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

Enhanced four-wave mixing in $\mathcal{PT}$-symmetric optomechanical systems

Open Access Open Access

Abstract

We investigate the enhanced four-wave mixing (FWM) process in a parity-time ($\mathcal{PT}$)-symmetric optomechanical system, where an active cavity is coupled to a passive cavity supporting a mechanical mode. The passive cavity is optically driven by a strong control field and a weak probe field, and the mechanical mode is excited by a weak coherent driving field. By tuning the coupling strength between the two cavities with balanced gain and loss, we find that the FWM intensity can be significantly enhanced near the exceptional points (EPs) at low control power, which is about 12 orders of magnitude higher than that of the single-cavity case. Due to the interference effect induced by the optical and mechanical driving field, it is shown that the FWM intensity can be further enhanced or suppressed by tuning the amplitude and phase of the mechanical driving field. Moreover, the dependence of the FWM intensity on the frequency and power of the control field is also discussed. Our work provides a route to enhance the four-wave mixing process in a flexible way.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The nonlinear coupling between optical and mechanical modes via radiation pressure is the basis of cavity optomechanical systems [13], which has been exploited to realize ground state cooling of a mechanical oscillator [4,5], preparation of nonclassical mechanical states [6], generation of squeezed light [7,8], optomechanically induced transparency (OMIT) [911], and so on. Moreover, four-wave mixing (FWM) is an important third-order nonlinear optical process in a variety of media including semiconductor-doped glasses [12], atomic systems [13,14], and photonic cystal fibers [15], which can find a number of applications in optical telecommunications. Recently, radiation-pressure-induced FWM has been theoretically studied in several types of optomechanical sytems [1620]. However, the FWM intensity is typically weak in these systems, which makes it difficult to be observed in experiments.

In 1998, Bender and Boettcher proposed that non-Hermitian Hamiltonian having parity-time ($\mathcal {PT}$)-symmetry can possess entirely real spectra [21]. $\mathcal {PT}$-symmetric systems can exhibit a phase transition between the unbroken- and broken-symmetry phases at the exceptional points (EPs), where the eigenvalues and the corresponding eigenvectors simultaneously coalesce [22]. This phase transition has been demonstrated in various physical systems and can lead to many intriguing phenomena, such as loss-induced transparency [23], topological energy transfer [24], enhanced sensitivity [25,26], controllable matching conditions for the FWM [27], and nonlinearity enhancement [2830]. In addition, Zhang et al. have recently demonstrated that symmetry-breaking-induced second-order nonlinearity at a microcavity surface can be improved by 14 orders of magnitude higher than that of the non-enhancement case [31].

More recently, $\mathcal {PT}$-symmetric optomechanical systems have been under extensive exploration in various aspects, including phonon laser [32,33], chaos [34], optomechanically induced transparency [35,36], optomechanically induced absorption [37], enhanced ground state cooling [38,39], enhanced high-order sideband generation [40,41], enhanced sensitivity [42], and so on. However, third-order nonlinear effect has seldom been studied in this system. In the present paper, we mainly focus on the four-wave mixing process in $\mathcal {PT}$-symmetric optomechanical systems in the presence of a strong optical control field, a weak optical probe field, and a weak coherent mechanical driving field. We demonstrate that the FWM intensity is significantly enhanced near the phase transition point between the unbroken- and broken-symmetry. Moreover, the FWM intensity can be effectively controlled by the frequency and power of the optical control field as well as the amplitude and phase of the mechanical driving field. It is worth pointing out that mechanical driving field has been exploited in experiments to realize virtual exceptional points [43], injection locking [44], electro-optomechanically induced transparency [45], cascaded optical transparency [46], nonreciprocal mode conversion [47], and phase-sensitive parametric amplifier [48]. Theoretical studies also show that the additional mechanical driving field can be employed to control the optical response of the optomechanical systems due to the more complicated interference effects [4954]. We note that mechanical driving field may also play a vital role in tuning the optomechanical systems and in achieving one-way optical flow [5557]. Different from the four-wave mixing in a $\mathcal {PT}$-symmetric coupler in Ref. [27], the FWM process in $\mathcal {PT}$-symmetric optomechanical systems is induced by the radiation pressure, where two control photons at frequency $\omega _c$ can mix with a probe photon at frequency $\omega _p$ via the mechanical mode to yield an idler photon at frequency $2\omega _c-\omega _p$ [1620]. With the assistance of the mechanical driving field, the FWM process in this optomechanical systems can be controlled more flexibly.

This paper is organized as follows. In Sec. 2, we present the theoretical model of the $\mathcal {PT}$-symmetric optomechanical system and derive the analytical expression of the FWM intensity. In Sec. 3, we numerically study the dependence of FWM intensity on various parameters, including the cavity-cavity coupling strength, the gain-loss ratio, the amplitude and phase of the mechanical driving field, and the frequency and power of the optical control field. Finally, we summarize our work in Sec. 4.

2. Model and theory

We consider the $\mathcal {PT}$-symmetric optomechanical system schematically shown in Fig. 1, which consists of two directly coupled whispering-gallery-mode (WGM) microtoroidal cavities with coupling strength $J$. One of the cavities is passive with resonance frequency $\omega _1$ and loss rate $\kappa _1$, which supports a mechanical radial breathing mode with effective mass $m$, resonance frequency $\omega _m$ and damping rate $\gamma _m$. Another cavity is active with resonance frequency $\omega _2$ and tunable gain rate $\kappa _2$, which can be fabricated from Er$^{3+}$-doped silica and can emit photons in the 1550 nm band [28,29]. The passive cavity is simultaneously driven by a strong control field and a weak probe field with amplitudes $\varepsilon _{i}$, frequency $\omega _{i}$, and phase $\phi _{i}$ ($i=c,p$), respectively. In addition, a weak coherent mechanical driving field with amplitude $\varepsilon _m$, frequency $\Omega =\omega _p-\omega _c$, and phase $\phi _m$ is applied to excite the mechanical mode. It should be pointed out that mechanical driving of the radial breathing mode has been experimentally realized in a microtoroid optomechanical cavity with an integrated electrical interface [44], where an inertial force was applied to the microtoroid in the radial direction. Moreover, optomechanical interaction can also be realized by the evanescent coupling of nanobeams and nanomembranes to WGM resonators [58,59]. In this way, it is possible to realize mechanical mixing of different mechanical resonators [54,58].

 figure: Fig. 1.

Fig. 1. Schematic illustration of the $\mathcal {PT}$-symmetric optomechanical system. The passive cavity with loss rate $\kappa _1$ interacts with a mechanical mode with resonance frequency $\omega _m$ and damping rate $\gamma _m$ via radiation pressure. The active cavity with tunable gain rate $\kappa _2$ is coupled to the passive cavity, and the coupling strength $J$ can be adjusted by the distance between them. In addition, the passive cavity is driven by a strong control field at frequency $\omega _c$ and a weak probe field at frequency $\omega _p$, and the mechanical resonator is excited by a weak coherent mechanical driving field at frequency $\Omega =\omega _p-\omega _c$.

Download Full Size | PDF

In a rotating frame at the frequency $\omega _c$ of the control field, the Hamiltonian of this system is given by

$$\begin{aligned} H&=H_{0}+H_{\mathrm{I}}+H_{\mathrm{dr}},\\ H_{0}&=\hbar\Delta_1 a_1^{{\dagger}}a_1+\hbar\Delta_2 a_2^{{\dagger}}a_2+\frac{p^2}{2m}+\frac{1}{2}m\omega_m^2 x^2,\\ H_{\mathrm{I}}&=-\hbar J(a_1^{{\dagger}}a_2+a_1 a_2^{{\dagger}})-\hbar g_1 a_1^{{\dagger}}a_1 x,\\ H_{\mathrm{dr}}&=i\hbar\sqrt{\eta_c\kappa_1}[(\varepsilon_c+\varepsilon_p e^{{-}i\Omega t-i\phi_{pc}})a_1^{{\dagger}}-\mathrm{H.c.}]-x\varepsilon_m\mathrm{cos}(\Omega t+\phi_m). \end{aligned}$$
$H_0$ represents the energy of the two cavity modes and one mechanical mode, where $a_1~(a_2)$ and $a_1^{\dagger }~(a_2^{\dagger })$ are the annihilation and creation operators of the passive (active) cavity mode, $x$ and $p$ are the displacement and momentum operators of the mechanical mode. $\Delta _1=\omega _1-\omega _c$ and $\Delta _2=\omega _2-\omega _c$ are the detunings between the two cavity modes and the control field. The first term in the Hamiltonian $H_{\mathrm {I}}$ denotes the interaction between the two cavity modes, where the coupling strength $J$ can be controlled by the distance between them. The second term in $H_{\mathrm {I}}$ describes the radiation pressure induced interaction between the passive cavity mode and the mechanical mode with the coupling constant $g_1$. Moreover, the first term in the Hamiltonian $H_{\mathrm {dr}}$ includes the interaction between the passive cavity and the control (probe) field, where the coupling efficiency $\eta _c=\kappa _{ex,1}/\kappa _1$ with $\kappa _{ex,1}$ being the loss rate due to external coupling, and $\phi _{pc}=\phi _p-\phi _c$ is the relative phase of the probe field and the control field. The amplitudes $\varepsilon _{c,p}$ are related to the powers $P_{c,p}$ of the control (probe) fields by $\varepsilon _{c,p}=\sqrt {2P_{c,p}/(\hbar \omega _{c,p})}$. The interaction Hamiltonian between the mechanical mode and its driving field can be described by $-x\varepsilon _m\mathrm {cos}(\Omega t+\phi _m)$.

According to Eq. (1) and Heisenberg equations of motion, we can derive the time evolution of the system operators as follows:

$$\dot{a_1}={-}\left(\frac{\kappa_1}{2}+i\Delta_1-g_1x\right)a_1+iJa_2+\sqrt{\eta_c\kappa_1}(\varepsilon_c+\varepsilon_pe^{{-}i\Omega t-i\phi_p}),$$
$$\dot{a_2}=\left(\frac{\kappa_2}{2}-i\Delta_2\right)a_2+iJa_1,$$
$$\dot{x}=\frac{p}{m},$$
$$\dot{p}={-}m\omega_m^2x-\gamma_m p+\hbar g_1 a_1^{{\dagger}}a_1+\varepsilon_m\cos(\Omega t+\phi_m),$$
where the quantum and thermal noise terms have been neglected since we mainly study the mean response of the system to the probe field, and the corresponding loss and gain terms for the cavity and mechanical modes have been added phenomenologically. Note that $\kappa _1>0$ and $\kappa _2>0$ in Eqs. (2)–(3) correspond to a passive-active hybrid optomechanical system. In the weak optomechanical coupling regime, only 2-order EP exists in the two optical supermodes of this optomechanical system. With balanced gain and loss ($\kappa _1=\kappa _2$) and when $J>(\kappa _1+\kappa _2)/4$, the system can be $\mathcal {PT}$-symmetric. However, the $\mathcal {PT}$-symmetry is broken if $J<(\kappa _1+\kappa _2)/4$. At the exceptional points, i.e., $J=(\kappa _1+\kappa _2)/4$, pairs of eigenvalues and the corresponding eigenstates of the system coalesce. Recently, Jing et al. showed that when the passive cavity is driven by a red-detuned strong control field that matches the mechanical frequency $\omega _m$, the supermode structure of the system is changed, featuring the emergence of high-order EP [38]. In this work, we mainly focus on the case that the control field is resonant with the cavity field, and we will investigate the FWM process at high-order EP in future.

By setting all the time derivatives in Eqs. (2)–(5) to be zero, we can obtain the steady-state solutions of the system, which are given as follows:

$$\begin{aligned} &a_{1s}=\frac{\sqrt{\eta_c\kappa_1}\varepsilon_c(\kappa_2/2-i\Delta_2)}{(\kappa_1/2+i\Delta_1')(\kappa_2/2-i\Delta_2)-J^2},\\ &a_{2s}={-}\frac{iJ\sqrt{\eta_c\kappa_1}\varepsilon_c}{(\kappa_1/2+i\Delta_1')(\kappa_2/2-i\Delta_2)-J^2},\\ &x_s=\frac{\hbar g_1|a_{1s}|^2}{m\omega_m^2},\\ &p_s=0, \end{aligned}$$
where $\Delta _1'=\Delta _1-g_1 x_s$ is the effective cavity-control field detuning. Since the probe field and mechanical driving field are much weaker than the strong control field, each Heisenberg operator can be rewritten as a sum of its steady-state solution and a small fluctuation, i.e., $o=o_s+\delta o~(o=a_1,a_2,x,p)$. Substituting this ansatz into Eqs. (2)–(5) and neglecting the second-order small quantities such as $\delta o\delta o$ and $\delta o\delta o^{\dagger }$, we can derive the following linearized Heisenberg-Langevin equations:
$$\delta\dot{a_1}={-}(\kappa_1/2+i\Delta_1')\delta a_1+ig_1 a_{1s}\delta x+iJ\delta a_2+\sqrt{\eta_c\kappa_1}\varepsilon_p e^{{-}i\Omega t-i\phi_{pc}},$$
$$\delta\dot{a_2}=(\kappa_2/2-i\Delta_2)\delta a_2+iJ\delta a_1,$$
$$\delta\dot{x}=\frac{\delta p}{m},$$
$$\delta\dot{p}={-}m\omega_m^2\delta x-\gamma_m\delta p+\hbar g_1(a_{1s}^*\delta a_1+a_{1s}\delta a_1^\dagger)+\varepsilon_m\mathrm{cos}(\Omega t+\phi_m).$$
The system is stable only if the real parts of all the eigenvalues of the coefficient matrix associated with Eqs. (7)–(10) are negative, which will be checked numerically in this work. We further make the ansatz: $\delta a_1=a_{1+}e^{-i\Omega t}+a_{1-}e^{i\Omega t}$, $\delta a_2=a_{2+}e^{-i\Omega t}+a_{2-}e^{i\Omega t}$, $\delta x=x_{+}e^{-i\Omega t}+x_{-}e^{i\Omega t}$, $\delta p=p_{+}e^{-i\Omega t}+p_{-}e^{i\Omega t}$. Substituting the ansatz into Eqs. (7)–(10) and comparing the coefficients of $e^{-i\Omega t}$ and $e^{i\Omega t}$ lead to eight algebraic equations, which yields
$$a_{1+}=\frac{\chi_1(\Omega)}{1+if^*(-\Omega)-if(\Omega)}\left\{[1+if^*(-\Omega)]\sqrt{\eta_c\kappa_1}\varepsilon_p e^{{-}i\phi_p}+ig_1 a_{1s}\chi_m(\Omega)\frac{\varepsilon_m}{2}e^{{-}i\phi_m}\right\},$$
$$a_{1-}=\frac{\chi_1(-\Omega)}{1-if(-\Omega)+if^*(\Omega)}\left[if^*(\Omega)\frac{a_{1s}}{a_{1s}^*}\sqrt{\eta_c\kappa_1}\varepsilon_p e^{i\phi_p}+ig_1 a_{1s}\chi_m^*(\Omega)\frac{\varepsilon_m}{2}e^{i\phi_m}\right].$$
For convenience, here we have defined
$$\begin{aligned} &\chi_m(\Omega)=\frac{1}{m(\omega_m^2-\Omega^2-i\gamma_m\Omega)},\\ &\chi_2(\Omega)=\frac{1}{\kappa_2/2-i\Delta_2+i\Omega},\\ &\chi_1(\Omega)=\frac{1}{\kappa_1/2+i(\Delta_1'-\Omega)-J^2\chi_2(\Omega)},\\ &f(\Omega)=\frac{\hbar g_1^2|a_{1s}|^2\chi_m(\Omega)}{\kappa_1/2+i(\Delta_1'-\Omega)-J^2\chi_2(\Omega)}. \end{aligned}$$
Applying the standard input-output relation $a_{1,\mathrm {out}}=a_{1,\mathrm {in}}-\sqrt {\eta _c\kappa _1}a_1$ [60], the output field from cavity $a_1$ can be given by
$$a_{1,\mathrm{out}}(t)= \underbrace{(\varepsilon_c-\sqrt{\eta_c\kappa_1}a_{1s})e^{{-}i\omega_c t}}_{\mathrm{Control~ field}}+\underbrace{(\varepsilon_p e^{{-}i\phi_{pc}}-\sqrt{\eta_c\kappa_1}a_{1+})e^{{-}i\omega_p t}}_{\mathrm{Probe~field}} -\underbrace{\sqrt{\eta_c\kappa_1}a_{1-}e^{{-}i(2\omega_c-\omega_p)t}}_{\mathrm{FWM~field}}.\\$$
We can see from Eq. (14) that the output field contains the input control field at the frequency $\omega _c$ and the probe field at the frequency $\omega _p$. In addition, a new four-wave mixing field at the frequency $2\omega _c-\omega _p$ is generated by the interaction between two control photons and a probe photon via the mechanical mode. The transmission of the probe field is obtained as
$$t_p=\frac{\varepsilon_p e^{{-}i\phi_{pc}}-\sqrt{\eta_c\kappa_1}a_{1+}}{\varepsilon_p e^{{-}i\phi_{pc}}}=t_{1}+t_{2},$$
with
$$t_1=1-\frac{1+if^*(-\Omega)}{1+if^*(-\Omega)-if(\Omega)}\chi_1(\Omega)\eta_c\kappa_1,$$
$$t_2={-}\frac{ig_1a_{1s}\chi_1(\Omega)\chi_m(\Omega)\varepsilon_m/(2\varepsilon_p)} {1+if^*(-\Omega)-if(\Omega)}\sqrt{\eta_c\kappa_1}e^{{-}i\phi}.$$
Here $\phi =\phi _m-\phi _{pc}$ is the phase difference between the mechanical and optical driving fields. Similarly, the FWM field normalized to the probe field can be defined as
$$\mathrm{FWM_p}=\frac{\sqrt{\eta_c\kappa_1}a_{1-}}{\varepsilon_p e^{i\phi_p}}=\mathrm{FWM}_1+\mathrm{FWM}_2,$$
with
$$\mathrm{FWM_1}=\frac{if^*(\Omega)\chi_1(-\Omega)a_{1s}/a_{1s}^*}{1-if(-\Omega)+if^*(\Omega)}\eta_c\kappa_1,$$
$$\mathrm{FWM_2}=\frac{ig_1a_{1s}\chi_{1}(-\Omega)\chi_m^*\varepsilon_m/(2\varepsilon_p)} {1-if(-\Omega)+if^*(\Omega)}\sqrt{\eta_c\kappa_1}e^{i\phi}.$$
We can see from Eqs. (15) and (18) that both the probe transmission and FWM field contain two coherent parts, where $t_1$ $(\mathrm {FWM_1})$ are the expressions without mechanical driving field and $t_2~(\mathrm {FWM_2})$ are the modifications due to mechanical driving field. Interference between these two coherent parts determines the properties of the output fields, where the amplitude $\varepsilon _m$ and the phase difference $\phi$ are important control parameters.

3. Enhanced four-wave mixing

In previous works, the properties of the transmitted probe field have been extensively investigated in $\mathcal {PT}$-symmetric optomechanical systems [4954]. In this section, we mainly study the effects of $\mathcal {PT}$-symmetry and mechanical driving field on the properties of the four-wave-mixing field. In the numerical simulations, we choose the parameters from recent experiments [29,61]: $\omega _1=2\pi c/\lambda$ with $\lambda =1550$ nm, the radius of the passive cavity $R=15$ $\mu$m, $g_1=\omega _1/R$, $\kappa _1/2\pi =6$ MHz, $m=6.2$ ng, $\omega _m/2\pi =78$ MHz, $\gamma _m/2\pi =12$ kHz, and $\eta _c=0.4$.

In Fig. 2, we first investigate the influence of the cavity-cavity coupling strength $J$ on the four-wave mixing intensity. Figure 2(a) plots the FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ versus $(\Omega -\omega _m)/\gamma _m$ for different values of coupling strength $J$. It is shown that the FWM intensity reaches the maximum at $\Omega =\omega _m$, where optomechanically induced transparency can occur. In the absence of the direct coupling between the two cavities, the FWM intensity is very weak, which is about $10^{-4.5}$. However, if the cavity-cavity coupling is turned on, such as $J=0.24(\kappa _1+\kappa _2)$, the FWM intensity can be improved about 4 orders of magnitude. Most importantly, if the coupling strength $J$ is tuned to approach the exceptional points with $J=0.251(\kappa _1+\kappa _2)$, the FWM intensity can be greatly enhanced, which is about 8 orders of magnitude higher than that of the single cavity case. In addition, when the cavity-control field detuning $\Delta _1^{'}=\Delta _2=0,$ the FWM intensity at $\Omega =\omega _m$ as a function of the coupling strength $J$ is plotted in Fig. 2(b). The maximum FWM intensity at the exceptional points is enhanced by over 12 orders of magnitude compared with the case that the coupling between the passive cavity and the active cavity is turned off. Therefore, the FWM intensity in $\mathcal {PT}$-symmetric optomechanical systems can be significantly enhanced around the exceptional points. This can be explained intuitively as follows. If the coupling strength $J$ is weak, the energy is strongly localized in the active cavity and the $\mathcal {PT}$-symmetry is broken. If the coupling strength is increased to surpass the exceptional points, the energy in the active cavity can flow into the passive one to compensate the dissipation, and the two supermodes have different frequencies but with the same damping rate. With balanced gain and loss, the two supermodes will be lossless in the $\mathcal {PT}$-symmetric regime. At the exceptional points, the two supermodes are degenerate with zero loss. Consequently, the FWM intensity is significantly enhanced due to field localization and negligible loss in the vicinity of the exceptional points [29,30].

 figure: Fig. 2.

Fig. 2. FWM intensitiy $\log _{10}|\mathrm {FWM_p}|^2$ as a function of (a) $(\Omega -\omega _m)/\gamma _m$ for different values of coupling strength $J$ and (b) $J/(\kappa _1+\kappa _2)$ with $\Omega =\omega _m$. Other parameters are $\Delta _1^{'}=\Delta _2=0$, $\omega _1=2\pi c/\lambda$ with $\lambda =1550$ nm, $R=15$ $\mu$m, $g_1=\omega _1/R$, $\kappa _1/2\pi =6$ MHz, $\kappa _2=0.99\kappa _1$, $m=6.2$ ng, $\omega _m/2\pi =78$ MHz, $\gamma _m/2\pi =12$ kHz, $P_c=50~$nW, $\eta _c=0.4$, $\varepsilon _p=\varepsilon _c/1000$, and $\varepsilon _m=0.$

Download Full Size | PDF

Moreover, the optical response of the system can be further modified by the additional mechanical driving field. In Fig. 3, we plot the FWM intensity $|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as functions of the mechanical driving amplitude $\varepsilon _m$ and phase difference $\phi$ with (a) $\Delta _1^{'}=\Delta _2=0$ and (b) $\Delta _1^{'}=\Delta _2=0.5\omega _m$. With the increase of the driving amplitude $\varepsilon _m$, Fig. 3(a) shows that the dependence of the FWM intensity on the phase difference is evident. The maximum FWM intensity is obtained when the phase difference $\phi =\pi /2$, but the minimum value locates around $\phi =3\pi /2$. However, if the cavity-control field detuning is tuned to be $\Delta _1^{'}=\Delta _2=0.5\omega _m$, it can be seen from Fig. 3(b) that the FWM intensity is enhanced monotonically with increasing the amplitude $\varepsilon _m$, where the influence of the phase difference $\phi$ is negligible for fixed $\varepsilon _m$. In the following, we will focus on the case that the phase difference plays an important role in controlling the FWM intensity.

 figure: Fig. 3.

Fig. 3. Contour plots of FWM intensity $|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as functions of the amplitude $\varepsilon _m$ and phase difference $\phi /\pi$ with (a) $\Delta _1^{'}=\Delta _2=0$ and (b) $\Delta _1^{'}=\Delta _2=0.5\omega _m$. The other parameters are the same as those in Fig. 2 except $J=0.251(\kappa _1+\kappa _2)$.

Download Full Size | PDF

To explain the physical origin of the above phase-dependent effect induced by the mechanical driving field, we show $|\mathrm {FWM_1}|^2$, $|\mathrm {FWM_2}|^2$, and $|\mathrm {FWM_p}|^2$ as functions of the mechanical driving amplitude $\varepsilon _m$ in Fig. 4. It is shown in Eq. (18) that $\mathrm {FWM_p}$ can be divided into $\mathrm {FWM_1}$ and $\mathrm {FWM_2}$, where $|\mathrm {FWM_1}|$ is independent on $\varepsilon _m$ but $|\mathrm {FWM_2}|$ increases monotonically with $\varepsilon _m$. When the phase difference $\phi =\pi /2$, there is constructive interference between $\mathrm {FWM_1}$ and $\mathrm {FWM_2}$. We can see from Fig. 4(a) that perfect constructive interfere occurs if $|\mathrm {FWM_1}|^2=|\mathrm {FWM_2}|^2$ at $\varepsilon _m\approx 2.7$ pN, where $|\mathrm {FWM_p}|^2=4|\mathrm {FWM_1}|^2$. However, if the phase difference is tuned to be $\phi =3\pi /2$, there is destructive interference between $\mathrm {FWM_1}$ and $\mathrm {FWM_2}$. Figure 4(c) shows that $|\mathrm {FWM_p}|^2$ can be completely suppressed at $\varepsilon _m\approx 2.7$ pN due to perfect destructive interference. With further increasing the amplitude $\varepsilon _m$, the FWM intensity $|\mathrm {FWM_p}|^2$ will keep growing because of the mechanical driving. Figures 4(b) and (d) show that the FWM intensity can be significantly enhanced for both $\phi =\pi /2$ and $\phi =3\pi /2$ if the mechanical driving amplitude is strong enough, where the interference effect is weak since $|\mathrm {FWM_1}|^2\gg |\mathrm {FWM_2}|^2$. However, if the mechanical driving field is too strong, the fluctuation $\delta a$ becomes larger and the second-order sideband effects should be considered [40,41]. For the chosen parameters, we find that $|a_{1+}|$= $|a_{1-}|\approx 0.017 a_{1s}$ at $\varepsilon _m=50$ pN. Therefore, the it is still reasonable to use the linearized method and neglect the quadratic term in $\delta a$. In addition, the mechanical driving field has no influence on the steady-state solution $a_{1s}$ and the effective optomechanical coupling strength $g_1 a_{1s}$, which will not result in the shift of the exceptional points.

 figure: Fig. 4.

Fig. 4. Plots of $|\mathrm {FWM}_1|^2$, $|\mathrm {FWM}_2|^2$, and $|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as functions of the mechanical driving amplitude $\varepsilon _m$ with $\phi =\pi /2$ in (a) and (b) and $\phi =3\pi /2$ in (c) and (d). Here we choose $\Delta _1^{'}=\Delta _2=0$ and $J=0.251(\kappa _1+\kappa _2)$. The other parameters are the same as those in Fig. 2.

Download Full Size | PDF

The gain rate of the active cavity $a_2$ can be controlled in experiments [28,29]. In Fig. 5, we plot the FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ versus the gain-loss ratio $\kappa _2/\kappa _1$ with $J=0.251(\kappa _1+\kappa _2)$. In the regime $\kappa _2/\kappa _1<0$, the cavity $a_2$ corresponds to a passive cavity, and the FWM intensity decreases with the increase of the gain-loss ratio $\kappa _2/\kappa _1$. The minimum value locates at $\kappa _2=0$. When the cavity $a_2$ becomes active, the FWM intensity starts to strengthen again, and the maximum value locates at $\kappa _2/\kappa _1\approx 1$ for $\varepsilon _m=0$. Therefore, the FWM intensity can be greatly enhanced near the exceptional points. In the presence of the additional mechanical driving field, the FWM intensity can be further strengthened and the influence of the phase difference is obvious in the vicinity of $\kappa _2/\kappa _1=1$. With balanced gain and loss, the FWM intensity is enhanced for $\phi =\pi /2$ and suppressed for $\phi =3\pi /2$ when the amplitude $\varepsilon _m$ equals to 4 pN.

 figure: Fig. 5.

Fig. 5. FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as a function of gain-loss ratio $\kappa _2/\kappa _1$ for different values of $\varepsilon _m$ and $\phi$. The other parameters are the same as those in Fig. 4.

Download Full Size | PDF

We proceed to study the dependence of the FWM intensity on the frequency of the control field. In Fig. 6, the FWM intensity $\log _{10}{|\mathrm {FWM_p}|^2}$ is plotted versus the cavity-control field detuning $\Delta _1^{'}/\omega _m$ with $\Delta _1^{'}=\Delta _2$ for different values of the mechanical driving amplitude $\varepsilon _m$ and phase difference $\phi$. In the absence of the mechanical driving field, it can be seen that the FWM intensity reaches the maximum when the control field is nearly resonant with the passive cavity. Furthermore, the FWM intensity can be greatly strengthened by the additional mechanical driving field. It should be noted that the phase-dependent effect is only evident around $\Delta _1^{'}=0$ and $\Delta _1^{'}=\omega _m$, as shown in the insets of Fig. 6. For the chosen parameters, the FWM intensity can be enhanced for $\phi =\pi /2$ and suppressed for $\phi =3\pi /2$, which is consistent with the results in Fig. 3.

 figure: Fig. 6.

Fig. 6. FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as a function of $\Delta _{1}^{'}/\omega _m$ for different values of $\varepsilon _m$ and $\phi$. Here we fix $\Delta _2=\Delta _1^{'}$, and the other parameters are the same as those in Fig. 4.

Download Full Size | PDF

Finally, Fig. 7 shows the FWM intensity $\log _{10}|\mathrm {FWM_p}^2|$ versus $(\Omega -\omega _m)/\gamma _m$ for different values of the control power $P_c$. With the increase of the control power $P_c$, the FWM intensity can be enhanced, as shown in Fig. 7. Furthermore, the inset of Fig. 7 plots the FWM intensity at $\Omega =\omega _m$ as a function of the control power $P_c$. It can be seen that the maximum FWM intensity locates at $P_c\approx 51.1$ nW, which is the critical value between the stable and unstable regimes due to the gain of the cavity $a_2$. With further increasing the control power $P_c$, the system becomes unstable and the FWM intensity decreases monotonically. Therefore, the FWM intensity can be greatly strengthened by increasing the power of the control field as long as the $\mathcal {PT}$-symmetric optomechanical system operates in the stable regime.

 figure: Fig. 7.

Fig. 7. FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ as a function of $(\Omega -\omega _m)/\gamma _m$ for different control power $P_c$. The inset shows the FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ versus the control power $P_c$. The other parameters are the same as those in Fig. 4 except $\Delta _1^{'}=\Delta _2=\omega _m$, $\varepsilon _m=4$ pN and $\phi =\pi /2$.

Download Full Size | PDF

4. Conclusion

In summary, we have explored the enhancement of four-wave mixing process based on a $\mathcal {PT}$-symmetric optomechanical system driven by a strong optical control field, a weak optical probe field, and a weak coherent mechanical driving field. We study the dependence of the FWM intensity on $\mathcal {PT}$-symmetry by tuning the cavity-cavity coupling strength and the gain-loss ratio, and it is found that the FWM intensity is greatly enhanced near the exceptional points, which can be $10^{12}$ orders of magnitude higher than that of the case without the active cavity. We also demonstrate that the FWM intensity can be further enhanced by controlling the amplitude and phase of the mechanical driving field. Moreover, it is shown that the FWM intensity increases monotonically with the power of the control field in the stable regime of this optomechanical system. The results obtained herein may be useful for gaining further insight into the properties of $\mathcal {PT}$-symmetric optomechanical system and find potential applications in optical wavelength converter and optical switch.

Funding

National Natural Science Foundation of China (11874170); China Postdoctoral Science Foundation (2017M620593); Qinglan Project of Jiangsu Province of China; Natural Science Research of Jiangsu Higher Education Institutions of China (17KJA460004, 19KJA150011); Practice Innovation Training Program Projects for Jiangsu College Students (201910323021Z).

Disclosures

The authors declare no conflicts of interest.

References

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

2. F. Marquardt and S. M. Girvin, “Optomechanics,” Physics 2, 40 (2009). [CrossRef]  

3. H. Xiong, L. G. Si, X. Y. Lü, X. X. Yang, and Y. Wu, “Review of cavity optomechanics in the weak-coupling regime: from linearization to intrinsic nonlinear interactions,” Sci. China: Phys., Mech. Astron. 58(5), 1–13 (2015). [CrossRef]  

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

5. J. D. Teufel, T. Donner, D. Li, J. W. Harlow, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W. Simmonds, “Sideband cooling of micromechanical motion to the quantum ground state,” Nature 475(7356), 359–363 (2011). [CrossRef]  

6. S. Hong, R. Riedinger, I. Marinkovic, A. Wallucks, S. G. Hofer, R. A. Norte, M. Aspelmeyer, and S. Gröblacher, “Hanbury Brown and Twiss interferometry of single phonons from an optomechanical resonator,” Science 358(6360), 203–206 (2017). [CrossRef]  

7. A. H. Safavi-Naeini, S. Gröblacher, J. H. Hill, J. Chan, M. Aspelmeyer, and O. Painter, “Squeezed light from a silicon micromechanical resonator,” Nature 500(7461), 185–189 (2013). [CrossRef]  

8. T. P. Purdy, P.-L. Yu, R. W. Peterson, N. S. Kampel, and C. A. Regal, “Strong optomechanical squeezing of light,” Phys. Rev. X 3(3), 031012 (2013). [CrossRef]  

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

10. 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]  

11. 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]  

12. R. K. Jain and R. C. Lind, “Degenerate four-wave mixing in semiconductor-doped glasses,” J. Opt. Soc. Am. 73(5), 647–653 (1983). [CrossRef]  

13. Y. Q. Li and M. Xiao, “Enhancement of nondegenerate four-wave mixing based on electromagnetically induced transparency in rubidium atoms,” Opt. Lett. 21(14), 1064–1066 (1996). [CrossRef]  

14. Y. Wu, J. Saldana, and Y. F. Zhu, “Large enhancement of four-wave mixing by suppression of photon absorption from electromagnetically induced transparency,” Phys. Rev. A 67(1), 013811 (2003). [CrossRef]  

15. W. J. Wadsworth, N. Joly, J. C. Knight, T. A. Birks, F. Biancalana, and P. S. J. Russell, “Supercontinuum and four-wave mixing with Q-switched pulses in endlessly single-mode photonic crystal fibres,” Opt. Express 12(2), 299–309 (2004). [CrossRef]  

16. S. M. Huang and G. S. Agarwal, “Normal-mode splitting and antibunching in Stokes and anti-Stokes processes in cavity optomechanics: Radiation-pressure-induced four-wave-mixing cavity optomechanics,” Phys. Rev. A 81(3), 033830 (2010). [CrossRef]  

17. C. Jiang, Y. S. Cui, and H. X. Liu, “Controllable four-wave mixing based on mechanical vibration in two-mode optomechanical systems,” Europhys. Lett. 104(3), 34004 (2013). [CrossRef]  

18. X. W. Xu and Y. Li, “Controllable optical output fields from an optomechanical system with mechanical driving,” Phys. Rev. A 92(2), 023855 (2015). [CrossRef]  

19. Z. Y. Li, X. You, Y. M. Li, Y. C. Liu, and K. C. Peng, “Multimode four-wave mixing in an unresolved sideband optomechanical system,” Phys. Rev. A 97(3), 033806 (2018). [CrossRef]  

20. X. F. Wang and B. Chen, “Four-wave mixing response in a hybrid atom-optomechanical system,” J. Opt. Soc. Am. B 36(2), 162–167 (2019). [CrossRef]  

21. C. M. Bender and S. Boettcher, “Real spectra in non-Hermitian Hamiltonians having $\mathcal{PT}$-Symmetry,” Phys. Rev. Lett. 80(24), 5243–5246 (1998). [CrossRef]  

22. N. Moiseyev, Non-Hermitian Quantum Mechanics (Cambridge University, 2011).

23. A. Guo, G. J. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. A. Siviloglou, and D. N. Christodoulides, “Observation of PT-symmetry breaking in complex optical potentials,” Phys. Rev. Lett. 103(9), 093902 (2009). [CrossRef]  

24. H. Xu, D. Mason, L. Y. Jiang, and J. G. E. Harris, “Topological energy transfer in an optomechanical system with exceptional points,” Nature 537(7618), 80–83 (2016). [CrossRef]  

25. W. Chen, Ş. K. Özdemir, G. Zhao, J. Wiersig, and L. Yang, “Exceptional points enhance sensing in an optical microcavity,” Nature 548(7666), 192–196 (2017). [CrossRef]  

26. H. Hodaei, A. U. Hassan, S. Wittek, H. G.-Gracia, R. E.-Ganainy, D. N. Christodoulides, and M. Khajavikhan, “Enhanced sensitivity at higher-order exceptional points,” Nature 548(7666), 187–191 (2017). [CrossRef]  

27. T. Wasak, P. Szańkowski, V. V. Konotop, and M. Trippenbach, “Four-wave mixing in a parity-time ($\mathcal{PT}$)-symmetric coupler,” Opt. Lett. 40(22), 5291–5294 (2015). [CrossRef]  

28. L. Chang, X. Jiang, S. Hua, C. Yang, J. Wen, L. Jiang, G. Li, G. Wang, and M. Xiao, “Parity-time symmetry and variable optical isolation in active-passive-coupled microresonators,” Nat. Photonics 8(7), 524–529 (2014). [CrossRef]  

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

30. J. Zhang, B. Peng, Ş. K. Özdemir, Y.-x. Liu, H. Jing, X.-y. Lü, Y.-l. Liu, L. Yang, and F. Nori, “Giant nonlinearity via breaking parity-time symmetry: A route to low-threshold phonon diodes,” Phys. Rev. B 92(11), 115407 (2015). [CrossRef]  

31. X. Y. Zhang, Q. T. Cao, Z. Wang, Y. X. Liu, C. W. Qiu, L. Yang, Q. H. Gong, and Y. F. Xiao, “Symmetry-breaking-induced nonlinear optics at a microcavity surface,” Nat. Photonics 13(1), 21–24 (2019). [CrossRef]  

32. H. Jing, S. K. Özdemir, X.-Y. Lü, J. Zhang, L. Yang, and F. Nori, “PT-symmetric phonon laser,” Phys. Rev. Lett. 113(5), 053604 (2014). [CrossRef]  

33. B. He, L. Yang, and M. Xiao, “Dynamical phonon laser in coupled active-passive microresonators,” Phys. Rev. A 94(3), 031802 (2016). [CrossRef]  

34. X.-Y. Lü, H. Jing, J.-Y. Ma, and Y. Wu, “PT-symmetry-breaking chaos in optomechanics,” Phys. Rev. Lett. 114(25), 253601 (2015). [CrossRef]  

35. H. Jing, Ş. 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]  

36. W. Li, Y. Jiang, C. Li, and H. Song, “Parity-time-symmetry enhanced optomechanically-induced-transparency,” Sci. Rep. 6(1), 31095 (2016). [CrossRef]  

37. X. Y. Zhang, Y. Q. Guo, P. Pei, and X. X. Yi, “Optomechanically induced absorption in parity-time-symmetric optomechanical systems,” Phys. Rev. A 95(6), 063825 (2017). [CrossRef]  

38. H. Jing, Ş. K. Özdemir, H. Lü, and F. Nori, “High-order exceptional points in optomechanics,” Sci. Rep. 7(1), 3386 (2017). [CrossRef]  

39. Y. L. Liu and Y.-x. Liu, “Energy-localization-enhanced ground-state cooling of a mechanical resonator from room temperature in optomechanics using a gain cavity,” Phys. Rev. A 96(2), 023812 (2017). [CrossRef]  

40. 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]  

41. L. Y. He, “Parity-time-symmetry-enhanced sideband generation in an optomechanical system,” Phys. Rev. A 99(3), 033843 (2019). [CrossRef]  

42. Z. P. Liu, J. Zhang, S. K. Özdemir, B. Peng, H. Jing, X.-Y. Lü, C.-W. Li, L. Yang, F. Nori, and Y.-x. Liu, “Metrology with PT-symmetric cavities: enhanced sensitivity near the PT-phase transition,” Phys. Rev. Lett. 117(11), 110802 (2016). [CrossRef]  

43. P. Renault, H. Yamaguchi, and I. Mahboob, “Virtual exceptional points in an electromechanical system,” Phys. Rev. Appl. 11(2), 024007 (2019). [CrossRef]  

44. C. Bekker, R. Kalra, C. Baker, and W. P. Bowen, “Injection locking of an electro-optomechanical device,” Optica 4(10), 1196–1204 (2017). [CrossRef]  

45. J. Bochmann, A. Vainsencher, D. D. Awschalom, and A. N. Cleland, “Nanomechanical coupling between microwave and optical photons,” Nat. Phys. 9(11), 712–716 (2013). [CrossRef]  

46. L. Fan, K. Y. Fong, M. Poot, and H. X. Tang, “Cascaded optical transparency in multimode-cavity optomechanical systems,” Nat. Commun. 6(1), 5850 (2015). [CrossRef]  

47. D. B. Sohn, S. Kim, and G. Bahl, “Time-reversal symmetry breaking with acoustic pumping of nanophotonic circuits,” Nat. Photonics 12(2), 91–97 (2018). [CrossRef]  

48. D. Bothner, S. Yanai, A. Iniguez-Rabago, M. Yuan, Y. M. Blanter, and G. A. Steele, “Cavity electromechanics with parametric mechanical driving,” arXiv:1908.08496v1.

49. W. Z. Jia, L. F. Wei, Y. Li, and Y.-x. Liu, “Phase-dependent optical response properties in an optomechanical system by coherently driving the mechanical resonator,” Phys. Rev. A 91(4), 043843 (2015). [CrossRef]  

50. J. Y. Ma, C. You, L. G. Si, H. Xiong, J. H. Li, X. X. Yang, and Y. Wu, “Optomechanically induced transparency in the presence of an external time-harmonic-driving force,” Sci. Rep. 5(1), 11278 (2015). [CrossRef]  

51. C. Jiang, Y. S. Cui, X. T. Bian, F. Zuo, H. L. Yu, and G. B. Chen, “Phase-dependent multiple optomechanically induced absorption in multimode optomechanical systems with mechanical driving,” Phys. Rev. A 94(2), 023837 (2016). [CrossRef]  

52. Y. Li, Y. Y. Huang, X. Z. Zhang, and L. Tian, “Optical directional amplification in a three-mode optomechanical system,” Opt. Express 25(16), 18907–18916 (2017). [CrossRef]  

53. 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]  

54. 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]  

55. S. Maayani, R. Dahan, Y. Kligerman, E. Moses, A. U. Hassan, H. Jing, F. Nori, D. N. Christodoulides, and T. Carmon, “Flying couplers above spinning resonators generate irreversible refraction,” Nature 558(7711), 569–572 (2018). [CrossRef]  

56. C. Zhai, R. Huang, H. Jing, and L.-M. Kuang, “Mechanical switch of photon blockade and photon-induced tunneling,” Opt. Express 27(20), 27649–27662 (2019). [CrossRef]  

57. X. W. Xu, Y. J. Zhao, H. Wang, H. Jing, and A. X. Chen, “Quantum nonreciprocality in quadratic optomechanics,” Photonics Res. 8(2), 143–150 (2020). [CrossRef]  

58. G. Anetsberger, O. Arcizet, Q. P. Unterreithmeier, R. Rivière, A. Schliesser, E. M. Weig, J. P. Kotthaus, and T. J. Kippenberg, “Near-field cavity optomechanics with nanomechanical oscillators,” Nat. Phys. 5(12), 909–914 (2009). [CrossRef]  

59. R. Schilling, H. Schütz, A. H. Ghadimi, V. Sudhir, D. J. Wilson, and T. J. Kippenberg, “Near-field integration of a SiN nanobeam and a SiO2 microcavity for Heisenberg-limited displacement sensing,” Phys. Rev. Appl. 5(5), 054019 (2016). [CrossRef]  

60. A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, and R. J. Schoelkopf, “Introduction to quantum noise, measurement, and amplification,” Rev. Mod. Phys. 82(2), 1155–1208 (2010). [CrossRef]  

61. E. Verhagen, S. Deléglise, S. Weis, A. Schliesser, and T. J. Kippenberg, “Quantum-coherent coupling of a mechanical oscillator to an optical cavity mode,” Nature 482(7383), 63–67 (2012). [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 (7)

Fig. 1.
Fig. 1. Schematic illustration of the $\mathcal {PT}$-symmetric optomechanical system. The passive cavity with loss rate $\kappa _1$ interacts with a mechanical mode with resonance frequency $\omega _m$ and damping rate $\gamma _m$ via radiation pressure. The active cavity with tunable gain rate $\kappa _2$ is coupled to the passive cavity, and the coupling strength $J$ can be adjusted by the distance between them. In addition, the passive cavity is driven by a strong control field at frequency $\omega _c$ and a weak probe field at frequency $\omega _p$, and the mechanical resonator is excited by a weak coherent mechanical driving field at frequency $\Omega =\omega _p-\omega _c$.
Fig. 2.
Fig. 2. FWM intensitiy $\log _{10}|\mathrm {FWM_p}|^2$ as a function of (a) $(\Omega -\omega _m)/\gamma _m$ for different values of coupling strength $J$ and (b) $J/(\kappa _1+\kappa _2)$ with $\Omega =\omega _m$. Other parameters are $\Delta _1^{'}=\Delta _2=0$, $\omega _1=2\pi c/\lambda$ with $\lambda =1550$ nm, $R=15$ $\mu$m, $g_1=\omega _1/R$, $\kappa _1/2\pi =6$ MHz, $\kappa _2=0.99\kappa _1$, $m=6.2$ ng, $\omega _m/2\pi =78$ MHz, $\gamma _m/2\pi =12$ kHz, $P_c=50~$nW, $\eta _c=0.4$, $\varepsilon _p=\varepsilon _c/1000$, and $\varepsilon _m=0.$
Fig. 3.
Fig. 3. Contour plots of FWM intensity $|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as functions of the amplitude $\varepsilon _m$ and phase difference $\phi /\pi$ with (a) $\Delta _1^{'}=\Delta _2=0$ and (b) $\Delta _1^{'}=\Delta _2=0.5\omega _m$. The other parameters are the same as those in Fig. 2 except $J=0.251(\kappa _1+\kappa _2)$.
Fig. 4.
Fig. 4. Plots of $|\mathrm {FWM}_1|^2$, $|\mathrm {FWM}_2|^2$, and $|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as functions of the mechanical driving amplitude $\varepsilon _m$ with $\phi =\pi /2$ in (a) and (b) and $\phi =3\pi /2$ in (c) and (d). Here we choose $\Delta _1^{'}=\Delta _2=0$ and $J=0.251(\kappa _1+\kappa _2)$. The other parameters are the same as those in Fig. 2.
Fig. 5.
Fig. 5. FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as a function of gain-loss ratio $\kappa _2/\kappa _1$ for different values of $\varepsilon _m$ and $\phi$. The other parameters are the same as those in Fig. 4.
Fig. 6.
Fig. 6. FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ as a function of $\Delta _{1}^{'}/\omega _m$ for different values of $\varepsilon _m$ and $\phi$. Here we fix $\Delta _2=\Delta _1^{'}$, and the other parameters are the same as those in Fig. 4.
Fig. 7.
Fig. 7. FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ as a function of $(\Omega -\omega _m)/\gamma _m$ for different control power $P_c$. The inset shows the FWM intensity $\log _{10}|\mathrm {FWM_p}|^2$ at $\Omega =\omega _m$ versus the control power $P_c$. The other parameters are the same as those in Fig. 4 except $\Delta _1^{'}=\Delta _2=\omega _m$, $\varepsilon _m=4$ pN and $\phi =\pi /2$.

Equations (20)

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

H = H 0 + H I + H d r , H 0 = Δ 1 a 1 a 1 + Δ 2 a 2 a 2 + p 2 2 m + 1 2 m ω m 2 x 2 , H I = J ( a 1 a 2 + a 1 a 2 ) g 1 a 1 a 1 x , H d r = i η c κ 1 [ ( ε c + ε p e i Ω t i ϕ p c ) a 1 H . c . ] x ε m c o s ( Ω t + ϕ m ) .
a 1 ˙ = ( κ 1 2 + i Δ 1 g 1 x ) a 1 + i J a 2 + η c κ 1 ( ε c + ε p e i Ω t i ϕ p ) ,
a 2 ˙ = ( κ 2 2 i Δ 2 ) a 2 + i J a 1 ,
x ˙ = p m ,
p ˙ = m ω m 2 x γ m p + g 1 a 1 a 1 + ε m cos ( Ω t + ϕ m ) ,
a 1 s = η c κ 1 ε c ( κ 2 / 2 i Δ 2 ) ( κ 1 / 2 + i Δ 1 ) ( κ 2 / 2 i Δ 2 ) J 2 , a 2 s = i J η c κ 1 ε c ( κ 1 / 2 + i Δ 1 ) ( κ 2 / 2 i Δ 2 ) J 2 , x s = g 1 | a 1 s | 2 m ω m 2 , p s = 0 ,
δ a 1 ˙ = ( κ 1 / 2 + i Δ 1 ) δ a 1 + i g 1 a 1 s δ x + i J δ a 2 + η c κ 1 ε p e i Ω t i ϕ p c ,
δ a 2 ˙ = ( κ 2 / 2 i Δ 2 ) δ a 2 + i J δ a 1 ,
δ x ˙ = δ p m ,
δ p ˙ = m ω m 2 δ x γ m δ p + g 1 ( a 1 s δ a 1 + a 1 s δ a 1 ) + ε m c o s ( Ω t + ϕ m ) .
a 1 + = χ 1 ( Ω ) 1 + i f ( Ω ) i f ( Ω ) { [ 1 + i f ( Ω ) ] η c κ 1 ε p e i ϕ p + i g 1 a 1 s χ m ( Ω ) ε m 2 e i ϕ m } ,
a 1 = χ 1 ( Ω ) 1 i f ( Ω ) + i f ( Ω ) [ i f ( Ω ) a 1 s a 1 s η c κ 1 ε p e i ϕ p + i g 1 a 1 s χ m ( Ω ) ε m 2 e i ϕ m ] .
χ m ( Ω ) = 1 m ( ω m 2 Ω 2 i γ m Ω ) , χ 2 ( Ω ) = 1 κ 2 / 2 i Δ 2 + i Ω , χ 1 ( Ω ) = 1 κ 1 / 2 + i ( Δ 1 Ω ) J 2 χ 2 ( Ω ) , f ( Ω ) = g 1 2 | a 1 s | 2 χ m ( Ω ) κ 1 / 2 + i ( Δ 1 Ω ) J 2 χ 2 ( Ω ) .
a 1 , o u t ( t ) = ( ε c η c κ 1 a 1 s ) e i ω c t C o n t r o l   f i e l d + ( ε p e i ϕ p c η c κ 1 a 1 + ) e i ω p t P r o b e   f i e l d η c κ 1 a 1 e i ( 2 ω c ω p ) t F W M   f i e l d .
t p = ε p e i ϕ p c η c κ 1 a 1 + ε p e i ϕ p c = t 1 + t 2 ,
t 1 = 1 1 + i f ( Ω ) 1 + i f ( Ω ) i f ( Ω ) χ 1 ( Ω ) η c κ 1 ,
t 2 = i g 1 a 1 s χ 1 ( Ω ) χ m ( Ω ) ε m / ( 2 ε p ) 1 + i f ( Ω ) i f ( Ω ) η c κ 1 e i ϕ .
F W M p = η c κ 1 a 1 ε p e i ϕ p = F W M 1 + F W M 2 ,
F W M 1 = i f ( Ω ) χ 1 ( Ω ) a 1 s / a 1 s 1 i f ( Ω ) + i f ( Ω ) η c κ 1 ,
F W M 2 = i g 1 a 1 s χ 1 ( Ω ) χ m ε m / ( 2 ε p ) 1 i f ( Ω ) + i f ( Ω ) η c κ 1 e i ϕ .
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.