n_{x}})x\Big{)}\right\}\right. (165)
qnx{sin((qn¯x+qnx)x)+sin((qn¯xqnx)x)}]\displaystyle\hskip 76.82243pt\left.-q_{n_{x}}\left\{\sin\Big{(}(q_{\bar{n}_{x}}+q_{n_{x}})x\Big{)}+\sin\Big{(}(q_{\bar{n}_{x}}-q_{n_{x}})x\Big{)}\right\}\right]
×{cos((qn¯yqny)y)cos((qn¯y+qny)y)}\displaystyle\hskip 51.21495pt\times\left\{\cos\Big{(}(q_{\bar{n}_{y}}-q_{n_{y}})y\Big{)}-\cos\Big{(}(q_{\bar{n}_{y}}+q_{n_{y}})y\Big{)}\right\}
×{cos((qn¯zqnz)z)cos((qn¯z+qnz)z)},\displaystyle\hskip 51.21495pt\times\left\{\cos\Big{(}(q_{\bar{n}_{z}}-q_{n_{z}})z\Big{)}-\cos\Big{(}(q_{\bar{n}_{z}}+q_{n_{z}})z\Big{)}\right\},

and similar manner for j0μ(y)j_{0\mu}^{(y)} and j0μ(z)j_{0\mu}^{(z)}. Here, qnα=πnα/Lαq_{n_{\alpha}}=\pi n_{\alpha}/L_{\alpha} for α=x,y,z\alpha=x,y,z. Note that an assumed background refractive index nbgn_{\rm bg} modulates the light velocity to c/nbgc/n_{\rm bg}. For the calculation, we use the 𝒌\bm{k}-representation. Then,

(c/nbg)ρ~0μ(𝒌)\displaystyle(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k}) =\displaystyle= 1(2π)3/2𝑑𝒙(c/nbg)ρ0μ(𝒙)ei𝒌𝒙\displaystyle\frac{1}{(2\pi)^{3/2}}\int d\bm{x}(c/n_{\rm bg})\rho_{0\mu}(\bm{x})e^{-i\bm{k}\cdot\bm{x}} (166)
=\displaystyle= (c/nbg)e8LxLyLz1(2π)3/21i3\displaystyle\frac{(c/n_{\rm bg})e}{8L_{x}L_{y}L_{z}}\frac{1}{(2\pi)^{3/2}}\frac{1}{i^{3}}
×[2kxkx2Δnx,n¯x22kxkx2Qnx,n¯x2][2kyky2Δny,n¯y22kyky2Qny,n¯y2][2kzkz2Δnz,n¯z22kzkz2Qnz,n¯z2]\displaystyle\hskip 8.53581pt\times\left[\frac{2k_{x}}{{k_{x}}^{2}-{\Delta_{n_{x},\bar{n}_{x}}}^{2}}-\frac{2k_{x}}{{k_{x}}^{2}-{Q_{n_{x},\bar{n}_{x}}}^{2}}\right]\left[\frac{2k_{y}}{{k_{y}}^{2}-{\Delta_{n_{y},\bar{n}_{y}}}^{2}}-\frac{2k_{y}}{{k_{y}}^{2}-{Q_{n_{y},\bar{n}_{y}}}^{2}}\right]\left[\frac{2k_{z}}{{k_{z}}^{2}-{\Delta_{n_{z},\bar{n}_{z}}}^{2}}-\frac{2k_{z}}{{k_{z}}^{2}-{Q_{n_{z},\bar{n}_{z}}}^{2}}\right]
×(1(1)NxeikxLx)(1(1)NyeikyLy)(1(1)NzeikzLz),\displaystyle\hskip 8.53581pt\times\left(1-(-1)^{N_{x}}e^{-ik_{x}L_{x}}\right)\left(1-(-1)^{N_{y}}e^{-ik_{y}L_{y}}\right)\left(1-(-1)^{N_{z}}e^{-ik_{z}L_{z}}\right),
j~0μ(x)(𝒌)\displaystyle\tilde{j}_{0\mu}^{(x)}(\bm{k}) =\displaystyle= 1(2π)3/2𝑑𝒙j0μ(x)(𝒙)ei𝒌𝒙\displaystyle\frac{1}{(2\pi)^{3/2}}\int d\bm{x}j_{0\mu}^{(x)}(\bm{x})e^{-i\bm{k}\cdot\bm{x}} (167)
=\displaystyle= e2me18LxLyLz1(2π)3/21i3\displaystyle\frac{e\hbar}{2m_{\rm e}}\frac{1}{8L_{x}L_{y}L_{z}}\frac{1}{(2\pi)^{3/2}}\frac{1}{i^{3}}
×[2Δnx,n¯xQnx,n¯xkx2Δnx,n¯x22Qnx,n¯xΔnx,n¯xkx2Qnx,n¯x2][2kyky2Δny,n¯y22kyky2Qny,n¯y2][2kzkz2Δnz,n¯z22kzkz2Qnz,n¯z2]\displaystyle\hskip 8.53581pt\times\left[\frac{2\Delta_{n_{x},\bar{n}_{x}}Q_{n_{x},\bar{n}_{x}}}{{k_{x}}^{2}-{\Delta_{n_{x},\bar{n}_{x}}}^{2}}-\frac{2Q_{n_{x},\bar{n}_{x}}\Delta_{n_{x},\bar{n}_{x}}}{{k_{x}}^{2}-{Q_{n_{x},\bar{n}_{x}}}^{2}}\right]\left[\frac{2k_{y}}{{k_{y}}^{2}-{\Delta_{n_{y},\bar{n}_{y}}}^{2}}-\frac{2k_{y}}{{k_{y}}^{2}-{Q_{n_{y},\bar{n}_{y}}}^{2}}\right]\left[\frac{2k_{z}}{{k_{z}}^{2}-{\Delta_{n_{z},\bar{n}_{z}}}^{2}}-\frac{2k_{z}}{{k_{z}}^{2}-{Q_{n_{z},\bar{n}_{z}}}^{2}}\right]
×(1(1)NxeikxLx)(1(1)NyeikyLy)(1(1)NzeikzLz)\displaystyle\hskip 8.53581pt\times\left(1-(-1)^{N_{x}}e^{-ik_{x}L_{x}}\right)\left(1-(-1)^{N_{y}}e^{-ik_{y}L_{y}}\right)\left(1-(-1)^{N_{z}}e^{-ik_{z}L_{z}}\right)

with following definitions: Nαn¯α+nαN_{\alpha}\equiv\bar{n}_{\alpha}+n_{\alpha}, Qnα,n¯αqnα+qn¯αQ_{n_{\alpha},\bar{n}_{\alpha}}\equiv q_{n_{\alpha}}+q_{\bar{n}_{\alpha}}, and Δnα,n¯αqnαqn¯α\Delta_{n_{\alpha},\bar{n}_{\alpha}}\equiv q_{n_{\alpha}}-q_{\bar{n}_{\alpha}}. Moreover, we find symmetric relations, ρ~μ0(𝒌)=ρ~0μ(𝒌)\tilde{\rho}_{\mu 0}(\bm{k})=\tilde{\rho}_{0\mu}(\bm{k}) and 𝒋~μ0(𝒌)=𝒋~0μ(𝒌)\tilde{\bm{j}}_{\mu 0}(\bm{k})=-\tilde{\bm{j}}_{0\mu}(\bm{k}).

The elements of the radiative correction matrix K¯\bar{K} in Eq. (85) is

K~μ0,0μ=Aμ,μ(2)(ω)+Aμ,μ(1)(ω)+Aμ,μ(0)\tilde{K}_{\mu^{\prime}0,0\mu}=A_{\mu^{\prime},\mu}^{(2)}(\omega)+A_{\mu^{\prime},\mu}^{(1)}(\omega)+A_{\mu^{\prime},\mu}^{(0)} (168)

with

Aμ,μ(0)\displaystyle A_{\mu^{\prime},\mu}^{(0)} =\displaystyle= μ0𝑑𝒌(c/nbg)ρ~μ0(𝒌)(1𝒌2)(c/nbg)ρ~0μ(𝒌)\displaystyle\mu_{0}\int d\bm{k}(c/n_{\rm bg})\tilde{\rho}_{\mu^{\prime}0}(-\bm{k})\left(\frac{1}{\bm{k}^{2}}\right)(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k}) (169)
=\displaystyle= μ0(L03LxLyLz)2(2π)3c2e2π5L0𝑑X𝑑Y𝑑Z\displaystyle\mu_{0}\left(\frac{{L_{0}}^{3}}{L_{x}L_{y}L_{z}}\right)^{2}\left(\frac{2}{\pi}\right)^{3}\frac{c^{2}e^{2}}{\pi^{5}L_{0}}\cdot\int_{-\infty}^{\infty}dX\int_{-\infty}^{\infty}dY\int_{-\infty}^{\infty}dZ
×(X2F¯nx,n¯x;nx,n¯x(X))(Y2F¯ny,n¯y;ny,n¯y(Y))(Z2F¯nz,n¯z;nz,n¯z(Z))1X2+Y2+Z21nbg2\displaystyle\hskip 28.45274pt\times\left(X^{2}\bar{F}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(Y^{2}\bar{F}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(Z^{2}\bar{F}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}(Z)\right)\frac{1}{X^{2}+Y^{2}+Z^{2}}\frac{1}{{n_{\rm bg}}^{2}}
Aμ,μ(1)(ω)\displaystyle A_{\mu^{\prime},\mu}^{(1)}(\omega) =\displaystyle= μ0𝑑𝒌(𝒋~μ0(𝒌))t(1𝒌2nbgω/c𝒌2+(nbgω/c)2𝒌)(c/nbg)ρ~0μ(𝒌)\displaystyle\mu_{0}\int d\bm{k}\left(\tilde{\bm{j}}_{\mu^{\prime}0}(-\bm{k})\right)^{\rm t}\left(-\frac{1}{\bm{k}^{2}}\frac{n_{\rm bg}\omega/c}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\bm{k}\right)(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k}) (170)
=\displaystyle= ηωμ0(L03LxLyLz)2(2π)3c2e2π5L0𝑑X𝑑Y𝑑Z\displaystyle\eta_{\omega}\mu_{0}\left(\frac{{L_{0}}^{3}}{L_{x}L_{y}L_{z}}\right)^{2}\left(\frac{2}{\pi}\right)^{3}\frac{c^{2}e^{2}}{\pi^{5}L_{0}}\cdot\int_{-\infty}^{\infty}dX\int_{-\infty}^{\infty}dY\int_{-\infty}^{\infty}dZ
×(Q¯nx,n¯xΔ¯nx,n¯x+Q¯ny,n¯yΔ¯ny,n¯y+Q¯nz,n¯zΔ¯nz,n¯z)\displaystyle\hskip 28.45274pt\times\left(\bar{Q}_{n^{\prime}_{x},\bar{n^{\prime}}_{x}}\bar{\Delta}_{n^{\prime}_{x},\bar{n^{\prime}}_{x}}+\bar{Q}_{n^{\prime}_{y},\bar{n^{\prime}}_{y}}\bar{\Delta}_{n^{\prime}_{y},\bar{n^{\prime}}_{y}}+\bar{Q}_{n^{\prime}_{z},\bar{n^{\prime}}_{z}}\bar{\Delta}_{n^{\prime}_{z},\bar{n^{\prime}}_{z}}\right)
×(X2F¯nx,n¯x;nx,n¯x(X))(Y2F¯ny,n¯y;ny,n¯y(Y))(Z2F¯nz,n¯z;nz,n¯z(Z))\displaystyle\hskip 28.45274pt\times\left(X^{2}\bar{F}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(Y^{2}\bar{F}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(Z^{2}\bar{F}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}(Z)\right)
×1X2+Y2+Z21(X2+Y2+Z2)nbg2ω¯2\displaystyle\hskip 170.71652pt\times\frac{1}{X^{2}+Y^{2}+Z^{2}}\frac{1}{(X^{2}+Y^{2}+Z^{2})-{n_{\rm bg}}^{2}\bar{\omega}^{2}}
Aμ,μ(2)(ω)\displaystyle A_{\mu^{\prime},\mu}^{(2)}(\omega) =\displaystyle= μ0𝑑𝒌(𝒋~μ0(𝒌))t(1𝒌2+(nbgω/c)2)𝒋~0μ(𝒌)\displaystyle\mu_{0}\int d\bm{k}\left(\tilde{\bm{j}}_{\mu^{\prime}0}(-\bm{k})\right)^{\rm t}\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{\bm{j}}_{0\mu}(\bm{k}) (171)
=\displaystyle= ηω2ω¯2μ0(L03LxLyLz)2(2π)3c2e2π5L0𝑑X𝑑Y𝑑Z\displaystyle-\frac{{\eta_{\omega}}^{2}}{\bar{\omega}^{2}}\mu_{0}\left(\frac{{L_{0}}^{3}}{L_{x}L_{y}L_{z}}\right)^{2}\left(\frac{2}{\pi}\right)^{3}\frac{c^{2}e^{2}}{\pi^{5}L_{0}}\cdot\int_{-\infty}^{\infty}dX\int_{-\infty}^{\infty}dY\int_{-\infty}^{\infty}dZ
×[(Π¯nx,n¯x;nx,n¯xF¯nx,n¯x;nx,n¯x(X))(Y2F¯ny,n¯y;ny,n¯y(Y))(Z2F¯nz,n¯z;nz,n¯z(Z))\displaystyle\hskip 28.45274pt\times\left[\left(\bar{\Pi}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}\bar{F}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(Y^{2}\bar{F}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(Z^{2}\bar{F}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}(Z)\right)\right.
+(X2F¯nx,n¯x;nx,n¯x(X))(Π¯ny,n¯y;ny,n¯yF¯ny,n¯y;ny,n¯y(Y))(Z2F¯nz,n¯z;nz,n¯z(Z))\displaystyle\hskip 28.45274pt+\left(X^{2}\bar{F}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(\bar{\Pi}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}\bar{F}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(Z^{2}\bar{F}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}(Z)\right)
+(X2F¯nx,n¯x;nx,n¯x(X))(Y2F¯ny,n¯y;ny,n¯y(Y))(Π¯nz,n¯z;nz,n¯zF¯nz,n¯z;nz,n¯z(Z))]\displaystyle\hskip 28.45274pt\left.+\left(X^{2}\bar{F}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(Y^{2}\bar{F}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(\bar{\Pi}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}\bar{F}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}(Z)\right)\right]
×1(X2+Y2+Z2)nbg2ω¯2.\displaystyle\hskip 199.16928pt\times\frac{1}{(X^{2}+Y^{2}+Z^{2})-{n_{\rm bg}}^{2}\bar{\omega}^{2}}.

Here, the functions for the integrands are

F¯nx,n¯x;nx,n¯x(X)\displaystyle\bar{F}_{n^{\prime}_{x},\bar{n^{\prime}}_{x};n_{x},\bar{n}_{x}}(X) =\displaystyle= (1X2Δ¯nx,n¯x21X2Q¯nx,n¯x2)(1X2Δ¯nx,n¯x21X2Q¯nx,n¯x2)\displaystyle\left(\frac{1}{X^{2}-{\bar{\Delta}_{n^{\prime}_{x},\bar{n^{\prime}}_{x}}}^{2}}-\frac{1}{X^{2}-{\bar{Q}_{n^{\prime}_{x},\bar{n^{\prime}}_{x}}}^{2}}\right)\left(\frac{1}{X^{2}-{\bar{\Delta}_{n_{x},\bar{n}_{x}}}^{2}}-\frac{1}{X^{2}-{\bar{Q}_{n_{x},\bar{n}_{x}}}^{2}}\right) (172)
×14(1+(1)Nx+Nx(1)NxeiπLxX/L0(1)NxeiπLxX/L0)\displaystyle\hskip 28.45274pt\times\frac{1}{4}\left(1+(-1)^{N_{x}+N^{\prime}_{x}}-(-1)^{N_{x}}e^{-i\pi L_{x}X/L_{0}}-(-1)^{N^{\prime}_{x}}e^{i\pi L_{x}X/L_{0}}\right)
Π¯nα,n¯α;nα,n¯α\displaystyle\bar{\Pi}_{n^{\prime}_{\alpha},\bar{n^{\prime}}_{\alpha};n_{\alpha},\bar{n}_{\alpha}} =\displaystyle= Q¯nα,n¯αΔ¯nα,n¯αQ¯nα,n¯αΔ¯nα,n¯α.\displaystyle\bar{Q}_{n^{\prime}_{\alpha},\bar{n^{\prime}}_{\alpha}}\bar{\Delta}_{n^{\prime}_{\alpha},\bar{n^{\prime}}_{\alpha}}\bar{Q}_{n_{\alpha},\bar{n}_{\alpha}}\bar{\Delta}_{n_{\alpha},\bar{n}_{\alpha}}. (173)

F¯ny,n¯y;ny,n¯y(Y)\bar{F}_{n^{\prime}_{y},\bar{n^{\prime}}_{y};n_{y},\bar{n}_{y}}(Y) and F¯nz,n¯z;nz,n¯z(Z)\bar{F}_{n^{\prime}_{z},\bar{n^{\prime}}_{z};n_{z},\bar{n}_{z}}(Z) are defined in a similar manner for Eq. (172). The integral variables and several factors are normalized as dimensionless ones, X=kxL0/πX=k_{x}L_{0}/\pi, Y=kyL0/πY=k_{y}L_{0}/\pi, Z=kzL0/πZ=k_{z}L_{0}/\pi, Q¯nα,n¯α=Qnα,n¯αL0/π\bar{Q}_{n_{\alpha},\bar{n}_{\alpha}}=Q_{n_{\alpha},\bar{n}_{\alpha}}L_{0}/\pi, Δ¯nα,n¯α=Δnα,n¯αL0/π\bar{\Delta}_{n_{\alpha},\bar{n}_{\alpha}}=\Delta_{n_{\alpha},\bar{n}_{\alpha}}L_{0}/\pi, and

ω¯=ωcL0π.\bar{\omega}=\frac{\omega}{c}\frac{L_{0}}{\pi}. (174)

Here, an introduced factor

ηω=π2mcL0ω¯=1m/meω2mec2\eta_{\omega}=\frac{\pi}{2}\frac{\hbar}{m^{*}cL_{0}}\bar{\omega}=\frac{1}{m^{*}/m_{\rm e}}\frac{\hbar\omega}{2m_{\rm e}c^{2}} (175)

means a relativistic factor. The elements of the matrices L¯\bar{L}, M¯\bar{M}, and N¯\bar{N} are obtained by Eqs. (86), (87), and (88), respectively.

For the matrices related to the vector potential term in the current, the elements of S¯\bar{S} in Eq. (90) becomes

S~mβ,0μ=Bmβ,μ(2)+Bmβ,μ(1)\tilde{S}_{m^{\prime}\beta,0\mu}=B_{m^{\prime}\beta,\mu}^{(2)}+B_{m^{\prime}\beta,\mu}^{(1)} (176)

with

Bmx,μ(1)(ω)\displaystyle B_{m^{\prime}x,\mu}^{(1)}(\omega) =\displaystyle= μ0𝑑𝒌φ~m(𝒌)(1𝒌2nbgω/c𝒌2+(nbgω/c)2kx)(c/nbg)ρ~0μ(𝒌)\displaystyle\mu_{0}\int d\bm{k}\tilde{\varphi}_{m^{\prime}}(-\bm{k})\left(-\frac{1}{\bm{k}^{2}}\frac{n_{\rm bg}\omega/c}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}k_{x}\right)(c/n_{\rm bg})\tilde{\rho}_{0\mu}(\bm{k}) (177)
=\displaystyle= iμ0(L03LxLyLz)32(2π)92(c2e2π5L0)12ωL02cπ2\displaystyle-i\mu_{0}\left(\frac{{L_{0}}^{3}}{L_{x}L_{y}L_{z}}\right)^{\frac{3}{2}}\left(\frac{2}{\pi}\right)^{\frac{9}{2}}\left(\frac{c^{2}e^{2}}{\pi^{5}L_{0}}\right)^{\frac{1}{2}}\frac{\omega{L_{0}}^{2}}{c\pi^{2}}
×dXdYdZ(X2h¯mx;nx,n¯x(X))(Yh¯my;ny,n¯y(Y))(Zh¯mz;nz,n¯z(Z))\displaystyle\hskip 28.45274pt\times\int_{-\infty}^{\infty}dX\int_{-\infty}^{\infty}dY\int_{-\infty}^{\infty}dZ\left(X^{2}\bar{h}_{m^{\prime}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(Y\bar{h}_{m^{\prime}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(Z\bar{h}_{m^{\prime}_{z};n_{z},\bar{n}_{z}}(Z)\right)
×1X2+Y2+Z21(X2+Y2+Z2)nbg2ω¯2,\displaystyle\hskip 170.71652pt\times\frac{1}{X^{2}+Y^{2}+Z^{2}}\frac{1}{(X^{2}+Y^{2}+Z^{2})-{n_{\rm bg}}^{2}\bar{\omega}^{2}},
Bmx,μ(2)(ω)\displaystyle B_{m^{\prime}x,\mu}^{(2)}(\omega) =\displaystyle= μ0𝑑𝒌φ~m(𝒌)(1𝒌2+(nbgω/c)2)j~0μ(x)(𝒌)\displaystyle\mu_{0}\int d\bm{k}\tilde{\varphi}_{m^{\prime}}(-\bm{k})\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{j}^{(x)}_{0\mu}(\bm{k}) (178)
=\displaystyle= i(ηωω¯2)μ0(L03LxLyLz)32(2π)92(c2e2π5L0)12ωL02cπ2\displaystyle i\left(\frac{\eta_{\omega}}{\bar{\omega}^{2}}\right)\mu_{0}\left(\frac{{L_{0}}^{3}}{L_{x}L_{y}L_{z}}\right)^{\frac{3}{2}}\left(\frac{2}{\pi}\right)^{\frac{9}{2}}\left(\frac{c^{2}e^{2}}{\pi^{5}L_{0}}\right)^{\frac{1}{2}}\frac{\omega{L_{0}}^{2}}{c\pi^{2}}
×dXdYdZ(π¯nx,n¯xh¯mx;nx,n¯x(X))(Yh¯my;ny,n¯y(Y))(Zh¯mz;nz,n¯z(Z))\displaystyle\hskip 28.45274pt\times\int_{-\infty}^{\infty}dX\int_{-\infty}^{\infty}dY\int_{-\infty}^{\infty}dZ\left(\bar{\pi}_{n_{x},\bar{n}_{x}}\bar{h}_{m^{\prime}_{x};n_{x},\bar{n}_{x}}(X)\right)\left(Y\bar{h}_{m^{\prime}_{y};n_{y},\bar{n}_{y}}(Y)\right)\left(Z\bar{h}_{m^{\prime}_{z};n_{z},\bar{n}_{z}}(Z)\right)
×1(X2+Y2+Z2)nbg2ω¯2.\displaystyle\hskip 170.71652pt\times\frac{1}{(X^{2}+Y^{2}+Z^{2})-{n_{\rm bg}}^{2}\bar{\omega}^{2}}.

Here, the functions for the integrands are

Xh¯mx;nx,n¯x(X)\displaystyle X\bar{h}_{m^{\prime}_{x};n_{x},\bar{n}_{x}}(X) =\displaystyle= Q¯mxX2Q¯mx2[XX2Δ¯nx,n¯x2XX2Q¯nx,n¯x2]\displaystyle\frac{\bar{Q}_{m^{\prime}_{x}}}{X^{2}-{\bar{Q}_{m^{\prime}_{x}}^{2}}}\left[\frac{X}{X^{2}-{\bar{\Delta}_{n_{x},\bar{n}_{x}}}^{2}}-\frac{X}{X^{2}-{\bar{Q}_{n_{x},\bar{n}_{x}}}^{2}}\right] (179)
×14(1+(1)Nx+mx(1)NxeiπLxX/L0(1)mxeiπLxX/L0),\displaystyle\hskip 42.67912pt\times\frac{1}{4}\left(1+(-1)^{N_{x}+m^{\prime}_{x}}-(-1)^{N_{x}}e^{-i\pi L_{x}X/L_{0}}-(-1)^{m^{\prime}_{x}}e^{i\pi L_{x}X/L_{0}}\right),
π¯nα,n¯α\displaystyle\bar{\pi}_{n_{\alpha},\bar{n}_{\alpha}} =\displaystyle= Q¯nαn¯αΔ¯nα,n¯α,\displaystyle\bar{Q}_{n_{\alpha}\bar{n}_{\alpha}}\bar{\Delta}_{n_{\alpha},\bar{n}_{\alpha}}, (180)

with a dimensionless factor Q¯mα=qmαL0/π\bar{Q}_{m_{\alpha}}=q_{m_{\alpha}}L_{0}/\pi. Bmy,μ(j)(ω)B_{m^{\prime}y,\mu}^{(j)}(\omega) and Bmz,μ(j)(ω)B_{m^{\prime}z,\mu}^{(j)}(\omega) are obtained in the similar manner. The elements of T¯\bar{T}, U¯\bar{U}, and V¯\bar{V} in Eqs. (91)-(93) are also given by them.

Finally, for the matrix R¯\bar{R}, Eq. (94) becomes

Cm,m(ω)\displaystyle C_{m^{\prime},m}(\omega) =\displaystyle= (ωpc/nbg)2d3𝒌φ~m(𝒌)(1𝒌2+(nbgω/c)2)φ~m(𝒌)δαβ\displaystyle\left(\frac{\omega_{\rm p}^{\prime}}{c/n_{\rm bg}}\right)^{2}\int d^{3}\bm{k}\tilde{\varphi}_{m^{\prime}}(-\bm{k})\left(\frac{1}{-\bm{k}^{2}+(n_{\rm bg}\omega/c)^{2}}\right)\tilde{\varphi}_{m}^{\ast}(\bm{k})\delta_{\alpha\beta} (181)
=\displaystyle= L03LxLyLz(2π)6(ωpω)2(ωL0cπ)2𝑑X𝑑Y𝑑Z\displaystyle-\frac{{L_{0}}^{3}}{L_{x}L_{y}L_{z}}\left(\frac{2}{\pi}\right)^{6}\left(\frac{\omega_{\rm p}}{\omega}\right)^{2}\left(\frac{\omega L_{0}}{c\pi}\right)^{2}\cdot\int_{-\infty}^{\infty}dX\int_{-\infty}^{\infty}dY\int_{-\infty}^{\infty}dZ
×(f¯mx;mx(X))(f¯my;my(Y))(f¯mz;mz(Z))1(X2+Y2+Z2)nbg2ω¯2\displaystyle\hskip 28.45274pt\times\left(\bar{f}_{m^{\prime}_{x};m_{x}}(X)\right)\left(\bar{f}_{m^{\prime}_{y};m_{y}}(Y)\right)\left(\bar{f}_{m^{\prime}_{z};m_{z}}(Z)\right)\frac{1}{(X^{2}+Y^{2}+Z^{2})-{n_{\rm bg}}^{2}\bar{\omega}^{2}}

with

f¯mx;mx(X)=Q¯mxX2Q¯mx2Q¯mxX2Q¯mx2×14(1+(1)mx+mx(1)mxeiπLxX/L0(1)mxeiπLxX/L0).\bar{f}_{m^{\prime}_{x};m_{x}}(X)=\frac{\bar{Q}_{m^{\prime}_{x}}}{X^{2}-{\bar{Q}_{m^{\prime}_{x}}^{2}}}\frac{\bar{Q}_{m_{x}}}{X^{2}-{\bar{Q}_{m_{x}}^{2}}}\times\frac{1}{4}\left(1+(-1)^{m_{x}+m^{\prime}_{x}}-(-1)^{m_{x}}e^{-i\pi L_{x}X/L_{0}}-(-1)^{m^{\prime}_{x}}e^{i\pi L_{x}X/L_{0}}\right). (182)

References

  • (1) M. S. Tame, K. R. McEnery, Ş. K. Özdemir, J. Lee, S. A. Maier, and M. S. Kim, Nat. Phys. 9, 329 (2013).
  • (2) C. J. Powell and J. B. Swan, Phys. Rev. 116, 81 (1959).
  • (3) L. Novotny and B. Hecht, Principles of Nano-Optics Second Edition (Cambridge University Press, United Kingdom) 2012.
  • (4) A. Otto, Z. Phys. 216, 398 (1968).
  • (5) E. Kretschmann and H. Raether, Z. Naturforsch. 23A, 2135 (1968).
  • (6) P. J. Feibelman, Prog. Surf. Sci. 12, 287 (1982).
  • (7) R. R. Gerhardts and K. Kempa, Phys. Rev. B 30, 5704 (1984).
  • (8) M. A. Garcia, J. Phys. D: Appl. Phys. 45, 389501 (2011).
  • (9) S. Zeng, K.-T. Yong, I. Roy, X.-Q. Dinh, X. Yu, and F. Luan, Plasmonics 6, 491 (2011).
  • (10) M. S. A. Gandhi, S. Chu, K. Senthilnathan, P. R. Babu, K. Nakkeeran, and Q. Li, Appl. Sci. 9, 949 (2019).
  • (11) A. O. Govorov, H. Zhang, H. V. Demir, and Y. K. Gun’ko, Nano Today 9, 85 (2014).
  • (12) M. L. Brongersma, N. J. Halas, and P. Nordlander, Nat. Nanotech. 10, 25 (2015).
  • (13) T. P. White and K. R. Catchpole, Appl. Phys. Lett. 101, 073905 (2012).
  • (14) A. O. Govorov, H. Zhang, and Y. K. Gun’ko, J. Phys. Chem. C 117, 16616 (2013).
  • (15) P. V. Kumar, T. P. Rossi, D. Marti-Dafcik, D. Reichmuth, M. Kuisma, P. Erhart, M. J. Puska, and D. J. Norris, ACS Nano 13, 3188 (2019).
  • (16) K. Ueno, T. Oshikiri, and H. Misawa, Chem. Phys. Chem. 17, 199 (2016).
  • (17) W. Li and J. G. Valentine, Nanophotonics 6, 177 (2017).
  • (18) T. Tatsuma, H. Nishi, and T. Ishida, Chem. Sci. 8, 3325 (2017).
  • (19) C. Clavero, Nat. Photo. 8, 95 (2014).
  • (20) R. Sundararaman, P. Narang, A. S. Jermyn, W. A. Goddard III, and H. A. Atwater, Nat. Commun. 5, 5788 (2014).
  • (21) M. Bernardi, J. Mustafa, J. B. Neaton, and S. G. Louie, Nat. Commun. 6, 7044 (2015).
  • (22) H. Zhang and A. O. Govorov, J. Phys. Chem. C 118, 7606 (2014).
  • (23) M. B. Ross and G. C. Schatz, J. Phys. D: Appl. Phys. 48, 184004 (2015).
  • (24) L. V. Besteiro, X.-T. Kong, Z. Wang, G. Hartland, and A. O. Govorov, ACS Photonics 4, 2759 (2017).
  • (25) A. Manjavacas, J. G. Liu, V. Kulkarni, and P. Nordlander, ACS Nano 8, 7630 (2014).
  • (26) K. D. Chapkin, L. Bursi, G. J. Stec, A. Lauchner, N. J. Hogan, Y. Cui, P. Nordlander, and N. J. Halas, Proc. Natl. Acad. Sci. USA 115, 9134 (2018).
  • (27) J. Ma, Z. Wang, and L.-W. Wang, Nat. Commun. 6, 10107 (2015).
  • (28) X. You, S. Ramakrishana, and T. Seideman, J. Phys. Chem. Lett. 9, 141 (2018).
  • (29) R. Zhang, L. Bursi, J. D. Cox, Y. Cui, C. M. Krauter, A. Alabastri, A. Manjavacas, A. Calzolari, S. Corni, E. Molinari, E. A. Carter, F. J. G. de Abajo, H. Zhang, and P. Nordlander, ACS Nano 11, 7321 (2017).
  • (30) J. M. Fitzgerald, S. Azadi, and V. Giannini, Phys. Rev. B 95, 235414 (2017).
  • (31) R. D. Senanayake, D. B. Lingerfelt, G. U. Kuda-Singappulige, X. Li, and C. M. Aikens, J. Phys. Chem. C 123, 14734 (2019).
  • (32) R. Schira and F. Rabilloud, J. Phys. Chem. C 123, 6205 (2019).
  • (33) J. M. Pitarke, V. M. Silkin, E. V. Chulkov, and P. M. Echenique, Rep. Prog. Phys. 70, 1 (2007).
  • (34) J. M. McMahan, S. K. Gray, andG. C. Schatz, Phys. Rev. Lett. 103, 097403 (2009).
  • (35) S. Raza, G. Toscano, A.-P. Jauho, M. Wubs, and N. A. Mortensen, Phys. Rev. B 84, 121412(R) (2011).
  • (36) G. Toscano, S. Raza, A.-P. Jauho, N. A. Mortensen, and M. Wubs, Opt. Exp. 20, 4176 (2012).
  • (37) A. Moreau, C. Ciracì, and D. R. Smith, Phys. Rev. B 87, 045401 (2013).
  • (38) W. Wang and J. M. Kinaret, Phys. Rev. B 87, 195424 (2013).
  • (39) N. A. Mortensen, S. Raza, M. Wubs, T. Søndergaard, and S. I. Bozhevolnyi, Nat. Commun. 5, 3809 (2014).
  • (40) T. Christensen, W. Yan, Søren Raza. A.-P. Jauho, N. A. Mortensen, and M. Wubs, ACS Nano 8, 1745 (2014).
  • (41) M. K. Svendsen, C. Wolff, A.-P. Jauho, N. A. Mortensen, and C. Tserkezis, J. Phys.: Condens. Matter 32, 395702 (2020), and related reference therein.
  • (42) T. Inaoka, J. Phys.: Condens. Matter 8, 10241 (1996).
  • (43) K. Cho, Optical Response of Nanostructures: Microscopic Nonlocal Theory Springer Series in Solid-State Sciences (Springer-Verlag, Tokyo) 2003.
  • (44) E. Prodan and P. Nordlander, Chem. Phys. Lett. 349, 153 (2001).
  • (45) F. H. L. Koppens, D. E. Chang, and F. J. G. de Abajo, Nano Lett. 11, 3370 (2011).
  • (46) D. Lu, Z. Hou, X. Liu, B. Da, and Z. J. Ding, J. Phys. Chem. C 123, 25341 (2019).
  • (47) M. K. Svendsen, Y. Kurman, P. Schmidt, F. Koppens, I. Kaminer, and K. S. Thygesen, Nat. Commun. 12, 2778 (2021).
  • (48) Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959).
  • (49) A. Tonomura, N. Osakabe, T. Matsuda, T. Kawasaki, J. Endo, S. Yano, and H. Yamada, Phys. Rev. Lett. 56, 792(1986).
  • (50) R. Kubo, J. Phys. Soc. Jpn., 12, 570 (1957).
  • (51) Complete inclusion of the correlation and quantum fluctuation for the excited states is a task beyond the simple preparation of the exact eigenstates |μ|\mu\rangle including the Coulomb interaction at equilibrium. This point will be a subject of future research based on the present study.
  • (52) K. Cho, Reconstruction of Macroscopic Maxwell Equations: A Single Susceptibility Theory Springer Tracts in Modern Physics Volume 237 (Springer-Verlag, Berlin Heidelberg) 2010.