Direct space-time modeling of mechanically dressed dipole-dipole interactions with electromagnetically-coupled oscillating dipoles
Abstract
We study the radiative dynamics of coupled electric dipoles, modelled as Lorentz oscillators (LOs), in the presence of real-time mechanical oscillations. The dipoles are treated in a self-consistent way through a direct electromagnetic simulation approach that fully includes the dynamical movement of the charges, accounting for radiation reaction, emission and absorption. This allows for a powerful numerical solution of optomechanical resonances without any perturbative approximations for the mechanical motion. The scaled population (excitation) dynamics of the LOs are investigated as well as the emitted radiation and electromagnetic spectra, which demonstrates how the usual dipole-dipole resonances couple to the underlying Floquet states, yielding multiple spectral peaks that are separated from the superradiant and subradiant states by an integer number of the mechanical oscillation frequency. Moreover, we observe that when the mechanical amplitude and frequency are sufficiently large, these additional spectral peaks undergo further modification, including spectral splitting, spectral squeezing, or shifting. These observations are fully corroborated by a theoretical Floquet analysis conducted on two coupled harmonic oscillators.
I Introduction
The spontaneous emission (SE) process can be modelled from the viewpoint of vacuum fluctuations in quantum field theory or from radiation reaction [1], with the latter mapping on to classical and semiclassical field-matter theories, whereby the radiative dynamics can be captured through polarization dipoles and Lorentz oscillators (LOs) [2] coupled to classical fields.
In both classical and quantum field pictures, the SE rate is governed by the (projected) photonic local density of states (LDOS) [3, 4, 5, 6], which is related to the imaginary part of the system’s electric field Green function , through
(1) |
for an emitter at position , with dipole moment (assumed real) and resonance frequency . The classical Green function of the medium is defined from
(2) |
where is the dielectric constant of the medium, is the speed of light in vacuum, and is the unit tensor (dyad), and we assume a nonmagnetic medium.
For example, for a dipole in free space, as depicted in Fig. 1(a), we have , and
(3) |
is the free space SE decay rate, for an emitter with resonance frequency with . In a dielectric with , this rate is simply modified by , the background refractive index of the medium. Moreover, for any inhomogeneous structure, the SE rate can be controlled by the LDOS (or projected LDOS, as the dipole direction also matters), which is modified by the electromagnetic environment surrounding the emitters. The pioneering work of Edward Purcell first pointed out that the SE rate of quantum emitters could be altered by placing them inside a resonant cavity [7], causing enhancement or inhibition of the SE rate. Subsequent research has extensively explored this process across various physical systems, such as emitters within photonic crystals [8, 9, 10, 11, 12, 13, 14].
An alternative approach for modifying the SE dynamics involves coupling to other atoms, or to a collection of atoms, typically in the near-field regime (i.e., much less than their emission wavelength). This results in collective excitations that decay either faster (superradiance) or slower (subradiance) than the individual emitters, which has been demonstrated in a number of systems [15, 16, 17, 18, 19]. These coupled dipole studies reveal that the modified SE rate is dependent on interatomic distances, which can significantly impact the amplitude and width of the emission spectrum [17, 20].

For two atoms, with their center-of-mass (COM) at positions and with resonant energies , with , then we also have an incoherent photon transfer rate [21]111Here we also allow for complex dipole moments
(4) |
and a coherent coupling rate
(5) |
where a Markov approximation has been applied (e.g., the rates are assumed to be constant over the frequency range of interest, so all rates are instantaneous in time).
Note also that the above expressions can be derived classically or quantum mechanically, with a full classical-quantum correspondence (discussed in more detail in Appendix B). For example, in the quantum case, when the dipole bosonic operators and are introduced for each atom, the interaction Hamiltonian for dipole-dipole coupling is , and within a rotating-wave approximation, . With the appropriate initial condition, this can lead to the creation of superradiant states (Dicke states [22], specialized to two dipoles), which decay with the rate , or subradiant states, which decay at the rate . For simplicity, here we assume equal emitter strengths and resonance frequencies, so that , , and [see Fig. 1(b)].
Recently, the direct numerical modelling of dipole-dipole interactions, with no intrinsic approximations (such as the dipole approximation, rotating wave approximation or Markov approximation), was demonstrated by directly solving the self-consistent electrodynamics simulations of LOs and oscillating point charges [23]. This direct modelling approach can have enormous benefits, as it connects classical dynamical simulations to complex SE dynamics, often in ways that are intractable or difficult to account for in standard quantum theory (or standard classical theory). For example, one does not have to make a rotating wave approximation, and one can also go beyond a dipole approximation, two common approximations in quantized light-matter interaction theories.
A particularly intriguing way of modifying the SE dynamics is to move the COM of the emission dipole itself, e.g., modeled as a moving atom [24]. At relativistic velocities, this can yield very small corrections to [25, 26]. A more practical and significant way to modify the SE rate is through mechanical perturbations of the photonic environment, resulting in an optomechanical coupling, e.g., where the SE can be altered by a vibrating cavity wall [27, 28, 29]. As is well known, if the atom is close to a mirror (e.g., a half-open cavity), the emitted photon can be reflected by the mirror, which (depending on the angle of emission, frequency, and precise distance from the mirror) can either create a constructive or destructive interference at the position of the atom [4, 30, 16, 31, 32].
Indeed, it is now well known that the SE rate and the effective level (frequency) shift of a quantum emitter, such as a two-level atom, can be modified by its interaction with an oscillating mirror. This interaction manifests in a positional-dependent SE rate and effective level shift, scaling with and , respectively, where is the wave vector and is the atom-mirror separation [33, 34]. In Ref. [33], using a rotating wave approximation and one-dimensional quantized fields, the SE of a two-level system near an oscillating plate was investigated. The study focused on field modes propagating within a limited solid angle, towards the mirror and reflected back to the atom, influenced by the mirror oscillation amplitude. In a Markov approximation (instantaneous coupling), when , and (with the round-trip time, and and are the (mechanically) oscillating amplitude and frequency of the mirror), the modified SE rate was found to be
(6) |
and similarly for the atom frequency, , where is less than 1 and depends on the scattering geometry, and is the zeroth-order Bessel function of the first kind, which represents the mirror’s oscillation effect. Modulating the mirror with the dependence , yields the following amplitude for an initially excited atom:
(7) |
where is the time-dependent population. The excited atom population decays exponentially at a modified rate, , which includes an oscillating modulation, where the amplitude of this modulation decreases when the oscillation frequency of the mirror increases. In the resolved sideband regime, where , this modification is small.
In Ref. [34], the authors considered a more general model, where a full three-dimensional electromagnetic field is quantized with the time-dependent boundary conditions determined by the (adiabatically) oscillating mirror. A general orientation of the atomic dipole moment was considered, but the oscillating frequency is only a few GHz compared with the transition frequency of an atom ( THz for that work), and the oscillating amplitude was . Another work investigated the SE rate and emission spectrum of a TLS positioned within a dynamic photonic crystal, which showed how SE modifications relate to the time-dependent photonic density states [35]. These findings suggested that the dynamical control of the boundary conditions causes a further modification of the SE process [33, 35, 34]. This is a nontrivial regime as the SE process is now non-stationary because of the time-dependent boundary conditions.
Apart from the interest in fundamental optics, the investigation of vibrational atoms and molecules is also important in the context of Raman spectroscopy and Surface-Enhanced Raman Spectroscopy (SERS), which forms a model of molecular optomechanics [36, 37, 38]. Molecular optomechanics provides insights into the fundamental principles governing the interaction between light and matter, as well as the development of technologies or applications, such as ultrasensitive detectors that can measure tiny forces and displacements at atomic scale [38, 39, 40].

In this paper, we introduce a direct space-time model for coupling mechanical oscillations to radiatively coupled emitters, where we directly move (oscillate) the COM of the dipoles mechanically, e.g., instead of vibrating mirrors which cause phonons to couple with photons [33, 34]. The main numerical simulations are carried out using PyCharge, a recently developed Python package that can calculate the electromagnetic fields and (vector and scalar) potentials generated by moving point charges and can self-consistently simulate dipoles as LOs [23]. The electric dipoles (LOs) are formed by a coupled oscillating point charges with opposite signs, equal magnitude and a separated displacement [see Fig. 1(a)]. The charges oscillate around the COM of the electric dipole along the direction of polarization. The origin of the LOs can be set to stationary [see Fig. 1(b)] or move as a function of time [see Fig. 2(a)]. The LOs are naturally damped within the self-consistent solver, with radiative decay. The motion of the LOs is restricted to nonrelativistic velocities; however, PyCharge can model a wide range of effects in electromagnetism, including “Braking radiation” (Bremsstrahlung) [41].
There are several benefits to using PyCharge to model the dynamics of LOs. First, it has already been shown to be very accurate for obtaining the known rates of dipole-dipole coupling from QED theory (typical numbers are relative error, even for very close dipole separations, deep sub-wavelength) [23]. Secondly, PyCharge can explore effects that cannot be solved analytically or via the usual simulation software for electromagnetism. For instance, the numerical simulation of PyCharge does not rely on a Markovian approximation, dipole approximation, nor any rotating wave approximation, which are often required in other approaches, such as quantum master equations.
Specifically, in this work we use PyCharge to explore the modified SE decay dynamics from a pair of atoms (treated as -polarized dipoles) that are coupled through photon-coupling effects such as Förster coupling [42], when one of the atoms are subject to a mechanical oscillation on the COM. In general, an atom subject to a periodic oscillation will form Floquet states [43, 44, 45], which can then dress the usual dipole-dipole-induced superradiant and subradiant states to have multiple resonances, separated (in the perturbative limit) by ( integer). Thus, we also develop a rigorous Floquet analysis of two-coupled atoms (harmonic oscillators) with mechanical perturbations, to help with the interpretation and prediction of our direct time-dependent solution using PyCharge. In PyCharge, the dipoles are modeled as classical LOs [4], by directly solving the equation of motion for the LOs including the scattered fields responsible for modified SE.
The rest of our article is organized as follows: In Sec. II we present the main theoretical background for the direct electromagnetic simulations in PyCharge. In Sec. III, we then describe an analytical model of two coupled quantum harmonic oscillators (HOs), with periodic modulation the dipole-dipole coupling, and show how this can be solved using Floquet theory. In Sec. IV, we present several example results showing the electric dipole dynamics of two coupled LOs with and without mechanical oscillations. We also show results for the emitted spectra, for different driving parameters, and each example regime is corroborated using the analytical Floquet theory. Finally, in Sec. V, we give a summary and conclusions. In addition, we present two Appendices. In Appendix A, we present the field solutions from oscillating point dipole in a homogeneous medium. While in Appendix B, we present a quantum and classical field theory for dipole-dipole coupling, which we use in the main text to derive a Floquet solution with periodic oscillations in the dipole-dipole coupling.
II Two Oscillator Optomechanics: Theory and numerical implementation of the Direct Electromagnetic Simulation
To numerically model the radiative decay through radiation reaction, we start with one initially excited dipole. Figure 2 shows a schematic of the two coupled LOs with mechanical oscillations governed by Eq. (8), as well as the potential energy levels. The LOs are formed by two opposite point charges, and masses for each dipole [having the effective (reduced) mass ], with a separate displacement, which oscillates around the origin along the -axis (-polarized dipoles). We can then set an initial displacement for the dipoles, from . For example, we can set and , where nm. Thus, we consider Dipole 1 is initially excited, and Dipole 2 will become excited through interactions from the electric field generated by Dipole 1. As they oscillate, the LOs will radiate (dissipate) energy that causes the dipole moment to decrease [4], but potentially oscillate when coupled to nearby LOs.
Considering two spatially separated LOs with the same effective mass, , the same charge, , and the same transition frequency, , in free space, the coupled dipole equations of motion are
(8a) | |||
(8b) |
where is the component of the total electric field in the direction of polarization generated by all sources in the system (including scattered fields), which is the sum of Coulomb and radiative electric field contributions, given in Eqs. (25) and (26) of Appendix. A, respectively. In the coupled dipole equation of motion, is the position vector of the COM of Dipole 1 which is in periodically mechanical motion, (assumed fixed) is the position vector of the COM of Dipole 2. Since any relevant physics here depends on the relative distance, we define . Thus, the mechanical motion is defined from
(9) |
where we let , and , so that is the mechanical oscillating amplitude of Dipole 1, is the initial static distance between the COMs of the two dipoles along -axis, and is the mechanical oscillating frequency. Therefore, the relative time-dependent distance between the two LOs is simply [see Fig. 2(a)].
For our study, we will assume that the average mechanical velocity of the dipole is . In addition, we constrain the movement of dipoles and charges to non-relativistic velocities, so that the dissipation of charges due to radiation reaction damping is negligible. Thus, we set a reasonable maximum velocity of charge to be . Thus, the mechanical oscillation parameters and play the significant role in our study, and we primarily focus on their respective dimensionless ratios, denoted as which is constrained to values less than (this limitation is imposed to ensure that the dipoles remain spatially separate and do not come in direct contact with each other) and .
In the SE rate regime, the energy decay rate of an emitted photon is equivalent to the population decay rate, where the free-space decay rate, , can also be modelled as an Einstein A coefficient [46]. However, it is important to emphasize that the relationship between the dynamics of LOs and the population of two-level systems’ states can only be connected in a weak excitation regime (since the LO model does not account for a two-level system saturation effects).
To connect to the LO equivalent of population decay, one can also calculate the self-consistent dipole moment of the LOs at each time step by solving Eq. (8). The total energy of an oscillating dipole, is then given by
(10) |
which is the sum of the potential energy and kinetic energy of the dipole, respectively, and is the time-dependent dipole moment of the th dipole. We assume that the total energy is proportional to the population of the excited state (mapping on to a two-level system model), due to the equivalence of the energy decay rate and the population decay rate. Subsequently, we can write the scaled population of the excited states as the normalized total energy [23],
(11) |
for the th dipole.
III Floquet theory of two coupled oscillators with periodic coupling
As described in Appendix. B, the system Hamiltonian for two coupled quantum HOs, when considering the coupling between two electric dipoles in free space, with equal oscillator frequencies , takes on the form
(12) |
where () is the bosonic annihilation (creation) operator of the th dipole and we assume the dipoles are coupled through a near-field electromagnetic interaction, with , where More generally, this interaction strength can be obtained from Eq. (5), which depends on the real part of the Green function, . A correspondence between the classical dipole model can easily be made by connecting (see Appendix. B, and also Ref. [23]), and in terms a quantum dipole moment, one can also show that , introduced earlier (see also Refs. [47, 48]).
Below we will use the Hamiltonian for the quantum HO problem, but we stress that exactly the same resonances occur for a classical Hamiltonian system, which is easy to recover by simply replacing the harmonic oscillator operators by (classical) -numbers. This classical correspondence is possible since the system obeys a linear response, where the resonances yield the polariton frequencies. Interestingly, note that the specific form of the system Hamiltonian is similar to a quantum Hopfield-like model without the diamagnetic term [49, 50], which is needed for describing the coupling between a quantum dipole and a single quantized cavity mode, otherwise a phase transition (often ill-defined) can occur when . In the present case, such a limit cannot be reached, since the dipole approximation is only valid when , where is the size (length) of the dipole (for example, the dipole moment ).
In order to understand the dynamics of the system and calculate the normal modes or resonances (in the absence of dissipation), we write the Heisenberg equation of motion as , with , where
(13) |
and the normal modes of the hybrid system are , which are net positive since for dipole-dipole interactions within the dipole approximation.
We next present an analysis of the two coupled HOs with relative COM movement where the system is Floquet engineered through the time-dependent dipole-dipole coupling rate . This time-dependent character of the system is through the time-dependent factor to the interaction Hamiltonian via , which gives , where
(14) |
Due to the periodicity of the time-dependent coupling, with period , the Hamiltonian is also periodic: , and so is the dynamical matrix of the equation of motion: .
We can thus expand the time-dependent Hamiltonian as a Fourier series , with , yielding
(15) |
where
(16) |
Notably, we see that the static coupling rate, , is renormalized by the average of the time-dependent term [51]:
(17) |
and thus the mechanical drive renormalizes the dc Hamiltonian through an effective increase in , which is the undriven coupling term. Later, we will see such an effect in our simulations when is sufficiently large.
Note that separates into a time-independent part for and (for finite , including a shift due to ), and a time-dependent interaction. Thus, while is related to the time-dependent interaction, there is a static contribution from that term as well. Solving the time-dependent equation of motion (which is similar to the time-dependent Schrödinger equation with the Floquet theory), yields , where is the Floquet quasienergy [52], and the Floquet mode is -periodic [53, 54]. The Floquet solutions, , form a complete basis for any value of , thus , where , with . Resonances occur at differences between Floquet quasienergies [55].
To compute the Floquet modes [51], we use a Fourier series expansion of , where the Fourier coefficient vectors are Floquet sidebands. Although the dynamical matrix is so that one might expect nominally eigenenergies and thus 4 quasienergies confined within a energy range, we only see 2 lines corresponding to the two modes (the determinant has only two real roots, the polariton resonances) as the Floquet theory is applied to the equation of motion (not the Hamiltonian) representing the effective transitions–from the dressed polariton states to the dressed ground state, within the system [51]. Indeed, the coupling between these quasienergy lines is drive-dependent and for small enough , and , the drive might not have enough push to couple all the manifolds of the energy states [see the right-most part of Fig. 2(b)], and thus, only a limited number of them come into a Brillouin zone to make a coupling and transition possible. This will also reveal the dynamics of the two dressed polariton modes, that is equivalent in classical and quantum pictures [50].
Although we have made several key approximation in the theory above, we will see later that the resonances are in very good agreement with the full electrodynamic simulations. To be specific, PyCharge employs a full wave approach without any approximation based on nonperturbative, linear electromagnetic theory (derived from Maxwell’s equations), so it can capture the nondipolar, and non-Markovian aspects of the coupling, as well as the damping. However, the Floquet analysis above properly accounts for the main effect of the modulated dipole-dipole interaction including the intrinsic asymmetry stepping from the term, while giving a solid theoretical background to the main effects that we expect to see when we modulate the dipole-dipole coupling with a periodic drive. Moreover, the analysis can also be used to look at quantum effects for coupled two-level systems, which yields different spectral signatures than the coupled harmonic oscillator problem (primarily due to saturation effects and biexcitonic coupling). This will be the subject of future work.
IV Results and Simulations
In this section, we show the results of the PyCharge simulations in time and frequency domains, and show its agreement with the Floquet theory. We let the first dipole be initially excited (internal charges are within their maximum relative distance) and the second one initially in the ground state (internal charges are within their minimum relative distance). In Figs. 3, 4 and 5, we first investigate the time evolution of the scaled population of the excited states of the two coupled LOs (-dipoles) in their panels (a), then show the Fourier transformation of the time-dependent dipole moments that shows the spectral peaks of the transitions in their panels (b), and finally show their corresponding Floquet quasienergy diagrams in panels (c-d) to explain that the transitions occur in the differences of the quasienergies. We use a fixed value of the (static) coupling between the dipoles of , which is large enough to easily resolve the dipole-dipole splitting and coherent oscillations, since . In addition, we also choose various values for the mechanical drive’s frequency and dynamical coupling. Specifically, we first use and a rather low drive frequency of , which is already large enough to show clear new resonances, as shown in Fig. 3; in addition, we investigate and a higher drive frequency of in Fig. 4, and we increase the dynamical coupling by and a rather high drive frequency of in Fig. 5. Below we discuss each of these three regimes.

Figure 3(a) shows the PyCharge simulation result of the (scaled) population of the excited states of the two coupled dipoles with (Dipole 1′ and , solid lines) and without (Dipole 1 and 2, dashed lines) mechanical oscillations, plotted as a function of time with the following mechanical parameters: and (), where refers a cycle of mechanical oscillation period. The time is scaled by the free-space decay rate ( = 21.48 GHz). Dipole 1 and 1′ [nm] are initially excited, while Dipole 2 and 2′ are not [], corresponding to the schematic diagram in Fig. 2(a). All of the LOs have a natural angular frequency of 200 THz. To ensure very good accuracy, the numerical simulation runs over 10 million time iterations (125,663 periods of ) with a time step s []. The dipole-dipole initial separation is nm () along the -axis, with charge magnitude , effective mass and static coupling strength (, where represents the period of a cycle in which a photon is emitted and subsequently re-absorbed by the same dipole, or simply the round trip time in a classical picture). The time domain simulations show an oscillatory decay (caused by dipole-dipole interactions) of the populations to a steady state, where, with the onset of the mechanical oscillation, the phase of the oscillatory decay changes dynamically; thus, the time interval between the same peaks of the moving and stationary dipoles is sometimes shorter and sometimes longer. However, this oscillatory decay behavior for the mechanically oscillating dipoles is always faster than the stationary dipoles; this indicates the emergence of higher frequency peaks in the spectral analysis [cf. spectral peaks in panel (b) and also panel (c)].
Figure 3(b) displays the spectral peaks of the dipoles in the frequency domain (shown in a logarithmic scale) for the corresponding cases in Figs 3(a), centered at the dipoles’ transition frequency and normalized by . The spectra are obtained by a fast Fourier transform (FFT) of the dipole moment with a Hamming window function applied to enhance peak clarity and smoothness. We see that all of the resonance peaks from the field spectrum plots are separated by an integer number of the driving frequency ( integer) when , as shown in Fig. 3(c). The blue and red dot-dashed vertical lines in panel (b) are the same as the blue and red thin solid lines representing the Floquet quasienergies, calculated from Eq. (15), in panel (c), for the desired values of the parameter space pointed out by the vertical transparent gray line, and with a magnified view in panel (d). This spectral behavior contrasts significantly with that of two coupled LOs with a fixed COM, which produce only two frequency peaks (the well known superradiant and subradiant states, separated by , which also connect to Dicke states [22]). As anticipated, mechanical oscillation can modify these states to a manifold of dressed (Floquet) states, shown in Fig. 2, where energy levels are separated by integer multiples of .

Figure 3(c) shows the Floquet quasienergies near , scaled with , where , for two coupled HOs as a function of normalized oscillating amplitude () for the same regime as that of panels (a) and (b). Here, we include a truncated number of Brillouin zones (separated by the grey dotted lines). Construction of the Brillouin zones, which is dependent on the strength of the external drive, assists the coupling of the energy states by the external quanta; thus, with a limited (not enough) number of Brillouin zones, due to the small choices of drive frequency and amplitude, the complete intermixing of all the system’s eigenenergy states might not be possible within a Brillouin zone. Moreover, within the small range of the drive amplitude, we do not yet observe the creation of effective anticrossings in the dynamical picture between the quasienergy levels, which would only occur for sufficiently larger values of .
In coupling scenarios where , the spectral lines are initially separated by the integers of the driving frequency, as demonstrated in Fig. 3(c), where we normally expect a single peak structure within the range of one . We also see that at (and, in general, increasing the coupling strength via and ) we are well at the verge of entering into the regime of splitting the peak, as seen in panel (b), where also the initial splitting of the Floquet quasienergy lines almost begin to occur near the vertical grey line in panel (c); which is clearer in the magnified region in panel (d). However, a split in the spectral lines with an internal separation of is observed when (and, more effective for ), because the splitting of the quasienergy lines is now more effective and also they may undergo through an ac Stark shift. Additionally, the center of splits is still separated by an integer multiple of because still we encounter transition only between the self repetition manifolds (going back and forth in different Brillouin zones from a state) of the polariton states only.

Figure 4 is similar to Fig. 3, but with a larger driving frequency, (). Comparing the results with (primed dipoles) and without mechanical oscillation (unprimed dipoles), i.e., comparing panel (a)’s in Figs. 3 and 4, the driving frequency clearly alters the phase of the population oscillations in a different way due to time-dependent fluctuation in the dipole-dipole coupling strength. As shown in Fig. 4(a), we observe an overall faster exchange of energy between Dipole 1′ and 2′ (mechanical) compared with Dipole 1 and 2 (non-mechanical), when the driving frequency is greater than the dipole-dipole frequency shift, , indicating an increase of the frequency shift. In other words, for in Fig. 3(a), we see that the populations of mechanical dipoles (1’ and 2’) oscillate sometimes faster, sometimes slower than the non-mechanical cases (1 and 2). However, once we have a larger driving frequency, the population exchange of mechanically moving dipoles are always faster than the non-mechanical case. This can imply that for , one can expect an overall faster exchange of energy between Dipole 1′ and 2′ (mechanical).
Furthermore, with , we observe that the resonance peaks tend to exhibit a splitting into superradiant and subradiant states with internal separation around 2, as illustrated in Fig. 4(b). The inset of panel (b) provides an enlarged view of the peaks associated with the dressed polariton states. Additionally, the center of the splitting peaks is still separated by an integer of driving frequency. The vertical blue and red dot-dashed lines show these phenomena are aligned with our Floquet analysis, showing the excellent agreement between Floquet analysis and PyCharge simulations.
We next increase the oscillating amplitude that can bring the dipoles much closer to each other to explore the role of the oscillating strength (of the mechanical drive), again with , as shown in Fig. 5, where we now use a larger driving amplitude, . The change in the phase of population oscillations has not been significant, but the frequency of the population oscillations increases and decreases dramatically, as the effective is significantly reduced when the dipoles are brought closer together, and significantly increases when the dipoles are farther apart. This causes the resonances to be modified dynamically, as shown in Fig. 5(b-d).
We also find that the internal separation between resonance peaks got further split when is large enough. Our study indicates that the oscillating amplitude is able to influence the positioning of resonant frequencies, causing a splitting of the original peaks, as well as a change in the separation between the split peaks (when the amplitude is sufficiently large). The ensuing dynamic control is also corroborated by Floquet analysis, see panels (c) and (d) of Fig. 5, where we see the renormalization of the dc eigenenergies (dashed lines) and the ac Stark shift of quasienergies (thin solid lines) are becoming largely impactful and are well distinguishable from the dc eigenenergies (thick solid lines).
In Figs. 3, 4 and 5, Floquet quasienergies and states are obtained by analyzing the system’s Floquet dynamical matrix corresponding to the equation of motion, which was constructed in an extended Hilbert space incorporating time periodicity, similar to the approach in Ref. [51], and their correspondence to spectral features of the dipole moment . Hence, this approach confirms the presence of well-defined Floquet states, with spectral peaks appearing at integer multiples of the driving frequency. Additionally, the theoretical Floquet model was validated by the numerical simulations, providing insights into the effects of mechanical motion and dipole-dipole interactions on the system’s energy structure.
Additionally, a closer examination of the results in the Floquet analysis reveals that, when the parameter reaches sufficiently high values, the quasienergy states exhibit crossing/anticrossing behavior. The region of crossing and avoided crossing occurs more rapidly with increasing oscillating amplitude, resulting in a dispersion and merging of the resonances. This phenomenon is better illustrated in Fig. 5(c). The region of crossing/anticrossing shifts toward larger values of normalized driving amplitude as the frequency of the drive increases. Thus, the amplitude modulation may decreases as driving frequency increases, similar to the findings in [33], and the amplitude of this modulation decreases when the oscillation frequency of the mirror increases.
Finally, we remark that we have also investigated the influence of the mechanical oscillation phase, along with the consequences of altering the signs of sine and cosine functions. We found that these effects are minimal when the mechanical frequency is sufficiently high. However, they can become more pronounced at lower driving frequencies, which was not the domain of interest for this study.
V Conclusions
We have presented a direct numerical simulation of moving point charge potentials, using PyCharge, and studied how the (scaled) population dynamics and dipole spectra of two two-dipole coupled LOs can be significantly influenced by periodic mechanical oscillations, in a non-trivial way. We also investigated how the spectral characteristics and resonance peaks vary with the mechanical perturbation parameters, and , and showed how the resonance peaks positioning, spectral splitting of the peaks, as well as the separation between the split peaks change with an increased oscillating amplitude .
These results were fully corroborated with an intuitive and rigorous Floquet theory, where we added a periodic oscillation to the usual dipole-dipole coupling term (which can be derived from classical or quantum Maxwell equations), which couples two HOs through a time-dependent coherent coupling term. Both the direct numerical and Floquet analysis demonstrated a reciprocal relationship between and , with resonances showing a separation by an integer of driving frequency, namely , and these resonances further split into superradiant and subradiant states when drive frequency becomes larger than the static dipole-dipole coupling. Significantly, in scenarios with sufficiently large values of both and , there was a reduction in the radiative decay rate of the system, implying that mechanical effects can effectively enhance the coupling strength and extend the lifetime of the coupling system.
Related effects have been predicted in studies of an atom in front of an oscillating mirror (phonon-photon coupling) [33, 34], and time-modulated photonic crystals (time-modulated band-gap) [35], where the sidebands in the emitted spectrum are separated from the atomic transition frequency by the modulation frequency of the plate or photonic lattice, ; however, the lateral peaks here were strongly asymmetric due to the different density of states at the edges of the photonic band gap [35]. In contrast, for our case, which deals with finite-size dipoles in free space, the lateral peaks are symmetric because the photonic density of states (DOS) is essentially the same at frequencies of the peaks.
The advantage of PyCharge is that we do not need to make any of the usual approximations (dipole approximation and rotating wave approximation) and although everything is classical in nature, we are still able to show clear analogies with quantum optical systems (quantum emitters and quantized fields) [33, 34]. Additionally, we have demonstrated with a stronger oscillating strength or faster driving frequency, the resonance peaks can be further modified. We have gone beyond usual perturbative optomechanical studies, where they have linearized Taylor series expansion [56, 57, 58].
Overall, our findings provide a deeper understanding of the interplay between light and matter with mechanical motion, using coupled atoms described by LOs and the self-consistent fields. This work lays the foundation for exploring more complex optomechanical systems, potentially leading to novel applications in fundamental nano-optics, quantum information processing, and sensing and communication protocols (e.g., using coupled atoms as nano antennas). Moreover, it showcases the opportunities of using numerical solutions of retarded potentials and moving point charges and dipoles (using and extending the open-source PyCharge software), in contrast to the usual numerical Maxwell solvers with the physical fields (such as through FDTD and finite-element solvers).
Finally, we remark these classical studies can also be used to help develop more advanced quantum models, since a classical-quantum correspondence is only expected in some limit, e.g., linear response with no saturation effects (HO models). Thus, future work, for example, could look at the dynamical coupling between dipoles treated as two-level systems, where the eigenstates also contain a two quanta excitation, and saturation and nonlinear effects can become important.
Acknowledgements.
This work was supported by the Natural Sciences and Engineering Research Council of Canada (NSERC), the National Research Council of Canada (NRC), the Canadian Foundation for Innovation (CFI), and Queen’s University, Canada.References
- Milonni [1976] P. W. Milonni, Semiclassical and quantum-electrodynamical approaches in nonrelativistic radiation theory, Physics Reports 25, 1 (1976).
- Barnes et al. [2020] W. L. Barnes, S. A. R. Horsley, and W. L. Vos, Classical antennas, quantum emitters, and densities of optical states, Journal of Optics 22, 073501 (2020).
- Gerry et al. [2005] C. Gerry, P. Knight, and P. Knight, Introductory Quantum Optics (Cambridge University Press, 2005).
- Novotny and Hecht [2012] L. Novotny and B. Hecht, Principles of Nano-Optics, 2nd ed. (Cambridge University Press, 2012).
- Joulain et al. [2003] K. Joulain, R. Carminati, J.-P. Mulet, and J.-J. Greffet, Definition and measurement of the local density of electromagnetic states close to an interface, Phys. Rev. B 68, 245405 (2003).
- Forati [2022] E. Forati, Spontaneous emission rate and the density of states inside a one dimensional photonic crystal, IEEE Journal on Multiscale and Multiphysics Computational Techniques 7, 46 (2022).
- Purcell et al. [1946] E. M. Purcell, H. C. Torrey, and R. V. Pound, Resonance absorption by nuclear magnetic moments in a solid, Phys. Rev. 69, 37 (1946).
- Ogawa et al. [2004] S. Ogawa, M. Imada, S. Yoshimoto, M. Okano, and S. Noda, Control of light emission by 3D photonic crystals, Science 305, 227 (2004).
- Lodahl and van Driel et al. [2004] P. Lodahl and A. F. van Driel et al., Controlling the dynamics of spontaneous emission from quantum dots by photonic crystals., Nature 430, 654 (2004).
- Angelakis et al. [2004] D. Angelakis, P. Knight, and E. Paspalakis, Photonic crystals and inhibition of spontaneous emission: an introduction, Contemporary Physics 45, 303 (2004).
- Dignam et al. [2006] M. M. Dignam, D. P. Fussell, M. J. Steel, C. M. de Sterke, and R. C. McPhedran, Spontaneous emission suppression via quantum path interference in coupled microcavities, Phys. Rev. Lett. 96, 103902 (2006).
- Hughes [2007] S. Hughes, Coupled-cavity QED using planar photonic crystals, Phys. Rev. Lett. 98, 083603 (2007).
- Noda et al. [2007] S. Noda, M. Fujita, and T. Asano, Spontaneous-emission control by photonic crystals and nanocavities, Nature Photonics 1, 449 (2007).
- Lodahl et al. [2015] P. Lodahl, S. Mahmoodian, and S. Stobbe, Interfacing single photons and single quantum dots with photonic nanostructures, Rev. Mod. Phys. 87, 347 (2015).
- DeVoe and Brewer [1996] R. G. DeVoe and R. G. Brewer, Observation of superradiant and subradiant spontaneous emission of two trapped ions, Phys. Rev. Lett. 76, 2049 (1996).
- Eschner et al. [2001] J. Eschner, C. Raab, F. Schmidt-Kaler, and R. Blatt, Light interference from single atoms and their mirror images, Nature 413, 495 (2001).
- Mokhlespour et al. [2012] S. Mokhlespour, J. E. M. Haverkort, G. Slepyan, S. Maksimenko, and A. Hoffmann, Collective spontaneous emission in coupled quantum dots: Physical mechanism of quantum nanoantenna, Phys. Rev. B 86, 245322 (2012).
- Van Vlack et al. [2012] C. Van Vlack, P. T. Kristensen, and S. Hughes, Spontaneous emission spectra and quantum light-matter interactions from a strongly coupled quantum dot metal-nanoparticle system, Phys. Rev. B 85, 075303 (2012).
- Trebbia et al. [2022] J.-B. Trebbia, Q. Deplano, P. Tamarat, and B. Lounis, Tailoring the superradiant and subradiant nature of two coherently coupled quantum emitters, Nature Communications 13, 10.1038/s41467-022-30672-2 (2022).
- Holzinger et al. [2021] R. Holzinger, M. Moreno-Cardoner, and H. Ritsch, Nanoscale continuous quantum light sources based on driven dipole emitter arrays, Applied Physics Letters 119, 024002 (2021).
- Angelatos and Hughes [2015] G. Angelatos and S. Hughes, Entanglement dynamics and mollow nonuplets between two coupled quantum dots in a nanowire photonic-crystal system, Phys. Rev. A 91, 051803 (2015).
- Dicke [1954] R. H. Dicke, Coherence in Spontaneous Radiation Processes, Phys. Rev. 93, 99 (1954).
- Filipovich and Hughes [2022] M. J. Filipovich and S. Hughes, PyCharge: An open-source python package for self-consistent electrodynamics simulations of lorentz oscillators and moving point charges, Computer Physics Communications 274, 108291 (2022).
- Dalibard et al. [1992] J. Dalibard, E. d’Eté de Physique Théorique, J. Raimond, U. J. Fourier, J. Zinn-Justin, U. J. F. de Grenoble, and N. A. S. Institute, Fundamental Systems in Quantum Optics, Les Houches 1990 (North-Holland, 1992).
- Boussiakou et al. [2002] L. G. Boussiakou, C. R. Bennett, and M. Babiker, Quantum theory of spontaneous emission by real moving atoms, Phys. Rev. Lett. 89, 123001 (2002).
- Cresser and Barnett [2003] J. D. Cresser and S. M. Barnett, The rate of spontaneous decay of a moving atom, Journal of Physics B: Atomic, Molecular and Optical Physics 36, 1755 (2003).
- Aspelmeyer et al. [2014a] M. Aspelmeyer, T. Kippenberg, and F. Marquardt, Cavity Optomechanics: Nano- and Micromechanical Resonators Interacting with Light, Quantum Science and Technology (Springer Berlin Heidelberg, 2014).
- Aspelmeyer et al. [2014b] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity optomechanics, Rev. Mod. Phys. 86, 1391 (2014b).
- Kippenberg and Vahala [2007] T. Kippenberg and K. Vahala, Cavity opto-mechanics, Opt. Express 15, 17172 (2007).
- Dorner and Zoller [2002] U. Dorner and P. Zoller, Laser-driven atoms in half-cavities, Phys. Rev. A 66, 023816 (2002).
- Wilson et al. [2003] M. A. Wilson, P. Bushev, J. Eschner, F. Schmidt-Kaler, C. Becher, R. Blatt, and U. Dorner, Vacuum-field level shifts in a single trapped ion mediated by a single distant mirror, Physical Review Letters 91, 10.1103/physrevlett.91.213602 (2003).
- Beige et al. [2002] A. Beige, J. Pachos, and H. Walther, Spontaneous emission of an atom in front of a mirror, Physical Review A 66 (2002).
- Glaetzle et al. [2010] A. Glaetzle, K. Hammerer, A. Daley, R. Blatt, and P. Zoller, A single trapped atom in front of an oscillating mirror, Optics Communications 283, 758 (2010).
- Ferreri et al. [2019] A. Ferreri, M. Domina, L. Rizzuto, and R. Passante, Spontaneous emission of an atom near an oscillating mirror, Symmetry 11, 10.3390/sym11111384 (2019).
- Calajò et al. [2017] G. Calajò, L. Rizzuto, and R. Passante, Control of spontaneous emission of a single quantum emitter through a time-modulated photonic-band-gap environment, Phys. Rev. A 96, 023802 (2017).
- Roelli et al. [2015] P. Roelli, C. Galland, N. Piro, and T. J. Kippenberg, Molecular cavity optomechanics as a theory of plasmon-enhanced raman scattering, Nature Nanotechnology 11, 164–169 (2015).
- Dezfouli et al. [2019] M. K. Dezfouli, R. Gordon, and S. Hughes, Molecular optomechanics in the anharmonic cavity-QED regime using hybrid metal-dielectric cavity modes, ACS Photonics 6, 1400–1408 (2019).
- Esteban et al. [2022] R. Esteban, J. J. Baumberg, and J. Aizpurua, Molecular optomechanics approach to surface-enhanced raman scattering, Accounts of Chemical Research 55, 1889 (2022).
- Patra et al. [2023] B. Patra, B. Kafle, and T. G. Habteyes, Molecular optomechanics induced hybrid properties in soft materials filled plasmonic nanocavities, Nano Letters 23, 5108 (2023), pMID: 37225673.
- Batignani et al. [2020] G. Batignani, C. Ferrante, and T. Scopigno, Accessing excited state molecular vibrations by femtosecond stimulated raman spectroscopy, The Journal of Physical Chemistry Letters 11, 7805 (2020), pMID: 32841039.
- Filipovich and Hughes [2021] M. J. Filipovich and S. Hughes, Space-time computation and visualization of the electromagnetic fields and potentials generated by moving point charges, American Journal of Physics 89, 482 (2021).
- Förster [1948] T. Förster, Zwischenmolekulare energiewanderung und fluoreszenz, Annalen der Physik 437, 55 (1948).
- Bukov et al. [2015] M. Bukov, L. D’Alessio, and A. Polkovnikov, Universal high-frequency behavior of periodically driven systems: from dynamical stabilization to floquet engineering, Advances in Physics 64, 139 (2015).
- Oka and Kitamura [2019] T. Oka and S. Kitamura, Floquet engineering of quantum materials, Annual Review of Condensed Matter Physics 10, 387 (2019).
- Mori [2023] T. Mori, Floquet states in open quantum systems, Annual Review of Condensed Matter Physics 14, 35 (2023).
- Milonni and Eberly [2010] P. W. Milonni and J. H. Eberly, Laser Physics (John Wiley Sons, Ltd, 2010) Chap. 9, pp. 401–455.
- Wubs et al. [2003] M. Wubs, L. G. Suttorp, and A. Lagendijk, Multipole interaction between atoms and their photonic environment, Phys. Rev. A 68, 013822 (2003).
- Kristensen et al. [2011] P. T. Kristensen, J. Mørk, P. Lodahl, and S. Hughes, Decay dynamics of radiatively coupled quantum dots in photonic crystal slabs, Phys. Rev. B 83, 075305 (2011).
- Muniain et al. [2025] U. Muniain, J. Aizpurua, R. Hillenbrand, L. Martín-Moreno, and R. Esteban, Description of ultrastrong light–matter interaction through coupled harmonic oscillator models and their connection with cavity-qed hamiltonians, Nanophotonics doi:10.1515/nanoph-2024-0528 (2025).
- Hughes et al. [2024] S. Hughes, C. Gustin, and F. Nori, Reconciling quantum and classical spectral theories of ultrastrong coupling: role of cavity bath coupling and gauge corrections, Optica Quantum 2, 133 (2024).
- Akbari et al. [2025] K. Akbari, F. Nori, and S. Hughes, Floquet engineering the quantum Rabi model in the ultrastrong coupling regime, Phys. Rev. Lett. 134, 063602 (2025).
- Nikishov and Ritus [1964] A. I. Nikishov and V. I. Ritus, Quantum Processes in the Field of a Plane Electromagnetic Wave and in a Constant Field 1, Sov. Phys. JETP 19, 529 (1964).
- Floquet [1883] G. Floquet, Sur les équations différentielles linéaires à coefficients périodiques, Annales scientifiques de l’École Normale Supérieure 2e série, 12, 47 (1883).
- Chicone [2006] C. Chicone, Ordinary Differential Equations with Applications, Texts in Applied Mathematics (Springer New York, 2006).
- Shirley [1965] J. H. Shirley, Solution of the schrödinger equation with a hamiltonian periodic in time, Phys. Rev. 138, B979 (1965).
- Agarwal [2012] G. S. Agarwal, Quantum Optics (Cambridge University Press, 2012).
- Chiangga et al. [2024] S. Chiangga, T. Sunpatanon, and T. D. Frank, Perturbation theoretical approach to determine optomechanical entanglement in mirror-field systems, Europhysics Letters 146, 65001 (2024).
- Primo et al. [2020] A. G. Primo, N. C. Carvalho, C. M. Kersul, N. C. Frateschi, G. S. Wiederhecker, and T. P. M. Alegre, Quasinormal-mode perturbation theory for dissipative and dispersive optomechanics, Phys. Rev. Lett. 125, 233601 (2020).
- Jackson [1999] J. D. Jackson, Classical electrodynamics, 3rd ed. (Wiley, New York, NY, 1999).
- Wiechert [1901] E. Wiechert, Elektrodynamische elementargesetze, Annalen der Physik 309, 667 (1901).
- Griffiths [2013] D. J. Griffiths, Introduction to electrodynamics (Pearson, 2013).
- Dung et al. [2002] H. T. Dung, L. Knöll, and D.-G. Welsch, Resonant dipole-dipole interaction in the presence of dispersing and absorbing surroundings, Phys. Rev. A 66, 063810 (2002).
- Martin and Piller [1998] O. J. F. Martin and N. B. Piller, Electromagnetic scattering in polarizable backgrounds, Phys. Rev. E 58, 3909 (1998).
- Scheel et al. [1999] S. Scheel, L. Knöll, and D.-G. Welsch, Spontaneous decay of an excited atom in an absorbing dielectric, Phys. Rev. A 60, 4094 (1999).
- Cohen-Tannoudji et al. [1997] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Photons and Atoms - Introduction to Quantum Electrodynamics (Wiley-Interscience, New York, 1997).
- Hopfield [1958] J. J. Hopfield, Theory of the contribution of excitons to the complex dielectric constant of crystals, Phys. Rev. 112, 1555 (1958).
- De Liberato [2014] S. De Liberato, Light-matter decoupling in the deep strong coupling regime: The breakdown of the purcell effect, Phys. Rev. Lett. 112, 016401 (2014).
- Miao and Agarwal [2023] Q. Miao and G. S. Agarwal, Polaritonic ultrastrong coupling: Quantum entanglement in the ground state, Phys. Rev. Res. 5, 033136 (2023).
Appendix A Electrodynamics of interacting and oscillating dipoles in PyCharge
In classical electrodynamics, the analytical solution for the electric and magnetic fields at position from the COM of an idealized electric dipole at time , in a homogeneous medium, is given by [59, 4]
(18) |
and
(19) |
where , is the time-dependent harmonic oscillator dipole moment, , and is unit vector of . Note here that the two-body charges constituting the dipole oscillate but the COM of the dipole is at rest. In free space, then .
When the COM of a dipole is in motion, one still can use Eqs. (18) and (19) to calculate the electromagnetic fields from a moving charge density by changing provided that the motion is with a constant velocity throughout time (nonaccelerating). However, when the motion is with time-varying velocity , the contributions from the retarded and advanced time difference between the source and the observation point play a role and hence contribute to radiation. The EM field of an accelerated charge density can be calculated via well-known Liénard-Wiechert potentials [59].
The charge and current densities of a point charge at the position with velocity are, respectively,
(20) |
and
(21) |
The scalar and vector potentials associated with a moving point charge in the Lorenz gauge, known as the Linard-Wiechert potentials [60, 61], are derived from Maxwell’s equations, and can be written as
(22) |
and
(23) |
where is the distance from source points, and , such that is the unit vector from the position of the point charge to the field point and is the velocity of the point charge expressed as a fraction of the speed of light. The quantity in brackets, evaluated at the retarded time , is given by
(24) |
where the delay time can be solved numerically, e.g., by the secant method or Newton’s method.
We can subsequently calculate the electric and magnetic fields of a point charge in arbitrary motion, using the solution of the Linard-Wiechert potentials. The calculation of electromagnetic fields and potentials of dipoles in PyCharge applies the principle of superposition of classical electrodynamics, which is the sum of individual contributions of each source. Each contribution is calculated along discretized spatial grids at a specific time. The physical (gauge-invariant) relativistically-correct, time-varying total electric field is given by a sum of the electric Coulomb (velocity) and radiation (acceleration) fields [61, 59], , where
(25) |
and
(26) |
with the velocity of the charge, the acceleration of the charge, and
(27) |
where is the position of the charge at the retarded time, and is the current position of the charge.
Appendix B Dipole-coupled harmonic oscillator theory using two electric dipoles in the near field regime
B.1 Quantum optical approach for two coupled dipoles
For any general medium, the Born-Markov reduced master equation for two coupled dipoles (treated as quantized HOs), in the rotating wave approximation, is [62, 21]
(28) |
where we define , and assume equal dipole resonances and dipole moments, so that (including a trivial Lamb shift, ). Thus all the coupling rates above are implicitly evaluated at . These rates are defined in the main text, and explicitly, for the relevant coherent dipole-dipole coupling term, we have
(29) |
Neglecting the dissipation, then we obtain the relevant effective Hamiltonian,
(30) |
where . Furthermore, if not adopting a rotating wave approximation, then
(31) |
Using this to obtain the Heisenberg equation of motion, with , where , yields the main matrix equation in the main text, for which we use to construct a Floquet solution. Moreover, in the case of dipoles in free space, then
(32) |
which is justified in more detail below, along with a connection to a classical model of dipole-dipole coupling, where a direct correspondence is shown.
B.2 Direct classical approach from the Green’s function formalism with two dipole oscillators
The same result (resonances and Hamiltonian matrix form) can be obtained directly from the classical electromagnetic theory and the Green’s function formalism. The interaction between two dipolar charge distribution are often described via the reciprocating electrostatic Coulomb force between them identified by, within the dipole approximation (and near field regime), but one can also use the Green’s function formalism in the presence of scattering bodies to find the total electric field in either of the dipoles.
The exact dyadic Green’s function solution with the source point at position , two point-dipoles (scatterers) at positions and , and observation point at position , in free space, is obtained via the self-consistent solution to the Dyson equation [63], including the scattering properties of all scatterers in the background medium, via
(33) |
where is the Green’s tensor of the background medium and is the polarizability tensor of the th dipole. We are interested in finding the total electric field at one of the dipoles from the other one, e.g. , containing the scattering effect of both dipoles. The superscript ‘(2)’ indicates we have the solution with two dipoles includes.
Here, we have adopted an exact iterative approach to write a series of Dyson equations given by [48], as
(34) |
where
(35) |
and,
(36) |
with
(37) |
and also . One can put all these together, now that we have all of the contributions only in terms of the background Green’s tensor, and obtain a unified formula for the total two-body scattering Green’s tensor in terms of the background Green’s tensor and the dipoles’ polarizabilities only.
The dipoles of interest have a preferred polarization, such as -polarized in the main text, where . Hence, the desired Green’s tensor reads [48]
(38) |
where
(39) |
where , and
(40) |
is the semiclassical formula for the polarizability (units of volume) of a LO when using the dipole definition ; the connection between this polarizibility model and the classical mass model is obtained via , as also mentioned in the main text. Note that in the limit of neglecting the dipoles’ self-interaction (small Lamb shift and background radiative decay), then . It is now easy to obtain the spectral resonances for any medium described through .
For a homogeneous background system, the Green function is known analytically [64]:
(41) |
with and . Interestingly, we can regard our desired near-field coupling term as either coming from the total fields, transverse and longitudinal, or in terms of only the longitudinal fields. This is because , where
(42) |
and
(43) |
We highlight that these expressions include a sum over all modes and the two Green’s tensors satisfy a fundamental completeness relation, .
Next, for the interaction between the dipoles one can simply employ the dyadic Green’s function approach and write the interaction energy between the two dipole via the well-known relation of [65, 59, 48] , with , as
(44) |
in agreement with Eq. (5) of the main text, such that . For our interests here, we can neglect radiative loss and also considering two dipoles in close proximity, in free space []; then for two -polarized dipoles, we include the near-field contributions (only to the Green’s function), so that
(45) |
where . Therefore, the classical Hamiltonian simply becomes , where the interaction term , as denoted above, is only characterized by the position-position interaction via the Green’s tensor.
The normal mode frequencies for the hybrid system is then calculated via the roots of the determinant of the classical dynamical matrix for the equations of motion, yielding the solution
(46) |
which simplifies to when , as in the main text.
Lastly, the second quantization of the dipole-dipole Hamiltonian can simply be done by promoting the normal-mode complex amplitudes to operators via and to write
(47) |
precisely coinciding with Eq. (31) and Eq. (12) of the main text. As mentioned also in the main text, this form is identical to a Hopfield-like model Hamiltonian without the diamagnetic term [66, 67, 68], and is justified here rigorously from Maxwell’s equations (which of course are used to construct the Green’s function solutions above).