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

Generation and enhancement of the sum sideband under double radiation pressure

Open Access Open Access

Abstract

A theoretical scheme to enhance the sum sideband generation (SSG) via double radiation pressure is proposed. In this scheme, both sides of the double-cavity system are driven by red and blue detuned pump lasers and frequency components are generated at the sum sideband through optomechanical nonlinear interaction. The results show that the efficiency of SSG can be improved with orders of magnitude. We further investigate the properties of SSG in resolved and unresolved sideband regimes. The efficiencies of upper sum sideband generation (USSG) and lower sum sideband generation (LSSG) are the equivalent in the unresolved sideband regime when the threshold condition is satisfied. It is worth noting that with the increase of the ratio between the dissipation rate of the cavity field and the decay rate of the mechanical resonator (MR), the amplitude of the LSSG can be superior to that of the USSG. Our scheme may provide a potential application in realizing the measurement of high-precision weak forces and quantum-sensitive sensing.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Cavity optomechanics [1] has attracted great interest as an emerging cross-discipline between nanoscience and quantum optics, which focuses on exploring the nonlinear coupling between optical cavities and mechanical motion under optical radiation pressure. In recent years, optomechanical systems [2] have also received considerable theoretical discussion [38], and its research has become a cutting-edge topic and a popular issue nowadays. Its numerous potential applications have been discovered in research from a range of physical fields, including optical mechanical dark modes [9], optomechanically induced transparency (OMIT) [1015], gravitational wave detection [1618], cooling of mechanical oscillators [1923], precision force sensing [2426], and optical frequency comb generation [2730]. OMIT is an interesting phenomenon caused by the radiation pressure between the optical and mechanical modes, which is a derivative of electromagnetically induced transparency. Based on OMIT, there are numerous novel research phenomena which have been identified, such as higher-order OMIT [31], reverse OMIT [32,12], and multiple OMIT [33,34], which provides a new possibility to realize optical storage and optical delay.

Recently, nonlinear sideband effects in OMIT have aroused much attention. Such as second-order sidebands [3540], higher-order sidebands [25,4145], and chaos [46]. Especially, the existence of sum [47] or difference sidebands [48] are revealed for the first time in an optomechanical system driven by multiple probe fields. These may provide a new experimental solution for precision measurements of charge [49,50] and mass detection [51]. However, the nonlinear interactions arising from weak coupling are not easily detected. Many theoretical works have been completed in order to enhance the efficiency of sum sideband generation (SSG). For example, it was demonstrated that an enhancement of several orders of magnitude was achieved for the efficiency of lower sum sideband generation (LSSG) by nonlinear optomechanical interactions via two-phonon processes [52]. In addition, Wu et al. have proposed a method to generate Laguerre-Gaussian optics sum sideband by exchanging orbital angular momentum in Laguerre-Gaussian rotating systems [53].

The above works were completed in an optomechanical system driven by one pump field and two probe fields. Furthermore, it was shown that the quantum coherence process due to dual driving can completely restrain the decoherence caused by dissipation of the MR, and this dual driving device can achieve optical signal amplification [54]. According to the above reasons, we present a double-cavity system driven by red and blue detuned pump fields. The results show that the efficiency of the SSG enhances significantly with increases of the ratio between the dissipation rate of the cavity and the decay rate of the MR. Furthermore, the amplitudes of USSG and LSSG tend to the identical maximum in the case of the threshold condition is satisfied. It is worth noting that the amplitude of the LSSG can be superior to that of the USSG in the unresolved sideband regime. The enhanced sum sideband effect can be applied to quantum information processing and optical communications, and so on.

The paper is organized as follows. In Sec. 2, we give a theoretical description of the double-cavity system and provide a derivation of the Heisenberg-Langevin motion equation under double driving, where the steady-state solution of the system and the amplitude of sum sideband generation are obtained. In Sec. 3, we concretely discuss the effect of system parameters on the sum sideband and further analyze the characteristics of sum sideband generation in both the resolved and unresolved sideband regime. Finally, the conclusions are presented in Sec. 4.

2. Theoretical model and formulations

As shown in Fig. 1(a), the optomechanical system considered here contains a double-cavity field, which inserts a totally reflecting MR between two fixed partially transmitting mirrors. The eigen-frequencies and decay rate of MR are ${\omega _m}$ and ${\gamma _m}$. Two identical optical cavities with length L and frequency ${\omega _0}$ will be obtained when MR is at rest. Although the two optical cavities are respectively coupled to the MR, the tunneling coupling between the two cavities is weak and therefore negligible [54]. The Hamiltonian of this system can be written as below (in a frame rotating at ${\omega _c}$):

$$\begin{aligned} &H = \hbar {\Delta _1}a_1^{\dagger} {a_1} + \hbar {\Delta _2}a_2^{\dagger} {a_2} + \hbar g({a_2^{\dagger} {a_2} - a_1^{\dagger} {a_1}} )x + \left( {\frac{{{p^2}}}{{2m}} + \frac{1}{2}m\omega_m^2{x^2}} \right)\\ &+ i\hbar \sqrt {\eta \kappa } {\varepsilon _c}({a_1^{\dagger} - {a_1}} )+ i\hbar \sqrt {\eta \kappa } {\varepsilon _d}({a_2^{\dagger} - {a_2}} )\\ &+ i\hbar \sqrt {\eta \kappa } [{({a_1^{\dagger} {\varepsilon_1}{e^{ - i{\delta_1}t}} + a_1^{\dagger} {\varepsilon_2}{e^{ - i{\delta_2}t}}} )- \textrm{H}\textrm{.c}.} ] \end{aligned}$$
where ${\Delta _1} = {\omega _0} - {\omega _c}$, ${\Delta _2} = {\omega _0} - {\omega _d}$ and ${\delta _i} = {\omega _i} - {\omega _c}({i = 1,2} )$. $\textrm{H}\textrm{.c}\textrm{.}$ is the Hermitian conjugate. p and x are the momentum and position operators of the MR with effective mass m and angular frequency ${\omega _m}$. Here ${a_i}({a_i^{\dagger} } )$ describes the annihilation (creation) operator of the $i$th cavity field. The term $\hbar g({a_2^{\dagger} {a_2} - a_1^{\dagger} {a_1}} )x$ denotes the interaction between the cavity field and the MR. g is the optomechanical coupling constant. The two cavity fields have the same relaxation decay rate and the total loss rate of the cavity fields $\kappa $ consists of an intrinsic loss rate ${\kappa _0}$ and an external loss rate ${\kappa _{ex}}$, and the coupling parameter $\eta = {\kappa / {{\kappa _{ex}}}}$ is chosen to be the critical coupling value ${1 / 2}$ throughout the work [55]. ${\varepsilon _c} = \sqrt {{{{P_c}} / {\hbar {\omega _c}}}} $ and ${\varepsilon _d} = \sqrt {{{{P_d}} / {\hbar {\omega _d}}}} $ are the amplitudes of the red and blue detuned pumping fields, respectively, where ${P_c}$ and ${P_d}$ are the powers of the dual pump fields. ${\varepsilon _i} = \sqrt {{{{P_i}} / {\hbar {\omega _i}}}} ({i = 1,2} )$ is the amplitude of the probe field. ${P_1}$ and ${P_2}$ are the powers of the dual probe fields with frequencies ${\omega _1}$ and ${\omega _2}$. The Heisenberg-Langevin equations can be obtained as follows:
$$\begin{aligned} &{{\dot{a}}_1} ={-} [{i({{\Delta _1} - gx} )+ \kappa } ]{a_1} + \sqrt {\eta \kappa } {\varepsilon _c} + \sqrt {\eta \kappa } ({{\varepsilon_1}{e^{ - i{\delta_1}t}} + {\varepsilon_2}{e^{ - i{\delta_2}t}}} ),\\ &{{\dot{a}}_2} ={-} [{i({{\Delta _2} + gx} )+ \kappa } ]{a_2} + \sqrt {\eta \kappa } {\varepsilon _d},\\ &\dot{x} = {p / m},\\ &\dot{p} = \hbar g({a_2^{\dagger} {a_2} - a_1^{\dagger} {a_1}} )- m\omega _m^2x - {\gamma _m}p \end{aligned}$$
where the operators are reduced to their expected values, i.e., ${a_i}(t )\equiv \left\langle {{{\hat{a}}_i}(t )} \right\rangle ({i = 1,2} )$ and $x(t )\equiv \left\langle {\hat{x}(t )} \right\rangle $ [10]. Equation (2) can be simplified as follows:
$$\begin{aligned} &{{\dot{a}}_1} ={-} [{i({{\Delta _1} - gx} )+ \kappa } ]{a_1} + \sqrt {\eta \kappa } {\varepsilon _c} + \sqrt {\eta \kappa } ({{\varepsilon_1}{e^{ - i{\delta_1}t}} + {\varepsilon_2}{e^{ - i{\delta_2}t}}} ),\\ &{{\dot{a}}_2} ={-} [{i({{\Delta _2} + gx} )+ \kappa } ]{a_2} + \sqrt {\eta \kappa } {\varepsilon _d},\\ &\left( {m\frac{{{d^2}}}{{d{t^2}}} + m{\gamma_m}\frac{d}{{dt}} + m\omega_m^2} \right)x = \hbar g({a_2^ \ast {a_2} - a_1^ \ast {a_1}} ). \end{aligned}$$

 figure: Fig. 1.

Fig. 1. (a) Schematic diagram of a double-cavity optomechanical system formed by a MR embedded in the middle of two fixed lenses. The opposite sides of the double-cavity optomechanical system are driven by a pump field 1 with frequency ${\omega _\textrm{c}}$ and a pump field 2 with frequency ${\omega _d}$, respectively. Two relatively weak probe fields with frequencies ${\omega _1}$ and ${\omega _2}$ are injected into the left optical cavity. Both cavities have the same cavity length L and frequency ${\omega _0}$ under static, while coupling with MR separately. (b) Frequency spectrogram of sum sideband generation in a double-cavity optomechanical system driven by dual pump fields and dual probe fields. ${\Delta _1} = {\omega _0} - {\omega _\textrm{c}}$, ${\Delta _2} = {\omega _0} - {\omega _d}$, ${\delta _1} = {\omega _1} - {\omega _\textrm{c}}$, ${\delta _2} = {\omega _2} - {\omega _\textrm{c}}$.

Download Full Size | PDF

Assuming that the input pump laser field is much more intense than the probe field. The perturbation method can be applied to these equations by factorization of mean field approximation, i.e., $\left\langle {Qc} \right\rangle = \left\langle Q \right\rangle \left\langle c \right\rangle $ [47]. The overall solution of intracavity field and mechanical displacement in the dual pump field and dual probe field can be written as $O = \bar{O} + \delta O({O = {a_1},{a_2},x} )$. The steady state value of the system can be obtained:

$$\begin{aligned} {{\bar{a}}_1} &= \frac{{\sqrt {\eta \kappa } {\varepsilon _c}}}{{i{\Delta _3} + \kappa }},\\ {{\bar{a}}_2} &= \frac{{\sqrt {\eta \kappa } {\varepsilon _d}}}{{i{\Delta _4} + \kappa }},\\ \bar{x} &= \frac{{\hbar g({{{|{{{\bar{a}}_2}} |}^2} - {{|{{{\bar{a}}_1}} |}^2}} )}}{{m\omega _m^2}} \end{aligned}$$
where ${\Delta _3} = {\Delta _1} - g\bar{x}$, ${\Delta _4} = {\Delta _2} + g\bar{x}$.

Now we consider the perturbation caused by the double probe field. The Langevin equation of the quantum fluctuation is as follows:

$$\begin{aligned} &\frac{d}{{dt}}\delta {a_1} = ig({\overline {{a_1}} \delta x + \delta x\delta {a_1}} )- ({i{\Delta _3} + k} )\delta {a_1} + \sqrt {\eta \kappa } ({{\varepsilon_1}{e^{ - i{\delta_1}t}} + {\varepsilon_2}{e^{ - i{\delta_2}t}}} ),\\ &\frac{d}{{dt}}\delta {a_2} ={-} ig({\overline {{a_2}} \delta x + \delta x\delta {a_2}} )- ({i{\Delta _4} + k} )\delta {a_2},\\ &\left( {m\frac{{{d^2}}}{{d{t^2}}} + m{\gamma_m}\frac{d}{{dt}} + m\omega_m^2} \right)\delta x = \hbar g[{\overline {a_2^ \ast } \delta {a_2} - \overline {a_1^ \ast } \delta {a_1} + \overline {{a_2}} \delta {a_2} - \overline {{a_1}} \delta {a_1} + \delta a_2^ \ast \delta {a_2} - \delta a_1^ \ast \delta {a_1}} ]. \end{aligned}$$

We must consider the nonlinear terms $ig\delta x\delta {a_1}$, $ig\delta x\delta {a_2}$ and $\hbar g({\delta a_2^ \ast \delta {a_2} - \delta a_1^ \ast \delta {a_1}} )$ since they play an important role in the sum sideband generation. In order to calculate the amplitude of the sum sideband, we deal with SSG in the perturbed state by introducing the following equations that neglect second-order and higher-order terms [47]:

$$\begin{aligned} &\delta {a_1} = A_1^ + {e^{ - i{\delta _1}t}} + A_1^ - {e^{i{\delta _1}t}} + A_2^ + {e^{ - i{\delta _2}t}} + A_2^ - {e^{i{\delta _2}t}} + A_s^ + {e^{ - i{\Omega _ + }t}} + A_s^ - {e^{i{\Omega _ + }t}} + \ldots ,\\ &\delta {a_2} = B_1^ + {e^{ - i{\delta _1}t}} + B_1^ - {e^{i{\delta _1}t}} + B_2^ + {e^{ - i{\delta _2}t}} + B_2^ - {e^{i{\delta _2}t}} + B_s^ + {e^{ - i{\Omega _ + }t}} + B_s^ - {e^{i{\Omega _ + }t}} + \ldots ,\\ &\delta x = {x_1}{e^{ - i{\delta _1}t}} + x_1^ \ast {e^{i{\delta _1}t}} + {x_2}{e^{ - i{\delta _2}t}} + x_2^ \ast {e^{i{\delta _2}t}} + {x_s}{e^{ - i{\Omega _ + }t}} + x_s^ \ast {e^{i{\Omega _ + }t}} + \ldots , \end{aligned}$$
with the frequency components of ${\pm} {\Omega _ + } ={\pm} ({{\delta_1} + {\delta_2}} )$ being called sum sideband. We simplify the calculation by retaining the nonlinear second order quantities for the SSG components and by neglecting other higher-order nonlinear terms. By inserting Eq. (6) into Eq. (5), we can obtain two sets of algebraic equations.
  • (1) The first group describes the linear response of the probe field with frequency ${\omega _1}({{\omega_2}} )$,
    $$\begin{aligned} &({{R_1} - i{\delta_1}} )A_1^ +{-} ig{{\bar{a}}_1}{x_1} - \sqrt {\eta \kappa } {\varepsilon _1} = 0,\\ &({{R_1} + i{\delta_1}} )A_1^ -{-} ig{{\bar{a}}_1}x_1^ \ast{=} 0,\\ &({{R_2} - i{\delta_1}} )B_1^ +{-} ig{{\bar{a}}_2}{x_1} = 0,\\ &({{R_2} + i{\delta_1}} )B_1^ -{-} ig{{\bar{a}}_2}x_1^ \ast{=} 0,\\ &\sigma ({{\delta_1}} ){x_1} - \hbar g[{\bar{a}_2^ \ast B_1^ +{-} \bar{a}_1^ \ast A_1^ +{+} {{\bar{a}}_2}{{({B_1^ - } )}^ \ast } - {{\bar{a}}_1}{{({A_1^ - } )}^ \ast }} ]= 0,\\ &({{R_1} - i{\delta_1}} )A_2^ +{-} ig{{\bar{a}}_1}{x_2} - \sqrt {\eta \kappa } {\varepsilon _2} = 0,\\ &({{R_1} + i{\delta_1}} )A_2^ -{-} ig{{\bar{a}}_1}x_2^ \ast{=} 0,\\ &({{R_2} - i{\delta_1}} )B_2^ +{-} ig{{\bar{a}}_2}{x_2} = 0,\\ &({{R_2} + i{\delta_1}} )B_2^ -{-} ig{{\bar{a}}_2}x_2^ \ast{=} 0,\\ &\sigma ({{\delta_1}} ){x_2} - \hbar g[{\bar{a}_2^ \ast B_2^ +{-} \bar{a}_1^ \ast A_2^ +{+} {{\bar{a}}_2}{{({B_2^ - } )}^ \ast } - {{\bar{a}}_1}{{({A_2^ - } )}^ \ast }} ]= 0, \end{aligned}$$
    where ${R_1} = i{\Delta _3} + \kappa $, ${R_2} = i{\Delta _4} + \kappa $, $\sigma (x )= m\omega _m^2 - m{x^2} - im{\gamma _m}x$.
  • (2) The second group describes the process of SSG,
    $$\begin{aligned} &({{R_1} - i{\Omega _ + }} )A_s^ +{-} ig{{\bar{a}}_1}{x_s} - ig({{x_1}A_2^ +{+} {x_2}A_1^ + } )= 0,\\ &({{R_1} + i{\Omega _ + }} )A_s^ -{-} ig{{\bar{a}}_1}x_s^ \ast{-} ig({x_1^ \ast A_2^ -{+} x_2^ \ast A_1^ - } )= 0,\\ &({{R_2} - i{\Omega _ + }} )B_s^ +{-} ig{{\bar{a}}_2}{x_s} - ig({{x_1}B_2^ +{+} {x_2}B_1^ + } )= 0,\\ &({{R_2} + i{\Omega _ + }} )B_s^ -{-} ig{{\bar{a}}_2}x_s^ \ast{-} ig({x_1^ \ast B_2^ -{+} x_2^ \ast B_1^ - } )= 0,\\ &\sigma ({{\Omega _ + }} ){x_s} = \hbar g[{\bar{a}_2^ \ast B_s^ +{-} \bar{a}_1^ \ast A_s^ +{+} {{({B_1^ - } )}^ \ast }B_2^ +{-} {{({A_1^ - } )}^ \ast }A_2^ +{+} {{({B_2^ - } )}^ \ast }B_1^ +{-} {{({A_2^ - } )}^ \ast }A_1^ + } ]\\ &+ \hbar g[{{{\bar{a}}_2}{{({B_s^ - } )}^ \ast } - {{\bar{a}}_1}{{({A_s^ - } )}^ \ast }} ]. \end{aligned}$$

The first group describes the linear response of the probe field. The solution of Eq. (6) is obtained as follows:

$$\begin{aligned} &A_j^ +{=} \frac{{\sqrt {\eta \kappa } {\varepsilon _j}{\tau ^2}({{\delta_j}} ){\theta _2}({{\delta_j}} )}}{{{\theta _1}({{\delta_j}} ){\theta _2}({{\delta_j}} ){\tau ^2}({{\delta_j}} )- {\beta _1}{\beta _2}}},\textrm{ }A_j^ -{=} \frac{{ig{{\bar{a}}_1}x_j^ \ast }}{{{D_1}({{\delta_j}} )}},\\ &B_j^ +{=} \frac{{\sqrt {\eta \kappa } {\varepsilon _j}{\beta _2}\tau ({{\delta_j}} )}}{{{\theta _1}({{\delta_j}} ){\theta _2}({{\delta_j}} ){\tau ^2}({{\delta_j}} )- {\beta _1}{\beta _2}}},\textrm{ }B_j^ -{=} \frac{{ig{{\bar{a}}_2}x_j^ \ast }}{{{D_2}({{\delta_j}} )}},\\ &{x_j} = \frac{{\hbar g({\bar{a}_2^ \ast B_j^ +{-} \bar{a}_1^ \ast A_j^ + } )}}{{\tau ({{\delta_j}} )}} \end{aligned}$$
where ${D_j}(x )= {R_j} + ix$, ${F_j} = i\hbar {g^2}{|{{{\bar{a}}_j}} |^2}$, ${\theta _j}(x )= D_j^ \ast (x )+ {{{F_j}} / {\tau (x )}}({j = 1,2} )$, ${\beta _1} = i\hbar {g^2}{\bar{a}_1}\bar{a}_2^ \ast $, ${\beta _2} = i\hbar {g^2}{\bar{a}_2}\bar{a}_1^ \ast $, $\tau (x )= \sigma (x )+ {{{F_1}} / {D_1^ \ast (x )}} - {{{F_2}} / {D_2^ \ast (x )}}$.

The second group corresponds to the SSG process. The output amplitude of the frequency components at sum sideband and the amplitude of the MR are as follows:

$$\begin{aligned} &A_s^ +{=} \frac{{[{{Z_1}({{\Omega _ + }} ){Q_2}({{\Omega _ + }} )- {\beta_1}{Z_2}({{\Omega _ + }} )} ]\tau ({{\Omega _ + }} )}}{{[{{Q_1}({{\Omega _ + }} ){Q_2}({{\Omega _ + }} )+ {\beta_1}{\beta_2}} ]}},\\ &A_s^ -{=} \frac{{ig({{{\bar{a}}_1}x_s^ \ast{+} x_1^ \ast A_2^ -{+} x_2^ \ast A_1^ - } )}}{{{D_1}({{\Omega _ + }} )}},\\ &B_s^ +{=} \frac{{{\beta _2}A_s^ +{+} {Z_2}({{\Omega _ + }} )\tau ({{\Omega _ + }} )}}{{{Q_2}({{\Omega _ + }} )}},\\ &B_s^ -{=} \frac{{ig({{{\bar{a}}_2}x_s^ \ast{+} x_1^ \ast B_2^ -{+} x_2^ \ast B_1^ - } )}}{{{D_2}({{\Omega _ + }} )}},\\ &{x_s} = \frac{{\hbar g[{({{\xi_2}({{\Omega _ + }} )- {\xi_1}({{\Omega _ + }} )} )- ({\bar{a}_2^ \ast B_s^ +{-} \bar{a}_1^ \ast A_s^ + } )} ]}}{{\tau ({{\Omega _ + }} )}}, \end{aligned}$$
where ${Q_1}(x )= D_1^ \ast (x )\tau (x )- {F_1}$, ${Q_2}(x )= D_2^ \ast (x )\tau (x )+ {F_2}$, ${Z_1}(x )= \frac{{i\hbar {g^2}{{\bar{a}}_1}[{{\xi_2}(x )- {\xi_1}(x )} ]}}{{\tau (x )}} + ig({{x_1}A_2^ +{+} {x_2}A_1^ + } )$, ${Z_2}(x )= \frac{{i\hbar {g^2}{{\bar{a}}_2}[{{\xi_2}(x )- {\xi_1}(x )} ]}}{{\tau (x )}} + ig({{x_1}B_2^ +{+} {x_2}B_1^ + } )$, ${\xi _1}(x )= {({B_1^ - } )^ \ast }B_2^ +{+} {({B_2^ - } )^ \ast }B_1^ +{+} ig{{\bar{a}}_2}[{{x_1}{{({B_2^ - } )}^ \ast } + {x_2}{{({B_1^ - } )}^ \ast }}]/$ ${D_2^ \ast (x )}$, ${\xi _2}(x )= {({A_1^ - } )^ \ast }A_2^ +{+} {({A_2^ - } )^ \ast }A_1^ +{+} {{ig{{\bar{a}}_1}[{{x_1}{{({A_2^ - } )}^ \ast } + {x_2}{{({A_1^ - } )}^ \ast }} ]} / {D_1^ \ast (x )}}$.

Furthermore, based on the input-output relation [56], the output field of this double-cavity system can be given as follows:

$$\begin{aligned} &{s_{out}} = \left( {{\varepsilon_c} - \sqrt {\eta \kappa } {{\bar{a}}_1}} \right) + \left( {{\varepsilon_1} - \sqrt {\eta \kappa } A_1^ + } \right){e^{ - i{\delta _1}t}} - \sqrt {\eta \kappa } A_1^ - {e^{i{\delta _1}t}} + \left( {{\varepsilon_2} - \sqrt {\eta \kappa } A_2^ + } \right){e^{ - i{\delta _2}t}}\\ &- \sqrt {\eta \kappa } A_2^ - {e^{i{\delta _2}t}} - \sqrt {\eta \kappa } A_s^ + {e^{ - i{\Omega _ + }t}} - \sqrt {\eta \kappa } A_s^ - {e^{i{\Omega _ + }t}}. \end{aligned}$$

The term $\left( {{\varepsilon_c} - \sqrt {\eta \kappa } {{\bar{a}}_1}} \right)$ represents the frequency spectra of the pump field with ${\omega _c}$. $- \sqrt {\eta \kappa } A_j^ - {e^{i{\delta _j}t}}$ and $\left( {{\varepsilon_j} - \sqrt {\eta \kappa } A_j^ + } \right){e^{ - i{\delta _j}t}}({j = 1,2} )$ express the Stokes and anti-Stokes fields, respectively. Furthermore, the terms $- \sqrt {\eta \kappa } A_s^ + {e^{ - i{\Omega _ + }t}}$ and $- \sqrt {\eta \kappa } A_s^ - {e^{i{\Omega _ + }t}}$ describe the output fields with frequencies ${\omega _c} + {\Omega _ + }$ and ${\omega _c} - {\Omega _ + }$, which are correlated with the upper and lower sum sideband separately. As shown in Fig. 1(b), the upper sum sideband can be generated when the pump and probe fields satisfy certain detuning conditions ${\Delta _1} = {\omega _m}$, ${\Delta _2} ={-} {\omega _m}$ and ${\delta _1} + {\delta _2} = {\omega _m}$ with the cavity field. And the lower sum sideband is generated at ${\delta _1} + {\delta _2} ={-} {\omega _m}$.

3. Results and discussion

Here, we only consider sum sideband effects. The efficiency of the upper (lower) sum sideband can be given by $\eta _s^ +{=} \left|{{{ - \sqrt {\eta \kappa } A_s^ + } / {{\varepsilon_1}}}} \right|\left( {\eta_s^ -{=} \left|{{{ - \sqrt {\eta \kappa } A_s^ - } / {{\varepsilon_1}}}} \right|} \right)$, which represents the proportion between the amplitudes of the upper (lower) sum sideband and the first probe field. The specific parameters are as follows [47]: $m = 20ng $, ${G / {2\mathrm{\pi }}} = {{ - 12.0\textrm{GHz}} / {\textrm{nm}}}$, ${{{\gamma _m}} / {2\mathrm{\pi }}} = 41.0\textrm{kHz}$, ${\kappa / {\mathrm{2\pi }}} = 15.0\textrm{MHz}$, ${{{\omega _m}} / {\mathrm{2\pi }}} = 51.8\textrm{MHz}$, ${\textrm{P}_\textrm{1}} = {\textrm{P}_\textrm{2}} = \textrm{1}\;{\mathrm{\mu} \mathrm{W}}$, ${\lambda _c} = 532\textrm{nm}$.

In order to illustrate the properties of SSG under double radiation pressure, Figs. 2(a) and 2(b) interpret the first probe field frequency ${\delta _1}$ and red detuned pump field power ${P_\textrm{c}}$ versus the efficiency (in logarithmic form) of upper (lower) sum sideband generation. It is shown that from Fig. 2 the efficiencies of both USSG and LSSG increases with the enhancement of the pump field power ${P_c}$. It can be clearly seen that in this double-cavity system, the efficiency of SSG is several orders of magnitude higher than that in general optomechanical system [47]. As displayed in Fig. 2(a), there is only one peak in the USSG, which reaches its maximum only at ${\delta _1} = {\omega _m}$ and generates a new absorption peak corresponding to the position of ${\delta _1} = 1.05{\omega _m}$. The physical explanation of this phenomenon is that the anti-Stokes scattering of the strongly red detuned pump field leads to its transparency around the cavity resonance [57]. In contrast, as depicted in Fig. 2(b), there are two peaks in the LSSG and the point positions corresponding to the peaks are ${\delta _1} = {\omega _m}$ and ${\delta _1} = 1.05{\omega _m}$, respectively. And the peak value at ${\delta _1} = {\omega _m}$ is significantly higher than that at ${\delta _1} = 1.05{\omega _m}$. The physical reason for their different behavior is that the latter is more separated from the detuning conditions of the cavity ${\Delta _1} = {\omega _m}$ [47]. The particular value of ${\delta _1} = 1.05{\omega _m}$ is attributed to ${\delta _1} + {\delta _2} = {\omega _m}$, which is called the matching condition. The efficiencies of USSG and LSSG as a function of the detuning frequency of the first probe field are shown in Fig. 2(c). It can be seen that a transparent window appearing in the USSG corresponds to the absorption peak in Fig. 2(a). Figure 2(d) depicts the amplitude of the MR at the sum sideband versus the detuning frequencies ${\delta _1}$ and ${\delta _2}$. It has been shown that ${x_s}$ becomes prominent on the two lines of ${\delta _1} + {\delta _2} ={\pm} {\omega _m}$ and at two points of $({{\delta_1},{\delta_2}} )= ({0, \pm {\omega_m}} )$. In particular, the amplitude of MR ${x_s}$ is especially remarkable at the four points $({{\delta_1},{\delta_2}} )= ({ \pm {\omega_m},0} )$, $({{\delta_1},{\delta_2}} )= ({ - {\omega_m},{\omega_m}} )$ and $({{\delta_1},{\delta_2}} )= ({{\omega_m}, - {\omega_m}} )$, which is about 7 fm. In Fig. 2, the amplitude of the USSG in this double-cavity system is approximately two orders of magnitude than a typical optomechanical system, while the LSSG is increased by around three orders of magnitude.

 figure: Fig. 2.

Fig. 2. The diagram of the efficiencies (in logarithmic form) of (a) USSG and (b) LSSG versus pump power ${P_\textrm{c}}$ and detuning frequency ${\delta _1}$ for ${\delta _2} ={-} 0.05{\omega _m}$. (c) Plot of the efficiency of upper (lower) sum sideband generation vs ${\delta _1}$ with ${P_\textrm{c}} = 20{\mathrm{\mu} \mathrm{W}}$. (d) Color graph of the amplitude of the MR (in femtometers) as a function of ${\delta _1}$ and ${\delta _2}$ at sum sideband. The other parameters are $m = 20\textrm{ng}$, ${G / {2\pi }} ={-} 12.0\textrm{GHz/nm}$, ${{{\gamma _m}} / {2\pi }} = 41.0\textrm{kHz}$, ${\kappa / {2\pi }} = 15.0\textrm{MHz}$, ${{{\omega _m}} / {2\pi }} = 51.8\textrm{MHz}$, ${P_1} = {P_2} = 1{\mathrm{\mu} \mathrm{W}}$, ${\lambda _c} = 532\textrm{nm}$, ${\Delta _1} ={-} {\omega _m}$, ${\Delta _2} = {\omega _m}$.

Download Full Size | PDF

The dissipation rate of the optomechanical cavity is an important parameter in the cavity optomechanical system. Then we discuss the properties of the upper (lower) sum sideband in the resolved $({\kappa < {\omega_m}} )$ and unresolved $({\kappa > {\omega_m}} )$ sideband regime. Figure 3 plots the efficiency of SSG for different ${\kappa / {{\omega _m}}}$ versus ${\delta _1}$ in the unresolved sideband regime. We can see from Figs. 3(a)–3(c) that the efficiencies of USSG and LSSG increase as the ratio of ${\kappa / {{\omega _m}}}$ rises in a certain range $({{{1 < \kappa } / {{\omega_m} < 5}}} )$. And the USSG and LSSG have the same peak at ${\delta _1} = {\omega _m}$ when ${\kappa / {{\omega _m}}} = 5$. As shown in Fig. 3(d), the amplitude of LSSG is slightly higher than that of USSG at ${\delta _1} = {\omega _m}$ in the case of ${\kappa / {{\omega _m}}} = 20$. Besides, LSSG changed from bimodal to unimodal and the absorption peak of USSG disappeared. The result shows that the efficiency of SSG can be improved in the unresolved sideband regime by choosing an appropriate detuning frequency ${\delta _1}$.

 figure: Fig. 3.

Fig. 3. The efficiency curves of the USSG and LSSG in the unresolved sideband regime vary with ${\delta _1}$, where ${{{\omega _m}} / {2\pi }} = 51.8\textrm{MHz}$. The ratio of the decay rate of the cavity fields versus the frequency of the MR for (a), (b), (c) and (d) is ${\kappa / {{\omega _m}}} = 1$, ${\kappa / {{\omega _m}}} = 2$, ${\kappa / {{\omega _m}}} = 5$ and ${\kappa / {{\omega _m}}} = 20$, respectively. Other parameters are the same as those in Fig. 2.

Download Full Size | PDF

In Fig. 4, the efficiency of SSG for different ratios of ${\kappa / {{\omega _m}}}$ varies with ${\delta _1}$ in the resolved sideband regime. It can be seen from Fig. 4(a), the amplitude of USSG is about two orders of magnitude higher than that of LSSG at ${\delta _1} = {\omega _m}$. However, the LSSG is more efficient at ${\delta _1} = 1.05{\omega _m}$. Figure 4(b) clearly shows that the efficiency of USSG is approximately three orders of magnitude greater than that of LSSG at ${\delta _1} = {\omega _m}$, and the amplitude difference between the USSG and LSSG is decreasing with the ratio of ${\kappa / {{\omega _m}}}$ at ${\delta _1} = 1.05{\omega _m}$. The efficiency difference between USSG and LSSG reaches five orders of magnitude when ${\delta _1} = {\omega _m}$, as shown in Fig. 4(c). In addition, we find that the amplitudes of USSG and LSSG are equal in the case of ${\delta _1} = 1.05{\omega _m}$. The results indicate that the efficiency of the USSG is consistently superior to that of the LSSG when the detuning frequency ${\delta _1} = {\omega _m}$ is satisfied. In parallel, the amplitude difference between USSG and LSSG decreases at ${\delta _1} = {\omega _m}$ and increases at ${\delta _1} = 1.05{\omega _m}$ with the reduce of ${\kappa / {{\omega _m}}}$. With the decrease of the ratio of ${\kappa / {{\omega _m}}}$, the amplitudes of USSG and LSSG are also reducing in the resolved sideband regime, as shown in Figs. 4(a)–4(c). It is notable that the efficiency of SSG decreases significantly in the resolved sideband regime, while the peak structures of USSG and LSSG are almost unchanged.

 figure: Fig. 4.

Fig. 4. The efficiency diagrams of USSG and LSSG in resolved sideband regimes with ${\delta _1}$. Other parameters are the same as those in Fig. 3.

Download Full Size | PDF

To describe the SSG more specifically, Fig. 5 shows the dependence of the efficiency of SSG on the detuning frequency ${\delta _1}({{\delta_2}} )$ of the probe field. As shown in Figs. 5(a) and 5(b), the 3D images show the peak structure of the USSG and LSSG. From Fig. 5(a), we can see that two reinforcement belts similar to mountain range are produced in the USSG. As expected, the USSG is enhanced at ${\delta _1} ={\pm} {\omega _m}$, which corresponds to the matching condition in Fig. 2(d). It is clear from Fig. 5(c) that excluding the four points $({ \pm {\omega_m},0} )$ and $({0, \pm {\omega_m}} )$, the efficiency of USSG is significantly increased at the four straight lines ${\delta _1} ={\pm} {\omega _m}$ and ${\delta _2} ={\pm} {\omega _m}$, which is different from the matching condition of SSG in general optomechanical systems. In comparison, ten sharp peaks are generated for the LSSG in Fig. 5(b), which indicates that the LSSG can be dramatically reinforced at these ten points. The matching condition for the LSSG is ${\delta _1} ={\pm} {\omega _m}$, ${\delta _2} ={\pm} {\omega _m}$, ${\delta _1} + {\delta _2} ={-} {\omega _m}$ and ${\delta _1} + {\delta _2} = {\omega _m}$ as shown in Fig. 5(d). It is worth noting that the maximum efficiency of LSSG can be achieved at these four points $({ - {\omega_m},{\omega_m}} )$, $({{\omega_m}, - {\omega_m}} )$ and $({ \pm {\omega_m},0} )$. The physical interpretation of these phenomena is that these four points are close to the resonance conditions for the maximum amplitude of the MR.

 figure: Fig. 5.

Fig. 5. The efficiencies (in logarithmic form) of USSG and LSSG as a function of detuning frequency ${\delta _1}$ and ${\delta _2}$. Other parameters are the same as those in Fig. 2.

Download Full Size | PDF

4. Conclusion

In summary, we have studied the characteristics of sum sideband generation via double radiation pressure. The results demonstrate that the efficiency of the LSSG is enhanced with more than three orders of magnitude compared to that in the general optomechanical system. Meanwhile, the USSG is improved for two orders of magnitude and its matching conditions are also altered. We also investigate the optimal matching conditions for the SSG in the resolved and unresolved sideband regimes. A threshold condition is determined. Under this condition, the efficiencies of USSG and LSSG tend to the same maximum value, and the maximum amplitude of the LSSG can be increased by approximately five orders of magnitude. Furthermore, the peak structures of USSG and LSSG can be changed under specific parameter conditions, and the LSSG is more efficient than the USSG. This research will deepen the understanding of nonlinear optomechanical interactions in resolved and unresolved sideband regimes and provide a method for quantum information processing and precision measurements.

Funding

Open Research Fund Program of Key Laboratory of Opto-Electronic Information Acquisition and Manipulation of Ministry of Education (OEIAM202004); Interdisciplinary Innovation Fund of Nanchang University (9166-27060003-YB12); Foundation for Distinguished Young Scientists of Jiangxi Province (20162BCB23009); Opening Project of Shanghai Key Laboratory of Special Artificial Microstructure Materials and Technology (ammt2021A-4); Finance Science and Technology Special "contract system" Project of Nanchang University Jiangxi Province (ZBG20230418015); National Natural Science Foundation of China (62061028).

Acknowledgments

This project was supported by the National Natural Science Foundation of China (Grant No. 62061028), the Finance Science and Technology Special “contract system” Project of Nanchang University Jiangxi Province (Grant No. ZBG20230418015), the Opening Project of Shanghai Key Laboratory of Special Artificial Microstructure Materials and Technology (Grant No. ammt2021A-4), the Foundation for Distinguished Young Scientists of Jiangxi Province (Grant No. 20162BCB23009), the Interdisciplinary Innovation Fund of Nanchang University (Grant No. 9166-27060003-YB12), and the Open Research Fund Program of Key Laboratory of Opto-Electronic Information Acquisition and Manipulation of Ministry of Education (Grant No. OEIAM202004).

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

2. M. Aspelmeyer, P. Meystre, and K. Schwab, “Quantum optomechanics,” Phys. Today 65(7), 29–35 (2012). [CrossRef]  

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

4. C. K. Law, “Interaction between a moving mirror and radiation pressure: a Hamiltonian formulation,” Phys. Rev. A 51(3), 2537–2541 (1995). [CrossRef]  

5. C. Genes, D. Vitali, and P. Tombesi, “Emergence of atom-light-mirror entanglement inside an optical cavity,” Phys. Rev. A 77(5), 050307 (2008). [CrossRef]  

6. A. Farace and V. Giovannetti, “Enhancing quantum effects via periodic modulations in optomechanical systems,” Phys. Rev. A 86(1), 013820 (2012). [CrossRef]  

7. H. Shi and M. Bhattacharya, “Quantum mechanical study of a generic quadratically coupled optomechanical system,” Phys. Rev. A 87(4), 043829 (2013). [CrossRef]  

8. J.-Q. Liao, C. K. Law, and L.-M. Kuang, “Enhancement of mechanical effects of single photons in modulated two-mode optomechanics,” Phys. Rev. A 92(1), 013822 (2015). [CrossRef]  

9. A. Forbes, A. Dudley, and M. McLaren, “Creation and detection of optical modes with spatial light modulators,” Adv. Opt. Photonics 8(2), 200–227 (2016). [CrossRef]  

10. S. Weis, R. Rivière, and S. Deléglise, “Optomechanically induced transparency,” Science 330(6010), 1520–1523 (2010). [CrossRef]  

11. H. Jing, Ş. K. Özdemir, and Z. Geng, “Optomechanically-induced transparency in parity-time-symmetric microresonators,” Sci. Rep. 5(1), 9663 (2015). [CrossRef]  

12. H. Zhang, F. Saif, and Y. Jiao, “Loss-induced transparency in optomechanics,” Opt. Express 26(19), 25199–25210 (2018). [CrossRef]  

13. H. Jing, H. Lü, and S. K. Özdemir, “Nanoparticle sensing with a spinning resonator,” Optica 5(11), 1424–1430 (2018). [CrossRef]  

14. H. Lü, Y. Jiang, and Y.-Z. Wang, “Optomechanically induced transparency in a spinning resonator,” Photonics Res. 5(4), 367–371 (2017). [CrossRef]  

15. T.-X. Lu, Y.-F. Jiao, and H.-L. Zhang, “Selective and switchable optical amplification with mechanical driven oscillators,” Phys. Rev. A 100(1), 013813 (2019). [CrossRef]  

16. C. M. Caves, “Quantum-mechanical radiation-pressure fluctuations in an interferometer,” Phys. Rev. Lett. 45(2), 75–79 (1980). [CrossRef]  

17. A. Abramovici, W. E. Althouse, and R. W. P. Drever, “LIGO: The laser interferometer gravitational-wave observatory,” Science 256(5055), 325–333 (1992). [CrossRef]  

18. V. B. Braginsky and S. P. Vyatchanin, “Low quantum noise tranquilizer for Fabry–Perot interferometer,” Phys. Lett. A 293(5-6), 228–234 (2002). [CrossRef]  

19. A. Schliesser, R. Rivière, and G. Anetsberger, “Resolved-sideband cooling of a micromechanical oscillator,” Nat. Phys. 4(5), 415–419 (2008). [CrossRef]  

20. C. Genes, D. Vitali, and P. Tombesi, “Ground-state cooling of a micromechanical oscillator: comparing cold damping and cavity-assisted cooling schemes,” Phys. Rev. A 77(3), 033804 (2008). [CrossRef]  

21. X. Chen, Y.-C. Liu, and P. Peng, “Cooling of macroscopic mechanical resonators in hybrid atom-optomechanical systems,” Phys. Rev. A 92(3), 033841 (2015). [CrossRef]  

22. W.-J. Nie, A. Chen, and Y. Lan, “Cooling mechanical motion via vacuum effect of an ensemble of quantum emitters,” Opt. Express 23(24), 30970–30984 (2015). [CrossRef]  

23. D.-G. Lai, F. Zou, and B.-P. Hou, “Simultaneous cooling of coupled mechanical resonators in cavity optomechanics,” Phys. Rev. A 98(2), 023860 (2018). [CrossRef]  

24. J.-Q. Zhang, Y. Li, and M. Feng, “Precision measurement of electrical charge with optomechanically induced transparency,” Phys. Rev. A 86(5), 053806 (2012). [CrossRef]  

25. H. Xiong, L.-G. Si, and X.-Y. Lü, “Carrier-envelope phase-dependent effect of high-order sideband generation in ultrafast driven optomechanical system,” Opt. Lett. 38(3), 353–355 (2013). [CrossRef]  

26. J.-J. Li and K.-D. Zhu, “Spin-based optomechanics with carbon nanotubes,” Sci. Rep. 2(1), 903 (2012). [CrossRef]  

27. Q. Wang, J.-Q. Zhang, and P.-C. Ma, “Precision measurement of the environmental temperature by tunable double optomechanically induced transparency with a squeezed field,” Phys. Rev. A 91(6), 063827 (2015). [CrossRef]  

28. J. Li, S. Zhang, and R. Yu, “Enhanced optical nonlinearity and fiber-optical frequency comb controlled by a single atom in a whispering-gallery-mode microtoroid resonator,” Phys. Rev. A 90(5), 053832 (2014). [CrossRef]  

29. C. Cao, S.-C. Mi, and Y.-P. Gao, “Tunable high-order sideband spectra generation using a photonic molecule optomechanical system,” Sci. Rep. 6(1), 1–8 (2016). [CrossRef]  

30. W.-X. Yang, A.-X. Chen, and X.-T. Xie, “Enhanced generation of higher-order sidebands in a single-quantum-dot–cavity system coupled to a PT-symmetric double cavity,” Phys. Rev. A 96(1), 013802 (2017). [CrossRef]  

31. C. Kong, S. Li, and C. You, “Two-color second-order sideband generation in an optomechanical system with a two-level system,” Sci. Rep. 8(1), 1060 (2018). [CrossRef]  

32. H. Lü, C. Wang, and L. Yang, “Optomechanically induced transparency at exceptional points,” Phys. Rev. Appl. 10(1), 014006 (2018). [CrossRef]  

33. K. Ullah, H. Jing, and F. Saif, “Multiple electromechanically-induced-transparency windows and Fano resonances in hybrid nano-electro-optomechanics,” Phys. Rev. A 97(3), 033812 (2018). [CrossRef]  

34. R.-J. Xiao, G.-X. Pan, and L. Zhou, “Multiple optomechanically induced transparency in a ring cavity optomechanical system assisted by atomic media,” Int. J. Theor. Phys. 54(10), 3665–3675 (2015). [CrossRef]  

35. X. Xiao, Q.-H. Liao, and N.-R. Zhou, “Tunable optical second-order sideband effects in a parity-time symmetric optomechanical system,” Sci. China: Phys., Mech. Astron. 63(11), 114211 (2020). [CrossRef]  

36. H. Suzuki, E. Brown, and R. Sterling, “Nonlinear dynamics of an optomechanical system with a coherent mechanical pump: second-order sideband generation,” Phys. Rev. A 92(3), 033823 (2015). [CrossRef]  

37. B. Chen, L.-D. Wang, and J. Zhang, “Second-order sideband effects mediated by microwave in hybrid electro-optomechanical systems,” Phys. Lett. A 380(5-6), 798–802 (2016). [CrossRef]  

38. Q.-H. Liao, M.-L. Song, and W.-D. Bao, “Generation of second-order sideband and slow-fast light effects in a PT-symmetric optomechanical system,” Chaos, Solitons Fractals 166, 112978 (2023). [CrossRef]  

39. L.-D. Wang, J.-K. Yan, and X.-F. Zhu, “Tunable second-order sideband effects in a three-mode optomechanical system containing a single quantum well,” Phys. E (Amsterdam, Neth.) 89, 134–138 (2017). [CrossRef]  

40. Y.-F. Jiao, T.-X. Lu, and H. Jing, “Optomechanical second-order sidebands and group delays in a Kerr resonator,” Phys. Rev. A 97(1), 013843 (2018). [CrossRef]  

41. H. Xiong, L.-G. Si, and A.-S. Zheng, “Higher-order sidebands in optomechanically induced transparency,” Phys. Rev. A 86(1), 013815 (2012). [CrossRef]  

42. Y. Jiao, H. Lü, and J. Qian, “Nonlinear optomechanics with gain and loss: amplifying higher-order sideband and group delay,” New J. Phys. 18(8), 083034 (2016). [CrossRef]  

43. J. Li, J. Li, and Q. Xiao, “Giant enhancement of optical high-order sideband generation and their control in a dimer of two cavities with gain and loss,” Phys. Rev. A 93(6), 063814 (2016). [CrossRef]  

44. C. Kong, H. Xiong, and Y. Wu, “Coulomb-interaction-dependent effect of high-order sideband generation in an optomechanical system,” Phys. Rev. A 95(3), 033820 (2017). [CrossRef]  

45. L.-G. Si, L.-X. Guo, and H. Xiong, “Tunable high-order-sideband generation and carrier-envelope-phase–dependent effects via microwave fields in hybrid electro-optomechanical systems,” Phys. Rev. A 97(2), 023805 (2018). [CrossRef]  

46. F. Monifi, J. Zhang, and Ş. K. Özdemir, “Optomechanically induced stochastic resonance and chaos transfer between optical fields,” Nat. Photonics 10(6), 399–405 (2016). [CrossRef]  

47. H. Xiong, L.-G. Si, and X.-Y. Lü, “Optomechanically induced sum sideband generation,” Opt. Express 24(6), 5773–5783 (2016). [CrossRef]  

48. H. Xiong, Y.-W. Fan, and X. Yang, “Radiation pressure induced difference-sideband generation beyond linearized description,” Appl. Phys. Lett. 109(6), 061108 (2016). [CrossRef]  

49. H. Xiong, Z.-X. Liu, and Y. Wu, “Highly sensitive optical sensor for precision measurement of electrical charges based on optomechanically induced difference-sideband generation,” Opt. Lett. 42(18), 3630–3633 (2017). [CrossRef]  

50. H. Xiong, L.-G. Si, and Y. Wu, “Precision measurement of electrical charges in an optomechanical system beyond linearized dynamics,” Appl. Phys. Lett. 110(17), 171102 (2017). [CrossRef]  

51. B. Wang, Z.-X. Liu, and H. Xiong, “Highly sensitive mass sensing by means of the optomechanical nonlinearity,” IEEE Photonics J. 10(6), 1–8 (2018). [CrossRef]  

52. X.-Y. Wang, L.-G. Si, and X.-H. Lu, “Generation and enhancement of sum sideband in a quadratically coupled optomechanical system with parametric interactions,” Opt. Express 27(20), 29297–29308 (2019). [CrossRef]  

53. H. Xiong, Y.-M. Huang, and Y. Wu, “Laguerre-Gaussian optical sum-sideband generation via orbital angular momentum exchange,” Phys. Rev. A 103(4), 043506 (2021). [CrossRef]  

54. X.-B. Yan, W.-Z. Jia, and Y. Li, “Optomechanically induced amplification and perfect transparency in double-cavity optomechanics,” Front. Phys. 10(3), 351–357 (2015). [CrossRef]  

55. S. M. Spillane, T. J. Kippenberg, and O. J. Painter, “Ideality in a fiber-taper-coupled microresonator system for application to cavity quantum electrodynamics,” Phys. Rev. Lett. 91(4), 043902 (2003). [CrossRef]  

56. J. D. Thompson, B. M. Zwickl, A. M. Jayich, F. Marquardt, S. M. Girvin, and J. G. E. Harris, “Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane,” Nature 452(7183), 72–75 (2008). [CrossRef]  

57. X.-H. Lu, L.-G. Si, and X.-Y. Wang, “Exceptional points enhance sum sideband generation in a mechanical PT-symmetric system,” Opt. Express 29(4), 4875–4886 (2021). [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 (5)

Fig. 1.
Fig. 1. (a) Schematic diagram of a double-cavity optomechanical system formed by a MR embedded in the middle of two fixed lenses. The opposite sides of the double-cavity optomechanical system are driven by a pump field 1 with frequency ${\omega _\textrm{c}}$ and a pump field 2 with frequency ${\omega _d}$, respectively. Two relatively weak probe fields with frequencies ${\omega _1}$ and ${\omega _2}$ are injected into the left optical cavity. Both cavities have the same cavity length L and frequency ${\omega _0}$ under static, while coupling with MR separately. (b) Frequency spectrogram of sum sideband generation in a double-cavity optomechanical system driven by dual pump fields and dual probe fields. ${\Delta _1} = {\omega _0} - {\omega _\textrm{c}}$, ${\Delta _2} = {\omega _0} - {\omega _d}$, ${\delta _1} = {\omega _1} - {\omega _\textrm{c}}$, ${\delta _2} = {\omega _2} - {\omega _\textrm{c}}$.
Fig. 2.
Fig. 2. The diagram of the efficiencies (in logarithmic form) of (a) USSG and (b) LSSG versus pump power ${P_\textrm{c}}$ and detuning frequency ${\delta _1}$ for ${\delta _2} ={-} 0.05{\omega _m}$. (c) Plot of the efficiency of upper (lower) sum sideband generation vs ${\delta _1}$ with ${P_\textrm{c}} = 20{\mathrm{\mu} \mathrm{W}}$. (d) Color graph of the amplitude of the MR (in femtometers) as a function of ${\delta _1}$ and ${\delta _2}$ at sum sideband. The other parameters are $m = 20\textrm{ng}$, ${G / {2\pi }} ={-} 12.0\textrm{GHz/nm}$, ${{{\gamma _m}} / {2\pi }} = 41.0\textrm{kHz}$, ${\kappa / {2\pi }} = 15.0\textrm{MHz}$, ${{{\omega _m}} / {2\pi }} = 51.8\textrm{MHz}$, ${P_1} = {P_2} = 1{\mathrm{\mu} \mathrm{W}}$, ${\lambda _c} = 532\textrm{nm}$, ${\Delta _1} ={-} {\omega _m}$, ${\Delta _2} = {\omega _m}$.
Fig. 3.
Fig. 3. The efficiency curves of the USSG and LSSG in the unresolved sideband regime vary with ${\delta _1}$, where ${{{\omega _m}} / {2\pi }} = 51.8\textrm{MHz}$. The ratio of the decay rate of the cavity fields versus the frequency of the MR for (a), (b), (c) and (d) is ${\kappa / {{\omega _m}}} = 1$, ${\kappa / {{\omega _m}}} = 2$, ${\kappa / {{\omega _m}}} = 5$ and ${\kappa / {{\omega _m}}} = 20$, respectively. Other parameters are the same as those in Fig. 2.
Fig. 4.
Fig. 4. The efficiency diagrams of USSG and LSSG in resolved sideband regimes with ${\delta _1}$. Other parameters are the same as those in Fig. 3.
Fig. 5.
Fig. 5. The efficiencies (in logarithmic form) of USSG and LSSG as a function of detuning frequency ${\delta _1}$ and ${\delta _2}$. Other parameters are the same as those in Fig. 2.

Equations (11)

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

H = Δ 1 a 1 a 1 + Δ 2 a 2 a 2 + g ( a 2 a 2 a 1 a 1 ) x + ( p 2 2 m + 1 2 m ω m 2 x 2 ) + i η κ ε c ( a 1 a 1 ) + i η κ ε d ( a 2 a 2 ) + i η κ [ ( a 1 ε 1 e i δ 1 t + a 1 ε 2 e i δ 2 t ) H .c . ]
a ˙ 1 = [ i ( Δ 1 g x ) + κ ] a 1 + η κ ε c + η κ ( ε 1 e i δ 1 t + ε 2 e i δ 2 t ) , a ˙ 2 = [ i ( Δ 2 + g x ) + κ ] a 2 + η κ ε d , x ˙ = p / m , p ˙ = g ( a 2 a 2 a 1 a 1 ) m ω m 2 x γ m p
a ˙ 1 = [ i ( Δ 1 g x ) + κ ] a 1 + η κ ε c + η κ ( ε 1 e i δ 1 t + ε 2 e i δ 2 t ) , a ˙ 2 = [ i ( Δ 2 + g x ) + κ ] a 2 + η κ ε d , ( m d 2 d t 2 + m γ m d d t + m ω m 2 ) x = g ( a 2 a 2 a 1 a 1 ) .
a ¯ 1 = η κ ε c i Δ 3 + κ , a ¯ 2 = η κ ε d i Δ 4 + κ , x ¯ = g ( | a ¯ 2 | 2 | a ¯ 1 | 2 ) m ω m 2
d d t δ a 1 = i g ( a 1 ¯ δ x + δ x δ a 1 ) ( i Δ 3 + k ) δ a 1 + η κ ( ε 1 e i δ 1 t + ε 2 e i δ 2 t ) , d d t δ a 2 = i g ( a 2 ¯ δ x + δ x δ a 2 ) ( i Δ 4 + k ) δ a 2 , ( m d 2 d t 2 + m γ m d d t + m ω m 2 ) δ x = g [ a 2 ¯ δ a 2 a 1 ¯ δ a 1 + a 2 ¯ δ a 2 a 1 ¯ δ a 1 + δ a 2 δ a 2 δ a 1 δ a 1 ] .
δ a 1 = A 1 + e i δ 1 t + A 1 e i δ 1 t + A 2 + e i δ 2 t + A 2 e i δ 2 t + A s + e i Ω + t + A s e i Ω + t + , δ a 2 = B 1 + e i δ 1 t + B 1 e i δ 1 t + B 2 + e i δ 2 t + B 2 e i δ 2 t + B s + e i Ω + t + B s e i Ω + t + , δ x = x 1 e i δ 1 t + x 1 e i δ 1 t + x 2 e i δ 2 t + x 2 e i δ 2 t + x s e i Ω + t + x s e i Ω + t + ,
( R 1 i δ 1 ) A 1 + i g a ¯ 1 x 1 η κ ε 1 = 0 , ( R 1 + i δ 1 ) A 1 i g a ¯ 1 x 1 = 0 , ( R 2 i δ 1 ) B 1 + i g a ¯ 2 x 1 = 0 , ( R 2 + i δ 1 ) B 1 i g a ¯ 2 x 1 = 0 , σ ( δ 1 ) x 1 g [ a ¯ 2 B 1 + a ¯ 1 A 1 + + a ¯ 2 ( B 1 ) a ¯ 1 ( A 1 ) ] = 0 , ( R 1 i δ 1 ) A 2 + i g a ¯ 1 x 2 η κ ε 2 = 0 , ( R 1 + i δ 1 ) A 2 i g a ¯ 1 x 2 = 0 , ( R 2 i δ 1 ) B 2 + i g a ¯ 2 x 2 = 0 , ( R 2 + i δ 1 ) B 2 i g a ¯ 2 x 2 = 0 , σ ( δ 1 ) x 2 g [ a ¯ 2 B 2 + a ¯ 1 A 2 + + a ¯ 2 ( B 2 ) a ¯ 1 ( A 2 ) ] = 0 ,
( R 1 i Ω + ) A s + i g a ¯ 1 x s i g ( x 1 A 2 + + x 2 A 1 + ) = 0 , ( R 1 + i Ω + ) A s i g a ¯ 1 x s i g ( x 1 A 2 + x 2 A 1 ) = 0 , ( R 2 i Ω + ) B s + i g a ¯ 2 x s i g ( x 1 B 2 + + x 2 B 1 + ) = 0 , ( R 2 + i Ω + ) B s i g a ¯ 2 x s i g ( x 1 B 2 + x 2 B 1 ) = 0 , σ ( Ω + ) x s = g [ a ¯ 2 B s + a ¯ 1 A s + + ( B 1 ) B 2 + ( A 1 ) A 2 + + ( B 2 ) B 1 + ( A 2 ) A 1 + ] + g [ a ¯ 2 ( B s ) a ¯ 1 ( A s ) ] .
A j + = η κ ε j τ 2 ( δ j ) θ 2 ( δ j ) θ 1 ( δ j ) θ 2 ( δ j ) τ 2 ( δ j ) β 1 β 2 ,   A j = i g a ¯ 1 x j D 1 ( δ j ) , B j + = η κ ε j β 2 τ ( δ j ) θ 1 ( δ j ) θ 2 ( δ j ) τ 2 ( δ j ) β 1 β 2 ,   B j = i g a ¯ 2 x j D 2 ( δ j ) , x j = g ( a ¯ 2 B j + a ¯ 1 A j + ) τ ( δ j )
A s + = [ Z 1 ( Ω + ) Q 2 ( Ω + ) β 1 Z 2 ( Ω + ) ] τ ( Ω + ) [ Q 1 ( Ω + ) Q 2 ( Ω + ) + β 1 β 2 ] , A s = i g ( a ¯ 1 x s + x 1 A 2 + x 2 A 1 ) D 1 ( Ω + ) , B s + = β 2 A s + + Z 2 ( Ω + ) τ ( Ω + ) Q 2 ( Ω + ) , B s = i g ( a ¯ 2 x s + x 1 B 2 + x 2 B 1 ) D 2 ( Ω + ) , x s = g [ ( ξ 2 ( Ω + ) ξ 1 ( Ω + ) ) ( a ¯ 2 B s + a ¯ 1 A s + ) ] τ ( Ω + ) ,
s o u t = ( ε c η κ a ¯ 1 ) + ( ε 1 η κ A 1 + ) e i δ 1 t η κ A 1 e i δ 1 t + ( ε 2 η κ A 2 + ) e i δ 2 t η κ A 2 e i δ 2 t η κ A s + e i Ω + t η κ A s e i Ω + t .
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.