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

The universal model of strong coupling at the nonlinear parametric resonance in open cavity-QED systems

Mikhail Tokman Institute of Applied Physics, Russian Academy of Sciences, Nizhny Novgorod, 603950, Russia    Maria Erukhimova Institute of Applied Physics, Russian Academy of Sciences, Nizhny Novgorod, 603950, Russia    Qianfan Chen Department of Physics and Astronomy, Texas A&M University, College Station, TX, 77843 USA    Alexey Belyanin Department of Physics and Astronomy, Texas A&M University, College Station, TX, 77843 USA
Abstract

Many molecular, quantum-dot, and optomechanical nanocavity-QED systems demonstrate strong nonlinear interactions between electrons, photons, and phonon (vibrational) modes. We show that such systems can be described by a universal model in the vicinity of the nonlinear parametric resonance involving all three degrees of freedom. We solve the nonperturbative quantum dynamics in the strong coupling regime of the nonlinear resonance, taking into account quantization, dissipation, and fluctuations of all fields. We find analytic solutions for quantum states in the rotating wave approximation which demonstrate tripartite quantum entanglement once the strong coupling regime is reached. We show how the strong coupling at the nonlinear resonance modifies photon emission and vibrational spectra, and how the observed spectra can be used to extract information about relaxation rates and the nonlinear coupling strength in specific systems.

I Introduction

Nonlinear optical interactions acquire qualitatively new features in the strong-coupling regime of cavity quantum electrodynamics (QED), especially when utilizing an extreme field localization achievable in nanophotonic cavities. Even in the standard cavity QED scenario of the strong coupling at the two-wave resonance between only two degrees of freedom, e.g., between an electronic or vibrational transition in a molecule and an electromagnetic (EM) cavity mode, strong coupling has been shown to modify the properties of Raman scattering, generation of harmonics, four-wave mixing, and nonlinear parametric interactions, with applications in photochemistry, quantum information, and quantum sensing; see, e.g., schmidt2016 ; pino2015 ; pino2016 ; ramelow2019 ; neuman2020 ; ge2021 ; k-wang2021 ; narang2021 ; wang2021 ; zasedatelev2021 ; pscherer2021 ; hughes2021 .

The dynamics becomes more complicated but also more interesting when the strong coupling regime is realized at the nonlinear resonance between three or more degrees of freedom. We will call this nonlinear resonance “parametric” for lack of a better word, although this term is often applied to nonlinear processes involving only the EM fields, in which there is no energy exchange between the fields and the medium, for example parametric down-conversion in transparent nonlinear materials LLcont . In contrast, for all examples in this paper the energy exchange between the bosonic fields and fermionic quantum emitters is of principal importance; moreover, we are interested in the nonperturbative regime of strong coupling when the excitation of the fermionic degree of freedom is not small. This is the regime most interesting for quantum technology applications as it actively involves the fermionic “qubit” in the processes of writing, reading, and transferring the information encoded in a quantum state.

In this paper we will focus on the nonlinear three-wave resonance between three degrees of freedom for the sake of simpler algebra, although the formalism does not have this limitation. One possible example where it can be realized is a molecule or an ensemble of molecules placed in a photonic or plasmonic nanocavity; e.g. benz2016 ; park2016 ; chikkaraddy2016 ; pscherer2021 ; hughes2021 . In this case the fermion system may comprise two or more electron states forming an optical transition at frequency ωe\omega_{e}, and the nonlinear parametric process may involve, e.g., a decay of the electron excitation into a cavity photon at frequency ω\omega and a phonon of a given vibrational mode of a molecule at frequency Ω\Omega, under the nonlinear resonance condition ωe=ω+Ω\omega_{e}=\omega+\Omega, or an absorption of a photon with simultaneous creation of electron and phonon excitations, given by the nonlinear resonance condition ω=ωe+Ω\omega=\omega_{e}+\Omega. When the strength of such a nonlinear three-wave interaction is higher than the dissipation rates, hybrid electron-photon-phonon states are formed. If the phonon mode is classical, the parametric process is simply the modulation of the electron-photon coupling by molecular vibrations which serve as an external driving force for the electron-photon quantum dynamics. If the phonon mode is quantized, the strong coupling between photon, phonon, and electron degrees of freedom near parametric resonance ωe=ω±Ω\omega_{e}=\omega\pm\Omega leads inevitably to the formation of tripartite entangled states belonging to the family of Greenberger-Horne-Zeilinger (GHZ) states tokman2021 .

Another route to the parametric resonance is within the framework of cavity optomechanics aspelmeyer2014 ; meystre2013 ; roelli2016 ; pirkkalainen2015 and quantum acoustics chu2017 ; hong2017 ; arriola2019 . It can occur in the situations where mechanical oscillations of a cavity parameter at frequency Ω\Omega modulate the resonance between the optical transition in a quantum emitter and a photon cavity mode. Here again the parametric resonance ωe=ω±Ω\omega_{e}=\omega\pm\Omega at strong coupling should give rise to tripartite entangled states of the electrons, photons, and mechanical vibrations liao2018 . While quantization of all three degrees of freedom in experiment remains an unsolved challenge, strong coupling and entanglement of acoustic phonons satzinger2018 ; bienfait2020 , resolving the energy levels of a nanomechanical oscillator arriola2019 , or cooling a macroscopic system into its motional ground state delic2020 have already been demonstrated.

Yet another situation leading to parametric resonance is when a quantum emitter such as a quantum dot or an optically active defect in a solid matrix is coupled to the EM cavity field and the phonon modes of a crystal lattice wilson2002 . In this case the coupling to phonons can be introduced via the same Huang-Rhys theory as in the molecular systems weiler2012 ; roy2015 ; denning2020 ; see the Hamiltonian Eq. (14) below. The exciton-photon coupling strength (Rabi frequency) in nanocavities can be even high enough to exceed the phonon frequency, which brings the system to the ultrastrong coupling regime denning2020 .

There is a great variety of models and formalisms describing these diverse physical systems, and attempts have been made to establish connections between different models. For example, it was shown in roelli2016 that plasmon-enhanced Raman scattering on individual molecules can be mapped onto a cavity optomechanics Hamiltonian when the plasmon frequency is far detuned from the electronic transition in a molecule, i.e., no electrons are excited. In this case the vibrational mode of a molecule is analogous to the mechanical oscillations of a cavity parameter. In pino2016 it was argued that Stokes Raman scattering in molecules under the condition of a strong coupling between the vibrational mode and the cavity mode is equivalent to the parametric decay of the pump photon into the Stokes photon and the vibrational quantum. In both cases a simple linear resonance ω=Ω\omega=\Omega was assumed and the EM field was far detuned from any electronic transition in a molecule, thus excluding any real electron excitation. In hughes2021 resonant Raman scattering of single molecules when the two-wave exciton-photon coupling frequency is comparable to the vibrational frequency was analyzed. The situation when the Rabi frequency becomes comparable to the vibrational frequency was also considered in tokman2021 .

In this paper we deal with the strong coupling at the nonlinear parametric resonance ωe=ω±Ω\omega_{e}=\omega\pm\Omega when the excitation of the electron transition and energy exchange between all three degrees of freedom are of principal importance. We show that in the RWA (i.e., excluding ultrastrong-coupling regimes) all physical models of electron-photon-vibrational coupling in molecular, optomechanical, and any other coupled three-mode system can be mapped onto the universal “parametric” Hamiltonian, independently on the specific physical mechanism of coupling. Note that the system must be in the RWA regime for the nonlinear parametric resonance and all associated physics to exist and make sense. Otherwise the three-wave and two-way resonances overlap tokman2021 and the GHZ-like entangled states cannot be created. Also, it becomes impossible to build a universal Hamiltonian. That is why the ultrastrong-coupling regimes are not of interest to us in this paper.

When solving for the quantum dynamics of the resulting nonlinear coupled system, we include the effects of decoherence and coupling of each dynamic subsystem (electrons, photons, and phonons) to its own reservoir in the Markov approximation. Previous works (see, e.g., roy2015 ; denning2020 ) included the non-Markovian effects in the coupling of the two-wave exciton-photon resonance to the phonon reservoir. In our case the phonon mode, which is strongly coupled to the exciton and photon modes through the nonlinear resonance, is part of the dynamical system. One could say that the phonon effect on the dynamics is “extremely non-Markovian”, except that this terminology ceases to have any meaning in this case. The Markov approximation is of course related only to (weak) coupling of all components of the dynamic system to their dissipative reservoirs.

Within the formalism of the stochastic equation of evolution for the state vector we are able to find the general analytic solution for the nonperturbative dynamics of the open quantum system. This approach is well known zoller1997 , but it is usually applied for numerical Monte-Carlo simulations scully1997 ; plenio1998 ; gisin1992 ; diosi1998 ; tannoudji1993 ; molmer1993 ; gisin1992-2 ; raedt2017 ; nathan2020 . We recently developed a version of stochastic Schrödinger equation suitable for analytic solutions in open strongly-coupled cavity QED problems tokman2021 ; chen2021 . We calculate the photon and phonon emission spectra to obtain the experimentally observable signatures of the strong coupling regime and tripartite quantum entanglement. It is remarkable that we are able to find analytic solutions for the quantum dynamics in systems of coupled electron, photon, and phonon excitations even including dissipation and fluctuation effects in all subsystems. This allows one to find the expressions for the time evolution of the state vector and observable quantities in the form which shows explicitly the dependence on all experimental parameters: transition energies and frequencies, matrix elements of the optical transitions, the spatial structure of the field modes, relaxation rates for all constituent subsystems, ambient temperatures etc. We believe that the results obtained in this paper will be useful for designing and interpreting the experiments on a broad range of cavity QED systems.

The paper is structured as follows. In Section II we present the Hamiltonian for coupled quantized fermion, photon, and phonon fields near the parametric resonance for one particular mechanism of three-wave coupling. In Section III we show that a large variety of different three-wave coupling mechanisms and physical systems are reduced to the same Hamiltonian which therefore can serve as a universal model of the parametric resonance. Section IV includes the effects of dissipation, decoherence, and fluctuations within the stochastic equation for the state vector which describes the evolution of an open system in contact with dissipative reservoirs. As compared to our recent work tokman2021 ; chen2021 , we develop a model of fluctuations and dissipative processes which includes all effects of phonon dissipation on the dynamics of the parametric process and the emission spectra. In Section V we describe the formation of entangled electron-photon-phonon states for an open system. Section VI calculates the emission spectra of photons and phonons resulting from the nonlinear parametric decay of an electron excitation. Appendix A contains the derivation details for some results in Sec. V.

II The universal Hamiltonian of the nonlinear parametric resonance

Consider a simple model of three interacting quantum subsystems which includes (1) an electron transition in a quantum emitter such as an atom, molecule, optically active impurity, quantum dot, etc., which we will model as a 2-level system, (2) a single-mode electromagnetic (EM) field in a cavity, and (3) a mode of mechanical, acoustic, or molecular vibrations (“phonons”). Although in this section we write the Hamiltonian for a specific model, in the next section we show that the same Hamiltonian describes the nonlinear parametric coupling in a variety of physical systems.

The generalization to many bosonic modes or fermionic degrees of freedom is straightforward and still allows analytic solution within the rotating wave approximation (RWA), but it leads to more cumbersome algebra (see, e.g., tokman2021-3 ), so we will keep only three degrees of freedom for clarity.

Figure 1 shows a generic model of parametric decay of an electron excitation in a quantum emitter (e.g., a molecule) into a photon of a cavity mode at frequency ω\omega and a phonon of a given vibrational mode at frequency Ω\Omega, under the condition of the parametric resonance ωeω+Ω\omega_{e}\approx\omega+\Omega.

Refer to caption
Figure 1: A sketch of nonlinear parametric resonance for a molecule in a cavity showing the decay of the electron excitation at frequency ωe\omega_{e} into a cavity mode photon at frequency ω\omega and a phonon of a given vibrational mode at frequency Ω\Omega. The relaxation rates of the electron, photon, and vibrational excitations are γ\gamma, μω\mu_{\omega}, and μΩ\mu_{\Omega}, respectively.

In the absence of coupling, the partial Hamiltonians are:

II.1 The 2-level fermion system

It is described by a standard effective Hamiltonian

H^e=ωeσ^σ^.\hat{H}_{e}=\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}. (1)

Here σ^=|01|\hat{\sigma}=\left|0\right\rangle\left\langle 1\right|, σ^=|10|\hat{\sigma}^{\dagger}=\left|1\right\rangle\left\langle 0\right| ; |0\left|0\right\rangle and |1\left|1\right\rangle are the eigenstates of an “atom” with energies 0 and ωe\hbar\omega_{e} respectively. The Hamiltonian (1) corresponds to the dipole moment operator

𝐝^=𝐝(σ^+σ^),\mathbf{\hat{d}}=\mathbf{d}\left(\hat{\sigma}^{\dagger}+\hat{\sigma}\right), (2)

where 𝐝=𝐞1|𝐫|0\mathbf{d=-e}\left\langle 1\right|\mathbf{r}\left|0\right\rangle, 𝐫\mathbf{r} is a coordinate for the finite motion of a bound electron.

II.2 The EM field

Here we consider a single-mode EM field for simplicity, although including many bosonic field modes does not present any principal difficulties. Besides, in a micro- or nanocavity other EM modes will be far detuned from the nonlinear resonance. The Hamiltonian is

H^em=ωc^c^.\hat{H}_{em}=\hbar\omega\hat{c}^{\dagger}\hat{c}. (3)

Here c^\hat{c} and c^\hat{c}^{\dagger} are standard bosonic annihilation and creation operators of photons or plasmons in the EM mode of frequency ω\omega . The electric field operator is

𝐄^=𝐄(𝐫)c^+𝐄(𝐫)c^.\mathbf{\hat{E}}=\mathbf{E}\left(\mathbf{r}\right)\hat{c}+\mathbf{E}^{\ast}\left(\mathbf{r}\right)\hat{c}^{\dagger}. (4)

The spatial structure of the normalization amplitude of the field 𝐄(𝐫)\mathbf{E}\left(\mathbf{r}\right) is determined by solving the boundary value problem. The normalization condition is

V[ω2ε(ω,𝐫)]ωω𝐄(𝐫)𝐄(𝐫)d3r=4πω.\int_{V}\frac{\partial\left[\omega^{2}\varepsilon\left(\omega,\mathbf{r}\right)\right]}{\omega\partial\omega}\mathbf{E}^{\ast}\left(\mathbf{r}\right)\mathbf{E}\left(\mathbf{r}\right)d^{3}r=4\pi\hbar\omega. (5)

Here VV is the quantization volume, ε(ω,𝐫)\varepsilon\left(\omega,\mathbf{r}\right) the dielectric function of the dispersive medium which fills in the resonator. Eq. (5) is derived, e.g., in tokman2016 ; tokman2013 ; tokman2015 ; tokman2018 .

II.3 The phonons

We again assume a single bosonic mode of a vibrational field for the same reasons,

H^p=Ωb^b^,\hat{H}_{p}=\hbar\Omega\hat{b}^{\dagger}\hat{b}, (6)

where b^\hat{b} and b^\hat{b}^{\dagger} are phonon annihilation and creation operators. Depending on the situation, they may define, e.g., the radius-vector of oscillations of the center of mass of an atom tokman2021 ; neuman2020 ; felipe2017 or a geometric parameter of the optomechanical cavity roelli2016 ; aspelmeyer2014 ,

𝐑^=𝐐b^+𝐐b^.\mathbf{\hat{R}}=\mathbf{Q}\hat{b}+\mathbf{Q}^{\ast}\hat{b}^{\dagger}. (7)

The normalization amplitude 𝐐\mathbf{Q} depends on the system; its absolute value can be expressed through an effective mass of the quantum mechanical oscillator aspelmeyer2014 : |𝐐|2=2meffΩ\left|\mathbf{Q}\right|^{2}=\frac{\hbar}{2m_{eff}\Omega}.

II.4 The coupling

The coupling between subsystems is strongest at resonance. Since usually ω,ωeΩ\omega,\omega_{e}\gg\Omega, two most relevant resonances are a two-wave resonance,

ωeω\omega_{e}\approx\omega (8)

and a three-wave (parametric) resonance,

ωeω±Ω\omega_{e}\approx\omega\pm\Omega (9)

There could be also harmonic resonances at the harmonics of the phonon frequency: ωeω±MΩ\omega_{e}\approx\omega\pm M\Omega, where MM is integer. The modulation of the system parameters by a classical phonon field at frequency Ω\Omega was studied, e.g., in chen2021 , and we don’t consider the classical field here.

The two-wave resonance in the RWA scully1997 is desrcibed by the Jaynes-Cummings (JC) Hamiltonian jaynes1963 ,

H^=ωc^c^+ωeσ^σ^+(ΩR(2)σ^c^+h.c.).\hat{H}=\hbar\omega\hat{c}^{\dagger}\hat{c}+\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}+\hbar\left(\Omega_{R}^{\left(2\right)}\hat{\sigma}^{\dagger}\hat{c}+h.c.\right). (10)

The electric dipole coupling between the electron and EM subsystems is expressed here through the effective Rabi frequency for the two-wave coupling, ΩR(2)=\Omega_{R}^{\left(2\right)}= 𝐝𝐄(𝐫0)-\frac{\mathbf{d}\cdot\mathbf{E}\left(\mathbf{r}_{0}\right)}{\hbar}, where 𝐫0\mathbf{r}_{0} is the coordinate of a point-like atom.

A three-wave (parametric) resonance appears in many different scenarios. As we show in the next section, various models existing in the literature can be described with one universal Hamiltonian. One of the scenarios leading to the universal three-wave coupling Hamiltonian is when a quantized phonon (vibrational) mode modulates the coupling energy between the electron transition and the EM field mode. In this case the JC Hamiltonian is generalized to the following form tokman2021 ; pino2016 :

H^=ωc^c^+ωeσ^σ^+Ωb^b^+(ΩR(3)σ^c^b^+h.c.),\hat{H}=\hbar\omega\hat{c}^{\dagger}\hat{c}+\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}+\hbar\Omega\hat{b}^{\dagger}\hat{b}+\hbar\left(\Omega_{R}^{\left(3\right)}\hat{\sigma}^{\dagger}\hat{c}\hat{b}+h.c.\right), (11)

where ΩR(3)\Omega_{R}^{\left(3\right)} is the coupling parameter for the three-wave resonance, which depends on the specific coupling mechanism. The above interaction term is written for the decay of an electron transition into the photon and the phonon, i.e., assuming that the electron excitation energy is the largest of the three. This decay process corresponds to the upper (plus) sign in the resonant condition (9). If we choose the lower (minus) sign in Eq. (9), the three-wave coupling Hamiltonian will become (ΩR(3)c^b^σ^+h.c.)\hbar\left(\Omega_{R}^{\left(3\right)}\hat{c}^{\dagger}\hat{b}\hat{\sigma}+h.c.\right).

Note that the three-wave coupling term in Eq. (11) has the structure formally equivalent to the parametric down-conversion (PDC) Hamiltonian describing the parametric decay of the quantized pump field into quantized signal and idler modes couteau2018 ; tokman2021-2 . Of course one difference is that all fields in the photonic PDC process are described by bosonic operators whereas the electron excitation in Eq. (11) is described by fermionic operators, giving rise to its specific nonlinearities.

The strong coupling regime is realized when the three-wave coupling parameter in Eq. (11) is larger than a certain combination of the relaxation constants γ,μω,μΩ\gamma,\mu_{\omega},\mu_{\Omega} of all subsystems. The exact criterion can be retrieved from the analytic solution presented in Sections V and VI below.

The specific form of the parameter ΩR(3)\Omega_{R}^{\left(3\right)} depends on the nonlinear coupling mechanism. For example, a phonon mode can modulate the position of the center of mass of an “atom” within a spatially nonuniform distribution of the EM field of a cavity mode. It can be realized for all kinds of quantum emitters: an electron transition in a molecule, a quantum dot or defect in a solid matrix, an optomechanical system with a varying cavity parameter, etc. In this case, in the limit of a small amplitude of vibrations, one can obtain that tokman2021

ΩR(3)=1[𝐝(𝐐)𝐄]𝐫=𝐫0.\Omega_{R}^{\left(3\right)}=-\frac{1}{\hbar}\left[\mathbf{d}\left(\mathbf{Q\cdot\nabla}\right)\mathbf{E}\right]_{\mathbf{r}=\mathbf{r}_{0}}. (12)

The Hamiltonian in Eq. (11) is valid if the three-wave resonance is well separated from the two-wave one. The conditions for that are tokman2021

|ωeωΩ||ωeω|,|ΩR(2,3)|Ω.\left|\omega_{e}-\omega-\Omega\right|\ll\left|\omega_{e}-\omega\right|,\ \left|\Omega_{R}^{\left(2,3\right)}\right|\ll\Omega. (13)

In the next section we will see that if the conditions (13) are satisfied, other models of three-wave coupling can be reduced to the universal parametric Hamiltonian (11). Note that in plasmonic nanocavities the two-wave or/and three-wave Rabi frequency ΩR(2,3)\Omega_{R}^{(2,3)} can become higher than the vibrational or phonon frequency Ω\Omega, which would violate the last of inequalities (13); see hughes2021 ; denning2020 ; tokman2021 .

III The models described by the universal parametric Hamiltonian

In addition to the three-wave coupling mechanism considered in the previous section, there are other ways for phonons or any mechanical oscillations to affect the coupling between the EM cavity field and the quantum emitter. Here we give several examples, assuming without loss of generality that the amplitudes 𝐐\mathbf{Q} in the expression for the position displacement operator in Eq. (7) are real functions.

III.1 Phonons modulate the energy of the electron transition (see, e.g., neuman2020 ; felipe2017 )

H^=ωc^c^+ωeσ^σ^+Ωb^b^+(ΩR(2)σ^c^+h.c.)+SΩσ^σ^(b^+b^).\hat{H}=\hbar\omega\hat{c}^{\dagger}\hat{c}+\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}+\hbar\Omega\hat{b}^{\dagger}\hat{b}+\hbar\left(\Omega_{R}^{\left(2\right)}\hat{\sigma}^{\dagger}\hat{c}+h.c.\right)+\hbar\sqrt{S}\Omega\hat{\sigma}^{\dagger}\hat{\sigma}\left(\hat{b}+\hat{b}^{\dagger}\right). (14)

Here SS is the Huang–Rhys factor, which determines the dependence of the transition energy on the dimensionless amplitude b^+b^\hat{b}+\hat{b}^{\dagger} of the phonon oscillations. Let’s call Eq. (14) the “molecular” Hamiltonian, although it can also describe the exciton-phonon coupling in quantum-dot systems weiler2012 ; roy2015 ; denning2020 .

III.2 Phonons modulate some geometric parameter of the cavity (see, e.g., roelli2016 ; aspelmeyer2014 )

H^=ωc^c^+ωeσ^σ^+Ωb^b^+(ΩR(2)σ^c^+h.c.)gc^c^(b^+b^).\hat{H}=\hbar\omega\hat{c}^{\dagger}\hat{c}+\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}+\hbar\Omega\hat{b}^{\dagger}\hat{b}+\hbar\left(\Omega_{R}^{\left(2\right)}\hat{\sigma}^{\dagger}\hat{c}+h.c.\right)-\hbar g\hat{c}^{\dagger}\hat{c}\left(\hat{b}+\hat{b}^{\dagger}\right). (15)

Here the factor gg determines the linear dependence of the resonant frequency of the cavity on some geometric parameter 𝐆\mathbf{G} modulated by mechanical vibrations:

g=𝐐ω𝐆.g=-\mathbf{Q}\frac{\partial\omega}{\partial\mathbf{G}}.

We will call Eq. (15) the “optomechanical” Hamiltonian.

The Hamiltonians (14) and (15) appear to be very different from Eq. (11). Indeed, they both contain the standard two-wave resonance and their three-wave coupling terms are different. We will show now that when the conditions (13) are satisfied, the “molecular” and “optomechanical” Hamiltonians are equivalent to the universal Hamiltonian in Eq. (11). To prove this statement, we write the Hamiltonian (11) in the interaction representation:

H^int=eiH^0tV^eiH^0t,\hat{H}_{int}=e^{i\hat{H}_{0}t}\hat{V}e^{-i\hat{H}_{0}t}, (16)

where

H^0=ωc^c^+ωeσ^σ^+Ωb^b^,V^=(ΩR(3)σ^c^b^+h.c.).\hat{H}_{0}=\hbar\omega\hat{c}^{\dagger}\hat{c}+\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}+\hbar\Omega\hat{b}^{\dagger}\hat{b},\ \ \hat{V}=\hbar\left(\Omega_{R}^{\left(3\right)}\hat{\sigma}^{\dagger}\hat{c}\hat{b}+h.c.\right).

This yields

H^int=(ΩR(3)σ^c^b^ei(ωeωΩ)t+h.c.).\hat{H}_{int}=\hbar\left(\Omega_{R}^{\left(3\right)}\hat{\sigma}^{\dagger}\hat{c}\hat{b}e^{i\left(\omega_{e}-\omega-\Omega\right)t}+h.c.\right). (17)

Next, we write the Hamiltonians in Eqs. (14) and (15) in the interaction representation by defining the unperturbed Hamiltonian as

H^0=ωc^c^+ωeσ^σ^+Ωb^b^+(ΩR(2)σ^c^+h.c.).\hat{H}_{0}=\hbar\omega\hat{c}^{\dagger}\hat{c}+\hbar\omega_{e}\hat{\sigma}^{\dagger}\hat{\sigma}+\hbar\Omega\hat{b}^{\dagger}\hat{b}+\hbar\left(\Omega_{R}^{\left(2\right)}\hat{\sigma}^{\dagger}\hat{c}+h.c.\right). (18)

This gives

V^=V^mol=SΩσ^σ^(b^+b^)\hat{V}=\hat{V}_{mol}=\hbar\sqrt{S}\Omega\hat{\sigma}^{\dagger}\hat{\sigma}\left(\hat{b}+\hat{b}^{\dagger}\right) (19)

for the “molecular” Hamiltonian and

V^=V^optom=gc^c^(b^+b^)\hat{V}=\hat{V}_{optom}=-\hbar g\hat{c}^{\dagger}\hat{c}\left(\hat{b}+\hat{b}^{\dagger}\right) (20)

for the optomechanical one.

The operator H^0\hat{H}_{0} given by Eq. (18) is not diagonal. In this case one should generally diagonalize H^0\hat{H}_{0}. However, sometimes a slightly different approach is simpler. Indeed, consider the Hamiltonian given by

H^=H^0(A^1,A^N)+V^(A^1,A^N),\hat{H}=\hat{H}_{0}\left(\hat{A}_{1,}\cdots\hat{A}_{N}\right)+\hat{V}\left(\hat{A}_{1,}\cdots\hat{A}_{N}\right), (21)

where A^i\hat{A}_{i} are certain operators related to coupled subsystems (here we assume that Hermitian conjugated operators are assigned different numbers “ii” ). For any interaction operator V^\hat{V}, which can be expanded in a series

V^=jkji=1Nj(A^i)nji\hat{V}=\sum_{j}k_{j}\prod_{i=1}^{N_{j}}\left(\hat{A}_{i}\right)^{n_{ji}}

(here njin_{ji} are positive integers), the following relationships are true:

H^int\displaystyle\hat{H}_{int} =\displaystyle= eiH^0tV^(A^1,A^N)eiH^0t\displaystyle e^{i\hat{H}_{0}t}\hat{V}\left(\hat{A}_{1,}\cdots\hat{A}_{N}\right)e^{-i\hat{H}_{0}t}
=\displaystyle= jkji=1NjeiH^0t(A^i)njieiH^0t\displaystyle\sum_{j}k_{j}\prod_{i=1}^{N_{j}}e^{i\hat{H}_{0}t}\left(\hat{A}_{i}\right)^{n_{ji}}e^{-i\hat{H}_{0}t}
=\displaystyle= jkji=1Nj(eiH^0tA^ieiH^0t)nji,\displaystyle\sum_{j}k_{j}\prod_{i=1}^{N_{j}}\left(e^{i\hat{H}_{0}t}\hat{A}_{i}e^{-i\hat{H}_{0}t}\right)^{n_{ji}},

from which one obtains

H^int=V^(A~^1,A~^N,),\hat{H}_{int}=\hat{V}\left(\widehat{\tilde{A}}_{1,}\cdots\widehat{\tilde{A}}_{N,}\right), (22)

where

A~^i(t,A^1,A^N)=eiH^0tA^ieiH^0t\widehat{\tilde{A}}_{i}\left(t,\hat{A}_{1,}\cdots\hat{A}_{N}\right)=e^{i\hat{H}_{0}t}\hat{A}_{i}e^{-i\hat{H}_{0}t} (23)

The operators A~^i\widehat{\tilde{A}}_{i} satisfy the Heisenberg equations

A~^it=i[H^0,A~^i]\frac{\partial\widehat{\tilde{A}}_{i}}{\partial t}=\frac{i}{\hbar}\left[\hat{H}_{0},\widehat{\tilde{A}}_{i}\right] (24)

for the initial conditions A~^i(t=0)=A^i\widehat{\tilde{A}}_{i}\left(t=0\right)=\hat{A}_{i}. In particular, for the Hamiltonian H^0\hat{H}_{0} given by Eq. (18) one obtains

σ~^t=iωeσ~^iΩR(2)c~^(12σ~^σ~^),\frac{\partial\widehat{\tilde{\sigma}}}{\partial t}=-i\omega_{e}\widehat{\tilde{\sigma}}-i\Omega_{R}^{\left(2\right)}\widehat{\tilde{c}}\left(1-2\widehat{\tilde{\sigma}}^{\dagger}\widehat{\tilde{\sigma}}\right), (25)
c~^t=iωc~^iΩR(2)σ~^,\frac{\partial\widehat{\tilde{c}}}{\partial t}=-i\omega\widehat{\tilde{c}}-i\Omega_{R}^{\left(2\right)\ast}\widehat{\tilde{\sigma}}, (26)

where we used the integral of motion σ~^σ~^+σ~^σ~^=1\widehat{\tilde{\sigma}}^{\dagger}\widehat{\tilde{\sigma}}+\widehat{\tilde{\sigma}}\widehat{\tilde{\sigma}}^{\dagger}=1. Taking into account the condition |ΩR(2)||ωeω|\left|\Omega_{R}^{\left(2\right)}\right|\ll\left|\omega_{e}-\omega\right| which follows from Eqs. (13), when solving for the operators c~^\widehat{\tilde{c}} and σ~^\widehat{\tilde{\sigma}} one can neglect the terms of the order of |ΩR(2)ωeω|2\left|\frac{\Omega_{R}^{\left(2\right)}}{\omega_{e}-\omega}\right|^{2}. For the initial conditions σ~^\widehat{\tilde{\sigma}} (t=0)=σ^\left(t=0\right)=\hat{\sigma} and c~^(t=0)=c^\widehat{\tilde{c}}\left(t=0\right)=\hat{c} the solution expanded in series in powers of the small parameter |ΩR(2)ωeω|\left|\frac{\Omega_{R}^{\left(2\right)}}{\omega_{e}-\omega}\right| takes the form

(σ~^c~^)=M^^(σ^c^),\left(\begin{array}[]{c}\widehat{\tilde{\sigma}}\\ \widehat{\tilde{c}}\end{array}\right)=\widehat{\widehat{M}}\left(\begin{array}[]{c}\hat{\sigma}\\ \hat{c}\end{array}\right), (27)

where

M^^\displaystyle\widehat{\widehat{M}} =\displaystyle= (eiωet00eiωt)\displaystyle\left(\begin{array}[]{cc}e^{-i\omega_{e}t}&0\\ 0&e^{-i\omega t}\end{array}\right) (33)
(0ΩR(2)ωeω(12σ^σ^)(eiωteiωet)ΩR(2)ωeω(eiωteiωet)0)+o(|ΩR(2)ωeω|2)\displaystyle-\left(\begin{array}[]{cc}0&\frac{\Omega_{R}^{\left(2\right)}}{\omega_{e}-\omega}\left(1-2\hat{\sigma}^{\dagger}\hat{\sigma}\right)\left(e^{-i\omega t}-e^{-i\omega_{e}t}\right)\\ \frac{\Omega_{R}^{\left(2\right)\ast}}{\omega_{e}-\omega}\left(e^{-i\omega t}-e^{-i\omega_{e}t}\right)&0\end{array}\right)+o\left(\left|\frac{\Omega_{R}^{\left(2\right)}}{\omega_{e}-\omega}\right|^{2}\right)

There is an exact solution for the operator b~^\widehat{\tilde{b}}:

b~^=eiH^0tb^eiH^0t=b^eiΩt.\widehat{\tilde{b}}=e^{i\hat{H}_{0}t}\hat{b}e^{-i\hat{H}_{0}t}=\hat{b}e^{-i\Omega t}. (34)

Now we substitute the three-wave coupling Hamiltonian (19) into Eq. (22) and use Eqs. (27)-(34). The conditions (13) combined with the RWA allow one to keep only slowly varying terms ei(ωeωΩ)t\propto e^{i\left(\omega_{e}-\omega-\Omega\right)t} in the final expression. Taking into account that σ^σ^=σ^σ^\hat{\sigma}^{\dagger}\hat{\sigma}^{\dagger}=\hat{\sigma}\hat{\sigma} =0=0 and taking 1ωeω1Ω\frac{1}{\omega_{e}-\omega}\approx\frac{1}{\Omega} in Eq. (33), we obtain the following expression for the “ molecular” Hamiltonian in the interaction picture:

(H^int)mol=(SΩR(2)σ^c^b^ei(ωeωΩ)t+h.c.)\left(\hat{H}_{int}\right)_{mol}=-\hbar\left(\sqrt{S}\Omega_{R}^{\left(2\right)}\hat{\sigma}^{\dagger}\hat{c}\hat{b}e^{i\left(\omega_{e}-\omega-\Omega\right)t}+h.c.\right) (35)

A similar derivation for the “ optomechanical” Hamiltonian given by Eq. (20) leads to the following result:

(H^int)optom=(gΩR(2)Ωσ^c^b^ei(ωeωΩ)t+h.c.)\left(\hat{H}_{int}\right)_{optom}=-\hbar\left(\frac{g\Omega_{R}^{\left(2\right)}}{\Omega}\hat{\sigma}^{\dagger}\hat{c}\hat{b}e^{i\left(\omega_{e}-\omega-\Omega\right)t}+h.c.\right) (36)

Clearly, in both cases the Hamiltonian has the same structure as the parametric Hamiltonian (17), in which

(ΩR(3))mol=SΩR(2)or(ΩR(3))optom=gΩΩR(2).\left(\Omega_{R}^{\left(3\right)}\right)_{mol}=-\sqrt{S}\Omega_{R}^{\left(2\right)}\ \ {\rm or}\ \ \left(\Omega_{R}^{\left(3\right)}\right)_{optom}=-\frac{g}{\Omega}\Omega_{R}^{\left(2\right)}. (37)

Therefore, one can use the universal parametric Hamiltonian (11) for all kinds of three-wave couplings after choosing an appropriate expression for the coupling parameter ΩR(3)\Omega_{R}^{\left(3\right)}.

We emphasize again that the universal character of the parametric Hamiltonian (11) holds as long as inequalities (13) are satisfied, which ensure that the three-wave nonlinear resonance can be separated from the two-wave resonance.

IV Including dissipation and fluctuations within the stochastic equation for the state vector

IV.1 Stochastic equation for the state vector

Many quantum information applications are based on the strong coupling regime in which the relaxation times τ\tau are much longer than the dynamical coupling times TT between the subsystems. For two- and three-wave couplings those times are determined by effective Rabi frequencies, TT -1 |ΩR(2,3)|\sim\left|\Omega_{R}^{\left(2,3\right)}\right|. As shown in tokman2021 ; chen2021 , the method of the stochastic equation for the state vector is often the most convenient way to describe the nonperturbative dynamics of open strongly coupled systems; it leads to simpler derivations for the observables and characterization of entanglement than the operator-valued Heisenberg-Langevin equations or the master equation for the density matrix.

Indeed, when applied to strongly coupled systems the Heisenberg approach leads to the nonlinear operator-valued equations even in the simplest case of a single two-level atom coupled to a single-mode field scully1997 . In contrast, the equations for the state vector components are always linear; they contain much fewer variables as compared to the density matrix equations and are split into low-dimensional blocks in the RWA, leading to analytic solutions for both two-wave scully1997 and three-wave tokman2021 ; chen2021 resonant coupling.

One potential difficulty with this approach is that dissipation and fluctuations may lead to the coupling between different blocks of equations for the state vector components that were uncoupled in a closed system. However, in the strong coupling regime the coupling through dissipative reservoirs is weak (scales as a small parameter T/τT/\tau) and can be taken into account perturbatively tokman2021 ; chen2021 .

Following tokman2021 ; chen2021 , we apply the stochastic equation for the state vector to derive analytic solution for the parametric coupling of a two-level fermionic quantum emitter to two boson fields. The stochastic equation has the following general form,

ddt|Ψ=iH^eff|Ψi|.\frac{d}{dt}\left|\Psi\right\rangle=-\frac{i}{\hbar}\hat{H}_{eff}\left|\Psi\right\rangle-\frac{i}{\hbar}\left|\mathfrak{R}\right\rangle. (38)

Here |Ψ\left|\Psi\right\rangle is the state vector; |\left|\mathfrak{R}\right\rangle is the noise term satisfying |¯=0\overline{\left|\mathfrak{R}\right\rangle}=0, where the bar means averaging over the noise statistics; H^eff=\hat{H}_{eff}= H^+H^(ah)\hat{H}+\hat{H}^{\left(ah\right)} is an effective Hamiltonian which is a non-Hermitian operator. Its non-Hermitian component H^(ah)\hat{H}^{\left(ah\right)} describes the effects of relaxation. The expressions for H^(ah)\hat{H}^{\left(ah\right)} and |\left|\mathfrak{R}\right\rangle must be consistent with each other to guarantee the conservation of the noise-averaged norm, Ψ(t)|Ψ(t)¯=1\overline{\left\langle\Psi\left(t\right)\right.\left|\Psi\left(t\right)\right\rangle}=1, and ensure that the system reaches a physically meaningful steady state in the absence of any external driving force. To calculate the observables from the state vector given by Eq. (38) one should apply a standard procedure but with an important extra step: averaging over the noise statistics, i.e., q=Ψ|q^|Ψ¯q=\overline{\left\langle\Psi\right|\hat{q}\left|\Psi\right\rangle} , where q^\hat{q} is a quantum-mechanical operator corresponding to the observable qq.

Perhaps the most popular version of the stochastic approach to derive the state vector, i.e., the stochastic Schrödinger equation (SSE), is its application for numerical Monte-Carlo simulations within the method of quantum jumps scully1997 ; plenio1998 ; gisin1992 ; diosi1998 ; tannoudji1993 ; molmer1993 ; gisin1992-2 ; raedt2017 ; nathan2020 . The stochastic equation in a different form, the Schrödinger-Langevin equation (SLE), was suggested to describe the Brownian motion of a quantum particle in a constant field kostin1972 ; katz2016 . Generally, using some version of the stochastic equation fits within the narrative of the Langevin method WM94 . Within the Langevin approach which describes the system with stochastic equations of evolution, the averaging over the reservoir degrees of freedom is equivalent to averaging over the statistics of the noise sources landau1965 . This paradigm allows one to describe open systems without relying on the density matrix.

IV.2 Comparison with the Lindblad formalism

It was shown in tokman2021 that one can choose the form of H^(ah)\hat{H}^{\left(ah\right)} and |\left|\mathfrak{R}\right\rangle in such a way that the observables calculated with Eq. (38) will coincide with those obtained by solving the master equation in the Lindblad approximation. The corresponding master equation has the form scully1997

ddtρ^=i[H^,ρ^]+L^(ρ^)\frac{d}{dt}\hat{\rho}=-\frac{i}{\hbar}\left[\hat{H},\hat{\rho}\right]+\hat{L}\left(\hat{\rho}\right) (39)

where L^(ρ^)\hat{L}\left(\hat{\rho}\right) is the relaxation operator (Lindbladian) which can be represented as

L^(ρ^)=i(H^(ah)ρ^ρ^H^(ah))+δL^(ρ^).\hat{L}\left(\hat{\rho}\right)=-\frac{i}{\hbar}\left(\hat{H}^{\left(ah\right)}\hat{\rho}-\hat{\rho}\hat{H}^{\left(ah\right)\dagger}\right)+\delta\hat{L}\left(\hat{\rho}\right). (40)

The equivalence (in the above sense) between the stochastic equation and the Lindblad approach exists if we substitute the anti-Hermitian part of the Hamiltonian from Eq. (40) into Eq. (38), and postulate the following correlation properties for the noise source:

|(t)(t′′)|¯=2δ(tt′′)δL^(ρ^)ρ^|ΨΨ|¯\overline{\left|\mathfrak{R}\left(t^{\prime}\right)\right\rangle\left\langle\mathfrak{R}\left(t^{\prime\prime}\right)\right|}=\hbar^{2}\delta\left(t^{\prime}-t^{\prime\prime}\right)\delta\hat{L}\left(\hat{\rho}\right)_{\hat{\rho}\Longrightarrow\overline{\left|\Psi\right\rangle\left\langle\Psi\right|}} (41)

IV.3 Parametric decay of the electron excitation into a photon and a phonon

We will describe the dynamics near the three-wave electron-photon-phonon resonance by the stochastic equation (38) with the parametric Hamiltonian (11). We will seek the state vector in the form

Ψ=n,α=0,(Cαn0|α|n|0+Cαn1|α|n|1),\Psi=\sum_{n,\alpha=0}^{\infty,\infty}\left(C_{\alpha n0}\left|\alpha\right\rangle\left|n\right\rangle\left|0\right\rangle+C_{\alpha n1}\left|\alpha\right\rangle\left|n\right\rangle\left|1\right\rangle\right),

where the order of indices corresponds to

Cphononphotonfermion|phonon|photon|fermion.C_{\mathrm{phonon\,photon\,fermion}}\left|phonon\right\rangle\left|photon\right\rangle\left|fermion\right\rangle.

Consider the initial state with an excited electron and no bosonic excitations, of the type |Ψ(0)=|0|0|1\left|\Psi\left(0\right)\right\rangle=\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle . Near the resonance ωeω+Ω\omega_{e}\approx\omega+\Omega the three-wave coupling leads to the excitation of the state |1|1|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle. In the zero-temperature limit, which is valid when the reservoir temperature is much lower than the transition frequency (for optical frequencies it is satisfied even at room temperature), relaxation processes could populate only the states with lower energies: |0|1|0\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle, |1|0|0\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle and |0|0|0\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle. Therefore, for this initial condition the state vector will have 5 components:

|Ψ(t)\displaystyle\left|\Psi\left(t\right)\right\rangle =\displaystyle= C000(t)|0|0|0+C010(t)|0|1|0+C100(t)|1|0|0\displaystyle C_{000}\left(t\right)\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle+C_{010}\left(t\right)\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle+C_{100}\left(t\right)\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle (42)
+C110(t)|1|1|0+C001(t)|0|0|1.\displaystyle+C_{110}\left(t\right)\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle+C_{001}\left(t\right)\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle.

Note that we are not restricting the basis in any way and only considering the initial conditions leading to single photon and phonon states to simplify algebra: see, for example, the Appendix in tokman2021-3 where arbitrary multiphoton states are considered in the same way, leading of course to more cumbersome expressions. Another reason to consider such initial conditions is that single-photon states are used in most applications, whereas generating multiphoton number states remains a major challenge.

To determine the anti-Hermitian part H^(ah)\hat{H}^{\left(ah\right)} of the Hamiltonian which describes relaxation and the correlator |(t)(t′′)|¯\overline{\left|\mathfrak{R}\left(t^{\prime}\right)\right\rangle\left\langle\mathfrak{R}\left(t^{\prime\prime}\right)\right|} describing fluctuations, we will use the expression for the total Lindbladian of the system including a 2-level “atom”, photons, and phonons scully1997 . For simplicity we will assume zero temperature of dissipative reservoirs, which is satisfied at TωT\ll\hbar\omega. The finite-temperature expressions are given in tokman2021 ; chen2021 . Then the Lindbladian is

L(ρ^)=Le(ρ^)+Lem(ρ^)+Lp(ρ^)L\left(\hat{\rho}\right)=L_{e}\left(\hat{\rho}\right)+L_{em}\left(\hat{\rho}\right)+L_{p}\left(\hat{\rho}\right) (43)
Le(ρ^)=γ2(σ^σ^ρ^+ρ^σ^σ^2σ^ρ^σ^)L_{e}\left(\hat{\rho}\right)=-\frac{\gamma}{2}\left(\hat{\sigma}^{\dagger}\hat{\sigma}\hat{\rho}+\hat{\rho}\hat{\sigma}^{\dagger}\hat{\sigma}-2\hat{\sigma}\hat{\rho}\hat{\sigma}^{\dagger}\right) (44)
Lem(ρ^)=μω2(c^c^ρ^+ρ^c^c^2c^ρ^c^)L_{em}\left(\hat{\rho}\right)=-\frac{\mu_{\omega}}{2}\left(\hat{c}^{\dagger}\hat{c}\hat{\rho}+\hat{\rho}\hat{c}\hat{c}^{\dagger}-2\hat{c}\hat{\rho}\hat{c}^{\dagger}\right) (45)
Lp(ρ^)=μΩ2(b^b^ρ^+ρ^b^b^2b^ρ^b^)L_{p}\left(\hat{\rho}\right)=-\frac{\mu_{\Omega}}{2}\left(\hat{b}^{\dagger}\hat{b}\hat{\rho}+\hat{\rho}\hat{b}\hat{b}^{\dagger}-2\hat{b}\hat{\rho}\hat{b}^{\dagger}\right) (46)

where γ\gamma, μω\mu_{\omega} and μΩ\mu_{\Omega} are relaxation rates of corresponding subsystems.

Then the stochastic equation for the state vector takes the form

(ddt000ddt+iω010+γ010000ddt+iω100+γ100)×(C000C010C100)=i(000010100),\left(\begin{array}[]{ccc}\frac{d}{dt}&0&0\\ 0&\frac{d}{dt}+i\omega_{010}+\gamma_{010}&0\\ 0&0&\frac{d}{dt}+i\omega_{100}+\gamma_{100}\end{array}\right)\times\left(\begin{array}[]{c}C_{000}\\ C_{010}\\ C_{100}\end{array}\right)=-\frac{i}{\hbar}\left(\begin{array}[]{c}\mathfrak{R}_{000}\\ \mathfrak{R}_{010}\\ \mathfrak{R}_{100}\end{array}\right), (47)
(ddt+iω110+γ110iΩR(3)iΩR(3)ddt+iω001+γ001)(C110C001)=i(110001),\left(\begin{array}[]{cc}\frac{d}{dt}+i\omega_{110}+\gamma_{110}&i\Omega_{R}^{\left(3\right)\ast}\\ i\Omega_{R}^{\left(3\right)}&\frac{d}{dt}+i\omega_{001}+\gamma_{001}\end{array}\right)\left(\begin{array}[]{c}C_{110}\\ C_{001}\end{array}\right)=-\frac{i}{\hbar}\left(\begin{array}[]{c}\mathfrak{R}_{110}\\ \mathfrak{R}_{001}\end{array}\right), (48)

where

αni=αni|;\mathfrak{R}_{\alpha ni}=\left\langle\alpha ni\right.\left|\mathfrak{R}\right\rangle;
ω010=ω,ω100=Ω,ω110=ω+Ω,ω001=ωe;\omega_{010}=\omega,\ \omega_{100}=\Omega,\ \omega_{110}=\omega+\Omega,\ \omega_{001}=\omega_{e};
γ010=12μω,γ100=12μΩ,γ110=12(μω+μΩ),γ001=12γ.\gamma_{010}=\frac{1}{2}\mu_{\omega},\ \gamma_{100}=\frac{1}{2}\mu_{\Omega},\ \gamma_{110}=\frac{1}{2}\left(\mu_{\omega}+\mu_{\Omega}\right),\ \gamma_{001}=\frac{1}{2}\gamma.

The correlators of the noise sources in Eqs. (47) and (48) are

αni(t)βmj(t′′)¯=2Dαni,βmj(t)δ(tt′′),\overline{\mathfrak{R}_{\alpha ni}^{\ast}\left(t^{\prime}\right)\mathfrak{R}_{\beta mj}\left(t^{\prime\prime}\right)}=\hbar^{2}D_{\alpha ni,\beta mj}(t^{\prime})\delta\left(t^{\prime}-t^{\prime\prime}\right), (49)

where the quantities Dαni,βmjD_{\alpha ni,\beta mj} are determined using Eqs. (41) and (43)-(46). For the diagonal elements of the correlators we obtain

D110,110\displaystyle D_{110,110} =\displaystyle= D001,001=0\displaystyle D_{001,001}=0
D100,100\displaystyle D_{100,100} =\displaystyle= μω|C110|2¯\displaystyle\mu_{\omega}\overline{\left|C_{110}\right|^{2}} (50)
D010,010\displaystyle D_{010,010} =\displaystyle= μΩ|C110|2¯\displaystyle\mu_{\Omega}\overline{\left|C_{110}\right|^{2}}
D000,000\displaystyle D_{000,000} =\displaystyle= γ|C001|2¯+μω|C010|2¯+μΩ|C100|2¯.\displaystyle\gamma\overline{\left|C_{001}\right|^{2}}+\mu_{\omega}\overline{\left|C_{010}\right|^{2}}+\mu_{\Omega}\overline{\left|C_{100}\right|^{2}}.

The off-diagonal elements are given by

Dαni,βmj\displaystyle D_{\alpha ni,\beta mj} =\displaystyle= Dβmj,αni\displaystyle D_{\beta mj,\alpha ni}^{\ast}
D110,αni\displaystyle D_{110,\alpha ni} =\displaystyle= D001,αni=D100,010=0\displaystyle D_{001,\alpha ni}=D_{100,010}=0 (51)
D000,100\displaystyle D_{000,100} =\displaystyle= μωC010C110¯\displaystyle\mu_{\omega}\overline{C_{010}^{\ast}C_{110}}
D000,010\displaystyle D_{000,010} =\displaystyle= μΩC100C110¯.\displaystyle\mu_{\Omega}\overline{C_{100}^{\ast}C_{110}}.

The dependence of quantities Dαni,βmjD_{\alpha ni,\beta mj} in the right-hand side of Eq. (49) on time tt^{\prime} is due to the time dependence of amplitudes CαniC_{\alpha ni} which enter Eqs. (50) and (51).

The derivation of the stochastic equation for the state vector including pure dephasing processes and the finite temperature of the reservoirs has been discussed in tokman2021 ; chen2021 .

Eqs. (48) describe the dynamic generation of an entangled state of the type |MIX=A(t)|0|0|1+B(t)|1|1|0\left|MIX\right\rangle=A\left(t\right)\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle+B\left(t\right)\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle whereas Eqs. (47) describe the relaxation dynamics leading to relaxation of populations to states with lower energies. The quantities D010,010D_{010,010} and D100,100D_{100,100} in Eqs. (50) are associated with processes of the relaxation to states |0|1|0\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle and |1|0|0\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle from the entangled state |MIX\left|MIX\right\rangle . The structure of the expression for D000,000D_{000,000} corresponds to the relaxation of the system from the entangled state to the ground state via both “direct” pathway |0|0|1|0|0|0\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle and multistep pathways |1|1|0|0|1|0|0|0|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle and |1|1|0|1|0|0|0|0|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle, as illustrated in Figs. 3 and 5 below.

IV.4 Expressions for noise sources in the stochastic equation

In addition to the expressions for noise correlators, it is convenient to know more detailed expressions for the random functions describing the noise sources αni(t)\mathfrak{R}_{\alpha ni}(t).

The effect of the reservoir on the dynamic system is characterized by the matrix elements of the operator which determines coupling to the reservoir. For a weak coupling these matrix elements are linear with respect to the matrix elements of the operators describing the dynamical system. Therefore, the functions αni(t)\mathfrak{R}_{\alpha ni}(t) should depend linearly on the components of the state vector of the system.

Here we again consider the low-temperature case when the relaxation processes can bring the populations only down, not up. When the reservoirs for each subsystem are statistically independent, one can try the following ansatz,

αni(t)=γCαn(i+1)(t)fe(t)+μωCα(n+1)i(t)fem(t)+μΩC(α+1)ni(t)fp(t),\mathfrak{R}_{\alpha ni}\left(t\right)=\hbar\sqrt{\gamma}C_{\alpha n\left(i+1\right)}\left(t\right)f_{e}\left(t\right)+\hbar\sqrt{\mu_{\omega}}C_{\alpha\left(n+1\right)i}\left(t\right)f_{em}\left(t\right)+\hbar\sqrt{\mu_{\Omega}}C_{\left(\alpha+1\right)ni}\left(t\right)f_{p}\left(t\right), (52)

where fe,em,p¯=0.\overline{f_{e,em,p}}=0. Here fe,em,p(t)f_{e,em,p}\left(t\right) are random functions that are determined by the statistics of noise in unperturbed electron, photon, and phonon reservoirs. The functions fe,em,pf_{e,em,p} should not depend on the set of variables CαniC_{\alpha ni} within the above approximations.

The linear dependence of the noise term on the state vector was also assumed in SLE kostin1972 ; katz2016 , in which the noise terms had the form

|=U^(𝐫,t)|Ψ,\left|\mathfrak{R}\right\rangle=\hat{U}\left(\mathbf{r},t\right)\left|\Psi\right\rangle, (53)

where U^(𝐫,t)\hat{U}\left(\mathbf{r},t\right) is the fluctuating component of the potential.

To ensure that Eq. (52) leads to the correlators Eqs. (49)-(51) that are consistent with the Lindblad master equation, one needs first to define the correlators of the random functions in Eq. (52) in the following way:

fκ(t)fλ(t′′)¯=δκλδ(tt′′),\overline{f_{\kappa}^{\ast}\left(t^{\prime}\right)f_{\lambda}\left(t^{\prime\prime}\right)}=\delta_{\kappa\lambda}\delta\left(t^{\prime}-t^{\prime\prime}\right), (54)

where κ,λ=e,em,p\kappa,\lambda=e,em,p, i.e. the fluctuations in different reservoirs are independent and Markovian. Second, one has to assume that the correlations are factorized when calculating the averages

Cαni(t)Cβmj(t′′)fκ(t)fλ(t′′)¯=Cαni(t)Cβmj(t′′)¯×fκ(t)fλ(t′′)¯.\overline{C_{\alpha ni}^{\ast}\left(t^{\prime}\right)C_{\beta mj}\left(t^{\prime\prime}\right)f_{\kappa}^{\ast}\left(t^{\prime}\right)f_{\lambda}\left(t^{\prime\prime}\right)}=\overline{C_{\alpha ni}^{\ast}\left(t^{\prime}\right)C_{\beta mj}\left(t^{\prime\prime}\right)}\times\overline{f_{\kappa}^{\ast}\left(t^{\prime}\right)f_{\lambda}\left(t^{\prime\prime}\right)}. (55)

Eq. (54) looks obvious, whereas the factorization in Eq. (55) is valid only in linear approximation with respect to relaxation constants γ\gamma and μω,Ω\mu_{\omega,\Omega}. However, the Lindbladian of the form given in Eqs. (42)-(45) is itself valid within the same approximation. Therefore, Eqs. (52), (54), and (55) lead to all expressions in Eqs. (50),(51). One also has to keep in mind that with our choice of our initial conditions the amplitudes Cαni=0C_{\alpha ni}=0 for all states with energies above those in states |1|1|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle and |0|0|1\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle.

V Dynamics of entangled fermion-photon-phonon states in a dissipative system

Here we write an explicit solution of Eqs (47) and (48) for the initial state vector |Ψ(0)=|0|0|1\left|\Psi\left(0\right)\right\rangle=\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle, when C001(0)=1C_{001}\left(0\right)=1, C000(0)=C010(0)=C100(0)=C110(0)=0C_{000}\left(0\right)=C_{010}\left(0\right)=C_{100}\left(0\right)=C_{110}\left(0\right)=0. Assuming exact resonance at ωe=ω+Ω\omega_{e}=\omega+\Omega, we obtain

C000(t)\displaystyle C_{000}\left(t\right) =\displaystyle= i0t000(t)𝑑t\displaystyle-\frac{i}{\hbar}\int_{0}^{t}\mathfrak{R}_{000}\left(t^{\prime}\right)dt^{\prime}
C010(t)\displaystyle C_{010}\left(t\right) =\displaystyle= i0t010(t)eiω(tt)γ010(tt)𝑑t\displaystyle-\frac{i}{\hbar}\int_{0}^{t}\mathfrak{R}_{010}\left(t^{\prime}\right)e^{-i\omega\left(t-t^{\prime}\right)-\gamma_{010}\left(t-t^{\prime}\right)}dt^{\prime} (56)
C100(t)\displaystyle C_{100}\left(t\right) =\displaystyle= i0t100(t)eiΩ(tt)γ100(tt)𝑑t,\displaystyle-\frac{i}{\hbar}\int_{0}^{t}\mathfrak{R}_{100}\left(t^{\prime}\right)e^{-i\Omega\left(t-t^{\prime}\right)-\gamma_{100}\left(t-t^{\prime}\right)}dt^{\prime},
C001(t)\displaystyle C_{001}\left(t\right) =\displaystyle= eiωeteγ110+γ0012t[cos(Ω~Rt)+γ110γ001Ω~Rsin(Ω~Rt)]+δC001\displaystyle e^{-i\omega_{e}t}e^{-\frac{\gamma_{110}+\gamma_{001}}{2}t}\left[\cos\left(\tilde{\Omega}_{R}t\right)+\frac{\gamma_{110}-\gamma_{001}}{\tilde{\Omega}_{R}}\sin\left(\tilde{\Omega}_{R}t\right)\right]+\delta C_{001} (57)
C110(t)\displaystyle C_{110}\left(t\right) =\displaystyle= eiωeteγ110+γ0012t(ieiθ)sin(Ω~Rt)+δC110\displaystyle e^{-i\omega_{e}t}e^{-\frac{\gamma_{110}+\gamma_{001}}{2}t}\left(ie^{-i\theta}\right)\sin\left(\tilde{\Omega}_{R}t\right)+\delta C_{110} (58)

where the effective Rabi frequency Ω~R=|ΩR(3)|2(γ110γ001)24\tilde{\Omega}_{R}=\sqrt{\left|\Omega_{R}^{\left(3\right)}\right|^{2}-\frac{\left(\gamma_{110}-\gamma_{001}\right)^{2}}{4}}, θ=Arg[ΩR(3)]\theta=\mathrm{Arg}\left[\Omega_{R}^{\left(3\right)}\right], and the terms δC100,110\delta C_{100,110} are linear with respect to random functions 001\mathfrak{R}_{001} and 110\mathfrak{R}_{110}.

The term proportional to γ110γ001Ω~R\frac{\gamma_{110}-\gamma_{001}}{\tilde{\Omega}_{R}} in the first of Eqs. (57) can be omitted when calculating most observables when the dissipation is weak, Ω~R\tilde{\Omega}_{R}\gg γani\gamma_{ani} (see, e.g., tokman2021 ). However, one has to keep in mind that this term is needed for Eqs. (57) to satisfy an exact integral of motion of Eqs. (48) :

ddt(|C001|2¯+|C110|2¯)=2γ001|C001|2¯2γ110|C110|2¯.\frac{d}{dt}\left(\overline{\left|C_{001}\right|^{2}}+\overline{\left|C_{110}\right|^{2}}\right)=-2\gamma_{001}\overline{\left|C_{001}\right|^{2}}-2\gamma_{110}\overline{\left|C_{110}\right|^{2}}.

Note that when averaged over the period 2πΩ~R\frac{2\pi}{\tilde{\Omega}_{R}} the integral is conserved even without this term.

The solution for the 5-component state vector can be written as

|Ψ\displaystyle\left|\Psi\right\rangle (59)
=\displaystyle= eiωeteγ110+γ0012t{[cos(Ω~Rt)+γ110γ0012Ω~Rsin(Ω~Rt)]|0|0|1+ieiθsin(Ω~Rt)|1|1|0}\displaystyle e^{-i\omega_{e}t}e^{-\frac{\gamma_{110}+\gamma_{001}}{2}t}\left\{\left[\cos\left(\tilde{\Omega}_{R}t\right)+\frac{\gamma_{110}-\gamma_{001}}{2\tilde{\Omega}_{R}}\sin\left(\tilde{\Omega}_{R}t\right)\right]\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle+ie^{-i\theta}\sin\left(\tilde{\Omega}_{R}t\right)\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\right\}
+δC001|0|0|1+δC110|1|1|0+C000|0|0|0+C100|1|0|0+C010|0|1|0,\displaystyle+\delta C_{001}\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle+\delta C_{110}\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle+C_{000}\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle+C_{100}\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle+C_{010}\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle,

where

δC001¯=δC110¯=C000¯=C100¯=C010¯=0,\overline{\delta C_{001}}=\overline{\delta C_{110}}=\overline{C_{000}}=\overline{C_{100}}=\overline{C_{010}}=0, (60)
|δC001|2¯=|δC110|2¯=0.\overline{\left|\delta C_{001}\right|^{2}}=\overline{\left|\delta C_{110}\right|^{2}}=0. (61)

It follows from Eq. (59) that in the entangled state |MIX=A(t)|0|0|1+B(t)|1|1|0\left|MIX\right\rangle=A\left(t\right)\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle+B\left(t\right)\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle the amplitudes A(t)A\left(t\right) and B(t)B\left(t\right) oscillate at the effective Rabi frequency and decay with the decay rate

γMIX=γ110+γ0012=14(μω+μΩ+γ).\gamma_{{}_{MIX}}=\frac{\gamma_{110}+\gamma_{001}}{2}=\frac{1}{4}\left(\mu_{\omega}+\mu_{\Omega}+\gamma\right).

The occupation probabilities |C001|2|C_{001}|^{2} and |C110|2|C_{110}|^{2} are plotted in Fig. 2 as a function of normalized time Ω~Rt\tilde{\Omega}_{R}t, along with the real parts of their eigenfrequencies obtained from Eqs. (48) as a function of detuning from the parametric resonance ω+Ωωe\omega+\Omega-\omega_{e}. Although the plots look like standard anticrossing behavior and decaying Rabi oscillations, one should keep in mind that (1) the anticrossing occurs not at the standard exciton-photon or phonon-photon resonance but at the nonlinear parametric resonance, which is controlled by the nonlinear coupling strength ΩR(3)\Omega_{R}^{(3)} and entangles three degrees of freedom; (2) the relaxation rates of each individual subsystem enter the analytic expressions plotted in Fig. 2 in a nontrivial way, as described above. This dependence will become even more nontrivial at high temperature of the reservoir. The presented solution provides the way to retrieve the analytic dependence of any observable on the relaxation and coupling parameters and determine correctly the criterion for observing the strong parametric coupling in frequency or time domain. Two obvious examples for such observables are photon and phonon emission spectra that are derived and plotted in Sec. VI, see Figs. 4 and 6.

Refer to caption
(a)
Refer to caption
(b)
Figure 2: (a) Real parts of eigenstate frequencies of Eqs. (48), shifted by the electron transition frequency ωe\omega_{e} and normalized by |ΩR(3)||\Omega_{R}^{(3)}|, as a function of detuning from the parametric resonance ω+Ωωe\omega+\Omega-\omega_{e} normalized by |ΩR(3)||\Omega_{R}^{(3)}|. The relaxation rates are μω=μΩ=0.3|ΩR(3)|\mu_{\omega}=\mu_{\Omega}=0.3|\Omega_{R}^{(3)}| and γ=0.2|ΩR(3)|\gamma=0.2|\Omega_{R}^{(3)}|. (b) Occupation probabilities |C001|2|C_{001}|^{2} and |C110|2|C_{110}|^{2} from Eqs. (57) and (58) as a function of normalized time Ω~Rt\tilde{\Omega}_{R}t for the same relaxation rates.

We also give the expressions for the occupation probabilities |C100|2¯\overline{\left|C_{100}\right|^{2}}, |C010|2¯\overline{\left|C_{010}\right|^{2}} and |C000|2¯\overline{\left|C_{000}\right|^{2}} that are valid under the condition Ω~R\tilde{\Omega}_{R}\gg γani\gamma_{ani}:

|C100|2¯μωμΩμωγ(eμΩ+μω+γ2teμΩt)\overline{\left|C_{100}\right|^{2}}\approx\frac{\mu_{\omega}}{\mu_{\Omega}-\mu_{\omega}-\gamma}\left(e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\Omega}t}\right) (62)
|C010|2¯μΩμωμΩγ(eμΩ+μω+γ2teμωt)\overline{\left|C_{010}\right|^{2}}\approx\frac{\mu_{\Omega}}{\mu_{\omega}-\mu_{\Omega}-\gamma}\left(e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\omega}t}\right) (63)
|C000|2¯γ(γμΩμω)γ2(μΩμω)2(1eμΩ+μω+γ2t)(μΩ1eμωtμωμΩγ+μω1eμΩtμΩμωγ)\overline{\left|C_{000}\right|^{2}}\approx\frac{\gamma\left(\gamma-\mu_{\Omega}-\mu_{\omega}\right)}{\gamma^{2}-\left(\mu_{\Omega}-\mu_{\omega}\right)^{2}}\left(1-e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}\right)-\left(\mu_{\Omega}\frac{1-e^{-\mu_{\omega}t}}{\mu_{\omega}-\mu_{\Omega}-\gamma}+\mu_{\omega}\frac{1-e^{-\mu_{\Omega}t}}{\mu_{\Omega}-\mu_{\omega}-\gamma}\right) (64)

It is easy to see that Eqs. (62)-(64) don’t have the divergence when [±(μωμΩ)γ]0\left[\pm\left(\mu_{\omega}-\mu_{\Omega}\right)-\gamma\right]\longrightarrow 0; for example,

lim(μΩμωγ)0[eμΩ+μω+γ2teμΩtμΩμωγ]=12teμΩt\lim_{\left(\mu_{\Omega}-\mu_{\omega}-\gamma\right)\longrightarrow 0}\left[\frac{e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\Omega}t}}{\mu_{\Omega}-\mu_{\omega}-\gamma}\right]=\frac{1}{2}te^{-\mu_{\Omega}t}

and so on. In the limit tt\longrightarrow\infty Eqs. (62)-(64) lead to |C100|2¯=|C010|2¯=0\overline{\left|C_{100}\right|^{2}}=\overline{\left|C_{010}\right|^{2}}=0, |C000|2¯=1\overline{\left|C_{000}\right|^{2}}=1.

The detailed derivation of Eqs. (62)-(64) is given in Appendix A. It is also shown there that in our case the quantities δC100\delta C_{100} and δC110\delta C_{110} do not contribute to the observables and therefore can be omitted.

In tokman2021 we used a simplified model to analyze the fermion-photon-phonon entanglement, in which all relaxation pathways to the ground state |0|0|0\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle are assumed to be “ direct”, which corresponds to taking all correlators equal to zero except D000,000D_{000,000}. This approach is essentially the Weisskopf-Wigner method, modified in order to conserve the norm of the state vector. It gives a correct result for the decay rate γMIX\gamma_{MIX} of the entangled state. At the same time, including multistep decay pathways is of principal importance when calculating and interpreting the emission spectra, as we will see in the next section.

VI Emission spectra of photons and phonons from the parametric decay of the electron excitation

VI.1 Derivation of the emission spectra from the solution of the stochastic equation for the state vector: a general scheme

Consider for definiteness the EM radiation out of a cavity. Its power spectrum received by the detector is given by scully1997 ; madsen2013

P(ν)=AS(ν),P\left(\nu\right)=A\cdot S\left(\nu\right),

where

S(ν)=1πRe0𝑑τeiντ0𝑑tK(t,τ),S\left(\nu\right)=\frac{1}{\pi}\mathrm{Re}\int_{0}^{\infty}d\tau e^{i\nu\tau}\int_{0}^{\infty}dtK\left(t,\tau\right), (65)

K=c^(t)c^(t+τ)K=\left\langle\hat{c}^{\dagger}\left(t\right)\hat{c}\left(t+\tau\right)\right\rangle, \left\langle\cdots\right\rangle is a quantum-mechanical averaging. The coefficient A includes the QQ-factor of a cavity, spatial structure of the outgoing field, and the position and properties of the detector.

To calculate the power spectrum one needs to know the solution of the Heisenberg-Langevin equations for the field operators c^(t)\hat{c}\left(t\right) and c^(t)\hat{c}^{\dagger}\left(t\right), then calculate the correlator, and average it over the statistics of Langevin noise: KK\Rightarrow c^(t)c^(t+τ)¯\overline{\left\langle\hat{c}^{\dagger}\left(t\right)\hat{c}\left(t+\tau\right)\right\rangle}. However, as we already discussed, the Heisenberg-Langevin equations become nonlinear in the strong-coupling regime. Therefore, it may be more convenient to obtain the spectra from the solution of the stochastic equation Eq.(33) for the state vector. The general procedure is as follows.

First, we need to transform the correlator

K(t,τ)=c^(t)c^(t+τ)=Ψ(0)|c^(t)c^(t+τ)|Ψ(0)K\left(t,\tau\right)=\left\langle\hat{c}^{\dagger}\left(t\right)\hat{c}\left(t+\tau\right)\right\rangle=\left\langle\Psi\left(0\right)\right|\hat{c}^{\dagger}\left(t\right)\hat{c}\left(t+\tau\right)\left|\Psi\left(0\right)\right\rangle (66)

to the Schrödinger picture without taking into account dissipation and fluctuations. If U^(t)\hat{U}\left(t\right) is the unitary operator of evolution of the system, one can write

K=Ψ(0)|U^(t)c^U^(t)U^(t+τ)c^U^(t+τ)|Ψ(0)=c^Ψ(t)|U^(t)U^(t+τ)|c^Ψ(t+τ),K=\left\langle\Psi\left(0\right)\right|\hat{U}^{\dagger}\left(t\right)\hat{c}^{\dagger}\,\hat{U}\left(t\right)\hat{U}^{\dagger}\left(t+\tau\right)\hat{c}\,\hat{U}\left(t+\tau\right)\left|\Psi\left(0\right)\right\rangle=\left\langle\hat{c}\,\Psi\left(t\right)\right|\hat{U}\left(t\right)\hat{U}^{\dagger}\left(t+\tau\right)\left|\hat{c}\,\Psi\left(t+\tau\right)\right\rangle, (67)

where c^\hat{c} is the Schrödinger’s (constant) operator which we will treat as an initial condition for the Heisenberg operator c^(t)\hat{c}\left(t\right) at t=0t=0. We will use the notation U^(t)U^t0(t)\hat{U}\left(t\right)\equiv\hat{U}_{t_{0}}\left(t^{\prime}\right), where we indicate explicitly the initial moment of time t0t_{0} and the duration of evolution t=tt0t^{\prime}=t-t_{0}. This will lead to the following replacements in Eq. (67): U^(t)U^0(t)\hat{U}\left(t\right)\Longrightarrow\hat{U}_{0}\left(t\right), U^(t+τ)U^0(t+τ)\hat{U}\left(t+\tau\right)\Longrightarrow\hat{U}_{0}\left(t+\tau\right). Furthermore, we obviously have

U^0(t+τ)=U^t(τ)U^0(t)\hat{U}_{0}\left(t+\tau\right)=\hat{U}_{t}\left(\tau\right)\hat{U}_{0}\left(t\right) (68)

which gives

K=c^Ψ(t)|U^0(t)(U^t(τ)U^0(t))|c^Ψ(t+τ)=c^Ψ(t)|U^0(t)U^0(t)U^t(τ)|c^Ψ(t+τ).K=\left\langle\hat{c}\,\Psi\left(t\right)\right|\hat{U}_{0}\left(t\right)\left(\hat{U}_{t}\left(\tau\right)\hat{U}_{0}\left(t\right)\right)^{\dagger}\left|\hat{c}\,\Psi\left(t+\tau\right)\right\rangle=\left\langle\hat{c}\,\Psi\left(t\right)\right|\hat{U}_{0}\left(t\right)\hat{U}_{0}^{\dagger}\left(t\right)\hat{U}_{t}^{\dagger}\left(\tau\right)\left|\hat{c}\,\Psi\left(t+\tau\right)\right\rangle.

Taking into account U^0(t)U^0(t)=1\hat{U}_{0}\left(t^{\prime}\right)\hat{U}_{0}^{\dagger}\left(t^{\prime}\right)=1, we obtain

K=U^t(τ)c^Ψ(t)|c^Ψ(t+τ).K=\left\langle\hat{U}_{t}\left(\tau\right)\hat{c}\,\Psi\left(t\right)\right.\left|\hat{c}\,\Psi\left(t+\tau\right)\right\rangle. (69)

Second, introducing the notations

ΨC^(t)=c^Ψ(t),Φ(t,τ)=U^t(τ)ΨC^(t),\Psi_{\hat{C}}\left(t\right)=\hat{c}\,\Psi\left(t\right),\ \ \Phi\left(t,\tau\right)=\hat{U}_{t}\left(\tau\right)\Psi_{\hat{C}}\left(t\right), (70)

we arrive at

K(t,τ)=Φ(t,τ)|ΨC^(t+τ).K\left(t,\tau\right)=\left\langle\Phi\left(t,\tau\right)\right.\left|\Psi_{\hat{C}}\left(t+\tau\right)\right\rangle. (71)

Therefore, in order to calculate the correlator KK through the solution of the equation for the state vector, one has to perform the following steps:

a) Find vector |ΨC^(t+τ)=\left|\Psi_{\hat{C}}\left(t+\tau\right)\right\rangle= c^Ψ(t+τ)\hat{c}\Psi\left(t+\tau\right), where Ψ(t+τ)\Psi\left(t+\tau\right) is the solution for the state vector |Ψ\left|\Psi\right\rangle at the time interval [0,t+τ]\left[0,t+\tau\right] with initial condition |Ψ(0)\left|\Psi\left(0\right)\right\rangle.

b) Find vector |Φ(t,τ)\left|\Phi\left(t,\tau\right)\right\rangle. To do that, one has to solve for the state vector |Ψ\left|\Psi\right\rangle at the time interval [t,t+τ]\left[t,t+\tau\right] with initial condition |ΨC^(t)\left|\Psi_{\hat{C}}\left(t\right)\right\rangle. The vector |ΨC^(t)\left|\Psi_{\hat{C}}\left(t\right)\right\rangle is the same as in part a), but instead of the time interval [0,t+τ]\left[0,t+\tau\right] one has to take the time interval [0,t]\left[0,t\right].

To check Eq. (71) for consistency, we note that in the absence of dissipation one can go back from this equation to the standard expression which follows directly from the initial equation (66):

K(t,τ)=Ψ(0)|eiH^tc^eiH^τc^eiH^(t+τ)|Ψ(0).K\left(t,\tau\right)=\left\langle\Psi\left(0\right)\right|e^{i\frac{\hat{H}}{\hbar}t}\hat{c}^{\dagger}e^{i\frac{\hat{H}}{\hbar}\tau}\hat{c}e^{-i\frac{\hat{H}}{\hbar}\left(t+\tau\right)}\left|\Psi\left(0\right)\right\rangle.

If the dynamic system is open, then a complete closed system “the dynamic system + reservoir” has its own unitary operator of evolution U^t0(t)\hat{U}_{t_{0}}\left(t^{\prime}\right). Therefore, Eq. (71) should be valid for a complete system as well which includes the reservoir variables. Now we apply the Langevin method which assumes that the averaging over the statistics of noise sources entering a stochastic equation (in this case Eq. (38)) is equivalent to averaging over the reservoir variables. Therefore, we can solve Eq. (38) and, following the above steps, find the functions ΨC^(t)\Psi_{\hat{C}}\left(t\right), ΨC^(t+τ)\Psi_{\hat{C}}\left(t+\tau\right) and Φ(t,τ)\Phi\left(t,\tau\right) which are now dependent on the noise sources. Then we substitute the latter two functions into Eq. (71) and perform averaging over the noise statistics. As a result, we obtain

K(t,τ)=Φ(t,τ)|ΨC^(t+τ)¯.K\left(t,\tau\right)=\overline{\left\langle\ \Phi\left(t,\tau\right)\right.\left|\Psi_{\hat{C}}\left(t+\tau\right)\right\rangle}. (72)

VI.2 Photon emission spectra for the parametric decay of an excited electron

Here we apply the general recipe of calculating K(t,τ)K\left(t,\tau\right) formulated in the previous section to a particular example of the parametric decay of an initially excited fermionic two-level system under strong coupling to a parametric electron-photon-phonon resonance.

a) Use the expressions (56)-(59) to find the vector |Ψ(t)\left|\Psi\left(t\right)\right\rangle.

b) Determine vectors |ΨC^(t)\left|\Psi_{\hat{C}}\left(t\right)\right\rangle and |ΨC^(t+τ)\left|\Psi_{\hat{C}}\left(t+\tau\right)\right\rangle, resulting in

|ΨC^(t)=c^|Ψ(t)=C110(t)|1|0|0+C010(t)|0|0|0,\left|\Psi_{\hat{C}}\left(t\right)\right\rangle=\hat{c}\left|\Psi\left(t\right)\right\rangle=C_{110}\left(t\right)\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle+C_{010}\left(t\right)\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle, (73)
|ΨC^(t+τ)=c^|Ψ(t+τ)=C110(t+τ)|1|0|0+C010(t+τ)|0|0|0.\left|\Psi_{\hat{C}}\left(t+\tau\right)\right\rangle=\hat{c}\left|\Psi\left(t+\tau\right)\right\rangle=C_{110}\left(t+\tau\right)\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle+C_{010}\left(t+\tau\right)\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle. (74)

c) To determine the vector |Φ(t,τ)\left|\Phi\left(t,\tau\right)\right\rangle we will use the solution of Eqs. (47),(48) at the time interval [t,t+τ]\left[t,t+\tau\right] where the initial condition |ΨC^(t)\left|\Psi_{\hat{C}}\left(t\right)\right\rangle is given by Eq. (73). In our case Eq. (73) determines the initial value of the state vector; its subsequent evolution is determined by a simple Eq. (47) for the amplitudes of states |1|0|0\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle and |0|0|0\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle. As a result, vector |Φ(t,τ)\left|\Phi\left(t,\tau\right)\right\rangle is given by

|Φ(t,τ)\displaystyle\left|\Phi\left(t,\tau\right)\right\rangle =\displaystyle= C100(Φ)(t,τ)|1|0|0+C000(Φ)(t,τ)|0|0|0\displaystyle C_{100}^{\left(\Phi\right)}\left(t,\tau\right)\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle+C_{000}^{\left(\Phi\right)}\left(t,\tau\right)\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle (75)
=\displaystyle= (eiωτγ100τC110(t)itt+τ100(Φ)(t,t)eiω(τ+tt)γ100(τ+tt)𝑑t)|1|0|0\displaystyle\left(e^{-i\omega\tau-\gamma_{100}\tau}C_{110}\left(t\right)-\frac{i}{\hbar}\int_{t}^{t+\tau}\mathfrak{R}_{100}^{\left(\Phi\right)}\left(t,t^{\prime}\right)e^{-i\omega\left(\tau+t-t^{\prime}\right)-\gamma_{100}\left(\tau+t-t^{\prime}\right)}dt^{\prime}\right)\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle
+(C010(t)itt+τ000(Φ)(t,t)𝑑t)|0|0|0,\displaystyle+\left(C_{010}\left(t\right)-\frac{i}{\hbar}\int_{t}^{t+\tau}\mathfrak{R}_{000}^{\left(\Phi\right)}\left(t,t^{\prime}\right)dt^{\prime}\right)\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle,

where functions C110(t)C_{110}\left(t\right) and C010(t)C_{010}\left(t\right) are determined by Eqs. (56), (57).

The superscript (Φ)\left(\Phi\right) in the terms αni\mathfrak{R}_{\alpha ni} in Eq. (75) means that the correlators of these noise terms correspond to the state vector |Φ\left|\Phi\right\rangle. The dependence on the initial time moment tt of the evolution in αni(Φ)(t,t)\mathfrak{R}_{\alpha ni}^{\left(\Phi\right)}\left(t,t^{\prime}\right) takes into account that the correlators of these random functions may depend on the value of tt as a parameter, because complex amplitudes Cαni(Φ)(t,τ)Cαni(Φ)(t,tt)C_{\alpha ni}^{\left(\Phi\right)}\left(t,\tau\right)\equiv C_{\alpha ni}^{\left(\Phi\right)}\left(t,t^{\prime}-t\right) depend on this parameter.

Next, we substitute the expressions for C110(t)C_{110}\left(t\right), C010(t)C_{010}\left(t\right), C110(t+τ)C_{110}\left(t+\tau\right) determined by Eqs. (56) and (57), into Eqs. (74) and (75); after that we substitute Eqs. (74) and (75) into Eq. (72). In the resulting expression we average over the noise statistics taking into account that the noise sources are delta-correlated. Omitting the terms that become zero after averaging, we obtain

K(t,τ)\displaystyle K\left(t,\tau\right) =\displaystyle= eiωτ(γ100+γ110+γ0012)τ(γ110+γ001)tsin(Ω~Rt)sin[Ω~R(t+τ)]\displaystyle e^{-i\omega\tau-\left(\gamma_{100}+\frac{\gamma_{110}+\gamma_{001}}{2}\right)\tau-\left(\gamma_{110}+\gamma_{001}\right)t}\sin\left(\tilde{\Omega}_{R}t\right)\sin\left[\tilde{\Omega}_{R}\left(t+\tau\right)\right] (76)
+eiωτγ010τ2γ010t0tD010,010(t)e2γ010t𝑑t\displaystyle+e^{-i\omega\tau-\gamma_{010}\tau-2\gamma_{010}t}\int_{0}^{t}D_{010,010}\left(t^{\prime}\right)e^{2\gamma_{010}t^{\prime}}dt^{\prime}
+eiω(t+τ)γ010(t+τ)2γ010ttt+τD~000,010(t,t)e(iω+γ010)t𝑑t.\displaystyle+e^{-i\omega\left(t+\tau\right)-\gamma_{010}\left(t+\tau\right)-2\gamma_{010}t}\int_{t}^{t+\tau}\tilde{D}_{000,010}\left(t,t^{\prime}\right)e^{\left(i\omega+\gamma_{010}\right)t^{\prime}}dt^{\prime}.

Here the quantity D010,010D_{010,010} is determined by Eqs. (50):

D010,010=2γ010|C110(t)|2¯=μΩ|C110(t)|2¯.D_{010,010}=2\gamma_{010}\overline{\left|C_{110}\left(t^{\prime}\right)\right|^{2}}=\mu_{\Omega}\overline{\left|C_{110}\left(t^{\prime}\right)\right|^{2}}. (77)

The function D~000,010(t,t)\tilde{D}_{000,010}\left(t,t^{\prime}\right) corresponds to the following correlator:

000(Φ)(t,t)010(t′′)¯=2D~000,010(t,t)δ(tt′′).\overline{\mathfrak{R}_{000}^{\left(\Phi\right)\ast}\left(t,t^{\prime}\right)\mathfrak{R}_{010}\left(t^{\prime\prime}\right)}=\hbar^{2}\tilde{D}_{000,010}\left(t,t^{\prime}\right)\delta\left(t^{\prime}-t^{\prime\prime}\right). (78)

To calculate the value of D~000,010(t,t)\tilde{D}_{000,010}\left(t,t^{\prime}\right) it is not enough to have expressions (50) and (51) because it is determined by correlations between the noise terms for different state vectors |Φ\left|\Phi\right\rangle and |Ψ\left|\Psi\right\rangle, which correspond to the solutions of the equation for the state vector with different initial conditions. We need to use the expression for the noise source obtained in Sec. IV.D. From Eqs. (52), (54), and (55) we obtain

D~000,010(t,t)=μΩC100(Φ)(t,tt)C110(t)¯.\tilde{D}_{000,010}\left(t,t^{\prime}\right)=\mu_{\Omega}\overline{C_{100}^{\left(\Phi\right)\ast}\left(t,t^{\prime}-t\right)C_{110}\left(t^{\prime}\right)}. (79)

Substituting here the appropriate term from Eq. (75) gives

D~000,010(t,t)=eiω(tt)γ100(tt)2γ010tC110(t)C110(t)¯.\tilde{D}_{000,010}\left(t,t^{\prime}\right)=e^{i\omega\left(t^{\prime}-t\right)-\gamma_{100}\left(t^{\prime}-t\right)-2\gamma_{010}t}\overline{C_{110}^{\ast}\left(t\right)C_{110}\left(t^{\prime}\right)}. (80)

Substituting Eqs. (77) and (80) into Eq. (76), we arrive at

K(t,τ)\displaystyle K\left(t,\tau\right) =\displaystyle= eiωτ(γ100+Γ2)τΓtsin(Ω~Rt)sin[Ω~R(t+τ)]\displaystyle e^{-i\omega\tau-\left(\gamma_{100}+\frac{\Gamma}{2}\right)\tau-\Gamma t}\sin\left(\tilde{\Omega}_{R}t\right)\sin\left[\tilde{\Omega}_{R}\left(t+\tau\right)\right] (81)
+eiωτγ010τμΩ[2Ω~R2γn(4Ω~R2+γn2)e2γ010t12γneΓt\displaystyle+e^{-i\omega\tau-\gamma_{010}\tau}\mu_{\Omega}\left[\frac{2\tilde{\Omega}_{R}^{2}}{\gamma_{n}(4\tilde{\Omega}_{R}^{2}+\gamma_{n}^{2})}e^{-2\gamma_{010}t}-\frac{1}{2\gamma_{n}}e^{-\Gamma t}\right.
14(2iΩ~Rγn)e2iΩ~RtΓt+14(2iΩ~R+γn)e2iΩ~RtΓt]\displaystyle\left.-\frac{1}{4\left(2i\tilde{\Omega}_{R}-\gamma_{n}\right)}e^{2i\tilde{\Omega}_{R}t-\Gamma t}+\frac{1}{4\left(2i\tilde{\Omega}_{R}+\gamma_{n}\right)}e^{-2i\tilde{\Omega}_{R}t-\Gamma t}\right]
+eiωτγ010τΓtμΩ4[e(γd+iΩ~R)τ1γd+iΩ~R(1e2iΩ~Rt)+c.c.].\displaystyle+e^{-i\omega\tau-\gamma_{010}\tau-\Gamma t}\frac{\mu_{\Omega}}{4}\left[\frac{e^{\left(-\gamma_{d}+i\tilde{\Omega}_{R}\right)\tau}-1}{-\gamma_{d}+i\tilde{\Omega}_{R}}\left(1-e^{2i\tilde{\Omega}_{R}t}\right)+c.c.\right].

where we denoted Γ=γ110+γ001\Gamma=\gamma_{110}+\gamma_{001}, γn=\gamma_{n}= Γ2γ010\Gamma-2\gamma_{010}, γd=γ100+Γ2γ010\gamma_{d}=\gamma_{100}+\frac{\Gamma}{2}-\gamma_{010}.

Now we have everything to determine the emission spectra given by Eq. (65). Using the values of γ100=μΩ2\gamma_{100}=\frac{\mu_{\Omega}}{2} and γ010=μω2\gamma_{010}=\frac{\mu_{\omega}}{2}, we obtain

S(ν)=1πRe0𝑑τeiντ0𝑑tK(t,τ)=S1(ν)+S2(ν)+S3(ν),S\left(\nu\right)=\frac{1}{\pi}\mathrm{Re}\int_{0}^{\infty}d\tau e^{i\nu\tau}\int_{0}^{\infty}dtK\left(t,\tau\right)=S_{1}\left(\nu\right)+S_{2}\left(\nu\right)+S_{3}\left(\nu\right), (82)

where

S1(ν)\displaystyle S_{1}\left(\nu\right) =\displaystyle= 2Ω~R2πΓ(4Ω~R2+Γ2)ReΓ+μΩ2i(νω)[γaci(νω)]2+Ω~R2\displaystyle\frac{2\tilde{\Omega}_{R}^{2}}{\pi\Gamma(4\tilde{\Omega}_{R}^{2}+\Gamma^{2})}\mathrm{Re}\frac{\Gamma+\frac{\mu_{\Omega}}{2}-i\left(\nu-\omega\right)}{[\gamma_{\mathrm{ac}}-i\left(\nu-\omega\right)]^{2}+\tilde{\Omega}_{R}^{2}}
S2(ν)\displaystyle S_{2}\left(\nu\right) =\displaystyle= Ω~R2πΓ(4Ω~R2+Γ2)μΩμω24+(νω)2\displaystyle\frac{\tilde{\Omega}_{R}^{2}}{\pi\Gamma(4\tilde{\Omega}_{R}^{2}+\Gamma^{2})}\frac{\mu_{\Omega}}{\frac{\mu_{\omega}^{2}}{4}+\left(\nu-\omega\right)^{2}} (83)
S3(ν)\displaystyle S_{3}\left(\nu\right) =\displaystyle= μΩΩ~R2πΓ(γd2+Ω~R2)(4Ω~R2+Γ2)×\displaystyle\frac{\mu_{\Omega}\tilde{\Omega}_{R}^{2}}{\pi\Gamma\left(\gamma_{d}^{2}+\tilde{\Omega}_{R}^{2}\right)(4\tilde{\Omega}_{R}^{2}+\Gamma^{2})}\times
{ReΓd[γaci(νω)]+2Ω~R2γdΓ[γaci(νω)]2+Ω~R2+Γdμω2μω24+(νω)2}\displaystyle\left\{-\mathrm{Re}\frac{\Gamma_{d}\left[\gamma_{\mathrm{ac}}-i\left(\nu-\omega\right)\right]+2\tilde{\Omega}_{R}^{2}-\gamma_{d}\Gamma}{\left[\gamma_{\mathrm{ac}}-i\left(\nu-\omega\right)\right]^{2}+\tilde{\Omega}_{R}^{2}}+\frac{\Gamma_{d}\frac{\mu_{\omega}}{2}}{\frac{\mu_{\omega}^{2}}{4}+\left(\nu-\omega\right)^{2}}\right\}

The parameters Γ\Gamma, γac\gamma_{\mathrm{ac}}, γd\gamma_{d}, and Γd\Gamma_{d} are expressed through the relaxation rates of the electron, photon, and phonon subsystems μω\mu_{\omega}, μΩ\mu_{\Omega} and γ\gamma as

Γ\displaystyle\Gamma =\displaystyle= γ110+γ001=12(μω+μΩ+γ),γac=γ100+Γ2=14(μω+γ)+34μΩ,\displaystyle\gamma_{110}+\gamma_{001}=\frac{1}{2}\left(\mu_{\omega}+\mu_{\Omega}+\gamma\right),\ \gamma_{\mathrm{ac}}=\gamma_{100}+\frac{\Gamma}{2}=\frac{1}{4}\left(\mu_{\omega}+\gamma\right)+\frac{3}{4}\mu_{\Omega},
γd\displaystyle\gamma_{d} =\displaystyle= γ100+Γ2γ010=14(γμω)+34μΩ,Γd=Γ+2γd=2μΩ+γ.\displaystyle\gamma_{100}+\frac{\Gamma}{2}-\gamma_{010}=\frac{1}{4}\left(\gamma-\mu_{\omega}\right)+\frac{3}{4}\mu_{\Omega},\ \Gamma_{d}=\Gamma+2\gamma_{d}=2\mu_{\Omega}+\gamma.

The expression for the power spectrum S(ν)S\left(\nu\right) contains the terms S2S_{2} and S3S_{3} which originate from the correlators D010,010D_{010,010} and D~000,010\tilde{D}_{000,010} respectively. The term S3(ν)S_{3}\left(\nu\right) consists of two terms which have the same spectral shapes as the functions S1(ν)S_{1}\left(\nu\right) and S2(ν)S_{2}\left(\nu\right) respectively. Therefore, including the term S3(ν)S_{3}\left(\nu\right) in Eq. (82) leads only to corrections to the amplitudes of the functions S1,2(ν)S_{1,2}\left(\nu\right); moreover, under the strong coupling conditions Ω~R\tilde{\Omega}_{R}\gg γani\gamma_{ani} these corrections are small: of the order of μΩ(Γ+γac)Ω~R2\sim\frac{\mu_{\Omega}\left(\Gamma+\gamma_{\mathrm{ac}}\right)}{\tilde{\Omega}_{R}^{2}} for the function S1(ν)S_{1}\left(\nu\right) and of the order of ΓdμωΩ~R2\sim\frac{\Gamma_{d}\mu_{\omega}}{\tilde{\Omega}_{R}^{2}} for the function S2(ν)S_{2}\left(\nu\right). For qualitative discussion we will neglect the contribution of S3(ν)S_{3}\left(\nu\right) and keep only the terms S1(ν)S_{1}\left(\nu\right) and S2(ν)S_{2}\left(\nu\right), although all terms are included in the spectra plotted in Fig. 4.

Refer to caption
Figure 3: Energy levels of |phonon|photon|electron\left|phonon\right\rangle\left|photon\right\rangle\left|electron\right\rangle states involved into the photon emission in the parametric decay of a single-electron excitation in a coupled phonon-photon-electron system. Bold red arrows indicate photon emission transitions with their peak frequencies labeled. Wavy purple arrows indicate various relaxation pathways.
Refer to caption
Figure 4: Normalized photon emission spectra Ω~R2S(ν)\tilde{\Omega}_{R}^{2}S(\nu) as a function of normalized detuning νωΩ~R\frac{\nu-\omega}{\tilde{\Omega}_{R}} from the cavity mode frequency ω\omega, for two different values of the phonon relaxation rate: μΩ=0.3\mu_{\Omega}=0.3 (red solid line) and μΩ=0.02\mu_{\Omega}=0.02 (blue dashed line). Other relaxation rates are μω=0.2\mu_{\omega}=0.2 and γ=0.1\gamma=0.1. All relaxation constants are in units of Ω~R\tilde{\Omega}_{R}.

Figure 3 indicates all transitions that give contributions to the photon emission. The function S1(ν)S_{1}\left(\nu\right) describes the emission spectrum at the transition |1|1|0|1|0|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle, which is split due to Rabi oscillations. The width of the peaks located at frequencies ν=ω±Ω~R\nu=\omega\pm\tilde{\Omega}_{R} is equal to γac=μΩ2+Γ2\gamma_{\mathrm{ac}}=\frac{\mu_{\Omega}}{2}+\frac{\Gamma}{2}. Here Γ2=14(μω+μΩ+γ)=γMIX\frac{\Gamma}{2}=\frac{1}{4}\left(\mu_{\omega}+\mu_{\Omega}+\gamma\right)=\gamma_{MIX} is the decay rate of the entangled state |MIX=A(t)|0|0|1+B(t)|1|1|0\left|MIX\right\rangle=A\left(t\right)\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle+B\left(t\right)\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle (see Eq. (59)), and μΩ2\frac{\mu_{\Omega}}{2} is the broadening of the state |1|0|0\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle due to relaxation. The spectrum given by S1(ν)S_{1}\left(\nu\right) agrees with the one obtained in tokman2021 .

The function S2(ν)S_{2}\left(\nu\right) describes the emission due to a two-step relaxation process described in section IV.C: |1|1|0|0|1|0|0|0|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle . The photons are emitted at the transition |0|1|0|0|0|0\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle, which is not affected by Rabi oscillations; see Fig. 3. Therefore, this contribution has a standard Lorentzian shape of an emitter at frequency ω\omega:

S2(ν)1γ0102+(νω)2=1μω24+(νω)2.S_{2}\left(\nu\right)\propto\frac{1}{\gamma_{010}^{2}+\left(\nu-\omega\right)^{2}}=\frac{1}{\frac{\mu_{\omega}^{2}}{4}+\left(\nu-\omega\right)^{2}}.

In the strong-coupling regime Ω~R\tilde{\Omega}_{R}\gg γani\gamma_{ani} the ratio of the amplitude of this central peak at frequency ν=ω\nu=\omega to the amplitudes of the split peaks at frequencies ν=ω±Ω~R\nu=\omega\pm\tilde{\Omega}_{R} is given by

S2(ω)S1(ω±Ω~R)μΩ(μω+γ+3μΩ)μω2.\frac{S_{2}\left(\omega\right)}{S_{1}\left(\omega\pm\tilde{\Omega}_{R}\right)}\approx\frac{\mu_{\Omega}\left(\mu_{\omega}+\gamma+3\mu_{\Omega}\right)}{\mu_{\omega}^{2}}. (84)

When μΩ0\mu_{\Omega}\rightarrow 0, Eq. (84) gives S2(ω)S1(ω±Ω~R)0\frac{S_{2}\left(\omega\right)}{S_{1}\left(\omega\pm\tilde{\Omega}_{R}\right)}\rightarrow 0: indeed without phonon relaxation the state |0|1|0\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle cannot be populated from |1|1|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle; therefore, the two-step radiation channel is suppressed. In the opposite limit of a fast phonon relaxation, when μΩ\mu_{\Omega}\gg μω,γ\mu_{\omega},\gamma, we obtain S2(ω)S1(ω±Ω~R)1\frac{S_{2}\left(\omega\right)}{S_{1}\left(\omega\pm\tilde{\Omega}_{R}\right)}\gg 1, i.e., the side peaks are weaker and more broadened than the central peak. The latter statement is true despite the large Rabi frequency Ω~R\tilde{\Omega}_{R}\gg μΩ\mu_{\Omega}. Indeed, when Ω~R\tilde{\Omega}_{R}\gg μΩ\mu_{\Omega} the interaction has the time to mix the states |1|1|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle and |0|0|1\left|0\right\rangle\left|0\right\rangle\left|1\right\rangle before the phonon relaxation kicks in. Nevertheless, if μΩ\mu_{\Omega}\gg μω,γ\mu_{\omega},\gamma the phonon relaxation is able to transfer population to state |0|1|0\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle faster than the radiative transition |1|1|0|1|0|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle corresponding to the Rabi-split spectrum.

This behavior is illustrated in Fig. 4 which shows photon emission spectra given by Eqs. (82) and (83) as a function of frequency detuning νω\nu-\omega from the cavity mode resonance, for two different values of the phonon relaxation rate: μΩ=1.5μω\mu_{\Omega}=1.5\mu_{\omega} (red solid line) and μΩ=0.1μω\mu_{\Omega}=0.1\mu_{\omega} (blue dashed line). The electron relaxation rate is kept at γ=0.1\gamma=0.1 and its exact value is not important for the overall shape of the spectra, although it affects absolute values and widths of the peaks. All quantities are normalized by Ω~R\tilde{\Omega}_{R}.

The relative magnitudes of the peaks and their widths depend sensitively on different combinations of the relaxation rates γ,μω,μΩ\gamma,\mu_{\omega},\mu_{\Omega}. The onset of the strong coupling regime in the frequency domain is determined by the visibility of nonlinear Rabi splitting between the side peaks in Fig. 4, i.e., the condition Ω~R>γac2\tilde{\Omega}_{R}>\frac{\gamma_{ac}}{2}. We point out again that the relaxation rates of the individual subsystems enter the quantum dynamics in a very nontrivial way, and one need to know all of them to evaluate the feasibility of strong coupling in any particular system. The reverse is also true: once the strong coupling regime is reached, measurements of the photoluminescence spectra yield both the relaxation rates and the nonlinear coupling strength in the system.

A more detailed discussion of the feasibility of strong coupling at the nonlinear resonance in particular systems can be found in tokman2021 . Here we only point out that in dielectric microcavities the photon relaxation rates can be very low, in the μ\mueV range, and the strong coupling threshold is likely to be determined by relaxation of the electron or vibrational transitions. In plasmonic nanocavities the photon relaxation rate can easily be tens of meV and will likely dominate the strong coupling threshold. On the other hand, the nonlinear coupling strength ΩR(3)\Omega_{R}^{(3)} is much higher in plasmonic nanocavities because of greatly enhanced electric field localization and electric field gradient. One can obtain the magnitude of ΩR(3)\Omega_{R}^{(3)} of the order of 100 meV for the field localization in the few nm range, which is now routinely demonstrated in plasmonic nanocavities.

VI.3 The phonon emission spectra

There is a complete symmetry for the two bosonic fields in the decay process close to the parametric resonance ωe=ω+Ω\omega_{e}=\omega+\Omega. Therefore, we can obtain the phonon emission spectrum from the expressions for the photon emission spectrum Eqs. (82), (83), after replacing

ωΩ,μωμΩ,γ010γ100.\omega\Longleftrightarrow\Omega,\ \mu_{\omega}\Longleftrightarrow\mu_{\Omega},\ \gamma_{010}\Longleftrightarrow\gamma_{100}.

This results in

Sp(ν)=S1p(ν)+S2p(ν)+S3p(ν),S_{p}\left(\nu\right)=S_{1p}\left(\nu\right)+S_{2p}\left(\nu\right)+S_{3p}\left(\nu\right), (85)

where

S1p(ν)\displaystyle S_{1p}\left(\nu\right) =\displaystyle= 2Ω~R2πΓ(4Ω~R2+Γ2)ReΓ+μΩ2i(νΩ)[γ~aci(νΩ)]2+Ω~R2\displaystyle\frac{2\tilde{\Omega}_{R}^{2}}{\pi\Gamma(4\tilde{\Omega}_{R}^{2}+\Gamma^{2})}\mathrm{Re}\frac{\Gamma+\frac{\mu_{\Omega}}{2}-i\left(\nu-\Omega\right)}{[\tilde{\gamma}_{\mathrm{ac}}-i\left(\nu-\Omega\right)]^{2}+\tilde{\Omega}_{R}^{2}}
S2p(ν)\displaystyle S_{2p}\left(\nu\right) =\displaystyle= Ω~R2πΓ(4Ω~R2+Γ2)μωμΩ24+(νΩ)2\displaystyle\frac{\tilde{\Omega}_{R}^{2}}{\pi\Gamma(4\tilde{\Omega}_{R}^{2}+\Gamma^{2})}\frac{\mu_{\omega}}{\frac{\mu_{\Omega}^{2}}{4}+\left(\nu-\Omega\right)^{2}} (86)
S3p(ν)\displaystyle S_{3p}\left(\nu\right) =\displaystyle= μωΩ~R2πΓ(γ~d2+Ω~R2)(4Ω~R2+Γ2)×\displaystyle\frac{\mu_{\omega}\tilde{\Omega}_{R}^{2}}{\pi\Gamma\left(\tilde{\gamma}_{d}^{2}+\tilde{\Omega}_{R}^{2}\right)(4\tilde{\Omega}_{R}^{2}+\Gamma^{2})}\times
{ReΓ~d[γ~aci(νΩ)]+2Ω~R2γ~dΓ[γ~aci(νΩ)]2+Ω~R2+Γ~dμΩ2μΩ24+(νΩ)2}\displaystyle\left\{-\mathrm{Re}\frac{\tilde{\Gamma}_{d}\left[\tilde{\gamma}_{\mathrm{ac}}-i\left(\nu-\Omega\right)\right]+2\tilde{\Omega}_{R}^{2}-\tilde{\gamma}_{d}\Gamma}{\left[\tilde{\gamma}_{\mathrm{ac}}-i\left(\nu-\Omega\right)\right]^{2}+\tilde{\Omega}_{R}^{2}}+\frac{\tilde{\Gamma}_{d}\frac{\mu_{\Omega}}{2}}{\frac{\mu_{\Omega}^{2}}{4}+\left(\nu-\Omega\right)^{2}}\right\}
γ~ac\displaystyle\tilde{\gamma}_{\mathrm{ac}} =\displaystyle= γ010+Γ2=14(μΩ+γ)+34μω,γ~d=γ010+Γ2γ100=14(γμΩ)+34μω\displaystyle\gamma_{010}+\frac{\Gamma}{2}=\frac{1}{4}\left(\mu_{\Omega}+\gamma\right)+\frac{3}{4}\mu_{\omega},\ \tilde{\gamma}_{d}=\gamma_{010}+\frac{\Gamma}{2}-\gamma_{100}=\frac{1}{4}\left(\gamma-\mu_{\Omega}\right)+\frac{3}{4}\mu_{\omega}
Γ~d\displaystyle\ \tilde{\Gamma}_{d} =\displaystyle= Γ+2γ~d=2μω+γ.\displaystyle\Gamma+2\tilde{\gamma}_{d}=2\mu_{\omega}+\gamma.

Similarly to the photon spectrum, the term S3p(ν)S_{3p}\left(\nu\right) is the sum of two terms which have the same spectral shape as S1p(ν)S_{1p}\left(\nu\right) and S2p(ν)S_{2p}\left(\nu\right), but much smaller magnitudes if Ω~R\tilde{\Omega}_{R}\gg γani\gamma_{ani}. Therefore we will again include only the first two terms in qualitative discussion, but include all terms when plotting the spectra.

Refer to caption
Figure 5: Energy levels of |phonon|photon|electron\left|phonon\right\rangle\left|photon\right\rangle\left|electron\right\rangle states involved into the phonon emission in the parametric decay of a single-electron excitation in a coupled phonon-photon-electron system. Bold red arrows indicate phonon emission transitions with their peak frequencies labeled. Wavy purple arrows indicate various relaxation pathways.
Refer to caption
Figure 6: Normalized phonon emission spectra Ω~R2Sp(ν)\tilde{\Omega}_{R}^{2}S_{p}(\nu) as a function of normalized detuning νΩΩ~R\frac{\nu-\Omega}{\tilde{\Omega}_{R}} from the vibrational mode frequency Ω\Omega, for two different values of the photon relaxation rate: μω=0.3\mu_{\omega}=0.3 (red solid line) and μω=0.02\mu_{\omega}=0.02 (blue dashed line). Other relaxation rates are μΩ=0.2\mu_{\Omega}=0.2 and γ=0.1\gamma=0.1. All relaxation constants are in units of Ω~R\tilde{\Omega}_{R}.

Figure 5 shows all transitions giving contributions to the phonon emission spectrum, with their peak frequencies indicated.

The function S1p(ν)S_{1p}\left(\nu\right) describes the phonon emission spectrum due to the transition |1|1|0|0|1|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|1\right\rangle\left|0\right\rangle, which demonstrates Rabi splitting. The width of the peaks centered at frequencies ν=Ω±Ω~R\nu=\Omega\pm\tilde{\Omega}_{R} is equal to γ~ac=μω2+Γ2\tilde{\gamma}_{\mathrm{ac}}=\frac{\mu_{\omega}}{2}+\frac{\Gamma}{2}. The function S2p(ν)S_{2p}\left(\nu\right) describes phonon emission due to a two-step relaxation |1|1|0|1|0|0|0|0|0\left|1\right\rangle\left|1\right\rangle\left|0\right\rangle\rightarrow\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle. The phonons are emitted at the second step, i.e. the transition |1|0|0|0|0|0\left|1\right\rangle\left|0\right\rangle\left|0\right\rangle\rightarrow\left|0\right\rangle\left|0\right\rangle\left|0\right\rangle, which is not affected by Rabi oscillations. Therefore, the spectrum due to this contribution is a standard Lorenzian line, similarly to the case of photons:

S2(ν)1γ1002+(νΩ)2=1μΩ24+(νΩ)2.S_{2}\left(\nu\right)\propto\frac{1}{\gamma_{100}^{2}+\left(\nu-\Omega\right)^{2}}=\frac{1}{\frac{\mu_{\Omega}^{2}}{4}+\left(\nu-\Omega\right)^{2}}.

The ratio of the amplitude of the central peak at ν=Ω\nu=\Omega to those of the side peaks at frequencies ν=Ω±Ω~R\nu=\Omega\pm\tilde{\Omega}_{R} at Ω~R\tilde{\Omega}_{R}\gg γani\gamma_{ani} is given by the expression equivalent to Eq. (84) after substituting ωΩ\omega\Longleftrightarrow\Omega\ and μωμΩ\mu_{\omega}\Longleftrightarrow\mu_{\Omega}:

S2p(Ω)S1p(Ω±Ω~R)μω(μΩ+γ+3μω)μΩ2.\frac{S_{2p}\left(\Omega\right)}{S_{1p}\left(\Omega\pm\tilde{\Omega}_{R}\right)}\approx\frac{\mu_{\omega}\left(\mu_{\Omega}+\gamma+3\mu_{\omega}\right)}{\mu_{\Omega}^{2}}. (87)

The phonon emission spectra are plotted in Fig. 6 as a function of frequency detuning νΩ\nu-\Omega from the cavity mode resonance. This time we keep the phonon relaxation rate fixed at μΩ=0.2\mu_{\Omega}=0.2 and plot the spectra for two values of the photon relaxation rate, greater and smaller than μΩ\mu_{\Omega}: μω=0.3\mu_{\omega}=0.3 (red solid line) and μω=0.02\mu_{\omega}=0.02 (blue dashed line). All quantities are normalized by Ω~R\tilde{\Omega}_{R}. The numbers are chosen to prove the point that the phonon and photon spectra are symmetric with respect to replacement indicated in the beginning of this section.

In experiment, measuring the ratios given by Eq. (84) and (87) allows one to determine the relationships between all relaxation rates μΩ\mu_{\Omega}, γ\gamma, and μω\mu_{\omega}. Indeed, one can obtain from Eqs. (84) and (87) that

ξΩx3+2x22xξω=0,y=ξΩx23x;\xi_{\Omega}x^{3}+2x^{2}-2x-\xi_{\omega}=0,\ y=\xi_{\Omega}x^{2}-3-x;

where x=μΩμωx=\frac{\mu_{\Omega}}{\mu_{\omega}}, y=y= γμω\frac{\gamma}{\mu_{\omega}}, ξω=S2(ω)S1(ω±Ω~R)\xi_{\omega}=\frac{S_{2}\left(\omega\right)}{S_{1}\left(\omega\pm\tilde{\Omega}_{R}\right)}, ξΩ=S2p(Ω)S1p(Ω±Ω~R)\xi_{\Omega}=\frac{S_{2p}\left(\Omega\right)}{S_{1p}\left(\Omega\pm\tilde{\Omega}_{R}\right)}.

VII Conclusions

We developed a universal model of three-wave nonlinear resonance which is applicable to the systems with coupled electron, photon, and vibrational degrees of freedom, for example in cavity QED with molecular quantum emitters or quantum dots and in cavity optomechanics systems. We obtained the analytic solution for the nonperturbative quantum dynamics of such systems in the vicinity of the nonlinear resonance, taking into account dissipation and fluctuations for all degrees of freedom in Markov approximation. We showed that the strong coupling regime can be reached at the nonlinear resonance when the nonlinear coupling strength exceeds a certain threshold which includes all relaxation rates. When all three degrees of freedom are quantized, strong coupling leads to the formation of tripartite entangled states. The presented solution can be used to derive the explicit analytic expression for any observable. As an example, we calculated photon and phonon emission spectra which have a characteristic three-peak form once the strong coupling is reached. We showed how the relative heights and widths of the peaks can be used to extract information about all relaxation rates in the system and the nonlinear coupling strength, or to establish the threshold for reaching the strong coupling regime.

Acknowledgements.
This work has been supported in part by the Air Force Office for Scientific Research Grant No. FA9550-21-1-0272, National Science Foundation Award No. 1936276, and Texas A&M University through STRP, X-grant and T3-grant programs. M.T. and M.E. acknowledge the support by the Center of Excellence “Center of Photonics” funded by The Ministry of Science and Higher Education of the Russian Federation, Contract No. 075-15-2020-906.

Appendix A The derivation of Eqs. (62)-(64)

A number of correlators of the random functions in Eq. (59) is zero due to Eqs. (50),(51):

δC110(t)αmi(t′′)¯=δC001(t)αmi(t′′)¯=0\overline{\delta C_{110}^{\ast}\left(t^{\prime}\right)\mathfrak{R}_{\alpha mi}\left(t^{\prime\prime}\right)}=\overline{\delta C_{001}^{\ast}\left(t^{\prime}\right)\mathfrak{R}_{\alpha mi}\left(t^{\prime\prime}\right)}=0 (88)
δC001(t)δC001(t′′)¯=δC110(t)δC110(t′′)¯=δC001(t)δC110(t′′)¯=0\overline{\delta C_{001}^{\ast}\left(t^{\prime}\right)\delta C_{001}\left(t^{\prime\prime}\right)}=\overline{\delta C_{110}^{\ast}\left(t^{\prime}\right)\delta C_{110}\left(t^{\prime\prime}\right)}=\overline{\delta C_{001}^{\ast}\left(t^{\prime}\right)\delta C_{110}\left(t^{\prime\prime}\right)}=0 (89)
δC001(t)C000(t′′)¯\displaystyle\overline{\delta C_{001}^{\ast}\left(t^{\prime}\right)C_{000}\left(t^{\prime\prime}\right)} =\displaystyle= δC110(t)C100(t′′)¯=δC001(t)C010(t′′)¯=δC110(t)C000(t′′)¯\displaystyle\overline{\delta C_{110}^{\ast}\left(t^{\prime}\right)C_{100}\left(t^{\prime\prime}\right)}=\overline{\delta C_{001}^{\ast}\left(t^{\prime}\right)C_{010}\left(t^{\prime\prime}\right)}=\overline{\delta C_{110}^{\ast}\left(t^{\prime}\right)C_{000}\left(t^{\prime\prime}\right)} (90)
=\displaystyle= δC110(t)C100(t′′)¯=δC110(t)C010(t′′)¯=0.\displaystyle\overline{\delta C_{110}^{\ast}\left(t^{\prime}\right)C_{100}\left(t^{\prime\prime}\right)}=\overline{\delta C_{110}^{\ast}\left(t^{\prime}\right)C_{010}\left(t^{\prime\prime}\right)}=0.

Eqs. (88)-(90) ensure that the variables δC100\delta C_{100} and δC110\delta C_{110} cannot contribute to the values of any observables and therefore can be omitted.

The other correlators are given by the equations that follow from Eqs. (56):

ddtC100C010¯=(γ100+γ010)C100C010¯+D100,010,\frac{d}{dt}\overline{C_{100}^{\ast}C_{010}}=-\left(\gamma_{100}+\gamma_{010}\right)\overline{C_{100}^{\ast}C_{010}}+D_{100,010},
ddtC100C000¯=γ100C100C000¯+D100,000,\frac{d}{dt}\overline{C_{100}^{\ast}C_{000}}=-\gamma_{100}\overline{C_{100}^{\ast}C_{000}}+D_{100,000},
ddtC010C000¯=γ010C010C000¯+D010,000,\frac{d}{dt}\overline{C_{010}^{\ast}C_{000}}=-\gamma_{010}\overline{C_{010}^{\ast}C_{000}}+D_{010,000},
ddt|C000|2¯=D000,000,\frac{d}{dt}\overline{\left|C_{000}\right|^{2}}=D_{000,000},
ddt|C010|2¯=2γ010|C010|2¯+D010,010,\frac{d}{dt}\overline{\left|C_{010}\right|^{2}}=-2\gamma_{010}\overline{\left|C_{010}\right|^{2}}+D_{010,010},
ddt|C100|2¯=2γ100|C100|2¯+D100,100.\frac{d}{dt}\overline{\left|C_{100}\right|^{2}}=-2\gamma_{100}\overline{\left|C_{100}\right|^{2}}+D_{100,100}.

Using Eqs. (50),(51), we arrive at

ddtC100C010¯\displaystyle\frac{d}{dt}\overline{C_{100}^{\ast}C_{010}} =\displaystyle= μω+μΩ2C100C010¯\displaystyle-\frac{\mu_{\omega}+\mu_{\Omega}}{2}\overline{C_{100}^{\ast}C_{010}}
ddtC100C000¯\displaystyle\frac{d}{dt}\overline{C_{100}^{\ast}C_{000}} =\displaystyle= μΩ2C100C000¯+μωC110C010¯\displaystyle-\frac{\mu_{\Omega}}{2}\overline{C_{100}^{\ast}C_{000}}+\mu_{\omega}\overline{C_{110}^{\ast}C_{010}} (91)
ddtC010C000¯\displaystyle\frac{d}{dt}\overline{C_{010}^{\ast}C_{000}} =\displaystyle= μω2C010C000¯+μΩC110C100¯\displaystyle-\frac{\mu_{\omega}}{2}\overline{C_{010}^{\ast}C_{000}}+\mu_{\Omega}\overline{C_{110}^{\ast}C_{100}}

ddt|C000|2¯\displaystyle\frac{d}{dt}\overline{\left|C_{000}\right|^{2}} =\displaystyle= γ|C001|2¯+μω|C010|2¯+μΩ|C100|2¯\displaystyle\gamma\overline{\left|C_{001}\right|^{2}}+\mu_{\omega}\overline{\left|C_{010}\right|^{2}}+\mu_{\Omega}\overline{\left|C_{100}\right|^{2}}
ddt|C010|2¯\displaystyle\frac{d}{dt}\overline{\left|C_{010}\right|^{2}} =\displaystyle= μω|C010|2¯+μΩ|C110|2¯\displaystyle-\mu_{\omega}\overline{\left|C_{010}\right|^{2}}+\mu_{\Omega}\overline{\left|C_{110}\right|^{2}} (92)
ddt|C100|2¯\displaystyle\frac{d}{dt}\overline{\left|C_{100}\right|^{2}} =\displaystyle= μΩ|C100|2¯+μω|C110|2¯\displaystyle-\mu_{\Omega}\overline{\left|C_{100}\right|^{2}}+\mu_{\omega}\overline{\left|C_{110}\right|^{2}}

Taking into account Eqs. (90), one can obtain that C110C010¯=C110C100¯\overline{C_{110}^{\ast}C_{010}}=\overline{C_{110}^{\ast}C_{100}} =0=0 in Eqs. (91); as a result, for our initial conditions Eqs. (91) yield

C100C010¯=C100C000¯=C010C000¯=0.\overline{C_{100}^{\ast}C_{010}}=\overline{C_{100}^{\ast}C_{000}}=\overline{C_{010}^{\ast}C_{000}}=0. (93)

The last two equations in Eqs. (92) give

|C010|2¯=μΩeμωt0teμωτ|C110|2¯𝑑τ,|C100|2¯=μωeμΩt0teμΩτ|C110|2¯𝑑τ.\overline{\left|C_{010}\right|^{2}}=\mu_{\Omega}e^{-\mu_{\omega}t}\int_{0}^{t}e^{\mu_{\omega}\tau}\overline{\left|C_{110}\right|^{2}}d\tau,\ \ \overline{\left|C_{100}\right|^{2}}=\mu_{\omega}e^{-\mu_{\Omega}t}\int_{0}^{t}e^{\mu_{\Omega}\tau}\overline{\left|C_{110}\right|^{2}}d\tau.

Substituting here the function from Eq. (57) and taking into account Eqs. (89),(90) results in

|C100|2¯=μωeμΩt0teμΩμωγ2τsin2(Ω~Rτ)𝑑τ,|C010|2¯=μΩeμωt0teμωμΩγ2τsin2(Ω~Rτ)𝑑τ;\overline{\left|C_{100}\right|^{2}}=\mu_{\omega}e^{-\mu_{\Omega}t}\int_{0}^{t}e^{\frac{\mu_{\Omega}-\mu_{\omega}-\gamma}{2}\tau}\sin^{2}\left(\tilde{\Omega}_{R}\tau\right)d\tau,\ \ \overline{\left|C_{010}\right|^{2}}=\mu_{\Omega}e^{-\mu_{\omega}t}\int_{0}^{t}e^{\frac{\mu_{\omega}-\mu_{\Omega}-\gamma}{2}\tau}\sin^{2}\left(\tilde{\Omega}_{R}\tau\right)d\tau;

For further integration we use the limit γαniΩ~R1\frac{\gamma_{\alpha ni}}{\tilde{\Omega}_{R}}\ll 1, leading to

|C100|2¯μωμΩμωγ(eμΩ+μω+γ2teμΩt)\overline{\left|C_{100}\right|^{2}}\approx\frac{\mu_{\omega}}{\mu_{\Omega}-\mu_{\omega}-\gamma}\left(e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\Omega}t}\right) (94)
|C010|2¯μΩμωμΩγ(eμΩ+μω+γ2teμωt)\overline{\left|C_{010}\right|^{2}}\approx\frac{\mu_{\Omega}}{\mu_{\omega}-\mu_{\Omega}-\gamma}\left(e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\omega}t}\right) (95)

Note that Eqs. (94),(95) do not contain any divergence when [±(μωμΩ)γ]0\left[\pm\left(\mu_{\omega}-\mu_{\Omega}\right)-\gamma\right]\longrightarrow 0. Indeed,

lim(μΩμωγ)0[eμΩ+μω+γ2teμΩtμΩμωγ]=12teμΩt,lim(μωμΩγ)0[eμΩ+μω+γ2teμωtμωμΩγ]=12teμωt.\lim_{\left(\mu_{\Omega}-\mu_{\omega}-\gamma\right)\longrightarrow 0}\left[\frac{e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\Omega}t}}{\mu_{\Omega}-\mu_{\omega}-\gamma}\right]=\frac{1}{2}te^{-\mu_{\Omega}t},\ \lim_{\left(\mu_{\omega}-\mu_{\Omega}-\gamma\right)\longrightarrow 0}\left[\frac{e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}-e^{-\mu_{\omega}t}}{\mu_{\omega}-\mu_{\Omega}-\gamma}\right]=\frac{1}{2}te^{-\mu_{\omega}t}.

Now we return to the first of Eqs. (92), which yields

|C000|2¯=0t(γ|C001|2¯+μω|C010|2¯+μΩ|C100|2¯)𝑑τ.\overline{\left|C_{000}\right|^{2}}=\int_{0}^{t}\left(\gamma\overline{\left|C_{001}\right|^{2}}+\mu_{\omega}\overline{\left|C_{010}\right|^{2}}+\mu_{\Omega}\overline{\left|C_{100}\right|^{2}}\right)d\tau.

We substitute Eqs. (94) and (95) into the second and third terms in the integrand and substitute the expression |C001|2¯\overline{\left|C_{001}\right|^{2}} which follows from Eqs. (57) into the first term in the integrand. Neglecting the small terms \propto γ110γ001Ω~R\frac{\gamma_{110}-\gamma_{001}}{\tilde{\Omega}_{R}}and γαniΩ~R\frac{\gamma_{\alpha ni}}{\tilde{\Omega}_{R}} the integration results in

|C000|2¯=γ(γμΩμω)γ2(μΩμω)2(1eμΩ+μω+γ2t)(μΩ1eμωtμωμΩγ+μω1eμΩtμΩμωγ)\overline{\left|C_{000}\right|^{2}}=\frac{\gamma\left(\gamma-\mu_{\Omega}-\mu_{\omega}\right)}{\gamma^{2}-\left(\mu_{\Omega}-\mu_{\omega}\right)^{2}}\left(1-e^{-\frac{\mu_{\Omega}+\mu_{\omega}+\gamma}{2}t}\right)-\left(\mu_{\Omega}\frac{1-e^{-\mu_{\omega}t}}{\mu_{\omega}-\mu_{\Omega}-\gamma}+\mu_{\omega}\frac{1-e^{-\mu_{\Omega}t}}{\mu_{\Omega}-\mu_{\omega}-\gamma}\right) (96)

References

  • (1) M. K. Schmidt, R. Esteban, A. Gonzalez-Tudela, G. Giedke, and J. Aizpurua, Quantum Mechanical Description of Raman Scattering from Molecules in Plasmonic Cavities, ACS Nano 10, 6291-6298 (2016).
  • (2) J. del Pino, J. Feist, and F. J. Garcia-Vidal, Signatures of Vibrational Strong Coupling in Raman Scattering, J. Phys. Chem. 119, 29132-29137 (2015).
  • (3) J. del Pino, F. J. Garcia-Vidal, and J. Feist, Exploiting vibrational strong coupling to make an optical parametric oscillator out of a Raman laser, Phys. Rev. Lett. 117, 277401 (2016).
  • (4) S. Ramelow, A. Farsi, Z. Vernon, S. Clemmen, X. Ji, J. E. Sipe, M. Liscidini, M. Lipson, and A. L. Gaeta, Strong Nonlinear Coupling in a Si3N4 Ring Resonator, Phys. Rev. Lett. 122, 153906 (2019).
  • (5) T. Neuman, J. Aizpurua and R. Esteban, Quantum theory of surface-enhanced resonant Raman scattering (SERRS) of molecules in strongly coupled plasmon-exciton systems, Nanophotonics 9, 295-308 (2020).
  • (6) F. Ge, X. Han, and J. Xu, Strongly coupled systems for nonlinear optics, Laser Photonics Rev. 15, 2000514 (2021).
  • (7) K. Wang, M. Seidel, K. Nagarajan, T. Chervy, C. Genet, and T. Ebbesen, Large optical nonlinearity enhancement under electronic strong coupling, Nat. Commun. 12, 1486 (2021).
  • (8) J. Flick, N. Rivera and P. Narang, Strong light-matter coupling in quantum chemistry and quantum photonics, Nanophotonics 7, 1479-1501 (2018).
  • (9) D. Wang, J. Flick, and P. Narang, Light-matter interaction of a molecule in a dissipative cavity from first principles, J. Chem. Phys. 154, 104109 (2021).
  • (10) A. V. Zasedatelev, A. V. Baranikov, D. Sannikov, D. Urbonas, F. Scafirimuto, V. Yu. Shishkov, E. S. Andrianov, Y. E. Lozovik, U. Scherf, T. Stoferle, R. F. Mahrt and P. G. Lagoudakis, Single-photon nonlinearity at room temperature, Nature 597, 493 (2021).
  • (11) A. Pscherer, M.Meierhofer, D. Wang, Hr. Kelkar, D. Martin-Cano, T. Utikal ,S. Gotzinger, and V. Sandoghdar, Single-Molecule Vacuum Rabi Splitting: Four-Wave Mixing and Optical Switching at the Single-Photon Level, Phys. Rev. Lett. 127, 133603 (2021).
  • (12) S. Hughes, A. Settineri, S. Savasta, and F. Nori, Resonant Raman scattering of single molecules under simultaneous strong cavity coupling and ultrastrong optomechanical coupling in plasmonic resonators: Phonon-dressed polaritons, Phys. Rev. B 104, 045431 (2021).
  • (13) L.D. Landau, E.M. Lifshitz, Electrodynamics of continuous media (Elsevier, Oxford, 1984).
  • (14) F. Benz, M. K. Schmidt, A. Dreismann, R. Chikkaraddy, Y. Zhang, A. Demetriadou, C. Carnegie, H. Ohadi, B. de Nijs, R. Esteban, J. Aizpurua, and J. J. Baumberg, Single-molecule optomechanics in picocavities, Science 354, 726 (2016).
  • (15) K.-D. Park, E. A. Muller, V. Kravtsov, P. M. Sass, J. Dreyer, J. M. Atkin, and M. B. Raschke, Variable-temperature tip-enhanced Raman spectroscopy of single-molecule fluctuations and dynamics, Nano Lett. 16, 479 (2016).
  • (16) R. Chikkaraddy, B. de Nijs, F. Benz, S. J. Barrow, O. A. Scherman, E. Rosta, A. Demetriadou, P. Fox, O. Hess, and J. J. Baumberg, Single-molecule strong coupling at room temperature in plasmonic nanocavities, Nature 535, 127 (2016).
  • (17) M. Tokman, M. Erukhimova, Y. Wang, Q. Chen, and A. Belyanin, Generation and dynamics of entangled fermion-photon-phonon states in nanocavities, Nanophotonics 10, 491 (2021).
  • (18) M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity optomechanics, Rev. Mod. Phys. 86, 1391 (2014).
  • (19) P. Meystre, A short walk through quantum optomechanics, Ann. Phys. 525, 215 (2013).
  • (20) P. Roelli, C. Galland, N. Piro and T. J. Kippenberg. Molecular cavity optomechanics as a theory of plasmon-enhanced Raman scattering, Nature Nano. 11, 164 (2016).
  • (21) J.-M. Pirkkalainen, S.U. Cho, F. Massel, J. Tuorila, T.T. Heikkila, P.J. Hakonen, and M.A. Sillanpaa, Cavity optomechanics mediated by a quantum two-level system, Nat. Comm. 6, 6981 (2015).
  • (22) Y. Chu, P. Kharel, W. H. Renninger, L. D. Burkhart, L. Frunzio, P. T. Rakich, R. J. Schoelkopf, Quantum acoustics with superconducting qubits, Science 358, 199 (2017).
  • (23) S. Hong, R. Riedinger, I. Marinkovic, A. Wallucks, S. G. Hofer, R. A. Norte, M. Aspelmeyer, and S. Groblacher, Hanbury Brown and Twiss interferometry of single phonons from an optomechanical resonator, Science 358, 203 (2017).
  • (24) P. Arrangoiz-Arriola, E. A. Wollack, Z. Wang, M. Pechal, W. Jiang, T. P. McKenna, J. D. Witmer, R. Van Laer, and A.H. Safavi-Naeini, Resolving the energy levels of a nanomechanical oscillator, Nature 571, 537 (2019).
  • (25) Q. Liao, Y. Ye, P. Lin, N. Zhou, and W. Nie, Tripartite entanglement in an atom-cavity-optomechanical system, Int. J. Theor. Phys. 57, 1319 (2018).
  • (26) K. J. Satzinger, Y. P. Zhong, H. Chang, et al., Quantum control of surface acoustic-wave phonons, Nature 563, 661 (2018).
  • (27) A. Bienfait, Y. P. Zhong, H.-S. Chang, M.-H. Chou, C. R. Conner, E. Dumur, J. Grebel, G. A. Peairs, R. G. Povey, K. J. Satzinger, and A. N. Cleland, Quantum Erasure Using Entangled Surface Acoustic Phonons, Phys. Rev. X 10, 021055 (2020).
  • (28) U. Delic, M. Reisenbauer, K. Dare, D. Grass, V. Vuletic, N. Kiesel, and M. Aspelmeyer, Cooling of a levitated nanoparticle to the motional quantum ground state, Science 367, 892 (2020).
  • (29) I. Wilson-Rae and A. Imamoglu, Quantum dot cavity-QED in the presence of strong electron-phonon interactions, Phys. Rev. B 65, 235311 (2002).
  • (30) S. Weiler, A. Ulhaq, S. M. Ulrich, D. Richter, M. Jetter, P. Michler, C. Roy, and S. Hughes, Phonon-assisted incoherent excitation of a quantum dot and its emission properties, Phys. Rev. B 86, 241304(R) (2012).
  • (31) K, Roy-Choudhury and S. Hughes, Quantum theory of the emission spectrum from quantum dots coupled to structured photonic reservoirs and acoustic phonons, Phys. Rev. B 92, 205406 (2015).
  • (32) E. V. Denning, M. Bundgaard-Nielsen, and J. Mork, Optical signatures of electron-phonon decoupling due to strong light-matter interactions, Phys. Rev. B 102, 235303 (2020).
  • (33) P. Zoller and C. W. Gardiner, Quantum Noise in Quantum Optics: the Stochastic Schrödinger Equation. Lecture Notes for the Les Houches Summer School LXIII on Quantum Fluctuations in July 1995, Elsevier Science Publishers B.V. 1997, edited by E. Giacobino and S. Reynaud; arXiv:quant-ph/9702030v1.
  • (34) M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University Press, 1997).
  • (35) M. B. Plenio and P. L. Knight, The quantum-jump approach to dissipative dynamics in quantum optics, Rev. Mod. Phys. 70, 101 (1998)
  • (36) N. Gisin and I. C. Percival, The quantum-state diffusion model applied to open systems, Journal of Physics A 25 , 5677 (1992).
  • (37) L. Diosi, N. Gisin, and W. T. Strunz, Non-Markovian quantum state diffusion, Phys. Rev. A 58, 1699-1712 (1998).
  • (38) C. Cohen-Tannoudji, B. Zambon and E. Arimondo, Quantum-jump approach to dissipative processes: application to amplification without inversion, J. Opt. Soc. Am. B 10, 2107-2120 (1993).
  • (39) K. Molmer, Y. Castin and J. Dalibard, Monte Carlo wave-function method in quantum optics, J. Opt. Soc. Am. B 10, 524-538 (1993).
  • (40) N. Gisin and I. C. Percival, Wave-function approach to dissipative processes: are there quantum jumps? Phys. Lett. A 167, 315-318 (1992).
  • (41) H. De Raedt, F. Jin, M. I. Katsnelson, and K. Michielsen, Relaxation, thermalization, and Markovian dynamics of two spins coupled to a spin bath, Phys. Rev. E 96, 053306 (2017).
  • (42) F. Nathan and M. S. Rudner, Universal Lindblad equation for open quantum systems, Phys. Rev. B 102, 115109 (2020).
  • (43) Q. Chen, Y. Wang, S. Almutairi, M. Erukhimova, M. Tokman, and A. Belyanin. Dynamics and control of entangled electron-photon states in nanophotonic systems with time-variable parameters, Phys. Rev. A 103, 013708 (2021).
  • (44) M. Tokman, Q. Chen, M. Erukhimova, Y. Wang, and A. Belyanin, Quantum dynamics of open many-qubit systems strongly coupled to a quantized electromagnetic field in dissipative cavities, https://arxiv.org/abs/2105.14674.
  • (45) M. Tokman, Y. Wang, I. Oladyshkin, A. R. Kutayiah, and A. Belyanin, Laser-driven parametric instability and generation of entangled photon-plasmon states in graphene, Phys. Rev. B 93, 235422 (2016).
  • (46) M. Tokman, X. Yao, and A. Belyanin, Generation of entangled photons in graphene in a strong magnetic field, Phys. Rev. Lett. 110, 077404 (2013).
  • (47) M. D. Tokman, M. A. Erukhimova, and V. V. Vdovin, The features of a quantum description of radiation in an optically dense medium, Ann. Phys. 360, 571 (2015).
  • (48) M. Tokman, Z. Long, S. AlMutairi, Y. Wang, M. Belkin, and A. Belyanin, Enhancement of the spontaneous emission in subwavelength quasi-two-dimensional waveguides and resonators, Phys. Rev. A. 97, 043801 (2018).
  • (49) F. Herrera, and F. C. Spano. Absorption and photoluminescence in organic cavity QED, Phys. Rev. A 95, 053867 (2017).
  • (50) E. T. Jaynes and F. W. Cummings, Comparison of quantum and semiclassical radiation theories with application to the beam maser, Proc. IEEE 51, 89 (1963).
  • (51) C. Couteau. Spontaneous parametric down-conversion, Cont. Phys. 59, 291-304 (2018).
  • (52) M. Tokman, Y. Wang, Q. Chen, L. Shterengas, and A. Belyanin, Generation of entangled photons via parametric down-conversion of lasing modes in semiconductor lasers and waveguides, https://arxiv.org/abs/2108.03528.
  • (53) M. D. Kostin, On the Schrödinger-Langevin Equation, J. Chem. Phys. 57, 3589 (1972).
  • (54) R. Katz, P.B. Gossiaux. The Schrödinger–Langevin equation with and without thermal fluctuations, Annals of Physics 368, 267-295 (2016).
  • (55) D.F. Walls and G. J. Milburn, Quantum Optics (Springer, Berlin, 1994).
  • (56) L.D. Landau, E.M. Lifshitz, Statistical Physics, Part 1 (Pergamon, Oxford, 1965)
  • (57) K. H. Madsen and P. Lodahl, Quantitative analysis of quantum dot dynamics and emission spectra in cavity quantum electrodynamics, New J. Phys. 15, 025013 (2013).