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

Generation and enhancement of sum sideband in a quadratically coupled optomechanical system with parametric interactions

Open Access Open Access

Abstract

We investigate theoretically the generation and enhancement of sum sideband in a quadratically coupled optomechanical system with parametric interactions. It is shown that the generation of frequency components at the sum sideband stems from the nonlinear optomechanical interactions via two-phonon processes in the quadratically coupled optomechanical system, while an optical parametric amplifier (OPA) inside the system can considerably improve the sum sideband generation (SSG). The dependence of SSG on the system parameters, including the power of the control field, the frequency detuning of the probe fields and the nonlinear gain of OPA are analyzed in detail. Our analytic calculation indicates that the SSG can be obtained even under weak driven fields and greatly enhanced via meeting the matching conditions. The effect of SSG may have potential applications for achieving measurement of electric charge (or other weak forces) with higher precision and on-chip manipulation of light propagation.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Benefiting from the rapid development of nanotechnology in recent years, cavity optomechanics as the cross field of nanophysics and quantum optics has received extensive attention and undergone rapid growth. The field of cavity optomechanics, which studies the nonlinear coupling interaction between the mechanical vibration and the optical cavity field via radiation pressure [1], has a wide range of applications in gravitational-wave detection [2], preparation of macroscopic-scale quantum entanglement [3,4], and monitoring of mechanical motion [5]. As is well known, in a coupled optomechanical system how the mechanical vibration interacts to the optical cavity field depends on the position of the mechanical resonator (MR) in the optical cavity field [6]. When the position of the MR is at the nodes of the cavity modes, the coupling between the mechanical vibration and the optical cavity field is linear [710]. Such the so-called linearly coupled optomechanical system has been extensively studied both experimentally [1113] and theoretically [1418] over the past few years, in which many interesting phenomena and effects such as optomechanically induced transparency (OMIT) [1923], slow light [2428], cooling of mechanical resonators [2931] and $\mathcal {PT}$-symmetry optomechanics [3235] have been investigated. Conversely, when the position of the MR is placed at the antinodes of the cavity modes in an optomechanical system, the optical cavity field could be quadratically coupled to the mechanical vibration [3639], which is the so-called quadratically coupled optomechanical system [6]. While the underlying physical processes in a quadratically coupled optomechanical system implies two-phonon processes, which are different from that in the linearly coupled optomechanical system in which the underlying physical processes involves one-phonon processes [15,40]. Using the similar procedure to deal with the light-mechanical interaction in the linearly coupled optomechanical system, many interesting phenomena and effects in a quadratically coupled optomechanical system are also verified, with examples such as photon blockade [41], higher-order sidebands generation [42,43] and optical amplification [44,45].

Recently, it is shown that in a linearly coupled optomechanical system driven by one strong control field and two weak probe fields, the output fields at the sum or difference sideband can be generated [46] when the intrinsic nonlinearity arising from optomechanical interactions are taken into account. Subsequently, the potential applications of difference sideband generation to achieve measurement of electric charge (or other weak forces) with higher precision is proposed [47]. Motivated by these works and in consideration of the underlying physical processes in the quadratically coupled optomechanical system being different from that in the linearly coupled optomechanical system, a question arises naturally: Could sum sideband generation (SSG) be obtained when the quadratically coupled optomechanical system is driven by one strong control field and two weak probe fields and the nonlinear terms of the intrinsic nonlinear optomechanical interactions are considered?

On the other hand, it has been reported that the optomechanical interaction can be enhanced in the linearly coupled optomechanical system assisted by an optical parametric amplifier (OPA), which can be used to improve the cooling of the MR [48], the normal mode splitting [49] and the strong mechanical squeezing [50]. Based on these reasons mentioned above, in this work, we propose a scheme to realize the generation and enhancement of sum sideband in a quadratically coupled optomechanical system with parametric interactions. Using experimentally achievable parameters, our results illustrate that not only the output fields at the sum sideband with frequencies $\pm \Omega _{+}$ (in a frame rotating at frequency $\omega _{c}$) can be generated, but also the nonlinear gain of OPA permits us to enhance the efficiencies of SSG. As we expected, even under the condition of weak driving, the existence of OPA makes the efficiencies of SSG significantly enhanced. It is worth noting that in the presence of OPA, the matching conditions of upper sum sideband generation (USSG) are modified. When the nonlinear gain of OPA increases, the linewidth and peak value of the sum sideband window become wider and larger, respectively. In addition, via adjusting the nonlinear gain of OPA, the efficiency of lower sum sideband generation (LSSG) can be enhanced by several orders of magnitude. From the viewpoint of application, enhancement of optomechanically induced SSG may be used for measurement of electric charge (or other weak forces) with higher precision [51] and on-chip manipulation of light propagation. Moreover, the existing mechanism of SSG may be applied to other similar systems, such as artificial molecules [52] and quantum well systems [53].

This paper is structured as follows: In Sec. 2, under the condition of considering the nonlinear terms of the optomechanical dynamics, the Langevin equations and the steady-state solutions of the system are derived. In addition, the amplitude of SSG is given by further analytic calculation with experimentally achievable parameters. In Sec. 3, we analyze the characteristics of SSG and discuss in detail the influence of the system parameters on SSG, especially the effect of the nonlinear gain of OPA on SSG. Finally, a conclusion is summarized in Sec. 4.

2. Physical model and dynamical equation

We consider such a model, a quadratically coupled optomechanical system with a membrane (treated as a MR) and an OPA inserted between two fixed mirrors. And the system is driven by two relatively weak probe fields with the frequencies $\omega _{1}$, $\omega _{2}$ and a strong control field with frequency $\omega _{c}$, as is schematically shown in Fig. 1(a). Generally, we suppose that $\Omega _{m}$ and $R$ are the angular frequency and limited reflectivity of the MR with the effective mass $m$ and the decay rate $\Gamma _{m}$. And $Q=\Omega _{m}/\Gamma _{m}$ is the mechanical quality factor. Moreover, the MR interacts with the environment at the thermal equilibrium temperature $T$. The resonant frequency of the cavity is $\omega _{o}$ which will be obtained when the MR in a balance position in the absence of additional excitation [1].

 figure: Fig. 1.

Fig. 1. (a) Schematic diagram of a quadratic optomechanical system consisting of an OPA and a MR inserted between two fixed mirrors. The system is excited by two relatively weak probe fields and a strong control field. The two probe fields are represented by probe field 1 with frequency $\omega _{1}$ and probe field 2 with frequency $\omega _{2}$, respectively. And $\omega _{c}$ is the frequency of the control field. The MR and cavity field are quadratically coupled via the radiation pressure. (b) Frequency spectrogram of sum sideband generation in the quadratically coupled optomechanical system with a control field and two probe fields driven.

Download Full Size | PDF

In a frame rotating at frequency $\omega _{c}$, the Hamiltonian of our scheme can therefore be written as [57].

$$\begin{aligned} H &=\hbar\Delta\hat{a}^{\dagger}\hat{a}+(\frac{\hat{p}^{2}}{2m}+\frac{1}{2}m\Omega_{m}^{2}\hat{x}^{2})+\hbar g\hat{a}^{\dagger}\hat{a}\hat{x}^2+i\hbar G(\hat{a}^{\dagger 2}e^{i\theta}-\hat{a}^{2}e^{{-}i\theta})+ \nonumber\\ &~~~~~~~~i\hbar\sqrt{\eta\kappa}[(\varepsilon_{c}+\varepsilon_{1}e^{{-}i\Omega_{1} t}+\varepsilon_{2}e^{{-}i\Omega_{2} t})\hat{a}^{\dagger}-H.c.], \end{aligned}$$
where, the first term represents the free Hamiltonian of the cavity field with $\Delta =\omega _{o}-\omega _{c}$ and $\hat {a}^{\dagger }$ ($\hat {a}$) are the detuning of the cavity field and the control field and creation (annihilation) operator. And the creation and annihilation operators satisfy the commutation relation $[\hat {a}, \hat {a}^{\dagger }]=1$. The second term denotes the energy of the MR with $\hat p$ and $\hat x$ are the momentum and position operators, respectively. The third term indicates the interaction between the cavity field and the MR via the radiation pressure with the optomechanical coupling strength $g$. And the quadratic coupling constant is ${g}$=${8\pi ^{2}c/(L\lambda _{c}^2)[R/(1-R)]^{\frac {1}{2}}}$ [6,57] with $c$ being the speed of light in a vacuum, $L$ being the length of the cavity and $\lambda _{c}$ being the wavelength of the control field. The fourth term describes the interaction between the cavity field and OPA. $\theta$ and $G$ are the phase of the field driving OPA and the nonlinear gain of OPA, which can be adjusted by the pump driving [48]. The last term describes the excitation of the external fields to the cavity mode with $\Omega _{1}=\omega _{1}-\omega _{c}$ ($\Omega _{2}=\omega _{2}-\omega _{c}$) being the detuning of the probe field 1 (probe field 2) and the control field. It consists of three fields: one strong control field and two relatively weak probe fields, and the corresponding amplitudes are $\varepsilon _{c}$, $\varepsilon _{1}$ and $\varepsilon _{2}$, respectively. Commonly, the amplitudes of the external driving fields satisfy the relations $\varepsilon _{c}=\sqrt {P_{c}/\hbar \omega _{c}}$ and $\varepsilon _{1(2)}=\sqrt {P_{1(2)}/\hbar \omega _{1(2)}}$. Here, $P_{c}$ and $P_{1(2)}$ are the powers of control field and probe fields, respectively. Besides, the coupling parameter $\eta =\kappa _{ex}/\kappa$ with $\kappa$ being the total decay rate of the cavity which contains an intrinsic decay $\kappa _{o}$ and an external decay rate $\kappa _{ex}$. And the coupling parameter $\eta$ can be continuously adjusted, and we adopt the critical coupling $1/2$ [58] .

The Heisenberg-Langevin equations of our scheme with Hamiltonian (1) can therefore be written as follows [14]:

$$\begin{aligned} &\partial_{t}\hat{x}=\frac{\hat{p}}{m}, \nonumber\\ &\partial_{t}\hat{p}={-}[m\Omega_{m}^{2}+2\hbar g\hat{a}^{\dagger}\hat{a}]\hat{x}-\Gamma_{m}\hat{p}+\hat{F}_{\mathrm{th}}, \nonumber\\ &\partial_{t}\hat{a}={-}[\frac{\kappa}{2}+i(\Delta+g\hat {x}^2)]\hat{a}+2Ge^{i\theta}\hat{a}^{\dagger}+\sqrt{\eta\kappa}(\varepsilon_{c}+\varepsilon_{1}e^{{-}i\Omega_{1} t}+\varepsilon_{2}e^{{-}i\Omega_{2} t})+\hat{a}_{in}, \end{aligned}$$
in which $\hat {F}_{\mathrm {th}}$ being the Langevin force arising from the interaction of the MR on the environment with the average value is zero [1]. And the Langevin force has the relation function $\langle \hat {F}_{\mathrm {th}}(t)\hat {F}_{\mathrm {th}}(t')\rangle =\Gamma _m\Omega _{m}(2\pi )^{-1}\int \exp [-i\omega (t-t')][1+\coth (\hbar \omega /2k_BT)]d\omega$ [37] with $k_B$ being the Boltzmann constant and $T$ being the temperature of the bath of the MR. The quantum noise of the cavity field is expressed by $\hat {a}_{in}$ with $\langle \hat {a}_{in}(t)\rangle =0$, $\langle \hat {a}_{in}^\dagger (t)\hat {a}_{in}(t')\rangle =[\exp (\hbar \omega _{o}/k_BT)-1]^{-1}\delta (t-t')$, $\langle \hat {a}_{in}(t)\hat {a}_{in}^\dagger (t')\rangle =\{[\exp (\hbar \omega _{o}/k_BT)-1]^{-1}+1\}\delta (t-t')$ [37].

Similar to the linearly coupled optomechanical system, we mainly study the mean response of the system to the probe fields in the quadratically coupled optomechanical system. However, we can obtain from Eqs. (2) that the mean values of $\hat {x}$ and $\hat {p}$ are zero which is the difference between the linearly coupled optomechanical system and the quadratically coupled optomechanical system. By further analyzing the evolution equations of the MR displacement and the cavity field amplitude in Eqs. (2), we can know that the MR displacement is decoupled from the cavity field amplitude in steady state. In other words, the mean value of the MR displacement has no effect on the output field. The reason for this feature is that the underlying physical process of a quadratically coupled optomechanical system is a two-phonon process [57]. Consequently, in the following works, we concentrate on calculating the evolution of the mean values of $\hat {a}$, $\hat {x}^2$, $\hat {p}^2$, and $\hat {x}\hat {p}+\hat {p}\hat {x}$, viz., $a(t)\equiv \langle \hat {a}(t)\rangle$, $X(t)\equiv \langle \hat {x}^2(t)\rangle$, $P(t)\equiv \langle \hat {p}^2(t)\rangle$, and $Q(t)\equiv \langle \hat {x}(t)\hat {p}(t)+\hat {p}(t)\hat {x}(t)\rangle$.

Using a so-called mean-field approximation, viz., $\langle abc\rangle =\langle a\rangle \langle b\rangle \langle c\rangle$ [1], the mean value equations can be written in the following forms:

$$\begin{aligned} &\partial_{t}X=\frac{Q}{m}, \nonumber\\ &\partial_{t}P={-}[m\Omega_{m}^{2}+2\hbar g a^{*}a]Q+\Gamma_{m}(1+2n)\hbar m\Omega_{m} -2\Gamma_{m}P, \nonumber\\ &\partial_{t}Q={-}[4\hbar ga^{*}a+2m\Omega_{m}^{2}]X-\Gamma_{m}Q+\frac{2P}{m}, \nonumber\\ &\partial_{t}a={-}[\frac{\kappa}{2}+i(\Delta+gX)]a+2Ge^{i\theta}\hat{a}^{*}+\sqrt{\eta\kappa}(\varepsilon_{c}+\varepsilon_{1}e^{{-}i\Omega_{1} t}+\varepsilon_{2}e^{{-}i\Omega_{2} t}), \end{aligned}$$
in which the constant $\Gamma _{m}(1+2n)\hbar m\Omega _{m}$ is arose from the interaction between the MR and the thermal environment with $n$=$[\exp (\hbar \Omega _m/k_BT)-1]^{-1}$ being the mean phonon occupation of energy $\hbar \Omega _{m}$ at temperature $T$. On the basis that the control field is much stronger than the probe fields, we adopt the perturbation method to deal with Eqs. (3). Generally, the steady-state solution of the system is provided by the control field, and the probe fields can be regarded as the perturbation of the control field. Therefore, the solution of Eqs. (3) can be written in such a form, $\Theta (t)=\bar {\Theta }+\delta \Theta$ with $\bar {\Theta }$ and $\delta \Theta$ are the mean value and perturbation term of each operator, where $\Theta (t)$ expresses any one of these quantities $a(t)$, $X(t)$, $P(t)$, or $Q(t)$. By inserting these expressions into Eqs. (3), in addition to first-order terms in the small quantities $\delta \Theta$, nonlinear terms such as $\delta a\delta X$, $\delta a\delta Q$, $\delta a^*\delta a$, $\delta a^*\delta a\delta X$ and $\delta a^*\delta a\delta Q$ will naturally appear in the time-evolution equations. And studies have shown that in a linearly coupled optomechanical system the similar nonlinear terms play a significant role in the generation of sum sideband [46], higher-order sidebands [54] and chaos [55]. Thus we analyze that it is not sufficiently to consider only the first-order terms of the optomechanical dynamics to generate sum sideband, and we should also consider the nonlinear terms for the SSG in the quadratically coupled optomechanical system. Therefore, in the perturbative region, we use the following ansatzs to analyze the SSG in the system [46]:
$$\begin{aligned} &{\delta}a=A_{{+}1}e^{{-}i\Omega_1 t} + A_{{-}1}e^{i\Omega_1 t}+A_{{+}2}e^{{-}i\Omega_2 t}+ A_{{-}2}e^{i\Omega_2 t}+A_{{+}s}e^{{-}i\Omega_+ t} + A_{{-}s}e^{i\Omega_+ t}, \nonumber\\ &{\delta}a^\ast{=}A_{{-}1}^\ast e^{{-}i\Omega_1 t} + A_{{+}1}^\ast e^{i\Omega_1 t}+A_{{-}2}^\ast e^{{-}i\Omega_2 t}+A_{{+}2}^\ast e^{i\Omega_2 t}+A_{{-}s}^\ast e^{{-}i\Omega_+ t}+A_{{+}s}^\ast e^{i\Omega_+ t}, \nonumber\\ &{\delta}Q=Q_1e^{{-}i\Omega_1 t} + Q_1^\ast e^{i\Omega_1 t}+Q_2e^{{-}i\Omega_2 t}+ Q_2^\ast e^{i\Omega_2 t}+Q_se^{{-}i\Omega_+ t}+Q_s^\ast e^{i\Omega_+ t}, \nonumber\\ &{\delta}X=X_1e^{{-}i\Omega_1 t} + X_1^\ast e^{i\Omega_1 t}+X_2e^{{-}i\Omega_2 t}+ X_2^\ast e^{i\Omega_2 t}+X_se^{{-}i\Omega_+ t}+X_s^\ast e^{i\Omega_+ t}, \nonumber\\ &{\delta}P=P_1e^{{-}i\Omega_1 t} + P_1^\ast e^{i\Omega_1 t}+P_2e^{{-}i\Omega_2 t}+ P_2^\ast e^{i\Omega_2 t}+P_se^{{-}i\Omega_+ t}+P_s^\ast e^{i\Omega_+ t}, \end{aligned}$$
with $\Omega _+=\Omega _{1}+\Omega _2$. The physical interpretation of the ansatzs of Eqs. (4) is that when the probe fields and the control field coexist, due to we take into account the nonlinear terms, so the output fields with frequencies $\pm \Omega _{+}$ (in a frame rotating at frequency $\omega _{c}$) are generated, which is very similar to sum frequency generation in a nonlinear medium [56]. The output fields with these frequency components are so-called SSG, where the frequency of the upper (lower) sum sideband is $\Omega _{+}$ ($-\Omega _{+}$) in a frame rotating at frequency $\omega _{c}$. It is worth noting that the underlying physical process is a two-phonon process. By inserting these expressions into the Heisenberg-Langevin Eqs. (3), the steady state values of Eqs. (3) can be obtained as follows:
$$\begin{aligned} {\bar{Q}} &= 0, \nonumber\\ {\bar{a}} &= \frac{(\beta^{*}+2Ge^{i\theta})}{\beta \beta^{*}+4G^2}\sqrt{\eta\kappa}\varepsilon_{c}, \nonumber\\ {\bar{X}} &= \frac{\bar{P}}{m^{2}\Omega^{2}_{m}(1+2\alpha)}, \nonumber\\ {\bar{P}} &= (1+2n)\frac{\hbar m\Omega_{m}}{2}, \end{aligned}$$
where $\bar \Delta =\Delta +g\bar {X}$, $\beta =\frac {\kappa }{2}+i{\bar \Delta }$, $\alpha =\frac {\hbar g|{\bar {a}}|^{2}}{m\Omega ^{2}_{m}}$, and the terms $\bar \Delta$ and $g\bar {X}$ are the effective cavity detuning and the frequency shift which stem from the quadratic coupling.

Considering that the process of SSG is a second order process, thus we can get the following equations:

(1) The first group represents the process of the probe field with frequency $\omega _{1}$ ($\omega _{2}$), where we ignore the nonlinear terms of the optomechanical dynamics,

$$\begin{aligned} &(\beta-i\Omega_j)A_{{+}j}={-}ig\bar{a}X_{j}+2Ge^{i\theta}A_{{-}j}^{*}+\sqrt{\eta\kappa}\varepsilon_{j} \nonumber\\ &(\beta+i\Omega_j)A_{{-}j}={-}ig\bar{a}X_{j}^\ast{+}2Ge^{i\theta}A_{{+}j}^{*} \nonumber\\ &(2\Gamma_m-i\Omega_j)P_{j}={-}(m\Omega^{2}_{m}+2\hbar g|{\bar{a}}|^{2})Q_{j} \nonumber\\ &(\Gamma_m-i\Omega_j)Q_{j}={-}4\hbar g[{\bar{a}}^\ast A_{{+}j}+{\bar{a}}A_{{-}j}^\ast]\bar{X} -2m\Omega^{2}_{m}(1+2\alpha)X_{j}+\frac{2P_{j}}{m} \nonumber\\ &Q_{j}+i\Omega_{1}mX_{j}=0. ~~~~~~~~(j=1,2) \end{aligned}$$
(2) The second group represents the process of SSG, where we take into account the nonlinear terms of the optomechanical dynamics,
$$\begin{aligned}&(\beta-i\Omega_+)A_{{+}s}={-}ig(\bar{a}X_{s}+X_{1}A_{{+}2}+X_{2}A_{{+}1})+2Ge^{i\theta}A_{{-}s}^{*} \nonumber\\ &(\beta+i\Omega_+)A_{{-}s}={-}ig(\bar{a}X_{s}^\ast{+}X_{1}^\ast A_{{-}2}+X_{2}^\ast A_{{-}1})+2Ge^{i\theta}A_{{+}s}^{*} \nonumber\\ &(2\Gamma_m-i\Omega_+)P_{s}={-}m\Omega^{2}_{m}(1+2\alpha)Q_{s}-2\hbar g[\bar{a}^\ast(A_{{+}1}Q_{2}+A_{{+}2}Q_{1})+\bar{a}(A_{{-}1}^\ast Q_{2}+A_{{-}2}^\ast Q_{1})] \nonumber\\ &(\Gamma_m-i\Omega_+)Q_{s}={-}4\hbar g[{\bar{a}}^\ast A_{{+}s}+\bar{a}A_{{-}s}^\ast{+}A_{{-}1}^\ast A_{{+}2}+A_{{-}2}^\ast A_{{+}1}]\bar{X}-4\hbar g[{\bar{a}}^\ast(A_{{+}1}X_{2}+A_{{+}2}X_{1}) \nonumber\\ &~~~~~~~~~~~~~~~~~~~~~~~~~~~+\bar{a}(A_{{-}1}^\ast X_{2}+A_{{-}2}^\ast X_{1})]-2m\Omega^{2}_{m}(1+2\alpha)X_{s}+\frac{2P_{s}}{m} \nonumber\\ &Q_{s}+i\Omega_{+}mX_{s}=0. \end{aligned}$$
The Eqs. (6) represent the linear process of the probe field with frequency $\omega _{1}$ ($\omega _{2}$), which has been studied in previous studies and used to study the two-phonon OMIT [57]. The solutions of these equations are:
$$\begin{aligned}&A_{{+}j}=\frac{\tau(\Omega_{j})X_{j}+\sqrt{\eta\kappa}\varepsilon_{j}(\beta^\ast{-}i\Omega_{j})}{s(\Omega_{j})}, \nonumber\\ &X_{j}=\frac{\sqrt{\eta\kappa}\varepsilon_{j}f(\Omega_{j})}{F(\Omega_{j})+d(\Omega_{j})+\sigma(\Omega_{j})+D(\Omega_{j})}, \nonumber\\ &A_{{-}j}=\frac{-ig\bar{a}X_{j}^\ast{+}2Ge^{i\theta}A_{{+}j}^{*}}{\beta+i\Omega_j}, ~~~~~~~~~~(j=1,2) \nonumber\\ &A_{{+}s}=\frac{\tau(\Omega_+)X_s+W}{s(\Omega_+)}, \nonumber\\ &A_{{-}s}=\frac{-ig\bar{a}X_s^\ast{-}ig(X_1^\ast A_{{-}2}+X_2^\ast A_{{-}1})+2Ge^{i\theta}A_{{+}s}^{*}}{\beta+i\Omega_+}, \nonumber\\ &X_s=\frac{f(\Omega_+)W+[\xi-X_2M(\Omega_2)-X_1M(\Omega_1)](\beta^\ast{-}i\Omega_+)}{[F(\Omega_+)+\sigma(\Omega_+)](\beta^\ast{-}i\Omega_+)-\tau(\Omega_+)f(\Omega_+)}, \end{aligned}$$
where
$$\begin{aligned}&s(\Omega_y)=(\beta-i\Omega_y)(\beta^\ast{-}i\Omega_y)-4G^2, \nonumber\\ &\sigma(\Omega_y)=4i\hbar g^2|\bar{a}|^2\bar{X}(2\Gamma_m-i\Omega_y)s(\Omega_y), \nonumber\\ &\tau(\Omega_y)={-}ig[\bar{a}(\beta^\ast{-}i\Omega_y)-2G\bar{a}^\ast e^{i\theta}], \nonumber\\ &d(\Omega_y)=4i\hbar g^2|\bar{a}|^2\bar{X}(2\Gamma_m-i\Omega_y)[-(\beta^\ast{-}i\Omega_y)^2+4G^2], \nonumber\\ &D(\Omega_y)=8i\hbar g^2G\bar{X}(2\Gamma_m-i\Omega_y)(\beta^\ast{-}i\Omega_y)[\bar{a}^{\ast2}e^{i\theta}-\bar{a}^{2}e^{{-}i\theta}],\nonumber\\ &f(\Omega_y)={-}4\hbar g\bar{X}(2\Gamma_m-i\Omega_y)(\beta^\ast{-}i\Omega_y) [\bar{a}^\ast(\beta^\ast{-}i\Omega_y)+2G\bar{a}e^{{-}i\theta}], \nonumber\\ &F(\Omega_y)=(\beta^\ast{-}i\Omega_y)s(\Omega_y)(\Gamma_m-i\Omega_y)[(2\Gamma_m-i\Omega_y) ({-}i\Omega_{y}m)+4m\Omega^{2}_{m}(1+2\alpha)], \nonumber\\ &M(\Omega_1)=4\hbar g(\beta^\ast{-}i\Omega_+)s(\Omega_+)(2\Gamma_m-i\Omega_{+}-i\Omega_{1}) (\bar{a}^\ast A_{{+}2}+\bar{a}A_{{-}2}^\ast ), \nonumber\\ &M(\Omega_2)=4\hbar g(\beta^\ast{-}i\Omega_+)s(\Omega_+)(2\Gamma_m-i\Omega_{+}-i\Omega_{2}) (\bar{a}^\ast A_{{+}1}+\bar{a}A_{{-}1}^\ast ), \nonumber\\ &\xi={-}4\hbar g\bar{X}(2\Gamma_m-i\Omega_+)s(\Omega_+)[ig\bar{a}(X_{1}A_{{-}2}^\ast{+}X_{2}A_{{-}1}^\ast)+(\beta^\ast{-}i\Omega_+)(A_{{-}1}^\ast A_{{+}2}+A_{{-}2}^\ast A_{{+}1})], \nonumber\\ &W=ig[-(\beta^\ast{-}i\Omega_+)(X_{1}A_{{+}2}+X_{2}A_{{+}1}) +2Ge^{i\theta}(X_{1}A_{{-}2}^\ast{+}X_{2}A_{{-}1}^\ast)], \end{aligned}$$
where in expressions $d(\Omega _y)$ and $D(\Omega _y)$, the value of $y$ is $y$ = 1, 2, and in other expressions the value of $y$ is $y$ = 1, 2, +.

Further, according to the input-output relation [37], we can get the output fields (in a frame rotating at frequency $\omega _{c}$) of the system as follows:

$$\begin{aligned} &s_{t-out}(t)=\sqrt{\eta\kappa}a(t)-s_{in}=(\sqrt{\eta\kappa}{\bar{a}}-\varepsilon_{c})+(\sqrt{\eta\kappa}A_{{+}1} -\varepsilon_{1})e^{{-}i\Omega_{1}t}+\sqrt{\eta\kappa}A_{{-}1}e^{i\Omega_{1}t}+ \nonumber\\ &~~~~~~~~~~~~~~~~~~~(\sqrt{\eta\kappa}A_{{+}2}-\varepsilon_{2})e^{{-}i\Omega_{2}t}+ \sqrt{\eta\kappa}A_{{-}2}e^{i\Omega_{2}t}+\sqrt{\eta\kappa}A_{{+}s}e^{{-}i\Omega_{+}t}+\sqrt{\eta\kappa}A_{{-}s}e^{i\Omega_{+}t}. \end{aligned}$$
The terms $(\sqrt {\eta \kappa }{\bar {a}}-\varepsilon _{c})$ and $(\sqrt {\eta \kappa }A_{+1}-\varepsilon _{1})e^{-i\Omega _{1}t}$ $(\sqrt {\eta \kappa }A_{+2}-\varepsilon _{2})e^{-i\Omega _{2}t}$ describe the output fields with the frequencies of $\omega _{c}$ and $\omega _{1}$ ($\omega _{2}$). And the terms $\sqrt {\eta \kappa }A_{-1}e^{i\Omega _{1}t}$ and $\sqrt {\eta \kappa }A_{-2}e^{i\Omega _{2}t}$ denote the Stokes process. Moreover, the terms $\sqrt {\eta \kappa }A_{+s}e^{-i\Omega _{+}t}$ and $\sqrt {\eta \kappa }A_{-s}e^{i\Omega _{+}t}$, representing the output fields with the frequencies $\omega _{c}+\Omega _{+}$ and $\omega _{c}-\Omega _{+}$, are related to the upper and lower sum sidebands, respectively. For a clearer understanding the generation of sum sideband in the quadratically coupled optomechanical system, we provide a frequency spectrogram of SSG in Fig. 1(b). As we all know, cavity optomechanics studies the radiation-pressure-based optomechanical interaction between the mechanical vibration and the optical cavity field [11]. When the optomechanical system is driven by a control field and the probe field 1, the radiation pressure oscillating with an effective frequency of $\Omega _{1}$ is generated. Similarly, when the optomechanical system is driven by a control field and the probe field 2, the radiation pressure oscillating with an effective frequency of $\Omega _{2}$ will be generated. Obviously, the simultaneous presence of the control field and two probe fields will generate a radiation pressure force oscillating with an effective frequency of $\Omega _{1}+\Omega _{2}$ or $\Omega _{1}-\Omega _{2}$. And here we focus on the radiation pressure force oscillating with an effective frequency of $\Omega _{1}+\Omega _{2}$. If this driving force oscillates close to the mechanical resonance frequency $\Omega _{m}$, the mechanical mode will start to oscillate coherently. This in turn leads to the emergency of the Stokes process and the anti-Stokes process in the optomechanical system. Consequently, the output fields with frequencies $\pm \Omega _{+}=\pm (\Omega _{1}+\Omega _{2})$ (in a frame rotating at frequency $\omega _{c}$ ) are generated, which are the so-called SSG, where the output fields with frequencies $+\Omega _{+}$ is called upper sum sideband generation (USSG) and the output fields with frequencies $-\Omega _{+}$ is called lower sum sideband generation (LSSG). It is worth noting that the underlying physical process of SSG in the quadratically coupled optomechanical system is a two-phonon process [57].

3. Results and discussion

In this section, we will numerically discuss the efficiency of SSG in a quadratically coupled optomechanical system with parametric interactions. In addition, in order to understand the characteristics of SSG more systematically, we also study the effect of system parameters on SSG, especially the effect of the nonlinear gain of OPA on SSG. The amplitude of the input probe field 1 is $\varepsilon _{1}$ , while the amplitude of the output field with the upper (lower) sum sideband is $|\sqrt {\eta \kappa } A_{+s}|$ ($|\sqrt {\eta \kappa }A_{-s}|$). We define $\eta _{s}^+=|\sqrt {\eta \kappa } A_{+s}/\varepsilon _{1}|$ ($\eta _{s}^-=|\sqrt {\eta \kappa }A_{-s}/{\varepsilon _{1}}|$) [46] which is dimensionless as the efficiency of generating the upper (lower) sum sideband. In this work, the parameters we used are similar to those in [37]. The specific parameters are as follows: $m=100$ pg, $\Omega _{m}=2\pi \times 0.1$ MHz, $R$=0.8, $\kappa =2\pi \times 75$ KHz, $T=50$ K, $Q=\Omega _{m}/\Gamma _{m}=3.14\times 10^4$, $L$=67 mm, $\lambda _{c}=2\pi c/\omega _{c}$=532 nm.

Firstly, we consider the effects of the power of the control field and the frequency detuning of the probe field 1 on SSG. Then we plot the efficiencies (in logarithmic form) of USSG and LSSG as a function of the power of the control field $P_{c}$ and the frequency detuning of the probe field 1 $\Omega _{1}$ in Fig. 2. It is shown that from Fig. 2 the peak structure appears for some special values of $\Omega _{1}$ in the efficiencies of SSG. This means that when the appropriate $\Omega _{1}$ is selected, the efficiencies of SSG will be enormously reinforced. It should be noted from Fig. 2 that the efficiencies of both USSG and LSSG increase monotonously with the increases of the power of the control field. The reason for this phenomenon is that the frequency difference between the two scattering paths is non-zero, so there is no interference between the two scattering paths. Accordingly, the efficiencies of SSG increase speedy with the increases of the power of the control field at first, and then the increase rate gradually slows down. From Fig. 2, we can also find that the efficiencies of SSG are quite small, which is caused by the intrinsic weak nonlinearity arising from optomechanical interaction. In addition to the phenomena mentioned above, we also find that the frequency detuning of the probe field 1 has a significant influence on SSG. From Fig. 2(a), we can see that there is only one peak in USSG, and the position of the point corresponding to the peak value is $\Omega _{1} = 2\Omega _{m}$. Unlike this, as can be seen from Fig. 2(b), there are two peaks in LSSG, and the positions of the point corresponding to the peak values are $\Omega _{1} = 2\Omega _{m}$ and $\Omega _{1} = 1.96\Omega _{m}$, respectively. The special value of $\Omega _{1} = 1.96\Omega _{m}$ is attributed to $\Omega _{1} + \Omega _{2} = 2\Omega _{m}$ with $\Omega _{2} = 0.02\times 2\Omega _{m}$. We also confirm that $\Omega _{1}$ varies with $\Omega _{2}$, which still satisfies the relation $\Omega _{1} + \Omega _{2} = 2\Omega _{m}$. The special value of $\Omega _{1}$ ($\Omega _{2}$) corresponding to these peak structures is called the matching conditions.

 figure: Fig. 2.

Fig. 2. (a) The efficiency ${\log _{10}\eta _{s}^+}$ (in logarithmic form) of a USSG and (b) the efficiency ${\log _{10}\eta _{s}^-}$ (in logarithmic form) of a LSSG as a function of the control field power $P_{c}$ and the probe field 1 detuning $\Omega _{1}$ for $\Omega _{2}=0.02\times 2\Omega _{m}$. The specific parameters are as follows: $m=100$ pg, $\Omega _{m}=2\pi \times 0.1$ MHz, $R$=0.8, $\kappa =2\pi \times 75$ KHz, $T=50$ K, $Q=\Omega _{m}/\Gamma _{m}=3.14\times 10^4$, $L$=67 mm, $\lambda _{c}=2\pi c/\omega _{c}$=532 nm, $P_{1}=P_{2}$=0.1 nW, $G=0$.

Download Full Size | PDF

Secondly, in order to understand the effect of OPA on SSG, the efficiencies (in logarithmic form) of SSG as a function of $\Omega _{1}$ for different values of $G$ are plotted in Fig. 3. We can clearly see from Fig. 3 that the efficiencies of SSG are greatly enhanced in the presence of OPA. To simplify, we assume that the phase $\theta$ of the field driving OPA is zero. From Fig. 3(a), we can see clearly that when $G=0$ which corresponds to the black line, there is only one peak in USSG. Interestingly, when $G\neq 0$ which corresponds to the blue (green, red) line in Fig. 3(a), it is shown that one peak appears at $\Omega _{1} = 2.02\Omega _{m}$ in USSG. And we also find that when OPA exists with $G=0.6\kappa$, the efficiency of USSG at $\Omega _{1} = 2.02\Omega _{m}$ is about 100 times higher than that under the condition without OPA. Moreover, the additional new peak makes the matching conditions of USSG modified. As can be seen from Fig. 3(b), whether $G$ exists or not, the peak structure of LSSG does not change. From Fig. 3(b) we also find that when OPA exists with $G=0.6\kappa$, the efficiency of LSSG at $\Omega _{1} = 2\Omega _{m}$ is about $0.1\%$ , while when OPA does not exist, the efficiency of LSSG is about $1\times 10^{-5}\%$. It is worth emphasizing that the efficiency of LSSG can be enhanced by several orders of magnitude with the increase of $G$. In other words, the existence of $G$ has a significant effect on the efficiencies of SSG. Associating Figs. 3(a) with 3(b), one can find that the efficiency of USSG is higher than that of LSSG.

 figure: Fig. 3.

Fig. 3. (a) The efficiency ${\log _{10}\eta _{s}^+}$ (in logarithmic form) of a USSG and (b) the efficiency ${\log _{10}\eta _{s}^-}$ (in logarithmic form) of a LSSG as a function of the probe field 1 detuning $\Omega _{1}$ for different values of $G$ with $P_c$=1 nW, $\Omega _{2}=-0.01\times 2\Omega _{m}$ . Other parameters are the same as those in Fig. 2.

Download Full Size | PDF

Finally, for the purpose of a more systematic understanding of SSG, we plot the efficiencies (in logarithmic form) of SSG as a function of detuning $\Omega _{1}$ and $\Omega _{2}$ in Fig. 4. We can clearly see from Figs. 4(a) or 4(b) the so-called matching conditions of SSG. And we also find that the efficiencies of USSG and LSSG exhibit peaks at some special values of $\Omega _{1}$ and $\Omega _{2}$. From Fig. 4(a), we can see that when $\Omega _{1}\rightarrow 2\Omega _{m}$, USSG becomes obvious, that is to say, the efficiency of USSG is enhanced at $\Omega _{1}=2\Omega _{m}$. In addition, we also find that USSG has another local maximum at $\Omega _{1}=-2\Omega _{m}$. It is worth noting that Fig. 4(a) is almost symmetrical to $\Omega _{1}$ and $\Omega _{2}$, which is caused by the fact that $\Omega _{1}$ and $\Omega _{2}$ play the same role in USSG. Therefore, USSG also has a local maximum at $\Omega _{2}=\pm 2\Omega _{m}$. What is more interesting is that the efficiency of USSG near the point of $(\Omega _{1}, \Omega _{2}) = (2\Omega _{m}, 2\Omega _{m})$ is much greater than that near the point of $(\Omega _{1}, \Omega _{2}) = (-2\Omega _{m}, -2\Omega _{m})$. The underlying physical explanation for this interesting phenomenon is that $(\Omega _{1}, \Omega _{2}) = (2\Omega _{m}, 2\Omega _{m})$ is close to the resonant condition of the cavity field $\bar {\Delta }=\Delta +g\bar {X}=2\Omega _{m}$. Through the above analysis, we conclude that the efficiency of USSG has local maximum values at $\Omega _{1} = \pm 2\Omega _{m}$ and $\Omega _{2} = \pm 2\Omega _{m}$. In other words, the matching condition for USSG is $\Omega _{1} = \pm 2\Omega _{m}$ and $\Omega _{2} = \pm 2\Omega _{m}$. As can be seen from Fig. 4(b), the efficiency of LSSG has local maximum not only at $\Omega _{1}=\pm 2\Omega _{m}$ and $\Omega _{2}=\pm 2\Omega _{m}$, but also local maximum at $\Omega _{1}+\Omega _{2}=\pm 2\Omega _{m}$. Therefore, compared with the matching conditions of USSG, the matching conditions of LSSG include $\Omega _{1}+\Omega _{2}=\pm 2\Omega _{m}$ in addition to $\Omega _{1}=\pm 2\Omega _{m}$ and $\Omega _{2}=\pm 2\Omega _{m}$. This additional matching condition is associated with two diagonals in Fig. 4(b). It is worth noting that in the intersection region the efficiency of LSSG is reinforced. It is precisely because of the existence of matching condition $\Omega _{1}+\Omega _{2}=\pm 2\Omega _{m}$ that there are two sidebands for lower sum sideband. On the contrary, there is only one sideband for upper sum sideband. The physical insight of matching conditions can be explained as follows: the special matching condition $\Omega _{1}+\Omega _{2}=\pm 2\Omega _{m}$ which can be understood as the interaction between the two probe fields generates a radiation pressure force oscillating at the mechanical resonance frequency, hence the mechanical mode starts to oscillate coherently. It is worth noting that the above underlying physical process in the quadratically coupled optomechanical system is a two-phonon process. Similarly, the common matching conditions $\Omega _{1}=\pm 2\Omega _{m}$ and $\Omega _{1}=\pm 2\Omega _{m}$ can also be explained as the interaction between the control field and one of the probe fields induced the radiation pressure force oscillating with frequency $\Omega _{1}$ and $\Omega _{2}$. If $\Omega _{1}$ and $\Omega _{2}$ are close to the mechanical resonance frequency, the significant mechanical oscillation will happen. Thus, the remarkable signals at the sum sideband can be generated via the Stokes- and anti-Stokes scattering of the cavity field. By comparing Figs. 4(a) with 4(c), we find that when $G\neq 0$, this means that when OPA exists, the matching conditions of USSG are modified. On the basis of matching conditions $\Omega _{1} = \pm 2\Omega _{m}$ and $\Omega _{2} = \pm 2\Omega _{m}$, the other matching condition $\Omega _{1}+\Omega _{2}=\pm 2\Omega _{m}$ appears, which corresponds to the two diagonals in Fig. 4(c). By combining Figs. 4(b) with 4(d), we know that the matching conditions of LSSG have not been modified when OPA exists. In addition, we also find that the existence of OPA increases the efficiency of LSSG by several orders of magnitude.

 figure: Fig. 4.

Fig. 4. (a), (c) The efficiency ${\log _{10}\eta _{s}^+}$ (in logarithmic form) of a USSG and (b), (d) the efficiency ${\log _{10}\eta _{s}^-}$ (in logarithmic form) of a LSSG as a function of the probe field 1 detuning $\Omega _{1}$ and the probe field 2 detuning $\Omega _{2}$ for $P_{c}$=1 nW. For the cases of (a), (b) we use $G=0$, and (c), (d) we use $G=0.6\kappa$. Other parameters are the same as those in Fig. 2.

Download Full Size | PDF

4. Conclusion

In summary, we have analyzed SSG in a quadratically coupled optomechanical system containing an OPA. The system is driven by one control field with frequency $\omega _{c}$ and two relatively weak probe fields with frequencies $\omega _{1}$ and $\omega _{2}$. We find that such a system can induce SSG when the nonlinear terms of the optomechanical dynamics are considered. By analytic calculation, we reveal that the amplitude of sun sideband can be well tuned by the power of the control field, the frequency detuning of the probe fields and the nonlinear gain of OPA. In particular, the SSG can be obtained even under weak driven fields and greatly enhanced via meeting the matching conditions. More interestingly, when $G\neq 0$, the efficiencies of SSG are greatly enhanced, especially the efficiency of LSSG can be enhanced by several orders of magnitude. However, it is worth noting that the matching conditions of USSG have been modified. This investigation may deepen the understanding of the nonlinear quadratically coupled optomechanical interactions within the perturbative regime and provide potential applications to the precision measurement and optical communications.

Funding

National Key Research and Development Program of China (2016YFA0301203); National Natural Science Foundation of China (11975103, 11875029, 11574104); Fundamental Research Funds for the Central Universities HUST (2018KFYYXJJ032).

References

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

2. D. E. McClelland, N. Mavalvala, Y. Chen, and R. Schnabel, “Advanced interferometry, quantum optics and optomechanics in gravitational wave detectors,” Laser Photonics Rev. 5, 677–696 (2011). [CrossRef]  

3. C. F. Ockeloen-Korppi, E. Damskägg, J.-M. Pirkkalainen, M. Asjad, A. A. Clerk, F. Massel, M. J. Woolley, and M. A. Sillanpää, “Stabilized entanglement of massive mechanical oscillators,” Nature 556(7702), 478–482 (2018). [CrossRef]  

4. X.-Y. Lü, G.-L. Zhu, L.-L. Zheng, and Y. Wu, “Entanglement and quantum superposition induced by a single photon,” Phys. Rev. A 97(3), 033807 (2018). [CrossRef]  

5. G. A. Brawley, M. R. Vanner, P. E. Larsen, S. Schmid, A. Boisen, and W. P. Bowen, “Nonlinear optomechanical measurement of mechanical motion,” Nat. Commun. 7(1), 10988 (2016). [CrossRef]  

6. M. Bhattacharya, H. Uys, and P. Meystre, “Optomechanical trapping and cooling of partially reflective mirrors,” Phys. Rev. A 77(3), 033819 (2008). [CrossRef]  

7. M. Paternostro, D. Vitali, S. Gigan, M. S. Kim, C. Brukner, J. Eisert, and M. Aspelmeyer, “Creating and Probing Multipartite Macroscopic Entanglement with Light,” Phys. Rev. Lett. 99(25), 250401 (2007). [CrossRef]  

8. 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]  

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

10. J. M. Dobrindt and T. J. Kippenberg, “Theoretical Analysis of Mechanical Displacement Measurement Using a Multiple Cavity Mode Transducer,” Phys. Rev. Lett. 104(3), 033901 (2010). [CrossRef]  

11. T. J. Kippenberg and K. J. Vahala, “Cavity Optomechanics: Back-Action at the Mesoscale,” Science 321(5893), 1172–1176 (2008). [CrossRef]  

12. Y.-S. Park and H. Wang, “Resolved-sideband and cryogenic cooling of an optomechanical resonator,” Nat. Phys. 5(7), 489–493 (2009). [CrossRef]  

13. I. Favero and K. Karrai, “Optomechanics of deformable optical cavities,” Nat. Photonics 3(4), 201–205 (2009). [CrossRef]  

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

15. S. Huang and G. S. Agarwal, “Electromagnetically induced transparency with quantized fields in optocavity mechanics,” Phys. Rev. A 83(4), 043826 (2011). [CrossRef]  

16. J.-Q. Liao, H. K. Cheung, and C. K. Law, “Spectrum of single-photon emission and scattering in cavity optomechanics,” Phys. Rev. A 85(2), 025803 (2012). [CrossRef]  

17. B. Chen, L. Shang, X.-F. Wang, J.-B. Chen, H.-B. Xue, X. Liu, and J. Zhang, “Atom-assisted second-order sideband generation in an optomechanical system with atom-cavity-resonator coupling,” Phys. Rev. A 99(6), 063810 (2019). [CrossRef]  

18. W.-A. Li and G.-Y. Huang, “Enhancement of optomechanically induced sum sideband using parametric interactions,” Phys. Rev. A 100(2), 023838 (2019). [CrossRef]  

19. H. Wang, X. Gu, Y.-X. Liu, A. Miranowicz, and F. Nori, “Optomechanical analog of two-color electromagnetically-induced transparency: photon transmission through an optomechanical device with a two-level system,” Phys. Rev. A 90(2), 023817 (2014). [CrossRef]  

20. H. Jing, S. K. Özdemir, Z. Geng, J. Zhang, X.-Y. Lü, B. Peng, L. Yang, and F. Nori, “Optomechanically-induced transparency in parity-time-symmetric microresonators,” Sci. Rep. 5(1), 9663 (2015). [CrossRef]  

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

22. H. Jing, H. Lü, S. K. özdemir, T. Carmon, and F. Nori, “Nanoparticle sensing with a spinning resonator,” Optica 5(11), 1424 (2018). [CrossRef]  

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

24. A. H. Safavi-Naeini, T. P. Mayer Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, “Electromagnetically induced transparency and slow light with optomechanics,” Nature 472(7341), 69–73 (2011). [CrossRef]  

25. C. Jiang, H.-X. Liu, Y.-S. Cui, X.-W. Li, G.-B. Chen, and B. Chen, “Electromagnetically induced transparency and slow light in two-mode optomechanics,” Opt. Express 21(10), 12165 (2013). [CrossRef]  

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

27. Z.-X. Liu, B. Wang, C. Kong, H. Xiong, and Y. Wu, “Magnetic-field-dependent slow light in strontium atom-cavity system,” Appl. Phys. Lett. 112(11), 111109 (2018). [CrossRef]  

28. 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]  

29. S.-H. Ouyang, J.-Q. You, and F. Nori, “Cooling a mechanical resonator via coupling to a tunable double quantum dot,” Phys. Rev. B 79(7), 075304 (2009). [CrossRef]  

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

31. S.-M. Huang and A.-X. Chen, “Improving the cooling of a mechanical oscillator in a dissipative optomechanical system with an optical parametric amplifier,” Phys. Rev. A 98(6), 063818 (2018). [CrossRef]  

32. C. M. Bender, M. Gianfreda, S. K. Özdemir, B. Peng, and L. Yang, “Twofold transition in $\mathcal {PT}$-symmetric coupled oscillators,” Phys. Rev. A 88(6), 062111 (2013). [CrossRef]  

33. B. Peng, S. K. Özdemir, F.-C. Lei, F. Monifi, M. Gianfreda, G.-L. Long, S.-H. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity-time-symmetric whispering-gallery microcavities,” Nat. Phys. 10(5), 394–398 (2014). [CrossRef]  

34. J.-H. Li, R. Yu, C.-L. Ding, and Y. Wu, “$\mathcal {PT}$-symmetry-induced evolution of sharp asymmetric line shapes and high-sensitivity refractive index sensors in a three-cavity array,” Phys. Rev. A 93(2), 023814 (2016). [CrossRef]  

35. Y.-L. Liu, R. Wu, J. Zhang, S. K. Özdemir, L. Yang, F. Nori, and Y.-X. Liu, “Controllable optical response by modifying the gain and loss of a mechanical resonator and cavity mode in an optomechanical system,” Phys. Rev. A 95(1), 013843 (2017). [CrossRef]  

36. F. Marquardt, J. P. Chen, A. A. Clerk, and S. M. Girvin, “Quantum theory of cavity-assisted sideband cooling of mechanical motion,” Phys. Rev. Lett. 99(9), 093902 (2007). [CrossRef]  

37. 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]  

38. A. Rai and G. S. Agarwal, “Quantum optical spring,” Phys. Rev. A 78(1), 013831 (2008). [CrossRef]  

39. J. C. Sankey, C. Yang, B. M. Zwickl, A. M. Jayich, and J. G. E. Harris, “Strong and tunable nonlinear optomechanical coupling in a low-loss system,” Nat. Phys. 6(9), 707–712 (2010). [CrossRef]  

40. M. Karuza, C. Biancofiore, M. Bawaj, C. Molinelli, M. Galassi, R. Natali, P. Tombesi, G. D. Giuseppe, and D. Vitali, “Optomechanically induced transparency in a membrane-in-the-middle setup at room temperature,” Phys. Rev. A 88(1), 013804 (2013). [CrossRef]  

41. J.-Q. Liao and F. Nori, “Photon blockade in quadratically coupled optomechanical systems,” Phys. Rev. A 88(2), 023853 (2013). [CrossRef]  

42. S.-P. Liu, W.-X. Yang, T. Shui, Z.-H. Zhu, and A.-X. Chen, “Tunable two-phonon higher-order sideband amplification in a quadratically coupled optomechanical system,” Sci. Rep. 7(1), 17637 (2017). [CrossRef]  

43. L.-G. Si, L.-X. Guo, H. Xiong, and Y. Wu, “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]  

44. L.-G. Si, H. Xiong, M. S. Zubairy, and Y. Wu, “Optomechanically induced opacity and amplification in a quadratically coupled optomechanical system,” Phys. Rev. A 95(3), 033803 (2017). [CrossRef]  

45. X.-Y. Wang, L.-G. Si, Z.-X. Liu, X.-H. Lu, and Y. Wu, “Tunable optical amplification arises from blue detuning in a quadratically coupled optomechanical system,” J. Opt. Soc. Am. B 36(5), 1355 (2019). [CrossRef]  

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

47. 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 (2017). [CrossRef]  

48. S. Huang and G. S. Agarwal, “Enhancement of cavity cooling of a micromechanical mirror using parametric interactions,” Phys. Rev. A 79(1), 013821 (2009). [CrossRef]  

49. S. Huang and G. S. Agarwal, “Normal-mode splitting in a coupled system of a nanomechanical oscillator and a parametric amplifier cavity,” Phys. Rev. A 80(3), 033807 (2009). [CrossRef]  

50. G. S. Agarwal and S. Huang, “Strong mechanical squeezing and its detection,” Phys. Rev. A 93(4), 043844 (2016). [CrossRef]  

51. S.-P Liu, B. Liu, and W.-X. Yang, “Highly sensitive mass detection based on nonlinear sum-sideband in a dispersive optomechanical system,” Opt. Express 27(4), 3909–3919 (2019). [CrossRef]  

52. H. R. Hamedi, “Transient absorption and lasing without inversion in an artificial molecule via Josephson coupling energy,” Laser Phys. Lett. 12(3), 035201 (2015). [CrossRef]  

53. A. Chen, “Coherent manipulation of spontaneous emission spectra in coupled semiconductor quantum well structures,” Opt. Express 22(22), 26991 (2014). [CrossRef]  

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

55. J. Ma, C. You, L.-G. Si, H. Xiong, J. Li, X. Yang, and Y. Wu, “Formation and manipulation of optomechanical chaos via a bichromatic driving,” Phys. Rev. A 90(4), 043839 (2014). [CrossRef]  

56. R. W. Boyd, “Nonlinear Optics,” 3rd ed., (Academic Press, Burlington, MA, 2008).

57. S. Huang and G. S. Agarwal, “Electromagnetically induced transparency from two-phonon processes in quadratically coupled membranes,” Phys. Rev. A 83(2), 023823 (2011). [CrossRef]  

58. S. M. Spillane, T. J. Kippenberg, O. J. Painter, and K. J. Vahala, “Ideality in a Fiber-Taper-Coupled Microresonator System for Application to Cavity Quantum Electrodynamics,” Phys. Rev. Lett. 91(4), 043902 (2003). [CrossRef]  

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 (4)

Fig. 1.
Fig. 1. (a) Schematic diagram of a quadratic optomechanical system consisting of an OPA and a MR inserted between two fixed mirrors. The system is excited by two relatively weak probe fields and a strong control field. The two probe fields are represented by probe field 1 with frequency $\omega _{1}$ and probe field 2 with frequency $\omega _{2}$, respectively. And $\omega _{c}$ is the frequency of the control field. The MR and cavity field are quadratically coupled via the radiation pressure. (b) Frequency spectrogram of sum sideband generation in the quadratically coupled optomechanical system with a control field and two probe fields driven.
Fig. 2.
Fig. 2. (a) The efficiency ${\log _{10}\eta _{s}^+}$ (in logarithmic form) of a USSG and (b) the efficiency ${\log _{10}\eta _{s}^-}$ (in logarithmic form) of a LSSG as a function of the control field power $P_{c}$ and the probe field 1 detuning $\Omega _{1}$ for $\Omega _{2}=0.02\times 2\Omega _{m}$. The specific parameters are as follows: $m=100$ pg, $\Omega _{m}=2\pi \times 0.1$ MHz, $R$=0.8, $\kappa =2\pi \times 75$ KHz, $T=50$ K, $Q=\Omega _{m}/\Gamma _{m}=3.14\times 10^4$, $L$=67 mm, $\lambda _{c}=2\pi c/\omega _{c}$=532 nm, $P_{1}=P_{2}$=0.1 nW, $G=0$.
Fig. 3.
Fig. 3. (a) The efficiency ${\log _{10}\eta _{s}^+}$ (in logarithmic form) of a USSG and (b) the efficiency ${\log _{10}\eta _{s}^-}$ (in logarithmic form) of a LSSG as a function of the probe field 1 detuning $\Omega _{1}$ for different values of $G$ with $P_c$=1 nW, $\Omega _{2}=-0.01\times 2\Omega _{m}$ . Other parameters are the same as those in Fig. 2.
Fig. 4.
Fig. 4. (a), (c) The efficiency ${\log _{10}\eta _{s}^+}$ (in logarithmic form) of a USSG and (b), (d) the efficiency ${\log _{10}\eta _{s}^-}$ (in logarithmic form) of a LSSG as a function of the probe field 1 detuning $\Omega _{1}$ and the probe field 2 detuning $\Omega _{2}$ for $P_{c}$=1 nW. For the cases of (a), (b) we use $G=0$, and (c), (d) we use $G=0.6\kappa$. Other parameters are the same as those in Fig. 2.

Equations (10)

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

H = Δ a ^ a ^ + ( p ^ 2 2 m + 1 2 m Ω m 2 x ^ 2 ) + g a ^ a ^ x ^ 2 + i G ( a ^ 2 e i θ a ^ 2 e i θ ) +                 i η κ [ ( ε c + ε 1 e i Ω 1 t + ε 2 e i Ω 2 t ) a ^ H . c . ] ,
t x ^ = p ^ m , t p ^ = [ m Ω m 2 + 2 g a ^ a ^ ] x ^ Γ m p ^ + F ^ t h , t a ^ = [ κ 2 + i ( Δ + g x ^ 2 ) ] a ^ + 2 G e i θ a ^ + η κ ( ε c + ε 1 e i Ω 1 t + ε 2 e i Ω 2 t ) + a ^ i n ,
t X = Q m , t P = [ m Ω m 2 + 2 g a a ] Q + Γ m ( 1 + 2 n ) m Ω m 2 Γ m P , t Q = [ 4 g a a + 2 m Ω m 2 ] X Γ m Q + 2 P m , t a = [ κ 2 + i ( Δ + g X ) ] a + 2 G e i θ a ^ + η κ ( ε c + ε 1 e i Ω 1 t + ε 2 e i Ω 2 t ) ,
δ a = A + 1 e i Ω 1 t + A 1 e i Ω 1 t + A + 2 e i Ω 2 t + A 2 e i Ω 2 t + A + s e i Ω + t + A s e i Ω + t , δ a = A 1 e i Ω 1 t + A + 1 e i Ω 1 t + A 2 e i Ω 2 t + A + 2 e i Ω 2 t + A s e i Ω + t + A + s e i Ω + t , δ Q = Q 1 e i Ω 1 t + Q 1 e i Ω 1 t + Q 2 e i Ω 2 t + Q 2 e i Ω 2 t + Q s e i Ω + t + Q s e i Ω + t , δ X = X 1 e i Ω 1 t + X 1 e i Ω 1 t + X 2 e i Ω 2 t + X 2 e i Ω 2 t + X s e i Ω + t + X s e i Ω + t , δ P = P 1 e i Ω 1 t + P 1 e i Ω 1 t + P 2 e i Ω 2 t + P 2 e i Ω 2 t + P s e i Ω + t + P s e i Ω + t ,
Q ¯ = 0 , a ¯ = ( β + 2 G e i θ ) β β + 4 G 2 η κ ε c , X ¯ = P ¯ m 2 Ω m 2 ( 1 + 2 α ) , P ¯ = ( 1 + 2 n ) m Ω m 2 ,
( β i Ω j ) A + j = i g a ¯ X j + 2 G e i θ A j + η κ ε j ( β + i Ω j ) A j = i g a ¯ X j + 2 G e i θ A + j ( 2 Γ m i Ω j ) P j = ( m Ω m 2 + 2 g | a ¯ | 2 ) Q j ( Γ m i Ω j ) Q j = 4 g [ a ¯ A + j + a ¯ A j ] X ¯ 2 m Ω m 2 ( 1 + 2 α ) X j + 2 P j m Q j + i Ω 1 m X j = 0.                 ( j = 1 , 2 )
( β i Ω + ) A + s = i g ( a ¯ X s + X 1 A + 2 + X 2 A + 1 ) + 2 G e i θ A s ( β + i Ω + ) A s = i g ( a ¯ X s + X 1 A 2 + X 2 A 1 ) + 2 G e i θ A + s ( 2 Γ m i Ω + ) P s = m Ω m 2 ( 1 + 2 α ) Q s 2 g [ a ¯ ( A + 1 Q 2 + A + 2 Q 1 ) + a ¯ ( A 1 Q 2 + A 2 Q 1 ) ] ( Γ m i Ω + ) Q s = 4 g [ a ¯ A + s + a ¯ A s + A 1 A + 2 + A 2 A + 1 ] X ¯ 4 g [ a ¯ ( A + 1 X 2 + A + 2 X 1 )                                                       + a ¯ ( A 1 X 2 + A 2 X 1 ) ] 2 m Ω m 2 ( 1 + 2 α ) X s + 2 P s m Q s + i Ω + m X s = 0.
A + j = τ ( Ω j ) X j + η κ ε j ( β i Ω j ) s ( Ω j ) , X j = η κ ε j f ( Ω j ) F ( Ω j ) + d ( Ω j ) + σ ( Ω j ) + D ( Ω j ) , A j = i g a ¯ X j + 2 G e i θ A + j β + i Ω j ,                     ( j = 1 , 2 ) A + s = τ ( Ω + ) X s + W s ( Ω + ) , A s = i g a ¯ X s i g ( X 1 A 2 + X 2 A 1 ) + 2 G e i θ A + s β + i Ω + , X s = f ( Ω + ) W + [ ξ X 2 M ( Ω 2 ) X 1 M ( Ω 1 ) ] ( β i Ω + ) [ F ( Ω + ) + σ ( Ω + ) ] ( β i Ω + ) τ ( Ω + ) f ( Ω + ) ,
s ( Ω y ) = ( β i Ω y ) ( β i Ω y ) 4 G 2 , σ ( Ω y ) = 4 i g 2 | a ¯ | 2 X ¯ ( 2 Γ m i Ω y ) s ( Ω y ) , τ ( Ω y ) = i g [ a ¯ ( β i Ω y ) 2 G a ¯ e i θ ] , d ( Ω y ) = 4 i g 2 | a ¯ | 2 X ¯ ( 2 Γ m i Ω y ) [ ( β i Ω y ) 2 + 4 G 2 ] , D ( Ω y ) = 8 i g 2 G X ¯ ( 2 Γ m i Ω y ) ( β i Ω y ) [ a ¯ 2 e i θ a ¯ 2 e i θ ] , f ( Ω y ) = 4 g X ¯ ( 2 Γ m i Ω y ) ( β i Ω y ) [ a ¯ ( β i Ω y ) + 2 G a ¯ e i θ ] , F ( Ω y ) = ( β i Ω y ) s ( Ω y ) ( Γ m i Ω y ) [ ( 2 Γ m i Ω y ) ( i Ω y m ) + 4 m Ω m 2 ( 1 + 2 α ) ] , M ( Ω 1 ) = 4 g ( β i Ω + ) s ( Ω + ) ( 2 Γ m i Ω + i Ω 1 ) ( a ¯ A + 2 + a ¯ A 2 ) , M ( Ω 2 ) = 4 g ( β i Ω + ) s ( Ω + ) ( 2 Γ m i Ω + i Ω 2 ) ( a ¯ A + 1 + a ¯ A 1 ) , ξ = 4 g X ¯ ( 2 Γ m i Ω + ) s ( Ω + ) [ i g a ¯ ( X 1 A 2 + X 2 A 1 ) + ( β i Ω + ) ( A 1 A + 2 + A 2 A + 1 ) ] , W = i g [ ( β i Ω + ) ( X 1 A + 2 + X 2 A + 1 ) + 2 G e i θ ( X 1 A 2 + X 2 A 1 ) ] ,
s t o u t ( t ) = η κ a ( t ) s i n = ( η κ a ¯ ε c ) + ( η κ A + 1 ε 1 ) e i Ω 1 t + η κ A 1 e i Ω 1 t +                                       ( η κ A + 2 ε 2 ) e i Ω 2 t + η κ A 2 e i Ω 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.