Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Tunable microwave-optical entanglement and conversion in multimode electro-opto-mechanics

Open Access Open Access

Abstract

We study tunable double-channel microwave-optical (M-O) entanglement and coherent conversion by controlling the quantum interference effect. This is realized in a two-mechanical-mode electro-opto-mechanical (EOM) system, in which two mechanical resonators (MRs) are coupled with each other by phase-dependent phonon-phonon interaction, and link the interaction between the microwave and optical cavity. It’s demonstrated that the mechanical coupling between two MRs leads to the interference of two pathways of electro-opto-mechanical interaction, which can generate the tunable double-channel phenomena in comparison with a typical three-mode EOM system. In particular, by tuning of phonon-phonon interaction and couplings between cavities with MRs, we can not only steer the switch from the M-O interaction with a single channel to that of the double-channel, but also modulate the entanglement and conversion characteristics in each channel. Moreover, our scheme can be extended to an N-mechanical-mode EOM system, in which N discrete channels will be observed and controlled. This study opens up prospects for quantum information transduction and storage with a wide bandwidth and multichannel quantum interface.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Cavity optomechanical systems [15] have attracted considerable attention in the past few years due to their potential applications in quantum information technology. Earlier researches have achieved dramatic progress in electromechanically or optomechanically induced transparency (EMIT/OMIT) [68], quantum mechanics in macroscopic objects [912], cooling [1318], optical filters [19], quantum sensors [2,20,21], etc. Furthermore, the nanomechanical resonators can be coupled to the microwave and optical cavities simultaneously, they can therefore entangle [22,23] or coherently convert [24,25] the photons at microwave and optical regions. Recent experiments [2641] are very encouraging to directly convert microwave and optical photons in either the classical or the quantum domain. Especially, it has been demonstrated that the cavity electro-opto-mechanical (EOM) system reaches the highest conversion efficiency of 0.47 [36]. However, it’s still challenging to meet the requirements of high conversion efficiency and low added noise. In contrast, the entanglement-based conversion [42,43], which first entangles microwave and optical photons [22,4446], and then completes the conversion with quantum teleportation, has been theoretically demonstrated to be a more efficient and practical method due to the introduction of classical communication channels. However, the performances on either entanglement-based conversion or direct conversion are both restricted by the moderate electromechanical and optomechanical couplings.

Multimode electro-opto-mechanical systems provide a possible approach to circumvent this difficulty. Experimentally, multimode optomechanical and electromechanical systems have been employed for circulators [4749], tunable optomechanically induced transparency [50,51], enhanced coherent entanglement or cooling [5254], etc. By introducing another mechanical resonator, the cooperativity between the photons and phonons can be enhanced or suppressed by dissipation engineering [53,55,56] or quantum interference [50,51,57]. However, these experiments mainly concerns the indirect coupling of mechanical modes, where an optical or microwave mode serves as an intermediary. It’s still blank in the multimode mechanical couplings on the EOM systems, in which mechanical modes links the interaction between microwave and optical modes.

In this paper, we theoretically study a two-mechanical-mode EOM system, where two mechanical resonators (MRs) are phase-dependent coupled with each other. The quantum interference of two pathways of electro-opto-mechanical interaction is studied in M-O entanglement and coherent conversion. The double-channel phenomena are observed and the modulation on entanglement or conversion in each channel can be realized by tuning the couplings. These results provide an approach to optimize the parameter settings to obtain the double-channel or single-channel optimal entanglement or conversion. Additionally, our scheme could be generalized to an N-mechanical-mode EOM system or other nonlinear interaction in multimode systems such as magnon-optics and atom ensembles.

2. System

We consider a two-mechanical-mode EOM system shown in Fig. 1(a), where two mechanical resonators (MRs) with frequencies ${\omega _{m1}}$ and ${\omega _{m2}}$ are coupled with each other, and MR1 serves as an intermediary to realize indirect coupling between microwave and optical modes ${\hat{a}_{e,o}}$. By modulating the mechanical coupling ${\lambda _r}$, we can further control the indirect interaction between ${\hat{a}_e}$ and ${\hat{a}_o}$. The two-mechanical-mode configuration can be realized in many systems, such as charged MRs [58], superconducting quantum circuit [54,59] and others [60,61]. Figure 1(b) shows schematic of a potentially experimental system, in which two MRs are Coulomb coupled. MR1 plays the role of a movable plate of a capacitor in LC circuit and movable mirror of optical cavity simultaneously, and thus modulates the resonance frequency of the microwave LC circuit as well as the optical cavity. The system Hamiltonian reads as

$$\begin{aligned} \hat{H} &= {{\hat{H}}_0} = \hbar {\omega _{m1}}\hat{b}_1^{\dagger} {{\hat{b}}_1} + \hbar {\omega _{m2}}\hat{b}_{_2}^{\dagger} {{\hat{b}}_2} + \hbar {\omega _e}\hat{a}_e^{\dagger} {{\hat{a}}_e} + \hbar {\omega _o}\hat{a}_o^{\dagger} {{\hat{a}}_o}\\ &+ \sum\limits_{j = e,o} {\hbar {g_{j1}}\hat{a}_j^{\dagger} {{\hat{a}}_j}({{{\hat{b}}_1} + \hat{b}_1^{\dagger} } )} + \hbar {\lambda _r}({{{\hat{b}}_1}\hat{b}_2^{\dagger} + \hat{b}_1^{\dagger} {{\hat{b}}_2}} )\\ &+ \sum\limits_{j = e,o} {i{\varepsilon _j}({\hat{a}_j^{\dagger} {e^{ - i{\omega_{d,j}}t}} - {{\hat{a}}_j}{e^{ - i{\omega_{d,j}}t}}} )} \end{aligned}, $$
where ${g_{e1}}$ (${g_{o1}}$) represents the coupling rate of the microwave (optical) cavity to MR1. ${\lambda _r} = {{\hbar \lambda } / {2\sqrt {{m_1}{m_2}{\omega _{m1}}{\omega _{m2}}} }}$ with $\lambda $ denotes the Coulomb coupling strength and ${m_l}(l = 1,2)$ is the mass of MRs. The microwave and optical pumps with frequency ${\omega _{d,j}}$ and amplitude ${\varepsilon _j}$ are applied to the system.

 figure: Fig. 1.

Fig. 1. (a) Schematic of a two-mechanical-mode EOM system. Potentially experimental systems: (b) The two MRs are coupled with each other via Coulomb interaction, which can be controlled by the bias voltages V1 and -V2 on the MR1 and MR2, respectively. (c) Schematic of a fully coupled two-mechanical-mode EOM system. $\zeta {e^{ {\pm} i\theta }}:\zeta ({e^{ - i\theta }}{\hat{b}_1}\hat{b}_2^{\dagger} + {e^{i\theta }}\hat{b}_1^{\dagger} {\hat{b}_2})$. (d) Two MRs are capacitive coupled to a superconducting charge qubit, and the phase-dependent phonon-hopping interaction is induced and controlled by gate voltages. $C_{1,2}^x$ represents the gate capacitance, which is varied with the displacement ${x_{1,2}}$ of each MR. JJ: Josephson junction.

Download Full Size | PDF

First, we consider two identical resonance frequency ${\omega _1} = {\omega _2} = {\omega _m}$ of MRs for simplicity. Defining two mechanical dressed modes ${\hat{b}_ + } = {{({{\hat{b}}_1} + {{\hat{b}}_2})} / {\sqrt 2 }}$ and ${\hat{b}_ - } = {{({{\hat{b}}_1} - {{\hat{b}}_2})} / {\sqrt 2 }}$, the Hamiltonian can be derived to

$$\begin{aligned}\hat{H} &= \hbar {\Delta _e}\hat{a}_e^{\dagger} {{\hat{a}}_e} + \hbar {\Delta _o}\hat{a}_o^{\dagger} {{\hat{a}}_o} + \hbar ({{\omega_m} - {\lambda_r}} )\hat{b}_ - ^{\dagger} {{\hat{b}}_ - } + \hbar ({{\omega_m} + {\lambda_r}} )\hat{b}_ + ^{\dagger} {{\hat{b}}_ + }\\ &+ \frac{1}{{\sqrt 2 }}\sum\limits_{j = e,o} {\hbar {g_{j1}}\hat{a}_j^{\dagger} {{\hat{a}}_j}({{{\hat{b}}_ + } + \hat{b}_ +^{\dagger} } )} + \frac{1}{{\sqrt 2 }}\sum\limits_{j = e,o} {\hbar {g_{j1}}\hat{a}_j^{\dagger} {{\hat{a}}_j}({{{\hat{b}}_ - } + \hat{b}_ -^{\dagger} } )} \\ &+ \sum\limits_{j = e,o} {i{\varepsilon _j}({\hat{a}_j^{\dagger} - {{\hat{a}}_j}} )} \end{aligned}. $$
It’s shown that ${\hat{b}_ + }$ and ${\hat{b}_ - }$ are decoupled with each other and the same resonance frequency of MRs is split to ${\omega _m} + {\lambda _r}$ and ${\omega _m} - {\lambda _r}$. Consequently, the exchange between excitations of microwave and optical modes is transmitted in two channels independently. The physics behind the double-channel transmission phenomenon can be understood from the level configuration in Fig. 2. Two pathways of electro-opto-mechanical interaction are interfered and hybridized into two independent pathways via Coulomb coupling, which can be controlled by changing the bias voltages of MRs. Additionally, a stronger coupling leads to a larger distance of two exchange channels. Therefore, the interaction between microwave and optical modes can be controlled by tuning the strength of Coulomb coupling. Note that the results under the case of ${\omega _1} \ne {\omega _2}$ are similar, where mechanical modes are hybridized into two decoupled dressed modes and lead to a double-channel in the same way.

 figure: Fig. 2.

Fig. 2. Schematic of the energy-level diagram in the four-mode EOM system. The $|{{n_e}} \rangle $, $|{{n_o}} \rangle $, $|{{m_1}} \rangle $ and $|{{m_2}} \rangle $ denote the number states of the microwave and optical cavity photon, and MR1 and MR2 phonons, respectively. $|{{n_o},{n_e},{m_1},{m_2}} \rangle $ \leftrightarrow |{{n_o},{n_e} + 1,{m_1},{m_2}} \rangle $and$|{{n_o},{n_e},{m_1},{m_2}} \rangle $ \leftrightarrow |{{n_o} + 1,{n_e},{m_1},{m_2}} \rangle $ transitions change the cavity fields.$|{{n_o},{n_e} + 1,{m_1},{m_2}} \rangle \leftrightarrow |{{n_o},{n_e},{m_1} + 1,{m_2}} \rangle $ and $|{{n_o} + 1,{n_e},{m_1},{m_2}} \rangle \leftrightarrow |{{n_o},{n_e},{m_1} + 1,{m_2}} \rangle $ transitions are caused by capacitive coupling and radiation-pressure coupling, respectively. $|{{n_o},{n_e},{m_1} + 1,{m_2}} \rangle \leftrightarrow |{{n_o},{n_e},{m_1},{m_2} + 1} \rangle $ transition is induced by the Coulomb coupling, which induces two mechanical dressed modes $|{{n_o},{n_e},{b_ + }} \rangle $ and $|{{n_o},{n_e},{b_ - }} \rangle $.

Download Full Size | PDF

In the following, we consider a more complete condition, where two MRs are both coupled to microwave and optical cavities simultaneously, and the MRs are coupled with each other via a phase-dependent phonon-hopping interaction shown in Fig. 1(c)–1(d). So, Hamiltonian (1) becomes

$$\begin{aligned} \hat{H} &= {{\hat{H}}_0} = \hbar {\omega _{m1}}\hat{b}_1^{\dagger} {{\hat{b}}_1} + \hbar {\omega _{m2}}\hat{b}_{_2}^{\dagger} {{\hat{b}}_2} + \hbar {\omega _e}\hat{a}_e^{\dagger} {{\hat{a}}_e} + \hbar {\omega _o}\hat{a}_o^{\dagger} {{\hat{a}}_o}\\ &+ \sum\limits_{j = e,o\atop l = 1,2} {\hbar {g_{jl}}\hat{a}_j^{\dagger} {{\hat{a}}_j}({{{\hat{b}}_l} + \hat{b}_l^{\dagger} } )} + \hbar \zeta ({{e^{ - i\theta }}{{\hat{b}}_1}\hat{b}_2^{\dagger} + {e^{i\theta }}\hat{b}_1^{\dagger} {{\hat{b}}_2}} )\\ &+ \sum\limits_{j = e,o} {i{\varepsilon _j}({\hat{a}_j^{\dagger} {e^{ - i{\omega_{d,j}}t}} - {{\hat{a}}_j}{e^{ - i{\omega_{d,j}}t}}} )} \end{aligned}. $$
Here, ${g_{jl}} = 0$ denotes a simplified case where the cavity mode ${\hat{a}_j}$ does not interact with the mechanical mode ${\hat{b}_l}$. By redefining the mechanical dressed modes
$$\begin{aligned} &{{\hat{b}}_ + } = f{{\hat{b}}_1} - {e^{i\theta }}h{{\hat{b}}_2}\\ &{{\hat{b}}_ - } = {e^{ - i\theta }}h{{\hat{b}}_1} + f{{\hat{b}}_2} \end{aligned}, $$
the Hamiltonian (3) becomes
$$\begin{aligned} \hat{H} &= \hbar {\Delta _e}\hat{a}_e^{\dagger} {{\hat{a}}_e} + \hbar {\Delta _o}\hat{a}_o^{\dagger} {{\hat{a}}_o} + \hbar {\omega _ + }\hat{b}_ + ^{\dagger} {{\hat{b}}_ + } + \hbar {\omega _ - }\hat{b}_ - ^{\dagger} {{\hat{b}}_ - }\\ &+ \sum\limits_{j = e,o} {\hbar {g_{j + }}\hat{a}_j^{\dagger} {{\hat{a}}_j}({{{\hat{b}}_ + } + \hat{b}_ +^{\dagger} } )} + \sum\limits_{j = e,o} {\hbar {g_{j - }}\hat{a}_j^{\dagger} {{\hat{a}}_j}({{{\hat{b}}_ - } + \hat{b}_ -^{\dagger} } )} \\ &+ \sum\limits_{j = e,o} {i{\varepsilon _j}({\hat{a}_j^{\dagger} - {{\hat{a}}_j}} )} \end{aligned}$$
with
$$\begin{array}{c} {\omega _ \pm } = \frac{1}{2}\left( {{\omega_{m1}} + {\omega_{m2}} \pm \sqrt {{{({{\omega_{m1}} - {\omega_{m2}}} )}^2} + 4{\zeta^2}} } \right)\\ f = {{|{{\omega_ - } - {\omega_{m1}}} |} / {\sqrt {{{({{\omega_ - } - {\omega_{m1}}} )}^2} + \zeta _{}^2} }},h = {{\zeta f} / {({{\omega_ - } - {\omega_{m1}}} )}}\\ {g_{e + }} = f{g_{e1}} - {e^{ - i\theta }}h{g_{e2}},{g_{e - }} = {e^{i\theta }}h{g_{e1}} + f{g_{e2}}\\ {g_{o + }} = f{g_{o1}} - {e^{i\theta }}h{g_{o2}},{g_{o - }} = {e^{ - i\theta }}h{g_{o1}} + f{g_{o2}} \end{array}. $$
Same with the interference depicted in Fig. 2, the hybridized modes ${\hat{b}_ + }$ and ${\hat{b}_ - }$ are decoupled and thus lead to the split channels with frequency ${\omega _ + }$ and ${\omega _ - }$. The coupling strengths of cavity and mechanical modes satisfy the following condition
$$g_{j + }^2 + g_{j - }^2 = g_{j1}^2 + g_{j2}^2$$
From Eqs. (5)-(7), it’s demonstrated that controlling of mechanical coupling between MRs could modulate the distance of two channels as well as the interaction between microwave and optical modes in each channel. In particular, it can also control the width and amplitude of each channel, and even the switching between the single working channel and a tunable double-channel.

3. Entanglement and conversion

3.1 Entanglement

In this section, we discuss the effect of the two-mechanical-mode on the entanglement of itinerant microwave and optical photons. We first consider a typical three-mode EOM system, which means ${\lambda _r} = 0,{g_{j2}} = 0$. As the driving of each cavity satisfies ${\Delta _e} ={-} {\Delta _o} = {\omega _{m1}}$ (${\Delta _j} = {\omega _j} - {\omega _{d,j}}$), the Hamiltonian in the interaction picture $\hbar {\omega _e}\hat{a}_e^{\dagger} {\hat{a}_e} + \hbar {\omega _o}\hat{a}_o^{\dagger} {\hat{a}_o}$ reduces to Eq. (8) after linearization and rotation approximation.

$${\hat{H}_I} = \hbar {G_{e1}}(\delta {\hat{a}_e}^{\dagger} {\hat{b}_1} + \hat{b}_1^{^{\dagger} }\delta {\hat{a}_e}) + \hbar {G_{o1}}(\delta \hat{a}_o^{\dagger} \hat{b}_1^{^{\dagger} } + {\hat{b}_1}\delta {\hat{a}_o}). $$
Here, $\delta {\hat{a}_j} = {\hat{a}_j} - \sqrt {{n_j}} $, and ${n_j}$ is the mean number of steady cavity mode. ${G_{j1}} = {g_{j1}}\sqrt {{n_j}} $ represents the multiphoton-phonon coupling rate. The first term represents a beam-splitter interaction to coherently exchange the excitation between the mechanical and microwave cavity mode, while the second term corresponds to a parametric down-conversion interaction entangling the mechanical mode and the optical cavity mode. The entanglement between the mechanical and optical cavity mode will be transmitted to the propagating optical and microwave mode if the optomechanical and electromechanical rates ${\Gamma _{o1}} = {{4G_{o1}^2} / {{\kappa _o}}}$, ${\Gamma _{e1}} = {{4G_{e1}^2} / {{\kappa _e}}}$ both exceed the decoherence rate of MR1 $r = {\gamma _{m1}}{\bar{n}_{m1}}$, where ${\kappa _e}$ (${\kappa _o}$) and ${\gamma _{m1}}$ represent the damping rate of the microwave (optical) cavity and MR1, respectively; ${\bar{n}_{m1}}$ is the Planck-law average photon number of MR1.

Utilizing quantum Langevin equations, the propagating modes can be written as [62]

$$\hat{E} = {A_e}\delta {\hat{a}_{e,{\mathop{\rm int}} }} - B\delta \hat{a}_{o,{\mathop{\rm int}} }^{\dagger} - {D_e}\delta {\hat{b}_1}, $$
$$\hat{O} = B\delta \hat{a}_{e,{\mathop{\rm int}} }^{\dagger} + {A_o}\delta {\hat{a}_{o,{\mathop{\rm int}} }} - {D_o}\delta \hat{b}_1^{\dagger} , $$
where $\hat{E}$ and $\hat{O}$ are the propagating microwave and optical fields respectively; $\delta {\hat{a}_{j,{\mathop{\rm int}} }}$ represents the intracavity quantum fluctuation, while $\delta {\hat{b}_1}$ is quantum Brownian noise of MR1; ${A_j}$, B, ${D_j}$ are parameters related with the quantum cooperativity parameters ${C_{j1}} = G_j^2/{\kappa _j}{\gamma _{m1}}$ ($j = e,o$). Detail expressions of ${A_j}$, B, ${D_j}$ can be found in Appendix. A.

One of the important issues is how to discuss entanglement quantitatively. Here, entanglement degree based on average number of propagating modes as well as their correlation is utilized to deal with this issue. Denoting the temperature of EOM as ${T_E}$, we can obtain the average photon or phonon number of quantum fluctuations in Eq. (9) and Eq. (10) at respective frequencies by Planck laws, which reads

$$\left\{ \begin{aligned} {{\bar{n}}_{e,{T_E}}} &= \frac{1}{{{e^{\hbar {\omega_e}/{k_B}{T_E}}} - 1}}\\ &{{\bar{n}}_{o,{T_E}}} = \frac{1}{{{e^{\hbar {\omega_o}/{k_B}{T_E}}} - 1}}\\ &{{\bar{n}}_{m1,{T_E}}} = \frac{1}{{{e^{\hbar {\omega_{m1}}/{k_B}{T_E}}} - 1}} \end{aligned} \right. . $$
Average photon number of propagating modes and their correlation read
$${\bar{N}_e} = \left\langle {{{\hat{E}}^{\dagger} }\hat{E}} \right\rangle = {|{{A_e}} |^2}{\bar{n}_{e,{T_E}}} + {|B |^2}({{{\bar{n}}_{o,{T_E}}} + 1} )+ {|{{D_e}} |^2}{\bar{n}_{m1,{T_E}}}, $$
$${\bar{N}_o} = \left\langle {{{\hat{O}}^{\dagger} }\hat{O}} \right\rangle = {|B |^2}({{{\bar{n}}_{e,{T_E}}} + 1} )+ {|{{A_o}} |^2}{\bar{n}_{o,{T_E}}} + {|{{D_o}} |^2}({{{\bar{n}}_{m1,{T_E}}} + 1} ), $$
$$\left\langle {\hat{E}\hat{O}} \right\rangle = {A_e}B({\bar{n}_{e,{T_E}}} + 1) - B{A_o}{\bar{n}_{o,{T_E}}} + {D_e}{D_o}({\bar{n}_{m1,{T_E}}} + 1). $$
The detail deductions of Eqs. (12)–(14) can be found in Appendix B. Hence, we can discuss entanglement metric by induced correlation as
$$\varepsilon = \frac{{\left|{\left\langle {\hat{E}\hat{O}} \right\rangle } \right|}}{{\sqrt {{{\bar{N}}_e}{{\bar{N}}_o}} }}. $$
As is stated above, only if $G_{e1}^2/{\kappa _e} > {\gamma _{m1}}{\bar{n}_{m1}}$ and $G_{o1}^2/{\kappa _o} > {\gamma _{m1}}{\bar{n}_{m1}}$ are satisfied, can propagating modes entangled. When referred to entanglement metric $\varepsilon $, the entanglement criterion can be stated as: if $\varepsilon $ is greater than 1, there exists entanglement between propagating modes; if not, there is no entanglement due to decoherence and other factors. In addition, the entanglement can be calculated by the logarithmic negativity [63]
$${E_N} = \max [{0, - \ln ({2\mu } )} ]$$
with
$$\mu = \sqrt {\frac{1}{2}} {\left[ {v_{11}^2 + v_{33}^2 + 2v_{13}^2 - \sqrt {{{({v_{11}^2 - v_{33}^2} )}^2} + 4v_{13}^2{{({{v_{11}} + {v_{33}}} )}^2}} } \right]^{{1 / 2}}}, $$
where ${v_{11}}$, ${v_{33}}$, ${v_{13}}$ are covariances for the quadratures of the microwave and optical cavities’ outputs (see Appendix C). Next, we consider the general case expressed by Eq. (3), where there are two channels to link the interaction between microwave and optical photons. The multiphoton-phonon coupling rates are derived as
$$\begin{aligned} {G_{e + }} &= f{G_{e1}} - {e^{ - i\theta }}h{G_{e2}},{G_{e - }} = {e^{i\theta }}h{G_{e1}} + f{G_{e2}}\\ &{G_{o + }} = f{G_{o1}} - {e^{i\theta }}h{G_{o2}},{G_{o - }} = {e^{ - i\theta }}h{G_{o1}} + f{G_{o2}} \end{aligned}, $$
with
$$G_{j + }^2 + G_{j - }^2 = G_{j1}^2 + G_{j2}^2. $$
Utilizing the updated parameters for ${\hat{b}_ + }$ and ${\hat{b}_ - }$ shown in Eq. (4), the interaction Hamiltonian can be written as
$$\begin{aligned} {{\hat{H}}_I} &= \hbar {G_{e + }}(\delta {{\hat{a}}_e}^{\dagger} \delta {{\hat{b}}_ + } + \delta \hat{b}_ + ^{^{\dagger} }\delta {{\hat{a}}_e}) + \hbar {G_{o + }}(\delta \hat{a}_o^{\dagger} \delta \hat{b}_ + ^{^{\dagger} } + \delta {{\hat{b}}_ + }\delta {{\hat{a}}_o})\\ &+ \hbar {G_{e - }}(\delta {{\hat{a}}_e}^{\dagger} \delta {{\hat{b}}_ - } + \delta \hat{b}_ - ^{^{\dagger} }\delta {{\hat{a}}_e}) + \hbar {G_{o - }}(\delta \hat{a}_o^{\dagger} \delta \hat{b}_ - ^{^{\dagger} } + \delta {{\hat{b}}_ - }\delta {{\hat{a}}_o}) \end{aligned}, $$
which denotes the entanglement generation in a double-channel. The first two terms represent the generation of M-O entanglement via the channel of ${\hat{b}_ + }$, while the last two terms correspond to that via the channel of ${\hat{b}_ - }$. The propagating modes are derived as
$${\hat{E}_ + } = {A_{e + }}\delta {\hat{a}_{e,{\mathop{\rm int}} + }} - {B_ + }\delta \hat{a}_{o,{\mathop{\rm int}} + }^{\dagger} - {D_{e + }}\delta {\hat{b}_ + }, $$
$${\hat{O}_ + } = {B_ + }\delta \hat{a}_{e,{\mathop{\rm int}} + }^{\dagger} + {A_{o + }}\delta {\hat{a}_{o,{\mathop{\rm int}} + }} - {D_{o + }}\delta \hat{b}_ + ^{\dagger} , $$
$${\hat{E}_ - } = {A_{e - }}\delta {\hat{a}_{e,{\mathop{\rm int}} - }} - {B_ - }\delta \hat{a}_{o,{\mathop{\rm int}} - }^{\dagger} - {D_{e - }}\delta {\hat{b}_ - }, $$
$${\hat{O}_ - } = {B_ - }\delta \hat{a}_{e,{\mathop{\rm int}} - }^{\dagger} + {A_{o - }}\delta {\hat{a}_{o,{\mathop{\rm int}} - }} - {D_{o - }}\delta \hat{b}_ - ^{\dagger} , $$
where the coefficients can be expressed by updated ${G_{e + }},{G_{e - }},{G_{o + }},{G_{o - }}$. ${\hat{E}_ + }$ (${\hat{E}_ - }$) and ${\hat{O}_ + }$ (${\hat{O}_ - }$) are quadrature entangled when the entanglement metric ${\varepsilon _ + }$ (${\varepsilon _ - }$) is larger than 1. Also, we can derive the logarithmic negativity ${E_{N \pm }}$ and bandwidth for each pathway, respectively.

In the following, we check the performance on entanglement and transmission of two-mechanical-mode EOM system. The feasible parameter settings are listed in Table 1. In Fig. 3(a), we plot the normalized output power ${P_o}$ of the optical cavity for ${\zeta / {2\pi }} = 0,0.16\textrm{MHz},0.28\textrm{MHz}$. $\delta = \omega - {\omega _e}$ represents the detuning of itinerant microwave photons and the resonance frequency of the microwave cavity. Other parameter settings are as follows: ${G_{e2}} = 0$, ${G_{o2}} = 0$ and $\theta = 0$. It’s obvious that there exists only one window to generate quadrature entangled optical and microwave photons if MR2 is independent, while two entanglement windows exist when the two MRs are coupled with each other. Besides, the distance of two windows is determined by the coupling strength $\zeta $. The Fig. 3(b) shows double-channel phenomena under the case of ${G_{e2}} > 0$, ${G_{o2}} > 0$. By controlling the electromechanical and optomechanical couplings, the amplitude and width of two windows are unequal. Physically, the width of each window is related to the effective mechanical damping rate given by ${\gamma _{m1}} + {\Gamma _{e \pm }} - {\Gamma _{o \pm }}$, while the amplitude is more related to the effective multiphoton-phonon coupling strength. Therefore, we can select the dominant channel by modulating the phase of mechanical coupling and electromechanical (optomechanical) coupling strength between the cavity with MR2.

 figure: Fig. 3.

Fig. 3. M-O Entanglement of the two-mechanical-mode EOM system. (a) The normalized output optical fields. (b) The normalized output optical fields for ${{{G_{e2}}} / {2\pi }} = 99.5\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 32.7\textrm{kHz}$ (blue lines), ${{{G_{e2}}} / {2\pi }} = 63.1\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 57.8\textrm{kHz}$ (red lines). (c) The normalized logarithmic negativity ${{{E_{N \pm }}} / {{{\bar{N}}_o}}}$ as a function of $\theta $ for ${{{G_{e2}}} / {2\pi }} = 99.5\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 32.7\textrm{kHz}$ (blue), ${{{G_{e2}}} / {2\pi }} = 63.1\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 57.8\textrm{kHz}$ (red). (d) ${{{E_{N + }}} / {{{\bar{N}}_o}}}$ as a function of ${G_{e2}}$ and ${G_{o2}}$.

Download Full Size | PDF

Tables Icon

Table 1. Parameter settings of the EOM systema

Since the most powerful entangled resources are expected to be those maximizing their quadrature correlations per photon emitted, we analyze the two-mechanical-mode EOM system in terms of the normalized logarithmic negativity ${{{E_N}} / {{{\bar{N}}_o}}}$. We first calculate the ${{{E_N}} / {{{\bar{N}}_o}}}$ as a function of $\zeta$. It’s shown that the coupling strength between MRs shows little effect on ${E_{N + }}$ and ${E_{N - }}$, but just change the distance of entanglement windows, as shown in Fig. 3(a) and 3(b). Corresponding results are listed in Appendix. D. In the following, we analyze the effect of $\theta $, ${G_{e2}}$ and ${G_{o2}}$ on entanglement under the condition of ${\zeta / {2\pi }} = 0.16\textrm{MHz}$, respectively. In Fig. 3(c), we plot the ${{{E_{N \pm }}} / {{{\bar{N}}_o}}}$ as a function of $\theta $ for different electromechanical and optomechanical couplings with MR2${{{G_{e2}}} / {2\pi }} = 99.5\textrm{kHz}$, ${{{G_{o2}}} / {2\pi }} = 32.7\textrm{kHz}$ and ${{{G_{e2}}} / {2\pi }} = 63.1\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 57.8\textrm{kHz}$. The variations of entanglement on two channels are opposite and the maximum of ${{{E_{N + }}} / {{{\bar{N}}_o}}}$ corresponds to the minimum of ${{{E_{N - }}} / {{{\bar{N}}_o}}}$, and vice versa. To further probe the effect of ${G_{e2}}$ and ${G_{o2}}$, we plot ${{{E_{N + }}} / {{{\bar{N}}_o}}}$ versus ${G_{e2}}$ and ${G_{o2}}$, which is shown in Fig. 3(d). There are two parts in the image, the stable (color region) and unstable (blank region) region. The boundary of two parts is given by Routh-Hurwitz criterion (See Appendix. E). In the color region, appropriate setting of ${G_{e2}}$ and ${G_{o2}}$ will increase the normalized entanglement and the maximum ${{{E_{N + }}} / {{{\bar{N}}_o}}}$ is obtained at certain values of ${G_{e1}}$ and ${G_{o1}}$. The blank region of the image indicates an unstable condition, which can be reduced by adjusting the couplings with MR1. Additionally, the ${{{E_{N - }}} / {{{\bar{N}}_o}}}$ versus ${G_{e2}}$ and ${G_{o2}}$ are similar with Fig. 3(d) due to the symmetry of the above parameters.

From Fig. 3, we can conclude that the M-O entanglement generated by a two-mechanical-mode EOM system is exhibited in a double-channel, and the entanglement characteristics in each channel are controllable by tuning the coupling strength and phase of phonon-phonon interaction. Physically, two pathways of electro-opto-mechanical interaction are interfered, which produces two discrete channels. In operator representation, the two mechanical modes coupled are hybridized into two decoupled modes and thus lead to a double-channel. When the cavities do not interact with MR2 directly, the two windows are equal in output fields and widths. Considering a more complicated case that two cavities are coupled with two MRs simultaneously, the amplitude and width of each entanglement window can be tuned by controlling $\zeta $, $\theta $, ${G_{e2}}$ and ${G_{o2}}$, which provides a more practical and feasible method to realize double-channel M-O entanglement.

3.1 Conversion

Coherent conversion between itinerant microwave and optical photons corresponds to the case of ${\Delta _e} = {\Delta _o} = {\omega _{m1}}$, and the Hamiltonian in the interaction picture is as follows:

$$\begin{aligned} {{\hat{H}}_C} &= \hbar {G_{e + }}(\delta {{\hat{a}}_e}\hat{b}_ + ^{\dagger} + {{\hat{b}}_ + }\delta \hat{a}_e^{^{\dagger} }) + \hbar {G_{o + }}(\delta {{\hat{a}}_o}\hat{b}_ + ^{\dagger} + {{\hat{b}}_ + }\delta {{\hat{a}}_ + }^{\dagger} )\\ &+ \hbar {G_{e - }}(\delta {{\hat{a}}_e}\hat{b}_ - ^{\dagger} + {{\hat{b}}_ - }\delta \hat{a}_e^{^{\dagger} }) + \hbar {G_{o - }}(\delta {{\hat{a}}_o}\hat{b}_ - ^{\dagger} + {{\hat{b}}_ - }\delta {{\hat{a}}_o}^{\dagger} ) \end{aligned},$$
where the first two terms represent the M-O conversion via the channel of ${\hat{b}_ + }$, while the last two terms correspond to that via the channel of ${\hat{b}_ - }$. Under this condition, the Heisenberg-Langevin equations can be calculated and the transmission coefficient $t(\omega )$ is written as [31]
$${t_k}(\omega ) = {\left. {\frac{{a_o^{out}(\omega )}}{{a_e^{in}(\omega )}}} \right|_k} = \frac{{\sqrt {{\Gamma _{ek}}{\Gamma _{ok}}} }}{{ - i({\omega - {\omega_{mk}}} )+ {{({{\Gamma _{ek}} + {\Gamma _{ok}} + {\kappa_m}} )} / 2}}}\sqrt {{R_e}{R_o}{\eta _e}{\eta _o}},$$
with ${R_e} = 1 + {({{{{\kappa_e}} / {4{\omega_m}}}} )^2}$, ${R_o} = 1 + {({{{{\kappa_o}} / {4{\omega_m}}}} )^2}$, ${\eta _e} = {{{\kappa _{e,ex}}} / {{\kappa _e}}}$ and ${\eta _o} = {{{\kappa _{o,ex}}} / {{\kappa _o}}}$. $k ={+} , - $ represents the two channels. ${\kappa _{e,ex}}({\kappa _{o,ex}})$ is the rate of energy transferred from the microwave (optical) cavity into propagating field. Generally, the total conversion efficiency is denoted by ${|{{t_k}(\omega )} |^2}$. Considering the weak coupling condition ${\kappa _e} > > {G_{ek}}$ and ${\kappa _o} > > {G_{ok}}$, the microwave (optical) mode is coherently converted to the mechanical mode at a rate of ${\Gamma _{ek}}$ (${\Gamma _{ok}}$) in the resolved sideband limit ($4{\omega _{mk}} > > {\kappa _e},{\kappa _o}$). Besides, the extraneous vibrations of MRs will introduce noise to the converted signals in the up-conversion, which is given by
$${n_{add,k}} = \frac{1}{{{R_e}{\eta _e}}}\left( {\frac{{{\kappa_m}{n_T}}}{{{\Gamma _{ek}}}} + ({{R_e} - 1} )+ ({{R_o} - 1} )} \right)$$
in the units of quanta of added noise for matched coupling rates. ${n_T} = {1 / {({{e^{{{\hbar {\omega_m}} / {{k_B}{T_m}}}}} - 1} )}}$ is the average photon number of mechanical thermal noise. ${k_B}$ is the Boltzmann's constant and ${T_m}$ is the environmental temperature of the MRs.

In Fig. 4, we investigate the effect of two-mechanical-mode on conversion from microwave to optical photons. Figure 4(a) shows the conversion efficiency ${|{{t_ + }(\omega )} |^2}$ in the higher-frequency channel as a function of ${G_{e2}}$ and ${G_{o2}}$ when ${\zeta / {2\pi }} = 0.08\textrm{MHz,}\theta = \pi $, while Fig. 4(b) corresponds to that in the lower-frequency channel. It’s shown that ${|{{t_ \pm }(\omega )} |^2}$ is tunable by controlling the electromechanical and optomechanical couplings. The results of ${|{{t_ + }(\omega )} |^2} = 0$ mean that the higher-frequency channel disappears and there is only lower-frequency channel to convert microwave photons. When we consider the case of $\theta = 0$ with the same other parameters, the variation of ${|{{t_ - }(\omega )} |^2}$ can be shown by Fig. 4(a), while Fig. 4(b) corresponds to that of ${|{{t_ + }(\omega )} |^2}$. Besides, the distance of two channels is mainly determined by the coupling strength $\zeta $ of MRs, which is shown in Fig. 4(c). Specifically, the conversion efficiency ${|{{t_ \pm }(\omega )} |^2}$ as a function of $\theta $ in the frequency domain is shown in Fig. 4(d). In the region $0 < \theta < \pi $, the left conversion window always become much broader and more efficient while the right one becomes weaker with the increasing of $\theta $. The left window becomes completely absorbed near $\theta = \pi $, which refers to the case of ${|{{t_ + }(\omega )} |^2} = 0$ shown in Fig. 4(a). The variation in $\pi < \theta < 2\pi $ is absolutely opposite with the case of $0 < \theta < \pi $. Thus, the switchable and tunable M-O conversion can be realized by modulating the mechanical coupling strength and phase of MRs.

 figure: Fig. 4.

Fig. 4. M-O conversion efficiency of the two-mechanical-mode EOM system. (a) The conversion efficiency ${|{{t_ + }(\omega )} |^2}$ of the itinerant microwave photons in the higher-frequency channel as a function of ${G_{e2}}$ and ${G_{o2}}$ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$, $\theta = \pi $. (b) ${|{{t_ - }(\omega )} |^2}$ under the same parameters as (a). (c) ${|{{t_ \pm }(\omega )} |^2}$ versus $\zeta$ and $\delta $ when $\theta = {\pi / 2}$. (d) ${|{{t_ \pm }(\omega )} |^2}$ versus $\theta $ and $\delta $ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$. Other parameters are the same as Fig. 3.

Download Full Size | PDF

In general, any extraneous vibrations of MRs and reflected photons will introduce noise to the output fields. In the up-conversion, the former is mainly determined by mechanical resonance frequency, environmental temperature and cooperativities, while the latter, the optical noise, is much less and negligible mainly since the resonance frequency of the optical cavity is seven orders of magnitude higher than that of MRs. As Eq. (27) shows that the optomechanical couplings have less effect on the added noise caused by extraneous vibrations of MRs in up-conversion, we only probe the effect of electromechanical and mechanical couplings. The results that ${n_{add, \pm }}$ as a function of ${G_{e2}}$ and $\theta $ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$ are shown in Fig. 5. Similar to the variation of conversion efficiency, the added noise is periodically controllable. In the region $0 < \theta < \pi $ ($\pi < \theta < 2\pi $), ${n_{add, + }}$ is increased while ${n_{add, - }}$ is reduced with the increasing of $\theta $. Moreover, we get the minimum values of added noise in the higher (lower)-frequency channel when $\theta = \pi $ ($\theta = 0$), and it can be further decreased by increasing ${G_{e2}}$.

 figure: Fig. 5.

Fig. 5. (a), (b) Added noise ${n_{add, \pm }}$ of the two-mechanical-mode EOM system in the higher-frequency and lower-frequency channels as a function of ${G_{e2}}$ and $\theta $ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$, respectively. Other parameters are the same as Fig. 3.

Download Full Size | PDF

Compared with a typical three-mode EOM system, the two-mechanical-mode EOM system can not only realize double-channel M-O conversion, but also control the conversion efficiency and added noise of each channel by modulating the strength and phase of mechanical coupling between MRs. Moreover, it’s due to the fact that there are three more tunable coupling parameters than a typical three-mode EOM system, the matching conditions of coupling rates are easier to reach. For the strong interference, the distance of double channels gets wider and may exceed the linewidth of microwave cavity.

4. Tunable entanglement and conversion in an N-mechanical-mode EOM system

In this section, we extend our scheme to investigate the tunable M-O entanglement and conversion in an N-mechanical-mode EOM system, where the nearest-neighboring mechanical modes are phase-dependent phonon-phonon coupled with each other and the microwave (optical) cavity mode is coupled to ${N_e}$ (${N_o},\textrm{ for }{N_e},{N_o} \le N$) mechanical modes simultaneously. Thus, the Hamiltonian in the interaction picture can be written as

$${\hat{H}_I} = \sum\limits_{l = 1}^N {\hbar {g_{el}}\hat{a}_e^{\dagger} {{\hat{a}}_e}({{{\hat{b}}_l} + \hat{b}_l^{\dagger} } )+ \hbar {g_{ol}}\hat{a}_o^{\dagger} {{\hat{a}}_o}({{{\hat{b}}_l} + \hat{b}_l^{\dagger} } )} + \sum\limits_{l = 1}^{N - 1} {\hbar {\zeta _l}({{e^{ - i{\theta_l}}}{{\hat{b}}_l}\hat{b}_{l + 1}^{\dagger} + {e^{i{\theta_l}}}\hat{b}_l^{\dagger} {{\hat{b}}_{l + 1}}} )}.$$
We set the parameters ${\omega _{ml}} = {\omega _m},{g_{el}} = {g_e},{g_{ol}} = {g_o},{\zeta _l} = \zeta .(l = 1,2,\ldots N)$ for simplicity. The Hamiltonian of these coupled mechanical modes can be diagonalized as [64]
$$\begin{aligned} {{\hat{H}}_m} &= \sum\limits_{l = 1}^N {\hat{b}_l^{\dagger} {{\hat{b}}_l}} + \zeta \sum\limits_{l = 1}^{N - 1} {({{e^{ - i{\theta_l}}}{{\hat{b}}_l}\hat{b}_{l + 1}^{\dagger} + {e^{i{\theta_l}}}\hat{b}_l^{\dagger} {{\hat{b}}_{l + 1}}} )} \\ &= \sum\limits_{k = 1}^N {{\Omega _k}\hat{B}_k^{\dagger} {{\hat{B}}_k}} \end{aligned},$$
with
$$\begin{aligned} {{\hat{b}}_1} &= \sum\limits_{k = 1}^N {{{\sin \left( {\frac{{k\pi }}{{N + 1}}} \right){{\hat{B}}_k}} / {\sqrt {{{({N + 1} )} / 2}} }}} \\ &{ {{{\hat{b}}_l}} |_{l \ge 2}} = {e^{ - i\sum\nolimits_{\nu = 1}^{l - 1} {{\theta _\nu }} }}\sum\limits_{k = 1}^N {{{\sin \left( {\frac{{lk\pi }}{{N + 1}}} \right){{\hat{B}}_k}} / {\sqrt {{{({N + 1} )} / 2}} }}} \end{aligned}.$$
Here, ${\hat{B}_k}$ is the $k\textrm{th}$ diagonalized mechanical mode with the resonance frequency
$${\Omega _k} = {\omega _m} + 2\zeta \cos \left( {\frac{{k\pi }}{{N + 1}}} \right).$$
Physically, N pathways of electro-opto-mechanical interaction are interfered and form N new independent pathways, which can be expressed by N line-coupled mechanical modes being hybridized into N decoupled mechanical modes. Correspondingly, the Hamiltonian in the interaction picture can be rewritten as
$${\hat{H}_I} = \sum\limits_{j = e,o} {\frac{{2{G_j}}}{{N + 1}}\sum\limits_{k = 1}^N {\left[ {\sin \left( {\frac{{k\pi }}{{N + 1}}} \right) + \sum\limits_{l = 2}^N {{e^{i\sum\nolimits_{\nu = 1}^{l - 1} {{\theta_\nu }} }}\sin \left( {\frac{{lk\pi }}{{N + 1}}} \right)} } \right]} } {\hat{a}_j}\hat{B}_k^{\dagger} + H.c.,$$
where every ${\hat{B}_k}$ are completely decoupled with other mechanical dressed modes, and lead to an N-discrete-channel M-O entanglement and conversion. In Fig. 6(a), we plot the output optical fields ${P_o}$ in M-O entanglement as a function of $\delta $ with $N = 3$. The Fig. 6(b) represents the ${|{t(\omega )} |^2}$ in M-O conversion with the same parameter settings. It’s obviously shown that there exist three channels to entangle or convert microwave and optical photons, and the performance on every channel can be controlled by tuning $\zeta$ and ${\theta _l}$. When $\zeta$ is so small that the multiple channels are undistinguished, we may get the entanglement or conversion with a wide bandwidth.

5. Conclusion

In this work, we theoretically studied a two-mechanical-mode EOM system and investigated the enhanced M-O entanglement as well as coherent conversion. It’s demonstrated that the quantum interference of two electro-opto-mechanical interaction pathways leads to the tunable double-channel phenomena. For a symmetric system where two MRs are coupled to the cavities identically, we can steer the switch from the M-O interaction with a single channel to that of the double-channel by tuning the coupling phase of two MRs, and the distance of two channels in frequency domain is linearly related with the coupling strength. For an asymmetric system, besides the tuning of phonon-phonon interaction, the entanglement and conversion characteristics of each channel can be controlled via the mismatch of electromechanical and optomechanical couplings. Specifically, the performance on one channel could be enhanced by sacrificing that of the other channel. Therefore, we can always find the optimal parameter settings to realize the best entanglement or conversion. Moreover, the matching conditions of couplings are easier to reach due to the three more tunable coupling parameters compared with a typical three-mode EOM system. Finally, we extend our scheme to investigate the M-O entanglement and conversion in an N-mechanical-mode EOM system, which shows N tunable discrete channels by appropriate parameter settings. Additionally, our scheme could be generalized to other nonlinear interaction in multimode systems such as magnon-optics and atom ensembles. The proposed multimode EOM system opens up prospects for quantum information transduction and storage with a wide bandwidth and multichannel quantum interface.

 figure: Fig. 6.

Fig. 6. Entanglement and conversion in a three-mechanical-mode EOM system. (a) In the entanglement generation, the output optical fields ${P_o}$ as a function of $\delta $ when ${\zeta / {2\pi }} = \textrm{0}\textrm{.08MHz},{\theta _1} = {\pi / 2},{\theta _2} = 0$ (blue solid line) and ${\zeta / {2\pi }} = \textrm{0}\textrm{.04MHz},{\theta _1} = {{3\pi } / 4},{\theta _2} = 0$ (red solid line), respectively. (b) In the conversion generation, the conversion efficiency ${|{t(\omega )} |^2}$ as a function of $\delta $ when ${\zeta / {2\pi }} = \textrm{0}\textrm{.08MHz},{\theta _1} = {\pi / 2},{\theta _2} = 0$ (blue solid line) and ${\zeta / {2\pi }} = \textrm{0}\textrm{.04MHz},{\theta _1} = {{3\pi } / 4},{\theta _2} = 0$ (red solid line), respectively. Other parameters are the same as Fig. 3.

Download Full Size | PDF

Appendix

A. Expressions of ${A_j}$, $B$, ${D_j}$ ($j = e,o$)

By applying quantum Langevin equations to involved cavity modes and substituting several variables as Ref. [62] did, parameters ${A_j}$, B, ${D_j}$ ($j = e,o$) can be expressed in terms of cooperativity parameter ${C_{j1}}$ ($j = e,o$) as

$${A_e} = \frac{{1 - ({C_{e1}} + {C_{o1}})}}{{1 + ({C_{e1}} - {C_{o1}})}},$$
$${A_o} = \frac{{1 + ({C_{e1}} + {C_{o1}})}}{{1 + ({C_{e1}} - {C_{o1}})}},$$
$$B = \frac{{2\sqrt {{C_{e1}}{C_{o1}}} }}{{1 + ({C_{e1}} - {C_{o1}})}},$$
$${D_e} = \frac{{2i\sqrt {{C_{e1}}} }}{{1 + ({C_{e1}} - {C_{o1}})}},$$
$${D_o} = \frac{{2i\sqrt {{C_{o1}}} }}{{1 + ({C_{e1}} - {C_{o1}})}}.$$
Considering the general case of two-mechanical-mode EOM system, whose Hamiltonian and multiphoton-phonon coupling rates are expressed by Eq. (5) and Eq. (18) in the main manuscript, the cooperativity parameters for ${\hat{b}_k}$ ($k ={+} , - $) reads as
$${C_{jk}} = \frac{{G_{jk}^2}}{{{\kappa _j}}}.$$
Hence, ${A_{jk}}$, ${B_k}$, ${D_{jk}}$ can be rewrite in terms of ${C_{jk}}$ like Eq. (3337).

B. Deductions of Eqs. (12)-(14)

The covariance matrixes of fluctuation operators are expressed as

$$\left\langle {\delta \hat{a}_i^{\dagger} \delta {{\hat{a}}_j}} \right\rangle = {\bar{n}_i}{\delta _{ij}},$$
$$\left\langle {\delta {{\hat{a}}_i}\delta \hat{a}_j^{\dagger} } \right\rangle = ({{{\bar{n}}_i} + 1} ){\delta _{ij}},$$
$$\left\langle {\hat{\delta }{a_i}\delta {{\hat{a}}_j}} \right\rangle = \left\langle {\delta \hat{a}_i^{\dagger} \delta \hat{a}_j^{\dagger} } \right\rangle = 0,$$
where $\delta {\hat{a}_i},\delta {\hat{a}_j} = \delta {\hat{a}_{e,{\mathop{\rm int}} }},\delta {\hat{a}_{o,{\mathop{\rm int}} }},\delta \hat{m}$, ${\bar{n}_i} = {\bar{n}_{e,{T_E}}},{\bar{n}_{o,{T_E}}},{\bar{n}_{m,{T_E}}}$, and ${\delta _{ij}} = \left\{ \begin{array}{cc} 1,i = j\\ &0,i \ne j \end{array} \right.$, we have
$$\begin{aligned} {{\bar{N}}_M} &= \left\langle {{{\hat{E}}^{\dagger} }\hat{E}} \right\rangle \\ &= \left\langle {({A_e^ \ast \delta \hat{a}_{e,{\mathop{\rm int}} }^{\dagger} - {B^ \ast }\delta {{\hat{a}}_{o,{\mathop{\rm int}} }} - C_e^ \ast \delta \hat{m}} )({A_e^{}\delta \hat{a}_{e,{\mathop{\rm int}} }^{} - B\delta \hat{a}_{o,{\mathop{\rm int}} }^{\dagger} - C_e^{}\delta \hat{m}} )} \right\rangle \\ &= {|{A_e^{}} |^2}\left\langle {\delta \hat{a}_{e,{\mathop{\rm int}} }^{\dagger} \delta \hat{a}_{e,{\mathop{\rm int}} }^{}} \right\rangle - A_e^ \ast B\left\langle {\delta \hat{a}_{e,{\mathop{\rm int}} }^{\dagger} \delta \hat{a}_{o,{\mathop{\rm int}} }^{\dagger} } \right\rangle - A_e^ \ast C_e^{}\left\langle {\delta \hat{a}_{e,{\mathop{\rm int}} }^{\dagger} \delta \hat{m}} \right\rangle \\ &- {B^ \ast }A_e^{}\left\langle {\delta {{\hat{a}}_{o,{\mathop{\rm int}} }}\delta {{\hat{a}}_{e,{\mathop{\rm int}} }}} \right\rangle + {|B |^2}\left\langle {\delta {{\hat{a}}_{o,{\mathop{\rm int}} }}\delta \hat{a}_{o,{\mathop{\rm int}} }^{\dagger} } \right\rangle + {B^ \ast }C_e^{}\left\langle {\delta {{\hat{a}}_{o,{\mathop{\rm int}} }}\delta \hat{m}} \right\rangle \\ &- C_e^ \ast A_e^{}\left\langle {\delta {{\hat{m}}^{\dagger} }\delta {{\hat{a}}_{e,{\mathop{\rm int}} }}} \right\rangle + C_e^ \ast B\left\langle {\delta {{\hat{m}}^{\dagger} }\delta \hat{a}_{o,{\mathop{\rm int}} }^{\dagger} } \right\rangle + {|{C_e^{}} |^2}\left\langle {\delta {{\hat{m}}^{\dagger} }\delta \hat{m}} \right\rangle \\ &= {|{A_e^{}} |^2}\left\langle {\delta \hat{a}_{e,{\mathop{\rm int}} }^{\dagger} \delta \hat{a}_{e,{\mathop{\rm int}} }^{}} \right\rangle + {|B |^2}\left\langle {\delta {{\hat{a}}_{o,{\mathop{\rm int}} }}\delta \hat{a}_{o,{\mathop{\rm int}} }^{\dagger} } \right\rangle + {|{C_e^{}} |^2}\left\langle {\delta {{\hat{m}}^{\dagger} }\delta \hat{m}} \right\rangle \\ &= {|{A_e^{}} |^2}{{\bar{n}}_{e,{T_E}}} + {|B |^2}({{{\bar{n}}_{o,{T_E}}} + 1} )+ {|{C_e^{}} |^2}{{\bar{n}}_{m,{T_E}}} \end{aligned}$$
Equation (42) yields Eq. (12), and Eqs. (13)-(14) can be yielded by similar operations.

C. Calculation details of the elements of $V(\omega )$

First, the quadrature operators of the microwave and optical signals can be denoted as ${\hat{X}_j},{\hat{Y}_j}$. We define the covariance matrix of our system by the quadratures of the microwave and optical cavities’ output in the frequency domain, which can be expressed as

$$\delta (\omega + \omega ^{\prime}){V_{ij}}(\omega ) = \frac{1}{2}\left\langle {{u_i}(\omega ){u_j}(\omega^{\prime}) + {u_j}(\omega^{\prime}){u_i}(\omega )} \right\rangle, $$
with ${\mathbf u} = {[{\hat{X}_e},{\hat{Y}_e},{\hat{X}_o},{\hat{Y}_o}]^T}$. Then, by using Eqs. (9), (10) and (43), we write the covariance matrix in a normal form as
$${\mathbf V}(\omega ) = \left( {\begin{array}{cccc} {{V_{11}}}&0&{{V_{13}}}&0\\ 0&{{V_{11}}}&0&{ - {V_{13}}}\\ {{V_{13}}}&0&{{V_{33}}}&0\\ 0&{ - {V_{13}}}&0&{{V_{33}}} \end{array}} \right), $$
where ${V_{11}}$ and ${V_{33}}$ represent the autocorrelation of microwave and optical signal, respectively, while ${V_{13}}$ corresponds to the cross-correlation. The elements of ${\mathbf V}(\omega )$ can be figured out by Eqs. (12)-(14) and expressed as
$${V_{11}} = \frac{{\left\langle {{{\hat{X}}_e}(\omega ){{\hat{X}}_e}(\omega^{\prime})} \right\rangle }}{{\delta (\omega + \omega ^{\prime})}} = {\bar{N}_e} + \frac{1}{2}, $$
$${V_{33}} = \frac{{\left\langle {{{\hat{X}}_o}(\omega ){{\hat{X}}_o}(\omega^{\prime})} \right\rangle }}{{\delta (\omega + \omega ^{\prime})}} = {\bar{N}_o} + \frac{1}{2}, $$
$${V_{13}} = \frac{{\left\langle {{{\hat{X}}_e}(\omega ){{\hat{X}}_o}(\omega^{\prime}) + {{\hat{X}}_o}(\omega^{\prime}){{\hat{X}}_e}(\omega )} \right\rangle }}{{\delta (\omega + \omega ^{\prime})}} = \left\langle {\hat{c}_e^{\textrm{out}}\hat{c}_o^{\textrm{out}}} \right\rangle. $$

D. Logarithmic negativity as a function of coupling strength of MRs

Table 2. The normalized logarithmic negativity as a function of coupling strength of MRs.

Tables Icon

Table 2. Normalized logarithmic negativity as a function of coupling strength of MRs.a

E. Expressions of $S$

By checking the Routh-Hurwitz stability conditions, we can know whether the system is stable. It’s expressed as

$${\kappa _e}{C_{em}} - {\kappa _o}{C_{om}} > K\max \left( {{\kappa_e} - {\kappa_o},\frac{{{\kappa_o}^2 - {\kappa_e}^2}}{{2{\gamma_m} + {\kappa_e} + {\kappa_o}}}} \right)$$
with $K = \frac{{{C_{em}}}}{{1 + {\kappa _o}/{\kappa _e}}} + \frac{{{C_{om}}}}{{1 + {\kappa _e}/{\kappa _o}}}$. The stability ratio is defined as Eq. (D.2) for convenience.
$$S = \frac{{{\kappa _e}{C_{em}} - {\kappa _o}{C_{om}}}}{{K\max \left( {{\kappa_e} - {\kappa_o},\frac{{{\kappa_o}^2 - {\kappa_e}^2}}{{2{\gamma_m} + {\kappa_e} + {\kappa_o}}}} \right)}}$$
The system is stable only if $S > 1$. $S = 1$ represents the boundary between stable region and unstable region.

Funding

National Natural Science Foundation of China (62073338).

Acknowledgments

This work is sponsored by the National Natural Science Foundation of China under Grant Nos. 62073338.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. T. J. Kippenberg and K. J. Vahala, “Cavity optomechanics: back-action at the mesoscale,” Science 321(5893), 1172–1176 (2008). [CrossRef]  

2. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86(4), 1391–1452 (2014). [CrossRef]  

3. M. Eichenfield, J. Chan, R. M. Camacho, K. J. Vahala, and O. Painter, “Optomechanical crystals,” Nature 462(7269), 78–82 (2009). [CrossRef]  

4. K. C. Balram, M. Davanço, J. Y. Lim, J. D. Song, and K. Srinivasan, “Moving boundary and photoelastic coupling in GaAs optomechanical resonators,” Optica 1(6), 414–420 (2014). [CrossRef]  

5. S. Gröblacher, K. Hammerer, M. R. Vanner, and M. Aspelmeyer, “Observation of strong coupling between a micromechanical resonator and an optical cavity field,” Nature 460(7256), 724–727 (2009). [CrossRef]  

6. G. S. Agarwal and S. M. Huang, “The electromagnetically induced transparency in mechanical effects of light,” Phys. Rev. A 81(4), 041803 (2010). [CrossRef]  

7. M. D. Eisaman, A. André, F. Massou, M. Fleischhauer, A. S. Zibrov, and M. D. Lukin, “Electromagnetically induced transparency with tunable single-photon pulses,” Nature 438(7069), 837–841 (2005). [CrossRef]  

8. S. Weis, R. Riviere, S. Deléglise, E. Gavartin, O. Arcizet, A. Schliesser, and T. J. Kippenberg, “Optomechanically induced transparency,” Science 330(6010), 1520–1523 (2010). [CrossRef]  

9. H. K. Lau and A. A. Clerk, “Macroscale entanglement and measurement,” Science 372(6542), 570–571 (2021). [CrossRef]  

10. L. M. de Lépinay, C. F. Ockeloen-Korppi, M. J. Woolley, and M. A. Sillanpää, “Quantum mechanics–free subsystem with mechanical oscillators,” Science 372(6542), 625–629 (2021). [CrossRef]  

11. S. Kotler, G. A. Peterson, E. Shojaee, F. Lecocq, K. Cicak, A. Kwiatkowski, S. Geller, S. Glancy, E. Knell, R. W. Simmonds, J. Aumentado, and J. D. Teufel, “Direct observation of deterministic macroscopic entanglement,” Science 372(6542), 622–625 (2021). [CrossRef]  

12. R. Riedinger, A. Wallucks, I. Marinkovic, C. Löschnauer, M. Aspelmeyer, S. Hong, and S. Gröblacher, “Remote quantum entanglement between two micromechanical oscillators,” Nature 556(7702), 473–477 (2018). [CrossRef]  

13. J. S. Zhang and A. X. Chen, “Large and robust mechanical squeezing of optomechanical systems in a highly unresolved sideband regime via Duffing nonlinearity and intracavity squeezed light,” Opt. Express 28(24), 36620–36631 (2020). [CrossRef]  

14. J. B. Clark, F. Lecocq, R. W. Simmonds, J. Aumentado, and J. D. Teufel, “Sideband cooling beyond the quantum backaction limit with squeezed light,” Nature 541(7636), 191–195 (2017). [CrossRef]  

15. H. K. Lau and A. A. Clerk, “Ground-state cooling and high-fidelity quantum transduction via parametrically driven bad-cavity optomechanics,” Phys. Rev. Lett. 124(10), 103602 (2020). [CrossRef]  

16. J. Chan, T. Alegre, A. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer, and O. Painter, “Laser cooling of a nanomechanical oscillator into its quantum ground state,” Nature 478(7367), 89–92 (2011). [CrossRef]  

17. X. Han, K. Y. Fong, and H. X. Tang, “A 10-GHz film-thickness-mode cavity optomechanical resonator,” Appl. Phys. Lett. 106(16), 161108 (2015). [CrossRef]  

18. M. Shen, J. Xie, C.-L. Zou, Y. Xu, W. Fu, and H. X. Tang, “High frequency lithium niobate film-thickness-mode optomechanical resonator,” Appl. Phys. Lett. 117(13), 131104 (2020). [CrossRef]  

19. P. B. Deotare, I. Bulu, I. W. Frank, Q. Quan, Y. Zhang, R. Ilic, and M. Loncar, “All optical reconfiguration of optomechanical filters,” Nat. Commun. 3(1), 846 (2012). [CrossRef]  

20. A. Krause, M. Winger, T. Blasius, Q. Lin, and O. Painter, “A high-resolution microchip optomechanical accelerometer,” Nat. Photonics 6(11), 768–772 (2012). [CrossRef]  

21. L. Richardson, A. Hines, A. Schaffer, B. P. Anderson, and F. Guzman, “Quantum hybrid optomechanical inertial sensing,” Appl. Opt. 59(22), G160–G166 (2020). [CrossRef]  

22. C. C. Zhong, X. Han, H. X. Tang, and L. Jiang, “Entanglement of microwave-optical modes in a strongly coupled electro-optomechanical system,” Phys. Rev. A 101(3), 032345 (2020). [CrossRef]  

23. Y. Y. Wang, R. Zhang, S. Chesi, and Y. D. Wang, “Reservoir-engineered entanglement in an unresolved-sideband optomechanical system,” Commun. Theor. Phys. 73(5), 055105 (2021). [CrossRef]  

24. J. T. Hill, A. H. Safavi-Naeini, J. Chan, and O. Painter, “Coherent optical wavelength conversion via cavity optomechanics,” Nat. Commun. 3(1), 1196 (2012). [CrossRef]  

25. Y. D. Wang and A. A. Clerk, “Using dark modes for high-fidelity optomechanical quantum state transfer,” New J. Phys. 14(10), 105010 (2012). [CrossRef]  

26. W. Hease, A. Rueda, R. Sahu, M. Wulf, G. Arnold, H. G. L. Schwefel, and J. M. Fink, “Bidirectional electro-optic wavelength conversion in the quantum ground state,” PRX Quantum 1(2), 020315 (2020). [CrossRef]  

27. J. Holzgrafe, N. Sinclair, D. Zhu, A. Shams-Ansari, M. Colangelo, Y. Hu, M. Zhang, K. K. Berggren, and M. Lončar, “Cavity electro-optics in thin-film lithium niobate for efficient microwave-to-optical transduction,” Optica 7(12), 1714–1720 (2020). [CrossRef]  

28. R. Hisatomi, A. Osada, Y. Tabuchi, T. Ishikawa, A. Noguchi, R. Yamazaki, K. Usami, and Y. Nakamura, “Bidirectional conversion between microwave and light via ferromagnetic magnons,” Phys. Rev. B 93(17), 174427 (2016). [CrossRef]  

29. N. Zhu, X. F. Zhang, X. Han, C. L. Zou, C. C. Zhong, C. H. Wang, L. Jiang, and H. X. Tang, “Waveguide cavity optomagnonics for microwave-to-optics conversion,” Optica 7(10), 1291–1297 (2020). [CrossRef]  

30. T. Bagci, A. Simonsen, S. Schmid, L. G. Villanueva, E. Zeuthen, J. Appel, J. M. Taylor, A. S. Sørensen, K. Usami, A. Schliesser, and E. S. Polzik, “Optical detection of radio waves through a nanomechanical transducer,” Nature 507(7490), 81–85 (2014). [CrossRef]  

31. R. W. Andrews, R. W. Peterson, T. P. Purdy, K. Cicak, R. W. Simmonds, C. A. Regal, and K. W. Lehnert, “Bidirectional and efficient conversion between microwave and optical light,” Nat. Phys. 10(4), 321–326 (2014). [CrossRef]  

32. M. Forsch, R. Stockill, A. Wallucks, I. Marinkovic, C. Gärtner, R. A. Norte, F. van Otten, A. Fiore, K. Srinivasan, and S. Gröblacher, “Microwave-to-optics conversion using a mechanical oscillator in its quantum ground state,” Nat. Phys. 16(1), 69–74 (2020). [CrossRef]  

33. M. Mirhosseini, A. Sipahigil, M. Kalaee, and O. Painter, “Superconducting qubit to optical photon transduction,” Nature 588(7839), 599–603 (2020). [CrossRef]  

34. M. Wu, E. Zeuthen, K. C. Balram, and K. Srinivasan, “Microwave-to-optical transduction using a mechanical supermode for coupling piezoelectric and optomechanical resonators,” Phys. Rev. Appl. 13(1), 014027 (2020). [CrossRef]  

35. K. C. Balram, M. I. Davanço, J. D. Song, and K. Srinivasan, “Coherent coupling between radiofrequency, optical and acoustic waves in piezo-optomechanical circuits,” Nat. Photonics 10(5), 346–352 (2016). [CrossRef]  

36. A. P. Higginbotham, P. S. Burns, M. D. Urmey, R. W. Peterson, N. S. Kampel, B. M. Brubaker, G. Smith, K. W. Lehnert, and C. A. Regal, “Harnessing electro-optic correlations in an efficient mechanical converter,” Nat. Phys. 14(10), 1038–1042 (2018). [CrossRef]  

37. G. Arnold, M. Wulf, S. Barzanjeh, E. Redchenko, A. Rueda, W. Hease, F. Hassani, and J. Fink, “Converting microwave and telecom photons with a silicon photonic nanomechanical interface,” Nat. Commun. 11(1), 4460 (2020). [CrossRef]  

38. X. Fernandez-Gonzalvo, Y. H. Chen, C. Yin, S. Rogge, and J. J. Longdell, “Coherent frequency up-conversion of microwaves to the optical telecommunications band in an Er:YSO crystal,” Phys. Rev. A 92(6), 062313 (2015). [CrossRef]  

39. B. Li, P. B. Li, Y. Zhou, S. L. Ma, and F. L. Li, “Quantum microwave-optical interface with nitrogen-vacancy centers in diamond,” Phys. Rev. A 96(3), 032342 (2017). [CrossRef]  

40. X. Fernandez-Gonzalvo, S. P. Horvath, Y. H. Chen, and J. J. Longdell, “Cavity-enhanced Raman heterodyne spectroscopy in Er3+:Y2SiO5 for microwave to optical signal conversion,” Phys. Rev. A 100(3), 033807 (2019). [CrossRef]  

41. K. Y. Xia, M. Johnsson, P. L. Knight, and J. Twamley, “Cavity-free scheme for nondestructive detection of a single optical photon,” Phys. Rev. Lett. 116(2), 023601 (2016). [CrossRef]  

42. Sh. Barzanjeh, M. Abdi, G. J. Milburn, P. Tombesi, and D. Vitali, “Reversible optical-to-microwave quantum interface,” Phys. Rev. Lett. 109(13), 130503 (2012). [CrossRef]  

43. J. Wu, C. H. Cui, L. R. Fan, and Q. T. Zhuang, “Deterministic microwave-optical transduction based on quantum teleportation,” Phys. Rev. Appl. 16(6), 064044 (2021). [CrossRef]  

44. A. Rueda, W. Hease, Sh. Barzanjeh, and J. M. Fink, “Electro-optic entanglement source for microwave to telecom quantum state transfer,” npj Quantum Inform. 5(1), 108 (2019). [CrossRef]  

45. D. Q. Zhu and P. B. Li, “Preparation of entangled states of microwave photons in a hybrid system via the electro-optic effect,” Opt. Express 25(23), 28305–28318 (2017). [CrossRef]  

46. C. C. Zhong, Z. X. Wang, C. L. Zou, M. Z. Zhang, X. Han, W. Fu, M. R. Xu, S. Shankar, M. H. Devoret, H. X. Tang, and L. Jiang, “Proposal for heralded generation and detection of entangled microwave–optical-photon pairs,” Phys. Rev. Lett. 124(1), 010511 (2020). [CrossRef]  

47. Z. Shen, Y. L. Zhang, Y. Chen, C. L. Zou, Y. F. Xiao, X. B. Zou, F. W. Sun, G. C. Guo, and C. H. Dong, “Experimental realization of optomechanically induced non-reciprocity,” Nat. Photonics 10(10), 657–661 (2016). [CrossRef]  

48. Y. Chen, Y. L. Zhang, Z. Shen, C. L. Zou, G. C Guo, and C. H Dong, “Synthetic gauge fields in a single optomechanical resonator,” Phys. Rev. Lett. 126(12), 123603 (2021). [CrossRef]  

49. N.R. Bernier, L.D. Tóth, A. Koottandavida, M. A. Ioannou, D. Malz, A. Nunnenkamp, A.K. Feofanov, and T. J. Kippenberg, “Nonreciprocal reconfigurable microwave optomechanical circuit,” Nat. Commun. 8(1), 604 (2017). [CrossRef]  

50. D. P. Lake, M. Mitchell, B. C. Sanders, and P. E. Barclay, “Two-colour interferometry and switching through optomechanical dark mode excitation,” Nat. Commun. 11(1), 2208 (2020). [CrossRef]  

51. D. G. Lai, X. Wang, W. Qin, B. P. Hou, F. Nori, and J. Q. Liao, “Tunable optomechanically induced transparency by controlling the dark-mode effect,” Phys. Rev. A 102(2), 023707 (2020). [CrossRef]  

52. C. Wang, Q. Lin, and B. He, “Breaking the optomechanical cooling limit by two drive fields on a membrane-in-the-middle system,” Phys. Rev. A 99(2), 023829 (2019). [CrossRef]  

53. Z. Shen, Y. L. Zhang, C. L. Zou, G. C. Guo, and C. H. Dong, “Dissipatively controlled optomechanical interaction via cascaded photon-phonon coupling,” Phys. Rev. Lett. 126(16), 163604 (2021). [CrossRef]  

54. D.-G. Lai, J.-F. Huang, X.-L. Yin, B.-P. Hou, W. Li, D. Vitali, F. Nori, and J.-Q. Liao, “Nonreciprocal ground-state cooling of multiple mechanical resonators,” Phys. Rev. A 102(1), 011502 (2020). [CrossRef]  

55. K. Stannigel, P. Rabl, and P. Zoller, “Driven-dissipative preparation of entangled states in cascaded quantum-optical networks,” New J. Phys. 14(6), 063014 (2012). [CrossRef]  

56. Y. L. Zhang, C. S. Yang, Z. Shen, C. H. Dong, G. C. Guo, C. L. Zou, and X. B. Zou, “Enhanced optomechanical entanglement and cooling via dissipation engineering,” Phys. Rev. A 101(6), 063836 (2020). [CrossRef]  

57. M. C. Kuzyk and H. L. Wang, “Controlling multimode optomechanical interactions via interference,” Phys. Rev. A 96(2), 023860 (2017). [CrossRef]  

58. P. C. Ma, J. Q. Zhang, Y. Xiao, M. Feng, and Z. M. Zhang, “Tunable double optomechanically induced transparency in an optomechanical system,” Phys. Rev. A 90(4), 043825 (2014). [CrossRef]  

59. F. Massel, S. U. Cho, J.-M. Pirkkalainen, P. J. Hakonen, T. T. Heikkilä, and M. A. Sillanpää, “Multimode circuit optomechanics near the quantum limit,” Nat. Commun. 3(1), 987 (2012). [CrossRef]  

60. P. Huang, L. Zhang, J. Zhou, T. Tian, P. Yin, C. Duan, and J. Du, “Nonreciprocal radio frequency transduction in a parametric mechanical artificial lattice,” Phys. Rev. Lett. 117(1), 017701 (2016). [CrossRef]  

61. H. Okamoto, A. Gourgout, C. Y. Chang, K. Onomitsu, I. Mahboob, E. Y. Chang, and H. Yamaguchi, “Coherent phonon manipulation in coupled mechanical resonators,” Nat. Phys. 9(8), 480–484 (2013). [CrossRef]  

62. Sh. Barzanjeh, S. Guha, C. Weedbrook, D. Vitali, J. H. Shapiro, and S. Pirandola, “Microwave quantum illumination,” Phys. Rev. Lett. 114(8), 080503 (2015). [CrossRef]  

63. M. B. Plenio, “Logarithmic negativity: a full entanglement monotone that is not convex,” Phys. Rev. Lett. 95(9), 090503 (2005). [CrossRef]  

64. W. Qin, C. Wang, and G. L. Long, “High-dimensional quantum state transfer through a quantum spin chain,” Phys. Rev. A 87(1), 012339 (2013). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. (a) Schematic of a two-mechanical-mode EOM system. Potentially experimental systems: (b) The two MRs are coupled with each other via Coulomb interaction, which can be controlled by the bias voltages V1 and -V2 on the MR1 and MR2, respectively. (c) Schematic of a fully coupled two-mechanical-mode EOM system. $\zeta {e^{ {\pm} i\theta }}:\zeta ({e^{ - i\theta }}{\hat{b}_1}\hat{b}_2^{\dagger} + {e^{i\theta }}\hat{b}_1^{\dagger} {\hat{b}_2})$. (d) Two MRs are capacitive coupled to a superconducting charge qubit, and the phase-dependent phonon-hopping interaction is induced and controlled by gate voltages. $C_{1,2}^x$ represents the gate capacitance, which is varied with the displacement ${x_{1,2}}$ of each MR. JJ: Josephson junction.
Fig. 2.
Fig. 2. Schematic of the energy-level diagram in the four-mode EOM system. The $|{{n_e}} \rangle $, $|{{n_o}} \rangle $, $|{{m_1}} \rangle $ and $|{{m_2}} \rangle $ denote the number states of the microwave and optical cavity photon, and MR1 and MR2 phonons, respectively. $|{{n_o},{n_e},{m_1},{m_2}} \rangle $ \leftrightarrow |{{n_o},{n_e} + 1,{m_1},{m_2}} \rangle $and$|{{n_o},{n_e},{m_1},{m_2}} \rangle $ \leftrightarrow |{{n_o} + 1,{n_e},{m_1},{m_2}} \rangle $ transitions change the cavity fields.$|{{n_o},{n_e} + 1,{m_1},{m_2}} \rangle \leftrightarrow |{{n_o},{n_e},{m_1} + 1,{m_2}} \rangle $ and $|{{n_o} + 1,{n_e},{m_1},{m_2}} \rangle \leftrightarrow |{{n_o},{n_e},{m_1} + 1,{m_2}} \rangle $ transitions are caused by capacitive coupling and radiation-pressure coupling, respectively. $|{{n_o},{n_e},{m_1} + 1,{m_2}} \rangle \leftrightarrow |{{n_o},{n_e},{m_1},{m_2} + 1} \rangle $ transition is induced by the Coulomb coupling, which induces two mechanical dressed modes $|{{n_o},{n_e},{b_ + }} \rangle $ and $|{{n_o},{n_e},{b_ - }} \rangle $.
Fig. 3.
Fig. 3. M-O Entanglement of the two-mechanical-mode EOM system. (a) The normalized output optical fields. (b) The normalized output optical fields for ${{{G_{e2}}} / {2\pi }} = 99.5\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 32.7\textrm{kHz}$ (blue lines), ${{{G_{e2}}} / {2\pi }} = 63.1\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 57.8\textrm{kHz}$ (red lines). (c) The normalized logarithmic negativity ${{{E_{N \pm }}} / {{{\bar{N}}_o}}}$ as a function of $\theta $ for ${{{G_{e2}}} / {2\pi }} = 99.5\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 32.7\textrm{kHz}$ (blue), ${{{G_{e2}}} / {2\pi }} = 63.1\textrm{kHz,}{{{G_{o2}}} / {2\pi }} = 57.8\textrm{kHz}$ (red). (d) ${{{E_{N + }}} / {{{\bar{N}}_o}}}$ as a function of ${G_{e2}}$ and ${G_{o2}}$.
Fig. 4.
Fig. 4. M-O conversion efficiency of the two-mechanical-mode EOM system. (a) The conversion efficiency ${|{{t_ + }(\omega )} |^2}$ of the itinerant microwave photons in the higher-frequency channel as a function of ${G_{e2}}$ and ${G_{o2}}$ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$, $\theta = \pi $. (b) ${|{{t_ - }(\omega )} |^2}$ under the same parameters as (a). (c) ${|{{t_ \pm }(\omega )} |^2}$ versus $\zeta$ and $\delta $ when $\theta = {\pi / 2}$. (d) ${|{{t_ \pm }(\omega )} |^2}$ versus $\theta $ and $\delta $ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$. Other parameters are the same as Fig. 3.
Fig. 5.
Fig. 5. (a), (b) Added noise ${n_{add, \pm }}$ of the two-mechanical-mode EOM system in the higher-frequency and lower-frequency channels as a function of ${G_{e2}}$ and $\theta $ when ${\zeta / {2\pi }} = 0.08\textrm{MHz}$, respectively. Other parameters are the same as Fig. 3.
Fig. 6.
Fig. 6. Entanglement and conversion in a three-mechanical-mode EOM system. (a) In the entanglement generation, the output optical fields ${P_o}$ as a function of $\delta $ when ${\zeta / {2\pi }} = \textrm{0}\textrm{.08MHz},{\theta _1} = {\pi / 2},{\theta _2} = 0$ (blue solid line) and ${\zeta / {2\pi }} = \textrm{0}\textrm{.04MHz},{\theta _1} = {{3\pi } / 4},{\theta _2} = 0$ (red solid line), respectively. (b) In the conversion generation, the conversion efficiency ${|{t(\omega )} |^2}$ as a function of $\delta $ when ${\zeta / {2\pi }} = \textrm{0}\textrm{.08MHz},{\theta _1} = {\pi / 2},{\theta _2} = 0$ (blue solid line) and ${\zeta / {2\pi }} = \textrm{0}\textrm{.04MHz},{\theta _1} = {{3\pi } / 4},{\theta _2} = 0$ (red solid line), respectively. Other parameters are the same as Fig. 3.

Tables (2)

Tables Icon

Table 1. Parameter settings of the EOM systema

Tables Icon

Table 2. Normalized logarithmic negativity as a function of coupling strength of MRs.a

Equations (49)

Equations on this page are rendered with MathJax. Learn more.

H ^ = H ^ 0 = ω m 1 b ^ 1 b ^ 1 + ω m 2 b ^ 2 b ^ 2 + ω e a ^ e a ^ e + ω o a ^ o a ^ o + j = e , o g j 1 a ^ j a ^ j ( b ^ 1 + b ^ 1 ) + λ r ( b ^ 1 b ^ 2 + b ^ 1 b ^ 2 ) + j = e , o i ε j ( a ^ j e i ω d , j t a ^ j e i ω d , j t ) ,
H ^ = Δ e a ^ e a ^ e + Δ o a ^ o a ^ o + ( ω m λ r ) b ^ b ^ + ( ω m + λ r ) b ^ + b ^ + + 1 2 j = e , o g j 1 a ^ j a ^ j ( b ^ + + b ^ + ) + 1 2 j = e , o g j 1 a ^ j a ^ j ( b ^ + b ^ ) + j = e , o i ε j ( a ^ j a ^ j ) .
H ^ = H ^ 0 = ω m 1 b ^ 1 b ^ 1 + ω m 2 b ^ 2 b ^ 2 + ω e a ^ e a ^ e + ω o a ^ o a ^ o + j = e , o l = 1 , 2 g j l a ^ j a ^ j ( b ^ l + b ^ l ) + ζ ( e i θ b ^ 1 b ^ 2 + e i θ b ^ 1 b ^ 2 ) + j = e , o i ε j ( a ^ j e i ω d , j t a ^ j e i ω d , j t ) .
b ^ + = f b ^ 1 e i θ h b ^ 2 b ^ = e i θ h b ^ 1 + f b ^ 2 ,
H ^ = Δ e a ^ e a ^ e + Δ o a ^ o a ^ o + ω + b ^ + b ^ + + ω b ^ b ^ + j = e , o g j + a ^ j a ^ j ( b ^ + + b ^ + ) + j = e , o g j a ^ j a ^ j ( b ^ + b ^ ) + j = e , o i ε j ( a ^ j a ^ j )
ω ± = 1 2 ( ω m 1 + ω m 2 ± ( ω m 1 ω m 2 ) 2 + 4 ζ 2 ) f = | ω ω m 1 | / ( ω ω m 1 ) 2 + ζ 2 , h = ζ f / ( ω ω m 1 ) g e + = f g e 1 e i θ h g e 2 , g e = e i θ h g e 1 + f g e 2 g o + = f g o 1 e i θ h g o 2 , g o = e i θ h g o 1 + f g o 2 .
g j + 2 + g j 2 = g j 1 2 + g j 2 2
H ^ I = G e 1 ( δ a ^ e b ^ 1 + b ^ 1 δ a ^ e ) + G o 1 ( δ a ^ o b ^ 1 + b ^ 1 δ a ^ o ) .
E ^ = A e δ a ^ e , int B δ a ^ o , int D e δ b ^ 1 ,
O ^ = B δ a ^ e , int + A o δ a ^ o , int D o δ b ^ 1 ,
{ n ¯ e , T E = 1 e ω e / k B T E 1 n ¯ o , T E = 1 e ω o / k B T E 1 n ¯ m 1 , T E = 1 e ω m 1 / k B T E 1 .
N ¯ e = E ^ E ^ = | A e | 2 n ¯ e , T E + | B | 2 ( n ¯ o , T E + 1 ) + | D e | 2 n ¯ m 1 , T E ,
N ¯ o = O ^ O ^ = | B | 2 ( n ¯ e , T E + 1 ) + | A o | 2 n ¯ o , T E + | D o | 2 ( n ¯ m 1 , T E + 1 ) ,
E ^ O ^ = A e B ( n ¯ e , T E + 1 ) B A o n ¯ o , T E + D e D o ( n ¯ m 1 , T E + 1 ) .
ε = | E ^ O ^ | N ¯ e N ¯ o .
E N = max [ 0 , ln ( 2 μ ) ]
μ = 1 2 [ v 11 2 + v 33 2 + 2 v 13 2 ( v 11 2 v 33 2 ) 2 + 4 v 13 2 ( v 11 + v 33 ) 2 ] 1 / 2 ,
G e + = f G e 1 e i θ h G e 2 , G e = e i θ h G e 1 + f G e 2 G o + = f G o 1 e i θ h G o 2 , G o = e i θ h G o 1 + f G o 2 ,
G j + 2 + G j 2 = G j 1 2 + G j 2 2 .
H ^ I = G e + ( δ a ^ e δ b ^ + + δ b ^ + δ a ^ e ) + G o + ( δ a ^ o δ b ^ + + δ b ^ + δ a ^ o ) + G e ( δ a ^ e δ b ^ + δ b ^ δ a ^ e ) + G o ( δ a ^ o δ b ^ + δ b ^ δ a ^ o ) ,
E ^ + = A e + δ a ^ e , int + B + δ a ^ o , int + D e + δ b ^ + ,
O ^ + = B + δ a ^ e , int + + A o + δ a ^ o , int + D o + δ b ^ + ,
E ^ = A e δ a ^ e , int B δ a ^ o , int D e δ b ^ ,
O ^ = B δ a ^ e , int + A o δ a ^ o , int D o δ b ^ ,
H ^ C = G e + ( δ a ^ e b ^ + + b ^ + δ a ^ e ) + G o + ( δ a ^ o b ^ + + b ^ + δ a ^ + ) + G e ( δ a ^ e b ^ + b ^ δ a ^ e ) + G o ( δ a ^ o b ^ + b ^ δ a ^ o ) ,
t k ( ω ) = a o o u t ( ω ) a e i n ( ω ) | k = Γ e k Γ o k i ( ω ω m k ) + ( Γ e k + Γ o k + κ m ) / 2 R e R o η e η o ,
n a d d , k = 1 R e η e ( κ m n T Γ e k + ( R e 1 ) + ( R o 1 ) )
H ^ I = l = 1 N g e l a ^ e a ^ e ( b ^ l + b ^ l ) + g o l a ^ o a ^ o ( b ^ l + b ^ l ) + l = 1 N 1 ζ l ( e i θ l b ^ l b ^ l + 1 + e i θ l b ^ l b ^ l + 1 ) .
H ^ m = l = 1 N b ^ l b ^ l + ζ l = 1 N 1 ( e i θ l b ^ l b ^ l + 1 + e i θ l b ^ l b ^ l + 1 ) = k = 1 N Ω k B ^ k B ^ k ,
b ^ 1 = k = 1 N sin ( k π N + 1 ) B ^ k / ( N + 1 ) / 2 b ^ l | l 2 = e i ν = 1 l 1 θ ν k = 1 N sin ( l k π N + 1 ) B ^ k / ( N + 1 ) / 2 .
Ω k = ω m + 2 ζ cos ( k π N + 1 ) .
H ^ I = j = e , o 2 G j N + 1 k = 1 N [ sin ( k π N + 1 ) + l = 2 N e i ν = 1 l 1 θ ν sin ( l k π N + 1 ) ] a ^ j B ^ k + H . c . ,
A e = 1 ( C e 1 + C o 1 ) 1 + ( C e 1 C o 1 ) ,
A o = 1 + ( C e 1 + C o 1 ) 1 + ( C e 1 C o 1 ) ,
B = 2 C e 1 C o 1 1 + ( C e 1 C o 1 ) ,
D e = 2 i C e 1 1 + ( C e 1 C o 1 ) ,
D o = 2 i C o 1 1 + ( C e 1 C o 1 ) .
C j k = G j k 2 κ j .
δ a ^ i δ a ^ j = n ¯ i δ i j ,
δ a ^ i δ a ^ j = ( n ¯ i + 1 ) δ i j ,
δ ^ a i δ a ^ j = δ a ^ i δ a ^ j = 0 ,
N ¯ M = E ^ E ^ = ( A e δ a ^ e , int B δ a ^ o , int C e δ m ^ ) ( A e δ a ^ e , int B δ a ^ o , int C e δ m ^ ) = | A e | 2 δ a ^ e , int δ a ^ e , int A e B δ a ^ e , int δ a ^ o , int A e C e δ a ^ e , int δ m ^ B A e δ a ^ o , int δ a ^ e , int + | B | 2 δ a ^ o , int δ a ^ o , int + B C e δ a ^ o , int δ m ^ C e A e δ m ^ δ a ^ e , int + C e B δ m ^ δ a ^ o , int + | C e | 2 δ m ^ δ m ^ = | A e | 2 δ a ^ e , int δ a ^ e , int + | B | 2 δ a ^ o , int δ a ^ o , int + | C e | 2 δ m ^ δ m ^ = | A e | 2 n ¯ e , T E + | B | 2 ( n ¯ o , T E + 1 ) + | C e | 2 n ¯ m , T E
δ ( ω + ω ) V i j ( ω ) = 1 2 u i ( ω ) u j ( ω ) + u j ( ω ) u i ( ω ) ,
V ( ω ) = ( V 11 0 V 13 0 0 V 11 0 V 13 V 13 0 V 33 0 0 V 13 0 V 33 ) ,
V 11 = X ^ e ( ω ) X ^ e ( ω ) δ ( ω + ω ) = N ¯ e + 1 2 ,
V 33 = X ^ o ( ω ) X ^ o ( ω ) δ ( ω + ω ) = N ¯ o + 1 2 ,
V 13 = X ^ e ( ω ) X ^ o ( ω ) + X ^ o ( ω ) X ^ e ( ω ) δ ( ω + ω ) = c ^ e out c ^ o out .
κ e C e m κ o C o m > K max ( κ e κ o , κ o 2 κ e 2 2 γ m + κ e + κ o )
S = κ e C e m κ o C o m K max ( κ e κ o , κ o 2 κ e 2 2 γ m + κ e + κ o )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.